Uploaded by Sasa Mongkolsermporn

616459204-ASH-2022-Education-Book-1

advertisement
Dr Prakash Singh Shekhawat
Hematology 2022
AMERICAN SOCIETY OF HEMATOLOGY EDUCATION PROGRAM
64th ASH® Annual Meeting and Exposition
Editors
Hanny Al-Samkari, MD
Executive Editor, Hematology 2022
Massachusetts General Hospital
Boston, Massachusetts
Alison R. Walker, MD, MPH, MBA
Deputy Editor, Hematology 2022
H. Lee Moffitt Cancer Center
Tampa, Florida
Rakhi P. Naik, MD, MHS
Deputy Editor, Hematology 2022
Johns Hopkins Medicine
Baltimore, Maryland
Alex F. Herrera, MD
Deputy Editor, Hematology 2022
City of Hope
Duarte, California
Ang Li, MD, MS
Special Section Editor, Hematology 2022
Baylor College of Medicine
Houston, Texas
AMERICAN SOCIETY OF HEMATOLOGY • WASHINGTON, DC
Dr Prakash Singh Shekhawat
© 2022 by The American Society of Hematology
ISSN 1520–4383 (online)
Hematology, the ASH Education Program, is published
annually by The American Society of Hematology (ASH) in
one volume per year.
Committee on Educational Affairs
All business correspondence and purchase and reprint
requests should be addressed to The American Society of
Hematology, 2021 L Street, NW, Suite 900, Washington, DC
20036, USA; phone: 202–776–0544; fax: 202–776–0545;
e-mail: ASH@hematology.org.
Marc J. Braunstein, MD, PhD (‘25)
Nicholas R. Burwick, MD (‘25)
Shruti Chaturvedi, MBBS, MS (‘25)
Faith Davies, MD (‘24)
Amy E. DeZern, MD (‘26)
James M. Foran, MD (‘22)
Nobuko Hijiya, MD (‘24)
Margaret Kasner, MD (‘26)
Kah Poh (Melissa) Loh, MD (‘23)
Martha P. Mims, MD, PhD (‘23)
Casey L. O’Connell, MD (‘23)
Heather A. O’Leary, PhD (‘23)
Nicki Panoskaltsis, MD, PhD (‘25)
Hematology, the ASH Education Program, is available on the
Internet at https://ashpublications.org/hematology.
© 2022 by The American Society of Hematology. All rights
reserved. Copyright is not claimed in any works of the United
States Government. Except as expressly permitted in this
statement, no part of this publication may be used (as here
in after defined) in any form or by any means now or hereafter
known, electronic or mechanical, without permission in
writing from the Publisher, The American Society of
Hematology. For purposes of this notice, the term “use”
includes but is not limited to reproduction, photocopying,
storage in a retrieval system, translation, and transmittal. The
Publisher hereby consents to the use of this publication or
any part hereof for any noncommercial educational purpose
within the health field such as classroom instruction and
clinical and residency training. This publication or any part
thereof may be used for educational purposes at conferences, continuing education courses, and other educational
activity, provided no fee or other compensation is charged
therefor. All materials so used must acknowledge the
Publisher’s copyright therein as “© 2022 by The American
Society of Hematology.” When requesting the Publisher’s
permission to use this publication or any part thereof, please
contact the Copyright Clearance Center at 222 Rosewood
Drive, Danvers, MA 01923, USA; phone: 978–750–8400; fax:
978–750–4470; or The American Society of Hematology
Publications Department at 2021 L Street, NW, Suite 900,
Washington, DC 20036, USA; phone: 202–776–0544;
fax: 202–776–0545.
Article Citations
Cite articles in this volume by listing Author(s), Title,
Hematology Am Soc Hematol Educ Program. 2022;2022:
beginning page number-ending page number.
Hematology Rights and Permission
To request permission to reprint or reuse Hematology
articles, figures, or tables, please visit the Hematology page
at the Copyright Clearance Center (https://marketplace.
copyright.com/rs-ui-web/mp/details/journal/123161909).
Jennifer R. Brown, MD, PhD (‘23) (Chair)
Anjali Advani, MD (‘23) (Vice Chair)
Ex Officio Members
Jane N. Winter, MD (‘22) – President
Robert A. Brodsky, MD (‘23) – President-Elect
Mohandas Narla, DSc (‘24) – Vice President
Liaisons
Hanny Al-Samkari, MD (‘24) – Executive Editor, Hematology
ASH Education Program
Jessica Altman, MD (‘21) – Senior Executive Editor, ASH-SAP 8
Laura M. De Castro, MD (‘23) – Member, Committee on
Practice
Cynthia E. Dunbar, MD (‘24) – Secretary
David Garcia, MD (‘22) – 2022 Education Program Co-Chair
Molly Weidner Mandernach, MD, MPH (‘22) – Junior Executive
Editor, ASH-SAP 8
Olatoyosi Odenike, MD (‘22) – 2022 Education Program
Co-Chair
Cover: Illustration of a blood clot with red blood cells (red),
white blood cells (blue), and activated platelets (purple).
Source: Thom Leach/Science Photo Library.
This activity is supported by educational grants from
AstraZeneca; GlaxoSmithKline; Incyte Corp.; Jazz
Pharmaceuticals, Inc.; Merck Sharp & Dohme Corp.; and
Novartis Pharmaceuticals Corp.
The Publisher disclaims responsibility for opinions expressed
by the authors.
ii
Dr Prakash Singh Shekhawat
Contents
ix
x
xi
Editors’ Message
Reviewers
Continuing Medical Education Information
Ham-Wasserman Lecture
1
Leukemogenesis in infants and young children with trisomy 21
IRENE ROBERTS
Acute Myeloid Leukemia: Improving Outcomes in Challenging Subsets
9
Achieving MRD negativity in AML: how important is this and how do we get there?
CHRISTOPHER S. HOURIGAN
15
Novel investigational approaches for high-risk genetic subsets of AML: TP53, KMT2A, FLT3
KIERAN D. SAHASRABUDHE AND ALICE S. MIMS
23
Optimizing outcomes in secondary AML
ANDREW MATTHEWS AND KEITH W. PRATZ
30
EVIDENCE-BASED MINIREVIEW
Clinical utilization of panel-based molecular testing for patients with AML
BRANDON J. AUBREY AND ANDREW BRUNNER
Anxiety Provoking Consultations: Mast Cells and Eosinophils
34
Available and emerging therapies for bona fide advanced systemic mastocytosis
and primary eosinophilic neoplasms
JASON GOTLIB
47
Approach to the patient with suspected hypereosinophilic syndrome
AMY D. KLION
55
How to evaluate the patient with a suspected mast cell disorder and how/when
to manage symptoms
CEM AKIN
Are Alternative Donors Now Mainstream in Allogeneic Transplant?
64
In 2022, which is preferred: haploidentical or cord transplant?
ARNON NAGLER AND MOHAMAD MOHTY
74
New strategies for mismatched unrelated donor (MMUD) hematopoietic cell transplant (HCT)
SHUKAIB ARSLAN AND MONZR M. AL MALKI
83
Hematology 2022—what is complete HLA match in 2022?
STEPHEN R. SPELLMAN
Dr Prakash Singh Shekhawat
Contents | iii
Autoimmune Hemolytic Anemias
90
Cold AIHA and the best treatment strategies
JENNY MCDADE DESPOTOVIC AND TAYLOR OLMSTED KIM
96
Evaluating patients with autoimmune hemolytic anemia in the transfusion service
and immunohematology reference laboratory: pretransfusion testing challenges
and best transfusion-management strategies
SUSAN T. JOHNSON AND KATHLEEN E. PUCA
105
Warm autoimmune hemolytic anemia and the best treatment strategies
DAVID J. KUTER
Beyond Routine Frontline Therapy of CML
114
Transplantation in CML in the TKI era: who, when, and how?
CHRISTIAN NIEDERWIESER AND NICOLAUS KRÖGER
123
Treatment of CML in pregnancy
HARRY F. ROBERTSON AND JANE F. APPERLEY
129
Management of TKI-resistant chronic phase CML
TIMOTHY P. HUGHES AND NARANIE SHANMUGANATHAN
Controversies in Aggressive NHL
138
CNS prophylaxis in aggressive B-cell lymphoma
MATTHEW R. WILSON, SABELA BOBILLO, AND KATE CWYNARSKI
146
Sequencing therapy in relapsed DLBCL
CHRISTOPHER R. FLOWERS AND OREOFE O. ODEJIDE
155
What is the role of up-front autologous stem cell transplantation in mantle cell lymphoma?
ANITA KUMAR
Immunotherapy in Multiple Myeloma
163
Antibodies and bispecifics for multiple myeloma: effective effector therapy
CHRISTOPHER CIPKAR, CHRISTINE CHEN, AND SUZANNE TRUDEL
173
Beyond the cell: novel noncellular immunotherapy approaches to multiple myeloma
SARAH A. HOLSTEIN
180
Cellular therapy for multiple myeloma: what’s now and what’s next
PAULA RODRIGUEZ-OTERO AND JESÚS F. SAN-MIGUEL
Improving Outcomes in ALL
190
New developments in ALL in AYA
NICOLAS BOISSEL
197
Optimal approach to T-cell ALL
KRISTEN M. O’DWYER
206
Ph+ ALL in 2022: is there an optimal approach?
MATTHEW J. WIEDUWILT
213
EVIDENCE-BASED MINIREVIEW
What is the opti­mal tyro­sine kinase inhib­i­tor for adults with newly diag­nosed
Philadelphia chro­mo­some–pos­i­tive acute lym­pho­blas­tic leu­ke­mia?
FADI G. HADDAD AND NICHOLAS J. SHORT
JAK/STAT Inhibition and Beyond in Ph-Negative MPNs
218
Hitting the brakes on accel­er­ated and blast-phase mye­lo­pro­lif­er­a­tive neo­plasms:
cur­rent and emerg­ing con­cepts
JAN PHILIPP BEWERSDORF AND RAAJIT K. RAMPAL
iv | Hematology 2022 | ASH Education Program
Dr Prakash Singh Shekhawat
225
Molecular prog­nos­ti­ca­tion in Ph-neg­a­tive MPNs in 2022
ALESSANDRO MARIA VANNUCCHI AND PAOLA GUGLIELMELLI
235
New approaches to tackle cytopenic mye­lo­fi­bro­sis
SAM­UEL B. REYNOLDS AND KRISTEN PETTIT
Long-Term Effects Monitoring for Survivors of Pediatric Hematologic Malignancies
245
Germline risk fac­tors for sec­ond malig­nant neo­plasms after treat­ment for pedi­at­ric
hema­to­logic malig­nan­cies
SMITA BHATIA
251
Mitigating, mon­i­tor­ing, and man­ag­ing long-term che­mo­ther­apy- and radi­a­tion-induced
car­diac tox­ic­ity
WENDY BOTTINOR AND ERIC J. CHOW
259
Risk fac­tors and screen­ing for neurocognitive impacts of ther­apy
KEVIN R. KRULL
Long-Term Health Effects of Curative Therapy for Sickle Cell Disease
266
Knowledge to date on sec­ond­ary malig­nancy fol­low­ing hema­to­poi­etic cell trans­plan­ta­tion
for sickle cell dis­ease
COURTNEY D. FITZHUGH
272
Long-term health out­comes fol­low­ing cura­tive ther­a­pies for sickle cell dis­ease
ROHINI CHAKRAVARTHY AND DEBRA L. FRIEDMAN
277
Organ func­tion indi­ca­tions and poten­tial improve­ments fol­low­ing cura­tive ther­apy
for sickle cell dis­ease
MONICA L. HULBERT, ALLISON A. KING, AND SHALINI SHENOY
283
EVIDENCE-BASED MINIREVIEW
How to uti­lize new ther­a­pies for sickle cell dis­ease
STEPHANIE GUARINO AND SOPHIE LANZKRON
Managing Thrombocytopenia in Challenging Situations
286
Thrombopoietin recep­tor ago­nists for che­mo­ther­apy-induced throm­bo­cy­to­pe­nia: a new solu­tion
for an old prob­lem
HANNY AL-SAMKARI
296
Thrombocytopenia and liver dis­ease: path­o­phys­i­­ol­ogy and periprocedural man­age­ment
HANA I. LIM AND ADAM CUKER
303
Thrombocytopenia in preg­nancy
ALLYSON M. PISHKO AND ARIELA L. MARSHALL
312
EVIDENCE-BASED MINIREVIEW
Full dose, mod­i­fied dose, or no anticoagulation for patients with can­cer and acute VTE
and throm­bo­cy­to­pe­nia
RUSHAD PATELL AND JEFFREY I. ZWICKER
Maximizing Outcomes in CLL
316
Novel ther­a­pies and com­bi­na­tions in CLL refrac­tory to BTK inhib­i­tors and venetoclax
LYDIA SCARFÒ
323
Selecting ini­tial ther­apy in CLL
INHYE E. AHN AND JENNIFER R. BROWN
329
Treatment of Richter’s syn­drome
PHILIP A. THOMPSON AND TANYA SIDDIQI
Multiple Myeloma: Assessing the Patient and the Disease
337
Fitness and frailty in mye­loma
CHARLOTTE PAWLYN, ABDULLAH M. KHAN, AND CIARA L. FREEMAN
Dr Prakash Singh Shekhawat
Contents | v
349
High or low? Assessing disease risk in multiple myeloma
TIM­O­THY MARTIN SCHMIDT
356
The bur­den of mye­loma: novel approaches to dis­ease assess­ment
MAT­THEW HO AND TAXIARCHIS KOURELIS
363
Mitigating the risk of venous throm­bo­em­bo­lism in patients with mul­ti­ple mye­loma receiv­ing
immu­no­mod­u­la­tory-based ther­apy
FAHRETTIN COVUT AND KRISTEN M. SANFILIPPO
Novel Approaches in MDS
368
New inves­ti­ga­tional com­bi­na­tions for higher-risk MDS
KRISTIN L. KOENIG AND UMA BORATE
375
Risk strat­i­fy­ing MDS in the time of pre­ci­sion med­i­cine
MARIO CAZZOLA
382
Targeting inflam­ma­tion in lower-risk MDS*
JESUS D. GONZALEZ-LUGO AND AMIT VERMA
Obstetric Management and Complications in Sickle Cell Disease in High- and Low-Income Countries
388
Acute pain epi­sodes, acute chest syn­drome, and pul­mo­nary throm­bo­em­bo­lism in preg­nancy
EUGENIA VICKY ASARE, MICHAEL R. DEBAUN, EDEGHONGHON OLAYEMI, THEODORE BOAFOR, AND SAM­UEL A. OPPONG
408
Evidence-based management of pregnant women with sickle cell disease in high-income countries
EUGENE OTENG-NTIM AND PANICOS SHANGARIS
414
Evidence-based obstet­ric man­age­ment of women with sickle cell dis­ease in low-income countries
BOSEDE B. AFOLABI, OCHUWA A. BABAH, AND TITILOPE A. ADEYEMO
Prophylactic Platelet Transfusions
421
Novel plate­let prod­ucts includ­ing cold-stored plate­lets
DANA V. DEVINE
424
Expanding the plate­let inven­tory to mit­i­gate the impact of severe short­ages
JAMES R. STUBBS, BETH H. SHAZ, RALPH R. VASSALLO, AND JOHN D. ROBACK
430
Platelet com­po­nents and bac­te­rial con­tam­i­na­tion: hos­pi­tal per­spec­tive 2022
ZBIGNIEW M. SZCZEPIORKOWSKI AND MONICA B. PAGANO
437
EVIDENCE-BASED MINIREVIEW
Strat­e­gies to man­age a severely HLA-alloimmunized patient with refrac­tory throm­bo­cy­to­pe­nia
DEBBIE JIANG AND SANDHYA R. PANCH
Reproductive and Sexual Health in Sickle Cell Disease
442
Incorporating gonadal health coun­sel­ing into pedi­at­ric care of sickle cell patients
LILLIAN R. MEACHAM, LYDIA H. PECKER, BEATRICE GEE, AND ADRIENNE MISHKIN
450
Epidemiology and treat­ment of priapism in sickle cell dis­ease
IBRAHIM M. IDRIS, ARTHUR L. BURNETT, AND MICHAEL R. DEBAUN
459
No crys­tal stair: supporting fer­til­ity care and the pur­suit of preg­nancy in women with sickle
cell dis­ease
LYDIA H. PECKER, ALECIA NERO, AND MINDY CHRISTIANSON
Thrombosis and Anticoagulation: Clinical Considerations in Selected Populations
467
Anticoagulant therapy for women: implications for menstruation, pregnancy, and lactation
EMMA DELOUGHERY AND BETHANY SAMUELSON BANNOW
474
Thrombosis and anticoagulation: clin­i­cal issues of spe­cial impor­tance to hema­tol­o­gists
who prac­tice in Asia
KOCHAWAN BOONYAWAT AND PANTEP ANGCHAISUKSIRI
vi | Hematology 2022 | ASH Education Program
Dr Prakash Singh Shekhawat
481
Thrombosis ques­tions from the inpa­tient wards
GEORGE GOSHUA, PAVAN K. BENDAPUDI, AND ALFRED IAN LEE
491
EVIDENCE-BASED MINIREVIEW
Should caplacizumab be used rou­tinely in unse­lected patients with immune throm­botic
throm­bo­cy­to­pe­nic pur­pura?
GEORGE GOSHUA AND PAVAN K. BENDAPUDI
Thrombosis Prevention and Treatment
495
Coming soon to a phar­macy near you? FXI and FXII inhib­i­tors to pre­vent or treat throm­bo­em­bo­lism
OMRI COHEN AND WALTER AGENO
506
To prophylax or not, and how much and how long? Controversies in VTE pre­ven­tion for med­i­cal
inpa­tients, includ­ing COVID-19 inpa­tients
ALEX C. SPYROPOULOS
515
The 5 most fre­quently asked ques­tions about fac­tor Xa inhib­i­tors
TZU-FEI WANG AND MARC CARRIER
To Transplant or Not to Transplant in Active or High-Risk Myeloid Disease
522
Transplant for TP53-mutated MDS and AML: because we can or because we should?
JURJEN VERSLUIS AND R. COLEMAN LINDSLEY
528
Transplant in AML with mea­sur­able resid­ual dis­ease: pro­ceed or defer?
CHARLES CRADDOCK
534
Allogeneic trans­plan­ta­tion for advanced acute leu­ke­mia
DAN­IEL WEISDORF
Treatment Approaches for the Multiple Myeloma Patient in 2022
539
Newly diag­nosed mul­ti­ple mye­loma: mak­ing sense of the menu
CAITLIN L. COSTELLO
551
The con­sul­tant’s guide to smol­der­ing mul­ti­ple mye­loma
SIGRUN THORSTEINSDOTTIR AND SIGURDUR YNGVI KRISTINSSON
560
The first relapse in mul­ti­ple mye­loma: how to pick the next best thing
SRINIVAS DEVARAKONDA, NIDHI SHARMA, AND YVONNE EFEBERA
Update in Hemophilia
569
Gene ther­apy for hemo­philia
AMIT C. NATHWANI
579
Long-term pro­phy­laxis: what are our options and how to define suc­cess?
MARILYN JEAN MANCO-JOHNSON AND BETH BOULDEN WARREN
586
Perioperative hemo­sta­sis for patients with hemo­philia
JACQUELINE N POSTON AND REBECCA KRUSE-JARRES
Updates in Targeted Therapy in Pediatric Leukemia
594
Clinical screen­ing for Ph-like ALL and the devel­op­ing role of TKIs
THAI HOA TRAN AND SARAH K. TASIAN
603
The evo­lu­tion of targeted ther­apy in pedi­at­ric AML: gemtuzumab ozogamicin, FLT3/IDH/BCL2
inhib­i­tors, and other ther­a­pies
LAUREN POMMERT AND KATHERINE TARLOCK
611
Updates in infant acute lym­pho­blas­tic leu­ke­mia and the poten­tial for targeted ther­apy
RISHI S. KOTECHA
Dr Prakash Singh Shekhawat
Contents | vii
Von Willebrand Disease
618
Diagnostic pit­falls and conun­drums in type 1 von Willebrand dis­ease
ROBERT F. SIDONIO AND MICHELLE LAVIN
624
Special con­sid­er­ations in GI bleed­ing in VWD patients
NICHOLAS L.J. CHORNENKI, EDWIN OCRAN, AND PAULA D. JAMES
631
What have we learned about the patient’s expe­ri­ence of von Willebrand dis­ease?
A focus on women
HEATHER VANDERMEULEN, SUMEDHA ARYA, SARAH NERSESIAN, NATALIE PHILBERT, AND MICHELLE SHOLZBERG
What is New in Classical Bone Marrow Failure Syndromes? (Focus on Management)
637
Dyskeratosis congenita and telo­mere biol­ogy dis­or­ders
SHARON A. SAVAGE
649
Modern man­age­ment of Fanconi ane­mia
CARLO DUFOUR AND FILOMENA PIERRI
658
Genetics of severe con­gen­i­tal neutropenia as a gate­way to per­son­al­ized ther­apy
JEAN DONADIEU AND CHRISTINE BELLANNÉ-CHANTELOT
What’s New in Indolent Lymphoma
666
Divergent paths: man­age­ment of early relapsed fol­lic­u­lar lym­phoma
RADHIKA TAKIAR, YASMIN KARIMI, AND TYCEL J. PHIL­LIPS
676
Management of mar­ginal zone lym­pho­mas
MICHELE MERLI AND LUCA ARCAINI
688
Biology of fol­lic­u­lar lym­phoma: insights and win­dows of clin­i­cal oppor­tu­nity
MEGAN PERRETT, CARINA EDMONDSON, AND JESSICA OKOSUN
695
EVIDENCE-BASED MINIREVIEW
When should autol­o­gous trans­plant or cel­lu­lar ther­apy be con­sid­ered for fol­lic­u­lar lym­phoma?
DAVID A. BOND AND AJAY K. GOPAL
Where Are We Headed in Hodgkin Lymphoma?
699
Do all­patients with pri­mary refrac­tory/first relapse of HL need autol­o­gous stem cell trans­plant?
ALISON J. MOSKOWITZ
706
Incorporating novel agents into front­line treat­ment of Hodgkin lym­phoma
SWETHA KAMBHAMPATI AND ALEX F. HERRERA
717
Individualized patient care in nod­u­lar lym­pho­cyte-pre­dom­i­nant Hodgkin lym­phoma
SVEN BORCHMANN
Errata
723
Perioperative consultative hematology: can you clear my patient for a procedure? Hematology Am
Soc Hematol Educ Program. 2021;2021:521-528.
BURNETT AE, RAGHEB B, KAATZ S.
724
Noninfectious complications of hematopoietic cell transplantation. Hematology Am Soc Hematol
Educ Program. 2021;2021:578-586.
WILLIAMS KM.
725
COVID-19 and thrombosis: searching for evidence. Hematology Am Soc Hematol Educ Program.
2021;2021:621-627.
THILAGAR B, BEIDOUN M, RHOADES R, KAATZ S.
*Note
The article “Targeting inflammation in lower-risk MDS” beginning on page 382 is not eligible for
continuing medical education (CME) credit.
viii | Hematology 2022 | ASH Education Program
Dr Prakash Singh Shekhawat
Editors’ Message
Welcome to the 64th Annual Meeting of the American Society of Hematology. This year marks
the second year of the Annual Meeting as a hybrid meeting, with many attendees thankful
to reconnect in person with colleagues from all over the world and others happy to attend
virtually. Whichever option you have chosen, you will hear the many striking scientific and
clinical advances in hematology over the past year and find a remarkable wealth of educational content covering all major areas within both classical and malignant hematology. To this
end, we have developed the Hematology ASH Education Program to be your primary and
indispensable resource. Your authors, with help from our invaluable peer reviewers and guidance from us, have worked to develop concise, readable, and informative articles filled with
useful tables and figures and headlined by a visual abstract. All of the articles either are based
directly on a talk given by the author at the meeting’s education program (using a case-based
format) or take the form of an evidence-based minireview exploring the current data for some
of the most challenging questions in modern hematology practice.
This year marks the first year of a new team of Hematology editors, and we are thrilled
to lead this venerable, over-5-decade-old ASH publication for the next 3 years. Of course,
Hematology is possible only thanks to the state-of-the-art education program developed
this year by Education Program Co-Chairs Dr. David Garcia and Dr. Olatoyosi Odenike, as well
as the tireless efforts of ASH staff, including Sheehan Misko, Erin Roberts, Jeremiah Murphy,
Brian Cannon, Glenn Landis, and Kenneth April. We are so thankful for all their efforts.
Happy reading, from your Hematology editors!
Hanny Al-Samkari, MD
Alison R. Walker,
MD, MPH, MBA
Rakhi P. Naik, MD, MHS
Dr Prakash Singh Shekhawat
Alex F. Herrera, MD
Ang Li, MD, MS
ix
Reviewers
The editors thank the session chairs and the reviewers who worked hard to ensure the reliability of the material presented in
Hematology 2022:
Anjali S. Advani
Ranjana Advani
Cem Akin
Monzr M. Al Malki
Aref Al-Kali
Frederick Appelbaum
Eugenia Vicky Asare
K. Scott Baker
Paul M. Barr
Nancy L. Bartlett
Renato Bassan
Shannon Bates
Lisa Baumann Kreuziger
Rafael Bejar
Jesus G. Berdeja
Smita Bhatia
Andrea Biondi
Larissa Bornikova
Jennifer R. Brown
Natalie S. Callander
Gunnar Cario
Shannon Carpenter
Marc Carrier
Giancarlo Castaman
Eric J. Chow
Adam D. Cohen
Claudia S. Cohn
Nathan T. Connell
Jorge Cortes
Charles Craddock
Stacy E. Croteau
Adam Cuker
Corey S. Cutler
Agnieszka Czechowicz
Anita D’Souza
Alexey V. Danilov
Faith E. Davies
Amy E. DeZern
Sunny H. Dzik
x
Yvonne A. Efebera
Matthew Ehrhardt
Toby A. Eyre
Joshua Field
Courtney D. Fitzhugh
Christopher R. Flowers
Annemarie E. Fogerty
Jacqueline S. Garcia
James N. George
Tracy I. George
Irene M. Ghobrial
Gerald J. Gleich
Victor R. Gordeuk
Jed B. Gorlin
Jason Gotlib
Erin M. Guest
Sumit Gupta
Sandra L. Haberichter
Claire N. Harrison
Andreas Hochhaus
Timothy P. Hughes
Ibrahim M. Idris
Tania Jain
Andra H. James
Paula D. James
Jill M. Johnsen
Moonjung Jung
Brad S. Kahl
Julie Kanter
Raj S. Kasthuri
Richard Kaufman
Alok A. Khorana
Adam S. Kittai
Jan-Henning Klusmann
Rami S. Komrokji
John Koreth
Johanna A. Kremer
Hovinga
Amrita Y. Krishnan
Michael H. Kroll
Rebecca Kruse-Jarres
Roshni Kulkarni
Kevin H. M. Kuo
John Kuruvilla
David J. Kuter
Andrew T. Kuykendall
Ola Landgren
Alfred I. Lee
Lai Heng Lee
Howard A. Liebman
R. Coleman Lindsley
Sagar Lonial
Izidore S. Lossos
Kami Maddocks
Paul Mansfield
John O. Mascarenhas
Michael J. Mauro
Eileen Merriman
Reid W. Merryman
Joseph Mikhael
Alice S. Mims
Alison J. Moskowitz
Kasiani C. Myers
Loretta J. Nastoupil
Robert S. Nickel
Omar Niss
Kristen M. O’Dwyer
Adam Jan Olszewski
Monica B. Pagano
Lydia H. Pecker
Kristen M. Pettit
Tycel J. Phillips
Allyson Pishko
Jerald P. Radich
Delphine Rea
Josep-Maria Ribera
Paul G. Richardson
Jacob M. Rowe
Dr Prakash Singh Shekhawat
Jeffrey E. Rubnitz
Sana Saif Ur Rehman
David A. Sallman
Kristen Sanfilippo
Santosh L. Saraf
Sharon A. Savage
Farzana Sayani
Mikkael A. Sekeres
Nina Shah
Urvi A. Shah
Joseph J. Shatzel
Bronwen E. Shaw
Shalini Shenoy
Nicholas J. Short
Sonali M. Smith
Gerald A. Soff
Stephen R. Spellman
Sarah K. Tasian
Jeffrey W. Taub
Philip A. Thompson
Darrell J. Triulzi
Suzanne Trudel
Lucie M. Turcotte
Geoffrey L. Uy
Emile Van Schaftingen
Sumithira Vasu
Tzu-Fei Wang
Jonathan Webster
Oliver Weigert
Matthew J. Wieduwilt
Tanya Wildes
Raymond Siu Ming
Wong
Jennifer C. Yui
Continuing Medical Education Information
The Hematology ASH Education Program is an annual publication that provides practicing hematologists with invaluable information on the most important areas of clinical progress.
Hematology 2022 is a peer-reviewed collection of articles
written by the 2022 ASH Education Program speakers and the
Ham-Wasserman Lecturer. The papers showcase groundbreaking advances and new concepts in 29 different fields. Every year,
the periodical provides an updated and comprehensive review
of each of the topics covered in the annual meeting education
sessions.
ABIM Maintenance of Certification
Successful completion of this CME activity enables the participant to earn up to 40
Knowledge Points in the American Board of Internal Medicine’s
(ABIM) Maintenance of Certification (MOC) program. Participants
will earn MOC points equivalent to the amount of CME credits
claimed for the activity. It is the CME activity provider’s responsibility to submit participant completion information to ACCME
for the purpose of granting ABIM MOC credit.
Claiming CME credits and ABIM points
Educational objectives
1. Employ the knowledge gained regarding the diagnosis and
treatment of malignant and nonmalignant hematologic disorders to improve patient care.
2. Discuss the state-of-the-art therapeutics in hematology.
3. Analyze the potential contribution of novel, not-yet-approved
modalities of therapy to current evidence-based management of malignant and nonmalignant hematologic disorders.
Date of release
December 2022
Date of expiration
On December 31, 2023, the ability to earn Continuing Medical Education credit and American Board of Internal Medicine
Maintenance of Certification Medical Knowledge points for this
product expires. This is the last date for users to claim credit
for this product. For questions about credit, please contact the
ASH Education Department at cme@hematology.org or call tollfree 866-828-1231 within the United States only; 1-202-776-0544
internationally.
The estimated time to complete this educational activity is 40
hours. To claim CME credit, users must complete an evaluation
of the product and a test of medical knowledge, both of which
are accessed through ASH Academy On Demand (academy.hema
tology.org). There is a one-time processing fee for claiming CME/
MOC credit. Users with scores of 80% or better on these selfassessment modules are eligible to claim credit for the activity.
To facilitate claiming of credit, the test for this product is
divided into two subtests, one of which focuses on malignant
hematology content with the other focusing on nonmalignant
content. Successful completion of each test earns the user 20
AMA Category 1 PRA CreditsTM. Users claim CME and/or ABIM
MOC credit for each test individually. You can take one or both
tests, depending on your CME and MOC needs.
The malignant hematology test consists of 31 questions, and
the nonmalignant hematology test consists of 21 questions. The
questions in each test can be answered in one sitting, or a user
can save their progress and return to complete the test at a
later time.
The malignant hematology test covers information presented
in the following sections:
•
Accreditation and Credit Designation
The American Society of Hematology is accredited
by the Accreditation Council for Continuing Medical Education to provide continuing medical education for physicians.
The American Society of Hematology designates this enduring material for a maximum of 40 AMA PRA Category 1 CreditsTM. Physicians should claim only the credit commensurate with the extent of their participation in the activity.
Physicians who participate in this CME activity but are not
licensed in the United States are also eligible for AMA Category
1 PRA CreditTM. To earn these credits, readers must pass two online tests (malignant and nonmalignant) based on chapters from
the book.
•
•
•
•
•
•
•
•
•
•
Acute Myeloid Leukemia: Improving Outcomes in Challenging Subsets
Are Alternative Donors Now Mainstream in Allogeneic
Transplant?
Beyond Routine Frontline Therapy of CML
Controversies in Aggressive NHL
Ham-Wasserman Lecture
Immunotherapy in Multiple Myeloma
Improving Outcomes in ALL
JAK/STAT Inhibition and Beyond in Ph-Negative MPNs
Long-Term Effects Monitoring for Survivors of Pediatric
Hematologic Malignancies
Maximizing Outcomes in CLL
Multiple Myeloma: Assessing the Patient and the Disease
Dr Prakash Singh Shekhawat
xi
•
•
•
•
•
Novel Approaches in MDS
To Transplant or Not to Transplant in Active or High-Risk
Myeloid Disease
Treatment Approaches for the Multiple Myeloma Patient
in 2022
Updates in Targeted Therapy in Pediatric Leukemia
Where Are We Headed in Hodgkin Lymphoma?
The nonmalignant hematology test covers information presented
in the following sections:
•
•
xii
Anxiety-Provoking Consultations: Mast Cells and Eosinophils
Autoimmune Hemolytic Anemias
•
•
•
•
•
•
•
•
•
Long-Term Effects of Curative Therapy for Sickle Cell Disease
Managing Thrombocytopenia in Challenging Situations
Obstetric Management and Complications in Sickle Cell
Disease in High- and Low-Income Countries
Prophylactic Platelet Transfusions
Reproductive and Sexual Health in Sickle Cell Disease
Thrombosis Prevention and Treatment
Update in Hemophilia
Von Willebrand Disease
What Is New in Classical Bone Marrow Failure Syndromes?
(Focus on Management)
Dr Prakash Singh Shekhawat
HAM - WASSERMAN LECTURE
Leukemogenesis in infants and young children
with trisomy 21
Department of Paediatrics and MRC Molecular Haematology Unit, MRC Weatherall Institute of Molecular Medicine, University of Oxford,
Headington, Oxford, UK
Children with Down syndrome (DS) have a greater than 100-fold increased risk of developing acute myeloid leukemia
(ML) and an approximately 30-fold increased risk of acute lymphoblastic leukemia (ALL) before their fifth birthday. ML-DS
originates in utero and typically presents with a self-limiting, neonatal leukemic syndrome known as transient abnormal
myelopoiesis (TAM) that is caused by cooperation between trisomy 21–associated abnormalities of fetal hematopoiesis
and somatic N-terminal mutations in the transcription factor GATA1. Around 10% of neonates with DS have clinical signs
of TAM, although the frequency of hematologically silent GATA1 mutations in DS neonates is much higher (~25%). While
most cases of TAM/silent TAM resolve without treatment within 3 to 4 months, in 10% to 20% of cases transformation
to full-blown leukemia occurs within the first 4 years of life when cells harboring GATA1 mutations persist and acquire
secondary mutations, most often in cohesin genes. By contrast, DS-ALL, which is almost always B-lineage, presents after
the first few months of life and is characterized by a high frequency of rearrangement of the CRLF2 gene (60%), often
co-occurring with activating mutations in JAK2 or RAS genes. While treatment of ML-DS achieves long-term survival
in approximately 90% of children, the outcome of DS-ALL is inferior to ALL in children without DS. Ongoing studies in
primary cells and model systems indicate that the role of trisomy 21 in DS leukemogenesis is complex and cell context
dependent but show promise in improving management and the treatment of relapse, in which the outcome of both
ML-DS and DS-ALL remains poor.
LEARNING OBJECTIVES
• Understand the risk of leukemia developing in children with Down syndrome
• Understand the natural history of myeloid leukemia in young children with Down syndrome
Background
Children with Down syndrome (DS) due to constitutional
trisomy 21 (T21) have a more than 50-fold increased risk
of developing acute leukemia before their fifth birthday
compared to children without DS.1,2 Few genetic disorders show such a strong link to leukemia and have offered
so many clues to clinicians and scientists about why this
should be so. While it is clear that the hematologic problems in DS must be underpinned by the presence of a
supernumerary copy of chromosome 21 (Hsa21), the more
we learn about potentially relevant mechanisms, the more
questions remain and the more enigmatic the answers
become. Nevertheless, clinical and mechanistic studies
have together significantly improved our understanding
of DS leukemias and the way we manage these disorders
in patients. Understanding DS leukemogenesis is important given that worldwide approximately 200 000 children every year are born with DS.3 Here, I describe how
observations made almost 40 years ago by Alvin Zipursky
have triggered a wealth of further studies that provide
insight into the prenatal origin of leukemias in DS as well as
the fundamental processes governing the biological properties and gene expression of hematopoietic cells.4
Leukemia and DS
Population-based studies show an increased frequency of
both acute myeloid leukemia (AML) and acute lymphoblastic leukemia (ALL) (Table 1).1,2 Given that the incidence of
nonhematologic cancers is half that reported in individuals without DS,1,2,5 this points to the specific susceptibility
of hematopoietic cells to the leukemogenic effects of T21
and also hints at effects on multiple lineages. A second distinctive feature of DS leukemias is the age at presentation,
which indicates, particularly for the specific subtype of
AML known as ML-DS, that there is a defined time window
during which T21-driven leukemic transformation occurs.
Dr Prakash Singh Shekhawat
Leukemogenesis in infants with trisomy 21 | 1
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/1/2021737/1roberts.pdf by guest on 09 December 2022
Irene Roberts
Table 1. Increased sus­cep­ti­bil­ity to leu­ke­mia in young chil­dren
with DS
Type of malig­nancy
Standardized inci­dence ratio
AML
Aged 0-4 years (ML-DS)
114
Aged 0-60 years
12
TAM and ML-DS: a mul­ti­step model of child­hood
leu­ke­mo­gen­e­sis
ALL
Aged 0-4 years
27
Aged 0-60 years
13
0.45
Solid tumorsa
mega­kar­yo­cytic mark­ers, an uncom­mon find­ing in chil­dren
with­out DS,8 while DS-ALL, unlike non-DS ALL, is almost exclu­
sively B-lin­e­age.9 Finally, DS leu­ke­mias are molec­u­larly dis­tinct,
as described in detail later.
Data reproduced with per­mis­sion from Marlow et al2 and Hasle et al.5
ML-DS is now known to orig­i­nate in utero and usu­ally pres­ents
with a self-lim­it­ing, neo­na­tal leu­ke­mic syn­drome known as tran­
sient abnor­mal myelopoiesis (TAM), or tran­sient mye­lo­pro­lif­er­a­
tive dis­or­der, that is vir­tu­ally unique to DS; while most cases of
TAM resolve within 3 to 4 months, sub­se­quent trans­for­ma­tion to
full-blown leu­ke­mia is lim­ited to the first 4 years of life (Figure 1).2,4-6
By con­trast, ALL is rare in infants with DS but oth­er­wise has
an age dis­tri­bu­tion sim­i­lar to ALL in chil­dren with­out DS.5,7 DS
leu­ke­mias also exhibit dis­tinct immunophenotypic char­ac­ter­is­
tics. Most nota­bly, blast cells in ML-DS coexpress ery­throid and
Transient leukemia
(TAM)
Fetal hematopoiec
stem/progenitor Cells
1. Trisomy 21 perturbs
fetal hematopoiesis
~25%
Leukemia
(ML-DS)
80-90%
Birth
+21
Spontaneous
remission
+21
GATA1s
2. Somac N-terminal
truncang GATA1 mutaons
(GATA1s) in T21 cells cause
neonatal leukemia (TAM)
unique to DS
10-20%
+21
GATA1s
+ other
Age
4 yrs
3. Acquision of addional
mutaons (eg in cohesin)
causes myeloid leukemia
(ML-DS) in persisng GATA1
mutant cells.
Figure 1. TAM and ML-DS, a multistep model of myeloid leukemogenesis in DS. Trisomy 21 causes an increase in the frequency of fetal
MK-erythroid stem and progenitor cells in tandem with a severe reduction of B-cell progenitors. On this cellular background, somatic
N-terminal truncating mutations in the GATA1 transcription factor gene that encode a shorter than normal GATA1 protein (GATA1s)
are acquired at a high frequency during fetal life. The expression of GATA1s causes a fetal/neonatal leukemic syndrome known as
TAM that is virtually unique to DS. Although TAM may be a severe disease, in 80% to 90% of cases it resolves spontaneously over the
first 4 months of life. Where GATA1s-producing blast cells persist, the acquisition of mutations in additional genes, particularly those
encoding cohesin complex proteins, leads to the development of ML-DS within the first 4 years of life.
2 | Hematology 2022 | ASH Education Program
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/1/2021737/1roberts.pdf by guest on 09 December 2022
a
Aged 0-60 years; the only solid tumor with an increased inci­dence
in DS is tes­tic­u­lar can­cer, which pres­ents in young adults, and the
only solid tumor in a child under the age of 5 years was a case of
ret­i­no­blas­toma.5
In the 2022 World Health Organization clas­si­fi­ca­tion, TAM and
ML-DS are col­lec­tively known as “mye­loid pro­lif­er­a­tions asso­ci­
ated with DS typ­i­cally asso­ci­ated with exon 2 or 3 GATA1 muta­
tions.”10 The key dis­cov­ery of the link between somatic N-ter­mi­nal
muta­tions in the tran­scrip­tion fac­tor gene GATA1 and TAM was
made by John Crispino’s lab, which had also first dem­on­strated
the same type of GATA1 muta­
tion in chil­
dren with DS acute
megakaryoblastic leu­ke­mia (now termed ML-DS).11,12 Further, his
lab showed that such muta­tions result in the exclu­sive pro­duc­
tion of a short GATA1 pro­tein (GATA1s) that lacks the N-ter­mi­nal
acti­va­tion domain. As the GATA1 gene is on the X chro­mo­some,
cells bear­ing GATA1s-caus­ing muta­tions also lose expres­sion of
the full-length GATA1 pro­tein, thus disrupting its nor­mal activ­
ity in reg­u­lat­ing hema­to­poi­etic cell devel­op­ment, par­tic­u­larly
of ery­throid and mega­kar­yo­cyte (MK) lin­e­ages, by alter­ing its
abil­ity to bind to and reg­u­late down­stream tar­gets.13-16 Although
almost all­cases of TAM occur in neo­na­tes with DS, an iden­ti­
cal clin­i­cal pre­sen­ta­tion can also affect neo­na­tes with mosaic
DS.17,18 These babies lack the char­ac­ter­is­tic phe­no­typic fea­tures
of DS, but periph­eral blood karyotyping shows that some or all­
of the hema­to­poi­etic cells har­bor T21, suggesting that both T21
and GATA1s-pro­duc­ing muta­tions are essen­tial for the devel­op­
ment of TAM/ML-DS. Indeed, in the sin­gle reported case of TAM
were also com­mon.27,28 While these sec­ond­ary events can occur
in other leu­ke­mias, their coacquisition with GATA1s is vir­tu­ally
unique to ML-DS. Thus, together these find­ings sup­port a mul­
ti­step model of leu­ke­mo­gen­e­sis in ML-DS in the vast major­ity
of cases (Figure 1) and pro­vide a frame­work both for under­
stand­ing the nat­u­ral his­tory of ML-DS and for inves­ti­gat­ing the
pre­cise mech­a­nisms respon­si­ble.
Clinical fea­tures and nat­u­ral his­tory of TAM and ML-DS
TAM is a fetal and neo­na­tal con­di­tion. Most cases pres­ent at or
just after birth and always by 3 months of age.25 Although TAM is
usu­ally a short-lived dis­ease that resolves with­out spe­cific ther­
apy, in fact it encompasses a wide spec­trum of sever­ity, from
clin­i­cally silent dis­ease to rap­idly fatal multiorgan fail­ure due to
the leu­ke­mic infil­tra­tion of mul­ti­ple tis­sues, par­tic­u­larly the liver
and lungs.17,18,29 While clin­i­cal stud­ies report that approx­i­ma­tely
10% of neo­na­tes with DS have TAM,4,6,17,18 pro­spec­tive stud­ies,
includ­ing our own, documenting the fre­quency of GATA1s muta­
tions in DS neo­na­tes using sen­si­tive next gen­er­a­tion sequenc­ing
indi­cate an even higher fre­quency of TAM, of approx­i­ma­tely 30%,
although many of these cases are clin­i­cally and hematologically
silent.6 As the esti­mated inci­dence of DS world­wide is 1 per 500
to 1 per 2000 live births,30 TAM is by far the most com­mon leu­ke­
mia in neo­na­tes. The char­ac­ter­is­tic hema­to­logic fea­tures of TAM
are key to the diag­no­sis and to our under­stand­ing of the path­
o­gen­e­sis of the dis­ease. While neo­na­tes with silent TAM due to
small GATA1s clones have no reli­able hema­to­logic mark­ers of dis­
ease, most cases of clin­i­cal TAM have leu­ko­cy­to­sis and increased
periph­eral blood blasts (>10%) (Table 2). While our group uses
a blast per­
cent­
age of greater than 10%, oth­
ers rec­
om­
mend
Table 2. Clinical and hema­to­logic fea­tures of TAM and silent TAM in neo­na­tes with DS
Condition
Clinical fea­tures
Hematologic fea­tures
TAM
GATA1s muta­tion; VAF var­i­able but usu­ally
<80%
• Hepatosplenomegaly 40%
• Skin rash 20%
• Pleural/peri­car­dial effu­sion ± asci­tes 10%
• Jaundice 70%-80%
• Blasts >10%a
• Hct usu­ally nor­mal
• Platelet count increased, nor­mal, or
reduced
• Leucocytosis usual and may exceed
100 000×109/l
• MK frag­ments usu­ally pres­ent on blood
smear
Silent TAM
GATA1s muta­tion; VAF usu­ally low, <10%
• No increase in fre­quency of hepatosplenomegaly,
jaun­dice, skin rash, pleu­ral/peri­car­dial effu­sions,
or asci­tes com­pared to neo­na­tes with DS who
lack GATA1s muta­tions
• Blasts ≤10%*
• Hct increased or nor­mal
• Platelets nor­mal or reduced
• WBC nor­mal
Neonate with DS and no GATA1s muta­tions
• Jaundice is com­mon in neo­na­tes with DS (~60%)
and is not a use­ful indi­ca­tor of TAM unless other
clin­i­cal signs are pres­ent.
• As up to 10% of neo­na­tes with DS have
hepatosplenomegaly, rash, pleu­ral/peri­car­dial
effu­sion, or asci­tes sec­ond­ary to med­i­cal
com­pli­ca­tions in the absence of GATA1 muta­tions,
care must be taken to con­firm the diag­no­sis
by blood smear review and/or GATA1 muta­tion
anal­y­sis.
Compared with neo­na­tes with­out DS:
• 98% have blasts on the blood smear
(<20%).
• Hct is often increased (24% are
poly­cy­the­mic, Hct >0.65).
• Platelet counts are lower (40% are
throm­bo­cy­to­pe­nic, <150 000×109/l).
• Leukocyte, neu­tro­phil, mono­cyte, and
baso­phil counts are increased.
Others have recommended defin­ing TAM as “the pres­ence of at least 5% blasts defined by immunophenotyping or mor­phol­ogy and/or the pres­ence
of a GATA1 muta­tion in a neo­nate with DS.”31
a
VAF, var­i­ant allele fre­quency; WBC, white blood cell count.
Data reproduced with per­mis­sion from Roberts et al6 and Tunstall et al.29
Dr Prakash Singh Shekhawat
Leukemogenesis in infants with tri­somy 21 | 3
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/1/2021737/1roberts.pdf by guest on 09 December 2022
with a GATA1 muta­tion and no evi­dence of T21, the GATA1 muta­
tion was a large dele­tion rather than a pro­ducer of GATA1s.19 In
addi­tion, inherited GATA1s-pro­duc­ing muta­tions are not leu­ke­
mo­genic in diso­mic indi­vid­u­als,20 although rare fam­i­lies with germline GATA1s muta­tions are reported where the devel­op­ment
of ML-DS–like acute leu­ke­mia was accom­pa­nied by acquired tri­
somy or tetrasomy 21,21,22 fur­ther supporting the impor­tance of
GATA1s/T21 as a potent onco­genic driver.
Studies on pre­na­tal and cord blood sam­ples from babies
with DS have shown that somatic GATA1 muta­tions are acquired
before birth, most likely dur­ing the sec­ond tri­mes­ter.23-25 Furthermore, mul­ti­ple GATA1 muta­tions occur in approx­i­ma­tely
25% of cases, suggesting a strong selec­
tive advan­
tage for
GATA1s in fetal cells.6,24 We pre­vi­ously showed that T21 itself
causes a strik­
ing increase in fetal liver mega­
kar­
yo­
cytic-ery­
throid stem and pro­gen­i­tor cells in DS in the absence of GATA1s
muta­tions, suggesting that it is the T21-medi­ated per­tur­ba­tion
of fetal hema­to­poi­e­sis that under­lies the leu­ke­mo­genic effects
of GATA1s in neo­na­tes with DS.26 Paired ML-DS and TAM sam­ples
show the same GATA1 muta­tions, confirming that ML-DS and
TAM are clon­ally linked con­di­tions and suggesting that other
fac­tors, in addi­tion to GATA1s, are needed to effect a full leu­
ke­mic trans­for­ma­tion from a tran­sient neo­na­tal syn­drome to
ML-DS, a con­di­tion that is fatal in the absence of che­mo­ther­
apy.24,27,28 Whole-genome and targeted sequenc­
ing in more
than 300 patient sam­ples not only con­firmed the pres­ence of
GATA1s muta­tions in ML-DS but also iden­ti­fied loss-of-func­tion
muta­tions in cohesin genes as the prin­ci­pal sec­ond­ary genetic
events in ML-DS; muta­tions in genes encoding epi­ge­netic reg­
u­la­tors and com­po­nents of tyro­sine kinase sig­nal­ing path­ways
defin­ing TAM on the basis of blasts greater than or equal to 5%.31
Importantly, how­ever, the blast per­cent­age can only serve as a
guide for the pres­ence of GATA1s muta­tions because more than
one-third of DS neo­na­tes with­out GATA1s muta­tions have blasts
greater than or equal to 5%, auto­mated counts are unre­li­able,
and man­ual counts require con­sid­er­able expe­ri­ence. Typical
blast cells in TAM resem­ble imma­ture or par­tially dif­fer­en­ti­ated
megakaryoblasts (Figure 2), but they are often pleo­mor­phic and
may dis­play fea­tures of ery­throid, baso­phil, or eosin­o­phil dif­fer­
en­ti­a­tion,32-34 reflecting the roles of GATA1 in mul­ti­ple lin­e­ages.35
Consistent with this, in addi­tion to CD117 (c-kit) TAM blasts coexpress var­i­able pro­por­tions of CD34, CD7, CD36, CD41/42b, and
CD235a (glycophorin A).8 Morphologic evi­dence of the per­tur­
ba­tion of megakaryopoiesis by GATA1s is almost always seen on
blood smears in TAM (Figure 2), although throm­bo­cy­to­pe­nia is
not a reli­able diag­nos­tic indi­ca­tor of TAM because the plate­let
count may be nor­mal, reduced, or even increased.29 While ane­
mia is uncom­mon, the median hemat­o­crit (Hct) is lower in neo­
na­tes with TAM com­pared with DS neo­na­tes with­out TAM (Table
2). This is con­sis­tent with some impair­ment of eryth­ro­poi­e­sis by
GATA1s, although less than seen in diso­mic mod­els of GATA1s
func­tion, per­haps due to com­pen­sa­tion by the under­ly­ing T21asso­ci­ated increase in eryth­ro­poi­e­sis in neo­na­tes with DS.6,13,14,29
For neo­
na­
tes with severe TAM (5%-20%), fac­
tors for early
death include pre­
term deliv­
ery, asci­
tes, a leu­
ko­
cyte count
greater than 100×109/l, hepa­to­meg­aly, and coagulopathy.17,18,29
For such cases the main­
stay of treat­
ment is sup­
port­
ive care
and the judi­cious use of cytarabine as recommended in recent
guide­lines.29,36 While cytarabine reduces the risk of early death in
severe TAM, there is no evi­dence yet that treat­ment pre­vents MLDS.37,38 In most neo­na­tes, TAM resolves within 3 to 4 months, even
when the dis­ease is severe. As 10% to 20% of cases of clin­i­cal
TAM sub­se­quently develop ML-DS, even after appar­ent com­plete
remis­sion, it seems sen­si­ble for chil­dren with TAM to be offered
reg­u­lar review until the age of 4 years, the time win­dow dur­ing
4 | Hematology 2022 | ASH Education Program
DS acute lym­pho­blas­tic leu­ke­mia
In con­trast to ML-DS, chil­dren with DS-ALL have an infe­rior out­
come to those with­out DS, both because of treat­ment-related
tox­ic­ity and intrin­sic dif­fer­ences in ALL biol­ogy.9,40 A recent large
inter­na­tional study matched for cyto­ge­netic sub­group showed
that com­pared to chil­dren with­out DS, those with DS had worse
5-year event-free sur­vival (75% vs 88%), over­all sur­vival (77% vs
94%), and post-induc­tion treat­ment-related mor­tal­ity (12.2% vs
2.7%).40 Survival after relapse is also infe­rior, even after trans­plan­
ta­tion, although chi­me­ric anti­gen recep­tor T-cell ther­apy has
shown prom­ise.41 While the clin­i­cal fea­tures are sim­i­lar to non-DS
ALL, the pat­tern of cyto­ge­netic abnor­mal­i­ties in DS-ALL is dis­tinct,
con­sis­tent with the leu­ke­mo­genic role of T21 in this dis­ease. The
fre­quency of “good prog­no­sis” cyto­ge­netic abnor­mal­i­ties, such
as ETV6-RUNX1 and high hyperdiploidy, is lower in DS-ALL, and
intrigu­ingly, BCR-ABL trans­lo­ca­tions and KMT2A gene rearrangements are rare or absent in most series.9,40 Instead, the most
com­mon geno­mic alter­ation in DS-ALL is the rearrangement of
CRLF2 (CRLF2r), which is found in approx­i­ma­tely 60% of cases,
half of which also have acti­vat­ing muta­tions in JAK2.42-44 A higher
pro­por­tion of RAS/MAPK path­way muta­tions is also reported in
DS-ALL, usu­ally in JAK2 wild-type cases or subclones.45,46 In addi­
tion, increas­ing evi­dence sug­gests the poten­tial impor­tance of
her­i­ta­ble risk loci (such as PAX5 or IKZF1 dele­tion) both in the
risk of devel­op­ing ALL in chil­dren with DS and also in treat­ment
out­come, par­tic­u­larly for IKZF1 dele­
tion in which a poten­
tial
mech­a­nism involv­ing decreased enhancer activ­ity, dif­fer­en­tial
pro­tein bind­ing, and effects on the cell growth of IKZF1 var­i­ants
has been reported.40,47 Finally, it is nota­ble that in chil­dren with­
out DS, T21 is one of the most fre­quent chro­mo­somal abnor­mal­
i­ties in B-ALL, par­tic­u­larly in high hyperdiploid ALL, where up to
100% of cases have 1 to 3 addi­tional cop­ies of Hsa21,48 supporting a key role for increased Hsa21 dos­age in lym­phoid leu­ke­mo­
gen­e­sis.
Although immunophenotypic and transcriptome anal­y­sis sug­
gests that DS-ALL blast transcriptomes are mainly shaped by their
molec­u­lar cyto­ge­netic sub­group,45,49 clin­i­cal and mech­a­nis­tic
stud­ies also sup­port an impor­tant role for T21 at sev­eral stages
in leu­ke­mo­gen­e­sis. Indeed, the acqui­si­tion of a super­nu­mer­ary
Hsa21 is always a non­ran­dom event restricted to cer­tain cyto­
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/1/2021737/1roberts.pdf by guest on 09 December 2022
Figure 2. Peripheral blood smear from a neonate with TAM.
Blood smear of a 3-day-old neonate with DS and an exon
2 mutation in the GATA1 gene showing pleomorphic blast cells
with features of immature and partially differentiated megakaryoblasts and a giant dysplastic platelet. May Grunwald Giemsa stain, ×100 magnification.
which these chil­dren are at risk of ML-DS.2,5,17,18,29 At pres­ent, no
clin­i­cal, hema­to­logic, or molec­u­lar fea­tures at birth reli­ably pre­
dict which chil­dren develop ML-DS, although molec­u­lar or flow
cytometric evi­dence of resid­ual dis­ease at 3 months con­fers an
increased risk of sub­se­quent ML-DS.37,38 In our expe­ri­ence, the
per­sis­tence of resid­ual GATA1s clones beyond 4 months makes
sub­se­quent ML-DS almost inev­i­ta­ble, although such clones are
usu­ally too small to pro­duce detect­able molec­u­lar, hema­to­logic,
or clin­i­cal evi­dence of their pres­ence. Although the median age
at diag­no­sis of ML-DS is 12 to 18 months, the evo­lu­tion of TAM to
ML-DS usu­ally man­i­fests as slow, pro­gres­sive pan­cy­to­pe­nia over
sev­eral months with occa­sional cir­cu­lat­ing blasts. However, in
some cases leu­ke­mic trans­for­ma­tion is much more acute, with
rap­idly increas­ing blasts, so any change in the blood count dur­
ing fol­low-up should prompt a care­ful review of the blood smear.
Adaptations to ML-DS che­mo­ther­apy reg­i­mens to min­i­mize
the increased tox­ic­ity of many drugs for patients with DS have
improved dis­ease-free sur­vival to 90%, although the out­come
for chil­dren who relapse remains poor.39
ge­netic sub­types of DS-ALL (eg, ETV-RUNX1 and P2RY8-CRLF2);
abnor­mal­i­ties of Hsa21 have also been esti­mated to affect up to
65% of all­child­hood B-ALL cases.50,51 The exis­tence of a pre­na­
tal stage of DS-ALL, although pos­tu­lated, has never been dem­
on­strated and might seem unlikely given the rar­ity of DS-ALL in
chil­dren under 12 months of age.9 On the other hand, there is
increas­ing evi­dence that T21 severely impairs fetal B-pro­gen­i­tor
devel­
op­
ment in humans, due both to intrin­
sic hema­
to­
poi­
etic
stem/pro­gen­i­tor defects and extrin­sic reg­u­la­tion through an
altered hema­to­poi­etic micro­en­vi­ron­ment,26,52 and that quan­ti­ta­
tive and qual­i­ta­tive defects in B cells are com­mon in chil­dren with
DS. This sug­gests that pre­na­tal per­tur­ba­tion of B lymphopoiesis
by T21 could shape the cel­lu­lar con­text for leu­ke­mia ini­ti­a­tion.
While clin­i­cal stud­ies pro­vide irre­fut­able evi­dence that T21 itself
is an impor­tant fac­tor in DS leu­ke­mo­gen­e­sis, model sys­tems have
been used to more pre­cisely deci­pher how the super­nu­mer­ary
copy of Hsa21 in hema­to­poi­etic cells pro­vi­des such a strong
leu­ke­mo­genic stim­u­lus. Hsa21 has approx­i­ma­tely 230 pro­teincod­ing genes, many of which have key roles in nor­mal and malig­
nant hema­to­poi­e­sis (such as DYRK1A, ERG, ETS2, and RUNX1),
as well as almost twice as many non–pro­
tein-cod­
ing genes,
includ­ing 5 microRNAs (miRNAs).3 Most stud­ies of DS leu­ke­mo­
gen­e­sis have focused on the dos­age of genes on Hsa21, either
using mouse mod­els based on seg­men­tal tri­so­mies or induced
plu­rip­o­tent stem cells (iPSC) (Table 3).14,53,54 While some Hsa21
genes and tis­sues exhibit the expected 1.5-fold higher expres­
sion lev­els than the cor­re­spond­ing diso­mic tis­sues, dif­fer­ences
in gene expres­sion are highly tis­sue-spe­cific and cell pop­u­la­tionspe­cific. This may explain why, even though the increased dos­
age of sev­eral Hsa21 genes has been shown to be leu­ke­mo­genic
in model sys­tems, direct links between altered gene expres­sion
and hema­to­poi­etic phe­no­types in DS have been so dif­fi­cult to
estab­lish and why none of these phe­no­types has so far been
explained by a sin­gle Hsa21 gene act­ing alone. Furthermore,
many stud­ies now show that the per­tur­ba­tion of gene expres­
sion by T21 is genome-wide.3,52,55 This sug­gests that epi­ge­netic
mech­a­nisms of cel­lu­lar adap­ta­tion to aneu­ploidy are likely to be
involved, par­tic­u­larly dur­ing fetal life when the exqui­site con­trol
of gene expres­sion is cru­cial for nor­mal devel­op­ment. In line with
this, recent work has found genome-wide effects on the DNA
Table 3. Principal genes on Hsa21 impli­cated in DS leu­ke­mo­gen­e­sis
Gene
Role in DS leu­ke­mia
DYRK1A
Promotes AMKL and B-ALL in murine mod­els.62,63
ERG
Cooperates with GATA1s to pro­mote a TAM-like dis­ease in a mouse model58; coop­er­ates with ETS2, RUNX1, and GATA1s to
pro­mote expan­sion of fetal MK pro­gen­i­tors in a T21 iPSC model.54
ETS2
Cooperates with ERG, RUNX1, and GATA1s to pro­mote expan­sion of fetal MK pro­gen­i­tors in a T21 iPSC model.54
HMGN1
Overexpression of HMGN1 pro­motes B-cell self-renewal, increased his­tone H3K27 acet­y­la­tion, and down­stream B-lin­e­age gene
expres­sion; coop­er­ates with BRC-ABL or CRLF2, JAK2, PAX5, and IKZF to pro­mote DS-ALL.64,65
RUNX1
Cooperates with ERG, ETS2, and GATA1s to pro­mote expan­sion of fetal MK pro­gen­i­tors in a T21 iPSC model.54
MicroRNAs
Overexpression of miR-125b-2 coop­er­ates with GATA1s to pro­mote murine fetal MK pro­gen­i­tor pro­lif­er­a­tion and self renewal in
vitro59; miR-99a, miR-125b-2, and miR-155 coop­er­ate with GATA1s and cohesin insuf­fi­ciency (STAG2 KO) to pro­mote AMKL in human
fetal cells in xeno­graft mod­els.61
AMKL, acute megakaryoblastic leu­ke­mia.
Dr Prakash Singh Shekhawat
Leukemogenesis in infants with tri­somy 21 | 5
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/1/2021737/1roberts.pdf by guest on 09 December 2022
Role of tri­somy 21 in leu­ke­mo­gen­e­sis in DS
meth­yl­a­tion of hema­to­poi­etic cells in cord blood from neo­na­
tes with DS as well as sig­nif­i­cant changes in the meth­yl­a­tion of
regions close to genes key to nor­mal hema­to­poi­etic devel­op­
ment, such as RUNX1 and FLI1.56
Clinical stud­ies indi­cate that leu­ke­mia ini­ti­a­tion in DS, at least
for ML-DS, crit­i­cally depends upon a fetal cell con­text as well
as the pres­ence of T21. Euploid mouse mod­els high­light devel­
op­men­tal-spe­cific dif­fer­ences in GATA1s func­tion, but these do
not develop leu­ke­mia.15 Similarly, sev­eral mouse mod­els based
on seg­men­tal tri­so­mies, iPSC, or the overexpression of a sin­
gle Hsa21 ortholog have not so far fully reca­pit­u­lated human
TAM/ML-DS despite iden­ti­fy­ing sev­eral potent onco­genic can­
di­dates (eg, DYRK1A, ERG, ETS2) that can trig­ger mye­lo­pro­lif­er­
a­tive or leu­ke­mia dis­or­ders,57,58 per­haps because they lack key
fea­tures of human fetal cells. In line with this, mod­els based
on inser­
tional muta­
gen­
e­
sis in murine fetal liver cells more
closely mimic the immunophenotypic and molec­u­lar char­ac­
ter­is­tics of TAM/ML-DS.27,59,60 Furthermore, Wagenblast and col­
leagues used clus­tered reg­u­larly interspaced short pal­in­dromic
repeats-Cas9 editing targeting GATA1 and STAG2 in human DS
fetal liver cells to inves­ti­gate the mech­a­nism of coop­er­a­tion
between T21, GATA1s, and cohesin loss in DS leu­ke­mo­gen­e­
sis and suc­cess­fully cre­ated a model that reca­pit­u­lated many
aspects of TAM and ML-DS.61 Using a com­bi­na­tion of genomewide anal­y­sis of GATA1-bind­ing sites, miRNA pro­fil­ing, and overexpression and knock­out (KO) exper­i­ments, they went on to
show that a sub­set of Hsa21 miRNAs (miR-99a, miR-125b-2, and
miR-155) was respon­si­ble, at least in part, for a GATA1s-induced
TAM-like leu­ke­mia in fetal cells,61 supporting ear­lier data show­
ing a role for Hsa21 miRNAs.59,60 Intriguingly, in their model, T21
appeared unnec­es­sary for the devel­op­ment of ML-DS as both
DS and nor­mal fetal liver cells could be transformed to a sim­i­
lar degree when dual edited to pro­duce GATA1s in the set­ting
of STAG2 KO. If T21 is mostly impor­tant for leu­ke­mia ini­ti­a­tion
in fetal cells in DS mye­loid leu­ke­mias, as these data sug­gest,
this may explain why TAM and the acqui­si­tion of GATA1s muta­
tions are con­fined to fetal and neo­na­tal life and why ML-DS is
restricted to a time win­dow dur­ing which fetal cells can still
sur­vive post­na­tally.
The role of T21 in leu­ke­mia ini­ti­a­tion and pro­gres­sion in DSALL has mainly been stud­ied using mouse mod­els in tan­dem
with the molec­
u­
lar anal­
y­
sis of pri­
mary patient sam­
ples.7,62-65
with BCR-ABL or a com­bi­na­tion of 4 DS-ALL cyto­ge­netic events
(CRLF2 overexpression, an acti­vat­ing JAK2 muta­tion (JAK2R683G),
Pax5 haploinsufficiency, and expres­sion of a dom­i­nant neg­a­tive
Ikzf isoform).64,65
Conclusion
Clinical and exper­i­men­tal stud­ies, only some of which are sum­
ma­rized here, have together yielded fas­ci­nat­ing insights into DS
leu­ke­mo­gen­e­sis and under­score the com­plex­ity of the role of
T21 in hema­to­poi­etic cells. The effects of T21 are devel­op­men­tal
stage–spe­cific and are most evi­dent before birth, when strik­ing
abnor­mal­i­ties of erythro-megakaryopoiesis and B lymphopoiesis
are pres­ent in the liver and bone mar­row (Figure 3). The pres­
ence of T21 and somatic GATA1s muta­tions in fetal hema­to­poi­etic
cells is essen­tial for leu­ke­mia ini­ti­a­tion in TAM/ML-DS. Whether
rewiring of gene expres­sion pro­grams in fetal early lym­phoid or
B pro­gen­i­tors per­sists into early child­hood and coop­er­ates with
aber­rant CRLF2/JAK2 and/or RAS func­tion to ini­ti­ate DS-ALL is
not known. Many ques­tions remain to be answered, not least
the rea­sons for the high fre­quency of somatic GATA1 muta­tions
in DS fetal blood cells and of CRLF2/JAK2 aber­ra­tions in DS-ALL.
Figure 3. Impact of T21 on hematopoiesis and leukemia in DS. Cartoon summarizing the putative mechanisms that link T21 with
altered hematopoiesis and leukemia in early childhood in children with DS. A trisomy 21–mediated genome-wide perturbation of
gene expression from early in embryonic/fetal development causes the expansion of a rapidly proliferating hematopoietic stem and
myeloid progenitor pool with erythroid/MK bias in fetal liver and BM (bone marrow). These effects are hematopoietic cell-intrinsic
but supported by T21-driven alterations in the microenvironment. The acquisition of somatic GATA1 mutations that encode a short
GATA1 protein (GATA1s), possibly as a mutagenic effect of T21, cause further expansion of erythro-megakaryocytic cells and selective
expansion of GATA1s clones, leading to the fetal/neonatal leukemia TAM. In 10% to 20% of cases of clinical TAM, ML-DS develops when
GATA1s clones persist and acquire secondary mutations, most often in cohesin genes. The expansion of fetal megakaryopoiesis in DS
occurs at the expense of B-progenitor development due to a T21-driven failure to properly activate the B-lineage molecular programs.
This may lead to postnatal expansion of a depleted abnormally programmed B-progenitor pool, perhaps secondary to infections in
early childhood, susceptible to transformation by aberrations in CRLF2, JAK2, and or RAS pathway signaling. HSC, hematopoietic
stem cell; HSPC, hematopoietic stem/progenitor cell.
6 | Hematology 2022 | ASH Education Program
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/1/2021737/1roberts.pdf by guest on 09 December 2022
Several genes on Hsa21 have been shown to be impor­tant for
nor­mal B-cell growth and dif­fer­en­ti­a­tion, includ­ing DYRK1A, ERG,
ETS2, HMGN1, and RUNX1; overexpression of DYRK1A and also of
HMGN1 have been directly linked to the path­o­gen­e­sis of B-ALL.6265
Increased expres­sion of DYRK1A has been reported in a num­
ber of leu­ke­mias, includ­ing B-ALL, and has recently been shown
to be nec­es­sary for the growth of B-ALL cells.63 Although not spe­
cif­i­cally inves­ti­gated in DS-ALL, this work is par­tic­u­larly inter­est­
ing because it showed that small-mol­e­cule inhib­i­tors targeting
DYRK1A were effec­tive in ani­mal mod­els of B-ALL.63 Also tak­ing
dif­fer­en­tial gene expres­sion in pri­mary patient cells as a starting
point, Lane et al. performed exper­i­ments in the Ts65Dn mouse,
one of the most well-char­ac­ter­ized seg­men­tal T21 mouse mod­
els, and not only narrowed down the puta­tive region of Hsa21
(21q22) to a region of 31 genes suf­fi­cient to alter B-cell self-renewal and dif­fer­en­ti­a­tion but also attrib­uted it to overexpression
of just 1 of these, HMGN1, which encodes a nucle­o­some-bind­ing
pro­tein.64 They went on to dem­on­strate that overexpression of
HMGN1 caused both a global increase in his­tone H3K27 acet­
y­
la­
tion and down­
stream B-lin­
e­
age gene expres­
sion and also
pro­
moted the devel­
op­
ment of B-ALL, albeit in coop­
er­
a­
tion
Evi­dence for a muta­genic phe­no­type or a gen­er­al­ized defect in
DNA repair in DS tis­sues is sparse66 and precisely how hema­to­
poi­etic cells to adapt to the poten­tially dev­as­tat­ing effects of
aneu­ploidy, and the extent to which this varies in dif­fer­ent cell
types, remains to be determined.
Conflict-of-inter­est dis­clo­sure
Irene Roberts: no com­pet­ing finan­cial inter­ests to declare.
Off-label drug use
Irene Roberts: nothing to disclose.
Correspondence
References
1.
Hasle H, Clemmensen IH, Mikkelsen M. Risks of leu­kae­mia and solid tumours
in indi­vid­u­als with Down’s syn­drome. Lancet. 2000;355(9199):165-169.
2. Marlow EC, Ducore J, Kwan ML, et al. Leukemia risk in a cohort of 3.9 mil­lion
chil­dren with and with­out Down syn­drome. J Pediatr. 2021;234(July):172180.e3180e3.
3. Antonarakis SE, Skotko BG, Rafii MS, et al. Down syn­drome. Nat Rev Dis
Primers. 2020;6(1):9.
4. Zipursky A. Transient leu­kae­mia—a benign form of leu­kae­mia in new­born
infants with tri­somy 21. Br J Haematol. 2003;120(6):930-938.
5. Hasle H, Friedman JM, Olsen JH, Rasmussen SA. Low risk of solid tumors in
per­sons with Down syn­drome. Genet Med. 2016;18(11):1151-1157.
6. Roberts I, Alford K, Hall G, et al; Oxford-Imperial Down Syndrome Cohort
Study Group. GATA1-mutant clones are fre­quent and often unsus­pected in
babies with Down syn­drome: iden­ti­fi­ca­tion of a pop­u­la­tion at risk of leu­ke­
mia. Blood. 2013;122(24):3908-3917.
7. Lee P, Bhansali R, Izraeli S, Hijiya N, Crispino JD. The biol­ogy, path­o­gen­e­sis
and clin­i­cal aspects of acute lym­pho­blas­tic leu­ke­mia in chil­dren with Down
syn­drome. Leukemia. 2016;30(9):1816-1823.
8. Brouwer N, Matarraz S, Nierkens S, et al. Immunophenotypic anal­y­sis of
acute megakaryoblastic leu­
ke­
mia: a EuroFlow study. Cancers (Basel).
2022;14(6):1583.
9. Buitenkamp TD, Izraeli S, Zimmermann M, et al. Acute lym­pho­blas­tic leu­
ke­mia in chil­dren with Down syn­drome: a ret­ro­spec­tive anal­y­sis from the
Ponte di Legno study group. Blood. 2014;123(1):70-77.
10. Khoury J, Solary E, Abla O, et al. The 5th Edition of The World Health Organization Classification of Hematolymphoid Tumors: Myeloid and his­
tio­
cytic/den­dritic neo­plasms. Leukemia. 2022;36(7):1703-1719.
11. Mundschau G, Gurbuxani S, Gamis AS, Greene ME, Arceci RJ, Crispino JD.
Mutagenesis of GATA1 is an ini­ti­at­ing event in Down syn­drome leu­ke­mo­
gen­e­sis. Blood. 2003;101(11):4298-4300.
12. Wechsler J, Greene M, McDevitt MA, et al. Acquired muta­tions in GATA1
in the megakaryoblastic leu­ke­mia of Down syn­drome. Nat Genet. 2002;
32(1):148-152.
13. Byrska-Bishop M, VanDorn D, Campbell AE, et al. Pluripotent stem cells
reveal ery­throid-spe­cific activ­i­ties of the GATA1N-ter­mi­nus. J Clin Invest.
2015;125(3):993-1005.
14. Ling T, Birger Y, Stankiewicz MJ, et al. Chromatin occu­pancy and epi­ge­
netic anal­y­sis reveal new insights into the func­tion of the GATA1N ter­mi­nus
in eryth­ro­poi­e­sis. Blood. 2019;134(19):1619-1631.
15. Li Z, Godinho FJ, Klusmann JH, Garriga-Canut M, Yu C, Orkin SH. Developmental stage-selec­tive effect of somat­i­cally mutated leu­ke­mo­genic tran­
scrip­tion fac­tor GATA1. Nat Genet. 2005;37(6):613-619.
16. Kadri Z, Shimizu R, Ohneda O, et al. Direct bind­ing of pRb/E2F-2 to GATA-1
reg­u­lates mat­u­ra­tion and ter­mi­nal cell divi­sion dur­ing eryth­ro­poi­e­sis. PLoS
Biol. 2009;7(6):e1000123.
17. Klusmann JH, Creutzig U, Zimmermann M, et al. Treatment and prog­nos­
tic impact of tran­sient leu­ke­mia in neo­na­tes with Down syn­drome. Blood.
2008;111(6):2991-2998.
Dr Prakash Singh Shekhawat
Leukemogenesis in infants with tri­somy 21 | 7
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/1/2021737/1roberts.pdf by guest on 09 December 2022
Irene Roberts, Department of Paediatrics, MRC Molecular Haematology Unit, MRC Weatherall Institute of Molecular Medicine, University of Oxford, Headington, Oxford OX3 9DS, United Kingdom;
e-mail: irene​­.roberts@imm​­.ox​­.ac​­.uk, irene​­.roberts@pae­di­at­rics​
­.ox​­.ac​­.uk.
18. Gamis AS, Alonzo TA, Gerbing RB, et al. Natural his­tory of tran­sient mye­
lo­pro­lif­er­a­tive dis­or­der clin­i­cally diag­nosed in Down syn­drome neo­na­
tes: a report from the Children’s Oncology Group Study A2971. Blood.
2011;118(26):6752-6759, quiz 6996.
19. Lukes J Jr, Danek P, Alejo-Valle O, et al. Chromosome 21 gain is dis­pens­able
for tran­sient mye­lo­pro­lif­er­a­tive dis­or­der driven by a novel GATA1 muta­tion.
Leukemia. 2020;34(9):2503-2508.
20. Hollanda LM, Lima CS, Cunha AF, et al. An inherited muta­tion lead­ing to
pro­duc­tion of only the short isoform of GATA-1 is asso­ci­ated with impaired
eryth­ro­poi­e­sis. Nat Genet. 2006;38(7):807-812.
21. Hasle H, Kline RM, Kjeldsen E, et al. Germline GATA1s gen­er­at­ing muta­tions
pre­dis­pose to leu­ke­mia with acquired tri­somy 21 and Down syn­drome-like
phe­no­type. Blood. 2021;139(21):3159-3165.
22. Parrella S, Aspesi A, Quarello P, et al. Loss of GATA-1 full length as a
cause of Diamond-Blackfan ane­
mia phe­
no­
type. Pediatr Blood Cancer.
2014;61(7):1319-1321.
23. Rainis L, Bercovich D, Strehl S, et al. Mutations in exon 2 of GATA1 are early
events in mega­kar­yo­cytic malig­nan­cies asso­ci­ated with tri­somy 21. Blood.
2003;102(3):981-986.
24. Ahmed M, Sternberg A, Hall G, et al. Natural his­tory of GATA1 muta­tions in
Down syn­drome. Blood. 2004;103(7):2480-2489.
25. Tamblyn JA, Norton A, Spurgeon L, et al. Prenatal ther­apy in tran­sient
abnor­mal myelopoiesis: a sys­tem­atic review. Arch Dis Child Fetal Neonatal
Ed. 2016;101(1):F67-F71.
26. Roy A, Cowan G, Mead AJ, et al. Perturbation of fetal liver hema­to­poi­etic
stem and pro­gen­i­tor cell devel­op­ment by tri­somy 21. Proc Natl Acad Sci
USA. 2012;109(43):17579-17584.
27. Yoshida K, Toki T, Okuno Y, et al. The land­scape of somatic muta­tions in
Down syn­drome-related mye­loid dis­or­ders. Nat Genet. 2013;45(11):12931299.
28. Labuhn M, Perkins K, Matzk S, et al. Mechanisms of pro­gres­sion of mye­loid
preleukemia to transformed mye­loid leu­ke­mia in chil­dren with Down syn­
drome. Cancer Cell. 2019;36(3):340.
29. Tunstall O, Bhatnagar N, James B, et al. Guidelines for the inves­ti­ga­tion and
man­age­ment of tran­sient leu­kae­mia of Down syn­drome. Br J Haematol.
2018;182(2):200-211.
30. de Graaf G, Buckley F, Skotko BG. Estimation of the num­ber of peo­ple with
Down syn­drome in Europe. Eur J Hum Genet. 2021;29(3):402-410.
31. Goemans BF, Noort S, Blink M, et al. Sensitive GATA1 muta­tion screen­ing
reli­ably identifies neo­na­tes with Down syn­drome at risk for mye­loid leu­ke­
mia. Leukemia. 2021;35(8):2403-2406.
32. Bozner P. Transient mye­lo­pro­lif­er­a­tive dis­or­der with ery­throid dif­fer­en­ti­a­
tion in Down syn­drome. Arch Pathol Lab Med. 2002;126(4):474-477.
33. Maroz A, Stachorski L, Emmrich S, et al. GATA1s induces hyperproliferation
of eosin­o­phil pre­cur­sors in Down syn­drome tran­sient leu­ke­mia. Leukemia.
2014;28(6):1259-1270.
34. Bain BJ. Basophilic dif­fer­en­ti­a­tion in tran­sient abnor­mal myelopoiesis. Am J
Hematol. 2016;91(8):847.
35. Katsumura KR, Bresnick EH; GATA Factor Mechanisms Group. The GATA fac­
tor rev­o­lu­tion in hema­tol­ogy. Blood. 2017;129(15):2092-2102.
36. Al-Kershi S, Golnik R, Flasinski M, et al. Recommendations for diag­no­sis
and treat­ment of chil­dren with tran­sient abnor­mal myelopoiesis (TAM) and
mye­loid leu­ke­mia in Down syn­drome (ML-DS). Klin Padiatr. 2021;233(6):267277.
37. Flasinski M, Scheibke K, Zimmermann M, et al. Low-dose cytarabine to pre­
vent mye­loid leu­ke­mia in chil­dren with Down syn­drome: TMD Prevention
2007 study. Blood Adv. 2018;2(13):1532-1540.
38. Yamato G, Deguchi T, Terui K, et al. Predictive fac­tors for the devel­op­ment
of leu­ke­mia in patients with tran­sient abnor­mal myelopoiesis and Down
syn­drome. Leukemia. 2021;35(5):1480-1484.
39. Hitzler J, Alonzo T, Gerbing R, et al. High-dose AraC is essen­tial for the
treat­ment of ML-DS inde­pen­dent of post-induc­tion MRD: results of the
COG AAML1531 trial. Blood. 2021;138(23):2337-2346.
40. Michels N, Boer JM, Enshaei A, et al. Minimal resid­ual dis­ease, long-term
out­come, and IKZF1 dele­tions in chil­dren and ado­les­cents with Down syn­
drome and acute lym­pho­cytic leu­kae­mia: a matched cohort study. Lancet
Haematol. 2021;8(10):e700-e710.
41. Laetsch TW, Maude SL, Balduzzi A, et al. Tisagenlecleucel in pedi­at­ric and
young adult patients with Down syn­drome-asso­ci­ated relapsed/refrac­tory
acute lym­pho­blas­tic leu­ke­mia. Leukemia. 2022;36(6):1508-1515.
42. Mullighan CG, Collins-Underwood JR, Phil­lips LA, et al. Rearrangement of
CRLF2 in B-pro­gen­i­tor- and Down syn­drome-asso­ci­ated acute lym­pho­
blas­tic leu­ke­mia. Nat Genet. 2009;41(11):1243-1246.
8 | Hematology 2022 | ASH Education Program
55. Liu B, Filippi S, Roy A, Roberts I. Stem and pro­gen­i­tor cell dys­func­tion in
human tri­so­mies. EMBO Rep. 2015;16(1):44-62.
56. Muskens IS, Li S, Jackson T, et al. The genome-wide impact of tri­somy 21
on DNA meth­yl­a­tion and its impli­ca­tions for hema­to­poi­e­sis. Nat Commun.
2021;12(1):821.
57. Grimm J, Heckl D, Klusmann JH. Molecular mech­a­nisms of the genetic pre­
dis­po­si­tion to acute megakaryoblastic leu­ke­mia in infants with Down syn­
drome. Front Oncol. 2021;11(June):636633.
58. Birger Y, Goldberg L, Chlon TM, et al. Perturbation of fetal hema­to­poi­e­sis in
a mouse model of Down syn­drome’s tran­sient mye­lo­pro­lif­er­a­tive dis­or­der.
Blood. 2013;122(6):988-998.
59. Klusmann JH, Li Z, Böhmer K, et al. miR-125b-2 is a poten­tial oncomiR on
human chro­mo­some 21 in megakaryoblastic leu­ke­mia. Genes Dev. 2010;
24(5):478-490.
60. Alejo-Valle O, Weigert K, Bhayadia R, et al. The mega­
kar­
yo­
cytic tran­
scrip­tion fac­tor ARID3A suppresses leu­ke­mia path­o­gen­e­sis. Blood. 2022;
139(5):651-665.
61. Wagenblast E, Araújo J, Gan OI, et al. Mapping the cel­lu­lar ori­gin and early
evo­lu­tion of leu­ke­mia in Down syn­drome. Science. 2021;373(6551).
62. Laurent AP, Kotecha RS, Malinge S. Gain of chro­mo­some 21 in hema­to­log­
i­cal malig­nan­cies: les­sons from study­ing leu­ke­mia in chil­dren with Down
syn­drome. Leukemia. 2020;34(8):1984-1999.
63. Bhansali RS, Rammohan M, Lee P, et al. DYRK1A reg­u­lates B cell acute lym­
pho­blas­tic leu­ke­mia through phos­phor­y­la­tion of FOXO1 and STAT3. J Clin
Invest. 2021;131(1):e135937.
64. Lane AA, Chapuy B, Lin CY, et al. Triplication of a 21q22 region con­trib­utes
to B cell trans­for­ma­tion through HMGN1 overexpression and loss of his­tone
H3 Lys27 trimethylation. Nat Genet. 2014;46(6):618-623.
65. Mowery CT, Reyes JM, Cabal-Hierro L, et al. Trisomy of a Down syn­drome
crit­i­cal region glob­ally amplifies tran­scrip­tion via HMGN1 overexpression.
Cell Rep. 2018;25(7):1898-1911.e51911e5.
66. Hasaart KAL, Bertrums EJM, Manders F, Goemans BF, van Boxtel R. Increased
risk of leu­kae­mia in chil­dren with Down syn­drome: a somatic evo­lu­tion­ary
view. Expert Rev Mol Med. 2021;23(April):e5.
© 2022 by The Amer­i­can Society of Hematology
DOI 10.1182/hema­tol­ogy.2022000395
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/1/2021737/1roberts.pdf by guest on 09 December 2022
43. Russell LJ, Capasso M, Vater I, et al. Deregulated expres­sion of cyto­kine
recep­tor gene, CRLF2, is involved in lym­phoid trans­for­ma­tion in B-cell pre­
cur­sor acute lym­pho­blas­tic leu­ke­mia. Blood. 2009;114(13):2688-2698.
44. Hertzberg L, Vendramini E, Ganmore I, et al. Down syn­drome acute lym­
pho­blas­tic leu­ke­mia, a highly het­ero­ge­neous dis­ease in which aber­rant
expres­sion of CRLF2 is asso­ci­ated with mutated JAK2: a report from the
International BFM Study Group. Blood. 2010;115(5):1006-1017.
45. Schwartzman O, Savino AM, Gombert M, et al. Suppressors and acti­va­tors
of JAK-STAT sig­nal­ing at diag­no­sis and relapse of acute lym­pho­blas­tic leu­
ke­mia in Down syn­drome. Proc Natl Acad Sci USA. 2017;114(20):E4030E4039.
46. Nikolaev SI, Garieri M, Santoni F, et al. Frequent cases of RAS-mutated
Down syn­drome acute lym­pho­blas­tic leu­kae­mia lack JAK2 muta­tions. Nat
Commun. 2014;5(August):4654.
47. Brown AL, de Smith AJ, Gant VU, et al. Inherited genetic sus­cep­ti­bil­ity to
acute lym­pho­blas­tic leu­ke­mia in Down syn­drome. Blood. 2019;134(15):12271237.
48. Paulsson K, Forestier E, Lilljebjörn H, et al. Genetic land­
scape of high
hyperdiploid child­hood acute lym­pho­blas­tic leu­ke­mia. Proc Natl Acad Sci
USA. 2010;107(50):21719-21724.
49. Kubota Y, Uryu K, Ito T, et al. Integrated genetic and epi­ge­netic anal­y­
sis revealed het­ero­ge­ne­ity of acute lym­pho­blas­tic leu­ke­mia in Down syn­
drome. Cancer Sci. 2019;110(10):3358-3367.
50. Schwab C, Harrison CJ. Advances in B-cell pre­cur­sor acute lym­pho­blas­tic
leu­ke­mia geno­mics. Hemasphere. 2018;2(4):e53.
51. Abbasi MR, Nebral K, Haslinger S, et al. Copy num­ber changes and allele
dis­tri­bu­tion pat­terns of chro­mo­some 21 in B cell pre­cur­sor acute lym­pho­
blas­tic leu­ke­mia. Cancers (Basel). 2021;13(18):4597.
52. Jardine L, Webb S, Goh I, et al. Blood and immune devel­op­ment in human
fetal bone mar­
row and Down syn­
drome. Nature. 2021;598(7880):327331.
53. MacLean GA, McEldoon J, Huang J et al. Downregulation of endothelin
recep­tor B con­trib­utes to defec­tive B cell lymphopoiesis in tri­somy 21 plu­
rip­o­tent stem cells. Sci Rep. 2018;8(1):8001.
54. Banno K, Omori S, Hirata K, et al. Systematic cel­lu­lar dis­ease mod­els reveal
syn­er­gis­tic inter­ac­tion of tri­somy 21 and GATA1 muta­tions in hema­to­poi­etic
abnor­mal­i­ties. Cell Rep. 2016;15(6):1228-1241.
ACUTE MYELOID LEUKEMIA: IMPROVING OUTCOMES IN CHALLENGING SUBSETS
Achieving MRD negativity in AML: how important
is this and how do we get there?
Laboratory of Myeloid Malignancies, Hematology Branch, National Heart, Lung, and Blood Institute, National Institutes of Health, Bethesda, MD
Multiple studies have demonstrated that patients with acute myeloid leukemia (AML) who have measurable residual disease (MRD) detected during or after treatment have higher relapse rates and worse survival than similar patients testing
negative. Updated response criteria for AML reflect the understanding that achievement of complete remission (CR) with
no detectable MRD using high-sensitivity tests represents a superior response over conventional cytomorphological CR
alone. Potential use cases for AML MRD testing are diverse and include patient selection for clinical trials and therapeutic
assignment, early relapse detection and intervention during sequential monitoring, and drug development, including
deep quantification for antileukemia efficacy and as a surrogate endpoint for overall survival in regulatory approvals.
Testing for AML MRD has not, however, been harmonized, and many technical and clinical questions remain. The implications of MRD test results for specific therapeutic combinations, molecular subsets, test types, treatment time points,
sample types, and patient characteristics remain incompletely defined. No perfect AML MRD test or testing strategy currently exists, and the evidence basis for clinical recommendations in this rare disease is sparse but growing. It is unproven
whether conversion of an MRD test result from positive to negative by additional therapeutic intervention improves
relapse risk and survival. Several national- and international-level consortia have recently been initiated to advance the
generation and collection of evidence to support the use of AML MRD testing in clinical practice, drug development, and
regulatory approvals.
LEARNING OBJECTIVES
• Explain why the updated AML response criteria now include a best possible response of CRMRD−
• Describe how tests validated for diagnostic profiling purposes only are likely insufficient for use in MRD
• Understand the objectives of ongoing national and international collaborative efforts to validate AML MRD testing
for multiple purposes
Introduction
CLINICAL CASE
A man in his 70s with significant medical comorbidities
presents to your clinic with AML. Flow cytometry dem­
onstrates abnormal CD34+ blasts in both blood (1%) and
marrow (~5%). Metaphase cytogenetics are reported
as 47XY, +8, inv(16)(p13.1q22), del(20)(q11.2q13.3), and
next generation sequencing (NGS) by a “myeloid panel”
reported 2 mutations in DNMT3A and 1 in TET2. You are
delighted when he achieves a complete remission (CR) by
cytomorphology after 1 cycle of treatment until your med­
ical student asks, “But what about measurable residual
disease [MRD]?” Your institution has the tests described
in the above diagnostic workup available. How, and why,
should you measure MRD?
Detectable disease after treatment is, by definition, refrac­
tory therapy­resistant disease. Patients and their doctors
generally strongly prefer no evidence of residual cancer.
With a blood cancer like AML, which is typically widely
disseminated throughout the body at initial diagnosis,
the key issue in assessing posttreatment response is how
accurately the evaluation of a small sample of the patient
reflects the total remaining leukemic burden in the body
with the capacity to lead to a subsequent clinically evident
“relapse.” An insufficient sample and/or suboptimal assess­
ment of a sample from a patient after treatment may lead
to false reassurance that a patient in “complete remission”
has achieved a state of disease clearance, a conceit quickly
shattered by the “relapse” that follows. While it has been
stated that the most pressing problem in AML is relapse,
Dr Prakash Singh Shekhawat
MRD testing in AML 2022 | 9
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/9/2021982/9hourigan.pdf by guest on 09 December 2022
Christopher S. Hourigan
When, where, and with what?
Given the strong evi­dence that test­ing for MRD in patients with
AML in remis­sion can strat­ify them into groups with higher and
lower risks of relapse and sur­vival, there is great appeal in devel­
op­ing a uni­ver­sal “best” sin­gle test by which to mon­i­tor AML
MRD. This poten­tially unre­al­is­tic aspi­ra­tion may be moti­vated by
the exam­ples of great suc­cess in MRD mon­i­tor­ing for other blood
can­cers, such as chronic mye­loid leu­ke­mia, acute promyelocytic
leu­ke­mia (both enti­ties with a sin­gle patho­gno­monic struc­tural
var­i­ant expressed at the tran­script level for track­ing by quan­
ti­ta­tive poly­mer­ase chain reac­tion), and the lym­phoid malig­
nan­cies and mul­ti­ple mye­loma (in which a dis­tinct cell sur­face
immunophenotype and a clon­ally rearranged immune recep­tor
pro­vide 2 ideal options for MRD mon­i­tor­ing). In con­trast, AML is
a name given to a genet­i­cally het­ero­ge­neous col­lec­tion of mye­
Table 1. Potential use cases for AML MRD
• Deep quan­ti­fi­ca­tion of antileukemia effi­cacy (eg, log reduc­tion after
2 cycles of ther­apy)
• Early relapse detec­tion and inter­ven­tion dur­ing sequen­tial
mon­i­tor­ing
• Therapeutic assign­ment (eg, selec­tion of trans­plant inten­sity where
oth­er­wise equi­poise)
• Patient selec­tion for clin­i­cal tri­als (eg, high-risk group of unmet
need)
• As a sur­ro­gate endpoint for over­all sur­vival for reg­u­la­tory approval
10 | Hematology 2022 | ASH Education Program
Figure 1. Some validated molecular targets for AML MRD testing. The frequency and co-occurrence of those molecular tar­
gets with evidence of utility for AML MRD testing, based on 200
adult AML patients from the Cancer Genome Atlas database.
loid malig­nan­cies, which can have chang­ing clonal pro­por­tions
even within 1 patient over time. While mul­ti­ple indi­vid­ual tar­gets
for the molec­u­lar mon­i­tor­ing of AML MRD have been described
(Figure 1), there remains great inter­
est in a “one-size-fits-all­
”
approach to MRD mon­i­tor­ing in this diverse set of blood can­
cers using, for exam­ple, mul­ti­pa­ram­e­ter flow cytom­e­try or NGS.
Currently, there is no “one best test” for all­cases of AML MRD.
Flow cytometry is widely avail­­able, is used rou­tinely in the
ini­tial diag­nos­tic workup, and has a poten­tially rapid turn­around
time. Limitations include the need for highly spe­cial­ized expert
inter­pre­ta­tion for best results, dif­fi­cul­ties in test har­mo­ni­za­tion
when not performed and, in par­
tic­
u­
lar, ana­
lyzed centrally,3,11
and sub­op­ti­mal relapse risk pre­dic­tion even in the best cen­ters
(a recent study of 743 con­sec­u­tive adults under­go­ing their first
alloHCT for AML in remis­sion at a sin­gle cen­ter showed that pre­
transplant flow cytom­e­try iden­ti­fied just 96 of 230 sub­se­quent
relapses).10,16 The ELN guide­lines cur­rently state that flow cytom­
e­try for AML MRD should be used when a val­i­dated molec­u­lar
test is not avail­­able.2,3
NGS of DNA is also widely avail­­able and used rou­tinely in the
ini­tial diag­nos­tic workup. A lon­ger turn­around time than other
meth­ods is bal­anced by objec­tive out­put that allows for a decen­
tralized inter­pre­ta­tion.17 NGS is not cur­rently ELN recommended
for detecting AML MRD as a stand-alone test, how­ever, due to
insuf­fi­cient data on appro­pri­ate tar­gets, per­for­mance char­ac­ter­
is­tics of opti­mal test­ing, and clin­i­cal util­ity.18 It is already clear
that the full spec­trum of somatic muta­tions detected at ini­tial
AML diag­no­sis are not all­use­ful as AML MRD tar­gets.2,12,19,20 The
con­cor­dance of AML MRD test­ing using flow cytom­e­try and NGS
has been observed to be incom­plete,19,21 with many poten­tial
expla­na­tions (Table 2). The remark­able oppor­tu­ni­ties presented
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/9/2021982/9hourigan.pdf by guest on 09 December 2022
a very rea­son­able coun­ter­ar­gu­ment is that the most press­ing
prob­lem is instead that AML ther­apy is sub­op­ti­mal and appears
ade­quate only due to insuf­fi­ciently strin­gent response cri­te­ria.
In rec­og­ni­tion of this, the Euro­pean LeukemiaNet (ELN) in 2017
updated the response cri­te­ria in AML with the addi­tion of a best
pos­si­ble response cat­e­gory of com­plete remis­sion with­out MRD
(CRMRD−).1
The case for MRD assess­
ments in AML response cri­
te­
ria is
clear, even if the logis­tics asso­ci­ated with wide­spread har­mo­
nized clin­i­cal implementation remain to be defined.2,3 For 40 years
the con­cept that MRD in AML exists and may be detect­able and
treat­able has been well under­stood.4-6 In the past 5 years, sys­
tem­atic reviews with meta-ana­ly­ses have dem­on­strated that,
regard­less of the MRD meth­od­ol­ogy used, patients test­ing pos­
i­tive, either at a spe­cific time point prior to allo­ge­neic hema­to­
poi­etic cell trans­plan­ta­tion (alloHCT) or more gen­er­ally at any
time dur­ing treat­ment,7,8 have worse sur­vival than those who test
MRD neg­a­tive. This strat­i­fi­ca­tion at the cohort level can assign
patients with AML in CR after treat­ment into groups with higher
and lower risks of sub­se­quent relapse. However, when cur­rently
mea­sured at a sin­gle time point, MRD test­ing is sub­op­ti­mal at an
indi­vid­ual patient level for relapse pre­dic­tion—par­tic­u­larly, per­
haps, when base­line patient and dis­ease char­ac­ter­is­tics are con­
sid­ered.9-12 Nevertheless, given the addi­tional insight pro­vided
there is increas­ing inter­est in using AML MRD test­ing results for a
vari­ety of poten­tial clin­i­cal use cases (Table 1).13-15 Achieving MRD
neg­a­tiv­ity in AML is an impor­tant sig­ni­fier of response to ini­tial
treat­ment, but addi­tional evi­dence is required to deter­mine if this
is the most appro­pri­ate goal and, if so, the best ways to get there
both in an indi­vid­ual patient and as an inter­na­tional com­mu­nity.
Table 2. Potential rea­sons for dis­crep­ancy between AML MRD test­ing meth­ods
AML has no appro­pri­ate muta­tion for track­ing (rare)
NGS test does not include muta­tion of inter­est (com­mon)
• Core bind­ing fac­tor leu­ke­mia with­out KIT or other RTK muta­tion
• Complex kar­yo­type AML with­out TP53 muta­tion
• Chromosomal aneu­ploidy or struc­tural var­i­ants
NGS test is insuf­fi­cient sen­si­tiv­ity/not val­i­dated for MRD tar­get (com­mon)
NGS true pos­i­tive, flow cytom­e­try false neg­a­tive
Lack of tar­get (No LAIP/insuf­fi­cient DfN) (rare)
Limit of detec­tion tech­nique mis­match (eg, NPM1-mutated AML) (com­mon)
Heterogeneous or unsta­ble immunophenotype, dif­fer­en­ti­ated cells after ther­apy (com­mon)
NGS false pos­i­tive, flow cytom­e­try true neg­a­tive
Technical error (NGS test not val­i­dated for MRD use)
Inappropriate NGS tar­get selec­tion (eg, iso­lated DNMT3A muta­tion)
Mutation germ line or pres­ent only in cells inca­pa­ble of caus­ing relapse (eg, lym­pho­cytes)
Flow cytometry false pos­i­tive, NGS true neg­a­tive
Regeneration after che­mo­ther­apy or allo­ge­neic trans­plant
DfN, dif­fer­ence from nor­mal; LAIP, leu­ke­mia-asso­ci­ated immunophenotype; RTK, recep­tor tyro­sine kinase.
Figure 2. Comparison of performance of tests used for AML diagnosis vs MRD. Cartoon approximation of wide differences in target
coverage (ie, number of features tracked, “breadth”) vs detection limit (ie, analytical sensitivity, “depth”) between tests validated for
use at initial diagnosis (blue) vs MRD tests in use or development. dPCR, digital PCR; UMI, unique molecular identifier.
by NGS mean it will almost cer­tainly play a large role in AML MRD
as appro­pri­ate tar­gets are val­i­dated and the per­for­mance char­
ac­ter­is­tics of dif­fer­ent approaches are bet­ter under­stood and
opti­mized (Figure 2). Other poten­tial forms of NGS, includ­ing
cell-free DNA, tran­script expres­sion, and meth­yl­a­tion sig­na­tures,
do not yet have suf­fi­cient evi­dence for AML MRD mon­i­tor­ing.
RNA-based NGS for the ELN-recommended AML MRD molec­u­
lar tar­gets has been described.22 Single-cell sequenc­ing has the
poten­tial to elu­ci­date clonal struc­ture at diag­no­sis to deter­mine
fea­tures asso­ci­ated spe­cif­i­cally with the malig­nant clone, includ­
ing linking geno­type with immunophenotype (the dif­fer­en­ti­a­tion
state of the cell with a detected muta­tion may have impli­ca­tions
for both ther­apy resis­tance and the abil­ity to lead to relapse).23-25
Beyond the spe­cific details of the cur­rent and future tech­nol­
ogy used for this pur­pose, an opti­mal AML MRD test­ing strat­egy for
relapse pre­dic­tion requires the opti­mi­za­tion of mul­ti­ple fac­tors,
includ­ing inter­vals between test­ing time points, the type (mar­
row vs blood) and amount of sam­ple tested, and an account­ing
for AML dis­ease biol­ogy and kinet­ics, as well as patient char­ac­
ter­is­tics that include age, ante­ced­ent dis­or­ders, time points of
treat­ment, and the inten­sity and nature of ther­apy.26-29
Is MRD clin­i­cally action­able, or does it just por­tend fate?
AML patients in a con­ven­tional, cytomorphological CR have a
higher risk of relapse and worse sur­vival if evi­dence of resid­ual
dis­ease is detected com­pared to those who test neg­a­tive—that
is, AML MRD test pos­i­tiv­ity is prog­nos­tic. Such patients are high
risk, are under­served by the cur­rent stan­dard of care, and should
be offered a clin­i­cal trial where pos­si­ble. It is, how­ever, impor­tant
to also say that MRD test neg­a­tiv­ity does not equal patient MRD
neg­a­tiv­ity. Patients test­ing MRD neg­a­tive by flow cytom­e­try
prior to alloHCT still have a 20% to 30% relapse rate,30 and dein­
tensification of stan­dard treat­ment based on an MRD test result
should only be attempted cau­tiously as part of a clin­i­cal trial.
Patients test­ing MRD-pos­i­tive may relapse, may die of a com­
pet­ing risk before relapse, be false pos­i­tive for tech­ni­cal (assay
Dr Prakash Singh Shekhawat
MRD test­ing in AML 2022 | 11
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/9/2021982/9hourigan.pdf by guest on 09 December 2022
Flow cytometry true pos­i­tive, NGS false neg­a­tive
autol­o­gous trans­plant if MRD neg­a­tive, and saw no dif­fer­ence
between groups, suggesting a ben­e­fit to treat­ment inten­si­fi­
ca­tion.34 Understanding if an MRD test result is just fate (ie, a
prog­nos­tic bio­marker) or if treat­ment mod­i­fi­ca­tion can change
out­comes (ie, also a pre­dic­tive bio­marker), and if so in which
patients and AML types, is cur­rently one of the most impor­tant
ques­tions in AML. Observational and reg­is­try stud­ies are poorly
suited to answer this ques­tion, high­light­ing the need to gen­er­
ate higher-qual­ity evi­dence for MRD in AML.
Recent ini­tia­tives gen­er­at­ing AML MRD evi­dence
AML is a rare dis­ease, fur­ther divided into sub­sets with diverse
genetic eti­­ol­o­gies and prog­no­ses and, increas­ingly, with mul­ti­
ple treat­ment alter­na­tives. In this con­text, sin­gle inves­ti­ga­tors
or insti­tu­tions can make only lim­ited gen­er­al­iz­able con­tri­bu­tions
to the evi­dence base supporting AML MRD test uses, moti­vat­ing
the for­ma­tion of sev­eral large national or mul­ti­na­tional coop­er­a­
tive efforts (Table 3).
Alongside updating the assess­ment cri­te­ria after AML treat­
ment in 2017 by includ­
ing a new, best pos­
si­
ble response of
CRMRD−,1 the ELN in 2016 also established a panel of inter­na­tional
experts (ini­tially 24 from 20 countries, includ­ing lab­o­ra­tory sci­
en­tists, pathol­o­gists, and leu­ke­mia phy­si­cians) to rec­om­mend
lab­o­ra­tory and clin­i­cal guide­lines for the use of AML MRD test­
ing. The first edi­tion of these con­sen­sus stan­dard of care guide­
lines was published in 2018 and pro­vided tech­ni­cal guid­ance for
performing molec­u­lar and flow cytom­e­try–based MRD test­ing
along with clin­i­cal rec­om­men­da­tions, includ­ing that AML with
mutated NPM1,35,36 core-bind­ing fac­tor AMLs, acute promyelo­
cytic leu­ke­mia, and the rare cases of AML with BCR-ABL1 should
be mon­i­tored with a val­i­dated molec­u­lar test, while all­ oth­ers
should be mon­i­tored by flow cytom­e­try.3 Updated guide­lines
were published in 2021, with a focus on the avail­­able evi­dence
base under­ly­ing these expert rec­om­men­da­tions and opti­mized
con­sen­sus gen­er­a­tion using a two-stage Delphi poll approach.2
This group, recently renamed ELN-DAVID, will likely con­tinue to
release updated con­sen­sus rec­om­men­da­tions every 2 to 4 years
as new high-qual­ity evi­dence becomes avail­­able.
Table 3. Some recent ini­tia­tives gen­er­at­ing evi­dence for MRD test­ing in AML
Initiative
Goal
Membership
ELN AML MRD guide­lines: Euro­pean
LeukemiaNet
Evidence-based clin­i­cal stan­dard of care
con­sen­sus guide­lines for AML MRD test­ing
International com­mit­tee of phy­si­cians and sci­en­tists
with exper­tise in AML MRD
MPAACT: Measurable Residual Disease
Partnership and Alliance in Acute
Myeloid Leukemia Clinical Treatment
Industry-led research alli­ance advanc­ing efforts to
estab­lish MRD as a sur­ro­gate endpoint for over­all
sur­vival in the treat­ment of AML
Founded in 2018 by Janssen, Genentech, Novartis,
and Celgene (now Bristol Myers Squibb), with recent
addi­tions of Amgen, AbbVie, and Kronos Bio.40
FNIH: Foundation of the NIH
Biomarkers Consortium for AML MRD
Establish and val­i­date new meth­ods for detecting
MRD in AML, includ­ing a library of ref­er­ence
stan­dards and evi­dence of clin­i­cal util­ity.
FDA, NIH, 2 aca­demic part­ners, and 15-25 pri­vate
sec­tor indus­try part­ners
Pre-MEASURE
NIH-led ret­ro­spec­tive pro­ject on >1000 patients
to deter­mine the impact of pre-alloHCT MRD
test­ing in CR1 blood using ultradeep NGS
In col­lab­o­ra­tion with the CIBMTR
MEASURE: Molecular Evaluation of AML
Patients After Stem Cell Transplant to
Understand Relapse Events
Prospective mul­ti­cen­ter pro­to­col to deter­mine
clin­i­cal util­ity of MRD test­ing in up to 1000 AML
patients under­go­ing alloHCT (NCT05224661)
National Marrow Donor Program, CIBMTR, NIH, with
ini­tially up to 16 US-based high-vol­ume alloHCT
cen­ters
NCI MyeloMATCH
Upcoming national pre­ci­sion med­i­cine mas­ter
pro­to­col for AML. Rapid drug effi­cacy screen­ing
using genetic assign­ments and MRD test­ing.
National Cancer Institute, ECOG-ACRIN, SWOG, The
Alliance, Cana­dian Cancer Trials Group Children’s
Oncology Group
12 | Hematology 2022 | ASH Education Program
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/9/2021982/9hourigan.pdf by guest on 09 December 2022
ana­lyt­i­cal fail­ure) or clin­i­cal (analyte detected but not asso­ci­ated
with relapse risk—eg, wrong tar­get or right tar­get but in the
wrong cel­lu­lar con­text) rea­sons, or may have their relapse risk
reduced by sub­se­quent anti­leu­ke­mic fac­tors (addi­tional ther­apy,
allo­ge­neic or autol­o­gous immune responses). Careful sys­tem­
atic study to under­stand the nature of false-neg­a­tive and falsepos­
i­
tive MRD tests are nec­
es­
sary to allow iter­
a­
tive improve­
ments to MRD tests and MRD test­ing strat­e­gies.
In addi­tion to MRD test­ing in AML being prog­nos­tic, it has
been shown to be pos­si­ble, in some cir­cum­stances, to con­vert
a patient test­ing pos­i­tive to a neg­a­tive test sta­tus by addi­tional
treat­ment (for exam­ple, alloHCT).31,32 The con­ver­sion of MRD test
result sta­tus from pos­i­tive to neg­a­tive, how­ever, does not nec­es­
sar­ily imply a clin­i­cal ben­e­fit in terms of increased over­all sur­vival
(from decreased relapse) or improved qual­ity of life. It is pos­si­
ble to imag­ine a worst-case sce­nario in which addi­tional ther­apy
made the bio­marker test turn neg­a­tive but with increased tox­
ic­ity and no sur­vival ben­e­fit. Some patients with MRD in remis­
sion are incom­pletely treated and have chemo-sen­si­tive dis­ease
left to treat. Alternatively, MRD may reflect the resid­ual chemoresis­tant clone, which may or may not be resis­
tant to novel
agents. Both these pos­si­bil­i­ties are test­able.
A large ret­ro­spec­tive Euro­pean Society for Blood and Mar­
row Transplantation reg­
is­
try suggested that myeloablative
(MAC) rather than reduced-inten­sity con­di­tion­ing (RIC) in youn­
ger patients miti­gated some of the risk asso­ci­ated with pre­
transplant MRD pos­i­tiv­ity.30 The phase 3 ran­dom­ized con­trolled
trial (RCT) BMT-CTN 0901 study (NCT01339910) dem­on­strated
reduced relapse and improved over­all sur­vival for youn­ger adults
in CR ran­dom­ized to MAC rather than RIC; sub­se­quent anal­y­
sis showed the greatest ben­e­fit for those who were NGS pos­
i­tive before con­di­tion­ing.20 The poor out­comes of those who
were MRD-pos­i­tive prior to RIC alloHCT were con­firmed by the
FIGARO trial, an RCT that dem­on­strated that addi­tional cytore­
ductive che­
mo­
ther­
apy prior to RIC (in patients inel­
i­
gi­
ble for
MAC due to age or comorbidity) did not improve out­comes.33
The GIMEMA AML1310 trial (NCT01452646) assigned youn­
ger,
inter­me­di­ate-risk patients who were MRD pos­i­tive to alloHCT, or
CLINICAL CASE (Con­t in­u ed)
Clinical NGS DNA sequenc­
ing “mye­
loid pan­
els” used for AML
diag­nos­tic pro­fil­ing are 10 to 500 times less sen­si­tive than AML
MRD tests and is a poor choice here. Additionally, which somatic
muta­tions detected in remis­sion are most asso­ci­ated with relapse
risk remain to be fully described. There is evi­dence that DNMT3A
and TET2 muta­tions should not be used for MRD test­ing; this was
reinforced by research show­ing that these muta­tions were only
Figure 3. The upcoming NCI Precision Medicine Initiative
MyeloMATCH. Schematic showing potential journey of a patient
enrolled in MyeloMATCH through the protocol tiers.
found in subclones unre­lated to the AML of this patient.24 Flow
cytometry performed to MRD stan­dards is a rea­son­able choice
but is not opti­mal given the pres­ence of a defin­ing molec­u­lar
fea­ture in this inver­sion-16 AML (CBFB-MYH11), quan­ti­fi­able using
a well-val­i­dated test.2,3 The patient achieved MRD neg­a­tiv­ity by
poly­
mer­
ase chain reac­
tion test­
ing and enjoyed a prolonged
remis­sion despite being inel­i­gi­ble for consolidative alloHCT.
Conclusion
Testing AML MRD neg­a­tive is pref­er­a­ble to test­ing pos­i­tive, all­
other fac­tors being equal. The phe­nom­ena of higher-sen­si­tiv­ity
tools allowing refined, but imper­
fect, prog­
nos­
ti­
ca­
tion for
patients with AML in remis­sion have been well described in the
lit­er­a­ture but incom­pletely trans­lated to the clinic. Because the
AML MRD test sta­tus reflects only the sam­ple that was tested, not
the entire patient, false-pos­i­tive and false-neg­a­tive results are
expected and have mul­ti­fac­to­rial causes. On an indi­vid­ual patient
level, AML MRD test sta­tus can help risk-strat­ify but is not a guar­
an­tee of fate; serial MRD mea­sure­ment kinet­ics are likely supe­
rior to sin­gle land­mark assess­ments. The com­ing years will see
the gen­er­a­tion of high-qual­ity evi­dence for the many poten­tialuse cases for AML MRD test­ing and col­lab­o­ra­tion on the first
logis­ti­cal steps toward a har­mo­nized national-level approach for
mea­sur­able patient ben­e­fit.
Acknowledgments
This work was supported by the Intramural Research Program
of the National Heart, Lung, and Blood Institute of the National
Institutes of Health.
The Visual Abstract, Figure 1, and Figure 2 are cour­tesy of Alan
Hoofring and Dr. Laura Dillon. Figure 3 is cour­tesy of Dr. Rich Little.
Conflict-of-inter­est dis­clo­sure
Christopher S. Hourigan: funding: Sellas, Foundation for the Na­
tional Institutes of Health Acute Myeloid Leukemia Measurable
Residual Disease Biomarkers Consortium.
Off-label drug use
Christopher S. Hourigan: nothing to disclose.
Dr Prakash Singh Shekhawat
MRD test­ing in AML 2022 | 13
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/9/2021982/9hourigan.pdf by guest on 09 December 2022
Two impor­tant ini­tia­tives in AML MRD from, or in part­ner­ship
with, the biopharmaceutical indus­try are wor­thy of com­ment.
First, fol­
low­
ing the US Food and Drug Administration (FDA)con­vened Duke Margolis Center Public Meeting on Minimal Resid­
ual Disease as a Surrogate Endpoint in Hematologic Cancer Trials
in 2016, a part­ner­ship of 4 phar­ma­ceu­ti­cal com­pa­nies was formed
(led ini­tially by Sharon McBain, then of Janssen, the part­ner­ship
was for­mal­ized and expanded in 2022 and is now known as
MPAACT) to advance efforts in this area, includ­ing the planned
per­for­mance of a meta-anal­y­sis of clin­i­cal trial data to eval­u­ate
the asso­ci­a­tion of MRD with over­all sur­vival. Complementary to
this, focus­ing on the gen­er­a­tion of new stan­dards, tests, and
data, rather than the anal­y­sis of existing data sets, in early 2022
the Foundation of the National Institutes of Health (NIH) Bio­
markers AML MRD Consortium was launched as a col­lab­o­ra­tion
between pub­lic-sec­tor (NIH, FDA), pri­vate-sec­tor (~20 phar­ma­
ceu­ti­cal, bio­tech­nol­ogy, research, or diag­nos­tic test­ing com­pa­
nies), and aca­demic (Fred Hutch and Dana Farber) part­ners.
AlloHCT is a key ther­a­peu­tic inter­ven­tion to reduce sub­se­
quent relapse risk for many patients with AML in CR, with good
evi­
dence that MRD before trans­
plant is prog­
nos­
tic.7,9,16,20,30,33,37
The NIH-funded Pre-MEASURE study eval­
u­
ated pretransplant
blood sam­ples from 1075 patients transplanted in first remis­sion
at one of 111 Center for International Blood and Marrow Trans­
plant Research (CIBMTR) sites between 2013 and 2019 to estab­
lish the clin­i­cal util­ity of NGS-based AML MRD test­ing for FLT3-ITD
and NPM1 muta­tions.38 Following this ret­ro­spec­tive study, the
National Marrow Donor Program and the CIBMTR spon­sored a
pro­spec­tive pro­to­col, MEASURE, at 16 major US trans­plant cen­
ters to estab­lish a national frame­work for intro­duc­ing MRD test­
ing into the clin­i­cal care of AML patients under­go­ing alloHCT
(https:​­/​­/clinicaltrials​­.gov​­/ct2​­/show​­/NCT05224661).
The National Cancer Institute pre­ci­sion med­i­cine ini­tia­tive
for patients with AML, MyeloMATCH, is also launching offi­cially
in early 2023. This national umbrella trial will test treat­ments
for AML, typ­i­cally in ran­dom­ized phase 2 designs com­par­ing
against the cur­rent best stan­dard of care ther­apy, eval­u­at­ing
early endpoint effi­cacy sig­nals in spe­cific molec­u­lar and clin­i­cal
risk groups. The novel design will assign a unique sin­gle patient
iden­ti­fier upon enroll­ment for ini­tial ther­apy, allowing sub­jects
to be followed through­out their treat­ment jour­ney while par­tici­
pat­ing in up to 4 dif­fer­ent RCTs based on sched­uled reassess­
ments (Figure 3). The intent is to use MRD test­ing as an effi­cacy
endpoint, as an inclu­sion cri­te­rion for sub­se­quent “tiers” of ther­
apy, while also facil­i­tat­ing the val­i­da­tion of novel, highly sen­si­tive
MRD assays such as duplex sequenc­ing.
Finally, reg­u­la­tory guid­ance for the use of MRD, includ­ing in
AML, for drug devel­op­ment has been published and pre­sum­ably
updated based on evi­dence resulting from the ini­tia­tives above.39
Correspondence
Christopher S. Hourigan, Laboratory of Myeloid Malignancies,
Hematology Branch, National Heart, Lung, and Blood Institute,
NIH Clinical Ctr, Room 10CRC 6-5142, 10 Center Dr, Bethesda,
MD 20892; e-mail: hourigan@nih​­.gov.
References
1.
14 | Hematology 2022 | ASH Education Program
DOI 10.1182/hema­tol­ogy.2022000323
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/9/2021982/9hourigan.pdf by guest on 09 December 2022
Döhner H, Estey E, Grimwade D, et al. Diagnosis and man­age­ment of AML
in adults: 2017 ELN rec­om­men­da­tions from an inter­na­tional expert panel.
Blood. 2017;129(4):424-447.
2. Heuser M, Freeman SD, Ossenkoppele GJ, et al. 2021 update mea­sur­able
resid­ual dis­ease in acute mye­loid leu­ke­mia: Euro­pean LeukemiaNet Work­
ing Party con­sen­sus doc­u­ment. Blood. 2021;138(26):2753-2767.
3. Schuurhuis GJ, Heuser M, Freeman S, et al. Minimal/mea­sur­able resid­ual
dis­ease in AML: a con­sen­sus doc­u­ment from the Euro­pean LeukemiaNet
MRD Working Party. Blood. 2018;131(12):1275-1291.
4. Löwenberg B, Hagenbeek A. Minimal Residual Disease in Acute Leukemia.
1st ed. Dordrecht: Springer; 1984.
5. Martens AC, Hagenbeek A. Detection of min­i­mal dis­ease in acute leu­ke­mia
using flow cytom­e­try: stud­ies in a rat model for human acute leu­ke­mia.
Cytometry. 1985;6(4):342-347.
6. Hagenbeek A, Martens AC. BCG treat­ment of resid­ual dis­ease in acute leu­
ke­mia: stud­ies in a rat model for human acute mye­lo­cytic leu­ke­mia (BNML).
Leuk Res. 1983;7(4):547-555.
7. Buckley SA, Wood BL, Othus M, et al. Minimal resid­ual dis­ease prior to
allo­ge­neic hema­to­poi­etic cell trans­plan­ta­tion in acute mye­loid leu­ke­mia: a
meta-anal­y­sis. Haematologica. 2017;102(5):865-873.
8. Short NJ, Zhou S, Fu C, et al. Association of mea­sur­able resid­ual dis­ease
with sur­vival out­comes in patients with acute mye­loid leu­ke­mia: a sys­tem­
atic review and meta-anal­y­sis. JAMA Oncol. 2020;6(12):1890-1899.
9. Othus M, Gale RP, Hourigan CS, Walter RB. Statistics and mea­sur­able resid­
ual dis­ease (MRD) test­ing: uses and abuses in hema­to­poi­etic cell trans­plan­
ta­tion. Bone Marrow Transplant. 2020;55(5):843-850.
10. Othus M, Wood BL, Stirewalt DL, et al. Effect of mea­sur­able (“min­i­mal”)
resid­ual dis­ease (MRD) infor­ma­tion on pre­dic­tion of relapse and sur­vival in
adult acute mye­loid leu­ke­mia. Leukemia. 2016;30(10):2080-2083.
11. Paiva B, Vidriales MB, Sempere A, et al; PETHEMA (Programa para el Estu­
dio de la Terapéutica en Hemopatías Malignas) Cooperative Study Group.
Impact of mea­sur­able resid­ual dis­ease by decentralized flow cytom­e­try: a
PETHEMA real-world study in 1076 patients with acute mye­loid leu­ke­mia.
Leukemia. 2021;35(8):2358-2370.
12. Murdock HM, Kim HT, Denlinger N, et al. Impact of diag­nos­tic genet­ics on
remis­sion MRD and trans­plan­ta­tion out­comes in older patients with AML.
Blood. 2022;139(24):3546-3557.
13. Gui G, Hourigan CS. Measurable resid­ual dis­ease assess­ment as a sur­ro­gate
marker in new drug devel­op­ment in acute mye­loid leu­ke­mia. Cancer J.
2022;28(1):73-77.
14. Spyridonidis A. How I treat mea­sur­able (min­i­mal) resid­ual dis­ease in acute
leu­ke­mia after allo­ge­neic hema­to­poi­etic cell trans­plan­ta­tion. Blood.
2020;135(19):1639-1649.
15. Hourigan CS, Gale RP, Gormley NJ, Ossenkoppele GJ, Walter RB. Mea­
surable resid­ual dis­ease test­ing in acute mye­loid leu­kae­mia. Leukemia.
2017;31(7):1482-1490.
16. Morsink LM, Sandmaier BM, Othus M, et al. Conditioning inten­sity, pretrans­plant flow cytometric mea­sur­able resid­ual dis­ease, and out­come in
adults with acute mye­loid leu­ke­mia under­go­ing allo­ge­neic hema­to­poi­etic
cell trans­plan­ta­tion. Cancers (Basel). 2020;12(9):2339.
17. Heuser M. Standardizing NGS-MRD in AML: an inter­na­tional study by the
ELN-MRD Working Party. Paper presented at: Euro­pean Hematology Asso­
ciation Annual Congress; 2021, vir­tual meet­ing.
18. Ghannam J, Dillon LW, Hourigan CS. Next-gen­er­a­tion sequenc­ing for mea­
sur­able resid­ual dis­ease detec­tion in acute mye­loid leu­kae­mia. Br J Hae­
matol. 2020;188(1):77-85.
19. Jongen-Lavrencic M, Grob T, Hanekamp D, et al. Molecular min­i­mal resid­ual
dis­ease in acute mye­loid leu­ke­mia. N Engl J Med. 2018;378(13):1189-1199.
20. Hourigan CS, Dillon LW, Gui G, et al. Impact of con­di­tion­ing inten­sity of
allo­ge­neic trans­plan­ta­tion for acute mye­loid leu­ke­mia with geno­mic evi­
dence of resid­ual dis­ease. J Clin Oncol. 2020;38(12):1273-1283.
21. Patkar N, Kakirde C, Shaikh AF, et al. Clinical impact of panel-based
error-corrected next gen­er­a­tion sequenc­ing ver­sus flow cytom­e­try to
detect mea­
sur­
able resid­
ual dis­
ease (MRD) in acute mye­
loid leu­
ke­
mia
(AML). Leukemia. 2021;35(5):1392-1404.
22. Dillon LW, Hayati S, Roloff GW, et al. Targeted RNA-sequenc­ing for the
quan­ti­fi­ca­tion of mea­sur­able resid­ual dis­ease in acute mye­loid leu­ke­mia.
Haematologica. 2019;104(2):297-304.
23. Bottomly D, Long N, Schultz AR, et al. Integrative anal­
y­
sis of drug
response and clin­i­cal out­come in acute mye­loid leu­ke­mia. Cancer Cell.
2022;40(8):850-864.e9864e9.
24. Dillon LW, Ghannam J, Nosiri C, et al. Personalized sin­gle-cell proteog­
enomics to dis­tin­guish acute mye­loid leu­ke­mia from non-malig­nant clonal
hema­to­poi­e­sis. Blood Cancer Discov. 2021;2(4):319-325.
25. Miles LA, Bowman RL, Merlinsky TR, et al. Single-cell muta­tion anal­y­sis of
clonal evo­lu­tion in mye­loid malig­nan­cies. Nature. 2020;587(7834):477-482.
26. Pratz KW, Jonas BA, Pullarkat V, et al. Measurable resid­ual dis­ease response
and prog­no­sis in treat­ment-naïve acute mye­loid leu­ke­mia with venetoclax
and azacitidine. J Clin Oncol. 2022;40(8):855-865.
27. Roboz GJ, Ravandi F, Wei AH, et al. Oral azacitidine pro­longs sur­vival of
patients with AML in remis­sion inde­pen­dently of mea­sur­able resid­ual dis­
ease sta­tus. Blood. 2022;139(14):2145-2155.
28. Maiti A, DiNardo CD, Wang SA, et al. Prognostic value of mea­sur­able resid­
ual dis­ease after venetoclax and decitabine in acute mye­loid leu­ke­mia.
Blood Adv. 2021;5(7):1876-1883.
29. Platzbecker U, Middeke JM, Sockel K, et al. Measurable resid­ual dis­easeguided treat­
ment with azacitidine to pre­
vent haematological relapse
in patients with myelodysplastic syn­
drome and acute mye­
loid leu­
kae­
mia (RELAZA2): an open-label, multicentre, phase 2 trial. Lancet Oncol.
2018;19(12):1668-1679.
30. Gilleece MH, Labopin M, Yakoub-Agha I, et al. Measurable resid­ual dis­
ease, con­di­tion­ing reg­i­men inten­sity, and age pre­dict out­come of allo­
ge­neic hema­to­poi­etic cell trans­plan­ta­tion for acute mye­loid leu­ke­mia in
first remis­sion: a reg­is­try anal­y­sis of 2292 patients by the Acute Leukemia
Working Party Euro­pean Society of Blood and Marrow Transplantation. Am
J Hematol. 2018;93(9):1142-1152.
31. Heuser M, Heida B, Büttner K, et al. Posttransplantation MRD mon­i­tor­ing in
patients with AML by next-gen­er­a­tion sequenc­ing using DTA and non-DTA
muta­tions. Blood Adv. 2021;5(9):2294-2304.
32. Paras G, Morsink LM, Othus M, et al. Conditioning inten­sity and peritrans­
plant flow cytometric MRD dynam­ics in adult AML. Blood. 2022;139(11):16941706.
33. Craddock C, Jackson A, Loke J, et al. Augmented reduced-inten­sity reg­i­
men does not improve postallogeneic trans­plant out­comes in acute mye­
loid leu­ke­mia. J Clin Oncol. 2021;39(7):768-778.
34. Venditti A, Piciocchi A, Candoni A, et al. GIMEMA AML1310 trial of riskadapted, MRD-directed ther­apy for young adults with newly diag­nosed
acute mye­loid leu­ke­mia. Blood. 2019;134(12):935-945.
35. Dillon R, Hills R, Freeman S, et al. Molecular MRD sta­tus and out­come after
trans­plan­ta­tion in NPM1-mutated AML. Blood. 2020;135(9):680-688.
36. Kapp-Schwoerer S, Weber D, Corbacioglu A, et al. Impact of gemtuzumab
ozogamicin on MRD and relapse risk in patients with NPM1-mutated AML:
results from the AMLSG 09-09 trial. Blood. 2020;136(26):3041-3050.
37. Thol F, Gabdoulline R, Liebich A, et al. Measurable resid­ual dis­ease mon­
i­tor­ing by NGS before allo­ge­neic hema­to­poi­etic cell trans­plan­ta­tion in
AML. Blood. 2018;132(16):1703-1713.
38. Hourigan CS, Dillon LW, Gui G, et al. Pre-MEASURE: mul­ti­cen­ter eval­u­a­tion
of the prog­nos­tic sig­nif­i­cance of mea­sur­able resid­ual dis­ease test­ing prior
to allo­ge­neic trans­plan­ta­tion for adult patients with AML in first remis­sion.
J Clin Oncol. 2022;40(suppl 16):7006.
39. US Food and Drug Administration. Hematologic malig­nan­cies: reg­u­la­tory
con­sid­er­ations for use of min­i­mal resid­ual dis­ease in devel­op­ment of drug
and bio­log­i­cal prod­ucts for treat­ment. https:​­/​­/www​­.fda​­.gov​­/regulatory​
­-information​­/search​­-fda​­-guidance​­-documents​­/hematologic​­-malignancies​
­-regulatory​­-considerations​­-use​­-minimal​­-residual​­-disease​­-development​
­-drug​­-and.
40. Johnson and Johnson. MPAACT consortium unites industry and academia
to establish measurable residual disease as a surrogate endpoint in acute
myeloid leukemia drug development. https://www.jnj.com/mpaact
-consortium-unites-industry-and-academia-to-establish-measurable
-residual-disease-as-a-surrogate-endpoint-in-acute-myeloid-leukemia
-drug-development.
ACUTE MYELOID LEUKEMIA: IMPROVING OUTCOMES IN CHALLENGING SUBSETS
Kieran D. Sahasrabudhe and Alice S. Mims
The James Cancer Center, The Ohio State University, Columbus, OH
The treatment landscape in acute myeloid leukemia (AML) is rapidly evolving, with multiple new therapies approved in
recent years. However, the prognosis for patients with high-risk genetic subsets of AML remains poor, and the development of more effective treatment options for these patients is ongoing. Three of these high-risk AML patient subsets
include TP53-mutated AML, FLT3-internal tandem duplication (ITD)-mutated AML, and AML harboring rearrangements
affecting the KMT2A locus (KMT2A-r AML). The prognosis for TP53-mutated AML remains poor with both intensive and
targeted regimens, including those incorporating the BCL-2 inhibitor, venetoclax. Allogeneic hematopoietic stem cell
transplantation is the only potentially curative therapy for these patients, but posttransplant relapse rates remain high.
Patients with FLT3-ITD-mutated AML continue to have suboptimal outcomes with standard therapies and experience
high rates of relapse following transplant. KMT2A-r AML is also associated with poor outcomes with current treatment
approaches, and effective standards of care are lacking for patients with relapsed/refractory disease. This article discusses current treatment approaches, along with the investigational agents being explored for the treatment of these 3
AML subsets, focusing primarily on agents that are further along in development.
LEARNING OBJECTIVES
• Recognize the high­risk subsets of TP53­mutated, FLT3­ITD mutated, and KMT2A­rearranged AML, in which better
treatment options are needed
• Understand that patients with TP53­mutated AML have poor outcomes with current standard therapies and
should be treated on clinical trials when possible
• Review the novel FLT3 inhibitors being studied in combination with many other agents in patients with FLT3­
mutated AML
• Learn about the novel inhibitors of the menin­KMT2A interaction that have demonstrated promise in the treatment
of KMT2A­rearranged AML
Current approaches
CLINICAL CASE 1
A 60­year­old man with a history of stage III colon can­
cer treated with surgery and adjuvant chemotherapy
8 years ago develops pancytopenia with circulating blasts.
A bone marrow biopsy reveals acute myeloid leukemia
(AML) with 30% myeloid blasts. Cytogenetic analysis
shows a complex karyotype including monosomy 7. A next
generation sequencing panel reveals a pathogenic TP53
mutation with a variant allele frequency of 30%. He is oth­
erwise feeling well without any functional limitations and
has no other notable comorbidities.
The TP53 protein is a transcription factor and tumor sup­
pressor that regulates many target genes with diverse
functions. The mechanism of action of TP53 along with the
therapeutic agents being investigated in TP53­mutated AML
is shown in Figure 1. TP53 mutations occur in 5% to 10% of
de novo AML and have a higher incidence in therapy­related
AML (t­AML), at 20% to 35%. TP53­mutated AML is also often
associated with a complex karyotype.1,2 Most AML cases har­
boring TP53 derangements involve biallelic inactivation of
the gene. TP53 derangements caused by mutations, dele­
tions, and other disruptions of the 17p13 locus are associated
with a very poor prognosis in AML with current approaches.
High­risk AML novel approaches: TP53, KMT2A, FLT3 | 15
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/15/2021947/15sahasrabudhe.pdf by guest on 09 December 2022
Novel investigational approaches for high-risk
genetic subsets of AML: TP53, KMT2A, FLT3
For patients treated with inten­sive che­mo­ther­apy, com­plete
remis­sion (CR) rates are less than 30%, and median over­all sur­
vival (OS) is less than 1 year.1,3 For patients who respond to induc­
tion ther­apy, sub­se­quent allo­ge­neic hema­to­poi­etic stem cell
trans­plan­ta­tion (allo-HSCT) remains the only poten­tially cura­tive
treat­ment. However, posttransplant out­comes also remain poor,
with a median posttransplant OS of less than 1 year and a 2-year
OS rate of less than 30%.4,5
There has been inter­est in less inten­sive treat­ments, includ­
ing hypomethylating agents (HMAs), in this patient pop­u­la­tion
given the poor prog­no­sis asso­ci­ated with inten­sive che­mo­ther­
apy. One sin­gle-arm study found that 10-day decitabine ther­apy
was asso­ci­ated with favor­able response rates in TP53-mutated
AML.6 However, this study included only 21 patients with TP53
muta­tions. Other stud­ies have found poor OS of less than 1 year
in patients with TP53 muta­tions treated with azacitidine, 5-day
decitabine, and 10-day decitabine.7,8 One of these was a ran­dom­
ized study involv­ing 71 patients, 30% of whom had TP53 muta­
tions, that com­pared 5-day vs 10-day decitabine. Among patients
with TP53 muta­tions, there was no dif­fer­ence in response rates
with 5-day vs 10-day decitabine. Median OS also did not dif­fer
16 | Hematology 2022 | ASH Education Program
sig­nif­i­cantly and was less than 6 months in both groups, indi­cat­
ing no advan­tage of 10-day over 5-day decitabine.
The addi­tion of venetoclax, a BCL-2 inhib­i­tor, to HMA ther­
apy has, unfor­tu­nately, not shown the improve­ment in patient
out­comes to the level seen in other genetic sub­types of AML.
Among patients with TP53-mutated dis­ease treated in the phase
3 VIALE-A study, the com­
bi­
na­
tion of azacitidine-venetoclax
yielded an encour­
ag­
ing CR with an incom­
plete hema­
to­
logic
recov­
ery (CRi) rate of 55% (95% CI, 38.3-71.4) com­
pared to
0% in patients treated with azacitidine plus pla­cebo (P<.001).9
However, a ret­ro­spec­tive review of patients with TP53-mutated
AML and poor-risk cyto­ge­net­ics treated both in VIALE-A and the
phase 1b study pre­ced­ing VIALE-A found that the com­bi­na­tion of
venetoclax plus an HMA did not trans­late into improved dura­tion
of response or OS com­pared to an HMA alone.10 Another study
of 10-day decitabine plus venetoclax in patients with newly
diag­nosed AML found that OS was sig­nif­i­cantly worse in TP53mutated vs wild-type patients, with a median OS of less than
6 months in patients with TP53 muta­tions.11 These results fur­ther
indi­cate the lack of improve­ment with 10-day decitabine with or
with­out venetoclax in this patient pop­u­la­tion. The over­all poor
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/15/2021947/15sahasrabudhe.pdf by guest on 09 December 2022
Figure 1. Mechanism of action of TP53 and agents under investigation for TP53-mutated AML. (A) Mechanism of action of TP53: cel­
lular stress and DNA damage activate the TP53 protein, which leads to downstream cell cycle arrest, DNA repair, and apoptosis. Cells
containing mutations affecting the TP53 protein are not able to respond to stress and DNA damage appropriately, which increases
the risk of neoplastic transformation. (B) Mechanism of action of magrolimab: the binding of CD47 on tumor cells to SIRPα on macro­
phages inhibits phagocytosis of tumor cells. Magrolimab is a monoclonal antibody that binds to CD47 and prevents it from binding
to SIRPα, which thereby facilitates tumor cell phagocytosis. (C) Mechanism of action of APR-246: APR-246 is converted to MQ under
physiological conditions. MQ binds to mutated TP53 and facilitates thermodynamic stabilization of the protein, thereby shifting equi­
librium to the active form of TP53. APR-246 also exhibits TP53-independent activity through glutathione depletion, which increases
lipid peroxidases and other reactive oxygen species, thereby promoting cell death through ferroptosis (not depicted). Created with
BioRender​.com.
out­comes with the cur­rent avail­­able stan­dards of care high­light
the need to develop addi­tional novel ther­a­pies beyond veneto­
clax to improve sur­vival and long-term out­comes in patients with
TP53-mutated AML.
Targeting CD47 in TP53-mutated AML
TP53 sta­bi­li­za­tion in TP53-mutated AML
APR-246 (epranetapopt) acts as a TP53 sta­bi­lizer and is a pro­drug
that con­verts to meth­y­lene quinuclidinone (MQ ) under phys­i­o­
log­i­cal con­di­tions. MQ binds to cys­te­ine res­i­dues in mutant TP53,
which leads to ther­mo­dy­namic sta­bi­li­za­tion of the pro­tein and
shifts equi­lib­rium toward the func­tional con­for­ma­tion (Figure 1C).15
APR-246 has also been shown to deplete glu­ta­thi­one. Glutathione
deple­tion increases lipid per­ox­i­dases and other reac­tive oxy­gen
spe­cies, which increases sus­cep­ti­bil­ity to ferroptosis.16 Ferropto­
sis induc­tion has been shown to be an impor­tant mech­a­nism of
the early leu­ke­mic cell death induced by APR-246 regard­less of
TP53 muta­tion sta­tus.17 A phase 1b/2 study of APR-246–azacitidine
in patients with TP53-mutated myelodysplastic syn­dromes (MDS)
or oligoblastic AML (20%-30% blasts) found an over­all response
rate (ORR) of 71%, with 44% of patients achiev­ing CR.18 Despite
the prom­is­ing results, a sub­se­quent phase 3 trial (NCT03745716)
failed to meet its pri­mary end point of a supe­rior CR rate with
APR-246–azacitidine com­pared to azacitidine alone in patients
with TP53-mutated MDS. However, a sub­se­quent phase 2 study of
APR-246–azacitidine as main­te­nance ther­apy fol­low­ing allo-HSCT
in patients with TP53-mutated AML or MDS showed prom­is­ing
results com­pared to his­tor­i­cal data, with a median relapse-free
sur­vival of 368 days and a median OS of 586 days.19 Another
phase 1 study of the tri­plet com­bi­na­tion of APR-246–venetoclax–
azacitidine for the pri­
mary treat­
ment of TP53-mutated AML
showed a CR rate of 37% and a CR/CRi rate of 53%, meet­ing the
Simon’s 2-stage effi­cacy cri­te­ria.20
The patient presented here is a fit 60-year-old man with TP53mutated, ther­apy-related AML and min­i­mal comorbidities. An
up-front dis­cus­sion regard­ing the lim­ited treat­ment options and
poor prog­no­sis is very impor­tant for shared deci­sion-mak­ing
that aligns with the patient’s treat­ment goals. If pos­si­ble, he
should be referred for a clin­i­cal trial focus­ing on TP53-mutated
AML. A sum­mary of our approach for the treat­ment of newly
diag­nosed TP53-mutated AML is presented in Figure 2.
CLINICAL CASE 2
A 50-year-old man with no major comorbidities and good per­
for­mance sta­tus pres­ents to dis­cuss treat­ment options for AML
diag­nosed in the con­text of fatigue and cir­cu­lat­ing blasts. A
bone mar­
row biopsy dem­
on­
strates 80% blasts. Polymerase
chain reac­tion with cap­il­lary elec­tro­pho­re­sis for FLT3 muta­tions
detects the pres­ence of an FLT3 inter­nal tan­dem dupli­ca­tion
(ITD) muta­tion. The kar­yo­type is nor­mal, and no other muta­
tions are detected on next gen­er­a­tion sequenc­ing.
Background
Mutations in the FMS-like tyro­sine kinase 3 (FLT3) pro­tein are
pres­ent in approx­i­ma­tely 30% of newly diag­nosed AML cases.
These muta­tions are clas­si­fied as ITD muta­tions affect­ing the
juxtamembrane domain or point muta­
tions in the tyro­
sine
kinase domain (TKD muta­tions). FLT3-ITD muta­tions have his­
tor­i­cally been established as a poor prog­nos­tic fac­tor.21 FLT3TKD muta­tions have not dem­on­strated a con­sis­tent prog­nos­tic
impact.22 The 2017 Euro­pean LeukemiaNet prog­nos­tic clas­si­fi­
ca­tion sys­tem sep­a­rated FLT3-ITD-mutated AML into dif­fer­ent
categories based on the ITD alle­lic ratio and the pres­ence of
a con­cur­rent NPM1 muta­tion.23 However, the updated 2022
ELN clas­si­fi­ca­tion includes all­ FLT3-ITD-mutated AML in the
inter­me­di­ate risk cat­e­gory.24 The cur­rent stan­dard of care for
most patients with FLT3-ITD muta­tions who are can­di­dates is
to undergo allo-HSCT in first CR. However, patients with FLT3ITD muta­tions are also at increased risk of relapse fol­low­ing
trans­plant.25 Recent efforts in FLT3-mutated AML have focused
on devel­op­ing novel FLT3 inhib­i­tors, com­bin­ing FLT3 inhib­i­
tors with other ther­a­pies to over­come ther­a­peu­tic resis­tance,
expanding treat­ment options for patients with relapsed/refrac­
tory (R/R) dis­ease, and study­ing posttransplant main­te­nance
ther­a­pies to reduce the risk of posttransplant relapse. The path­
o­gen­e­sis of FLT3-mutated AML and the mech­a­nism of action of
FLT3 inhib­i­tors is shown in Figure 3A.
Current approaches
The cur­rent stan­dard front­line ther­apy for patients with newly
diag­nosed FLT3-mutated AML who are eli­gi­ble for inten­sive che­
mo­ther­apy is induc­tion ther­apy with midostaurin com­bined with
tra­di­tional “7+3” followed by con­sol­i­da­tion with midostaurincytarabine. Midostaurin is a first-gen­er­a­tion type 1 FLT3 inhib­
i­
tor with activ­
ity against both ITD- and TKD-mutated recep­
tors. This approach was informed by the phase 3 RATIFY trial
in which midostaurin outperformed pla­
cebo when com­
bined
High-risk AML novel approaches: TP53, KMT2A, FLT3 | 17
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/15/2021947/15sahasrabudhe.pdf by guest on 09 December 2022
The CD47 pro­
tein is overexpressed in mye­
loid malig­
nan­
cies,
includ­ing in AML leukemic stem cells. The CD47 pro­tein acts as
a “do not eat me” sig­nal that inhib­its phago­cy­to­sis of tumor
cells by bind­ing to the sig­nal reg­u­la­tory pro­tein alpha on mac­
ro­phages. In pre­clin­i­cal stud­ies, CD47 block­ade over­comes this
inhi­bi­tion and stim­u­lates tumor cell kill­ing by mac­ro­phages
(Figure 1B).12 A phase 1b study com­bin­ing the CD47 anti­body mag­
rolimab with azacitidine for the treat­ment of patients with newly
diag­nosed AML inel­i­gi­ble for inten­sive che­mo­ther­apy showed
prom­is­ing effi­cacy in both TP53-mutated and TP53-wild-type
patients. The objec­tive response rate was 65% for all­com­ers
and 71% for TP53-mutated patients. Median OS was 18.9 months
for TP53-wild-type patients and 12.9 months for TP53-mutated
patients.13 A tri­plet com­bi­na­tion of magrolimab-azacitidinevenetoclax was sub­se­quently stud­ied in a phase 1b/2 clin­i­cal
trial. In patients with newly diag­nosed AML, the CR/CRi rate was
94% for all­patients and 100% for patients with TP53-mutated dis­
ease.14 The most com­mon grade 3/4 adverse events were pneu­
mo­nia, febrile neutropenia, hyperbilirubinemia, ele­vated ala­nine
ami­no­trans­fer­ase, and skin infec­tion. CD47 is expressed on red
blood cells, and it is com­mon to see a drop in hemo­glo­bin, par­
tic­u­larly early in the course of treat­ment. The ongo­ing phase 3
tri­als Enhance-2 (NCT04778397) and Enhance-3 (NCT05079230)
are fur­ther addressing the role of magrolimab in TP53-mutated
and TP53-wild-type dis­ease, respec­tively. Other CD47 antibod­
ies are cur­rently in devel­op­ment in ear­lier-phase clin­i­cal tri­als.
CLINICAL CASE 1 (Con­tin­ued)
with ­che­mo­ther­apy in patients aged 18 to 59 (haz­ard ratio for
death, 0.78; one-sided P=.009).26 Though the RATIFY study only
included patients aged 18 to 59, there is no age lim­i­ta­tion on the
US Food and Drug Administration’s approval.
For patients who are inel­i­gi­ble for inten­sive che­mo­ther­apy,
the com­bi­na­tion of venetoclax plus HMA ther­apy can be given
front­
line. The com­
pos­
ite CR rate was 72.4% in patients with
FLT3-mutated AML treated with front­line venetoclax-azacitidine
on the VIALE-A trial.9
The cur­
rent stan­
dard of care for patients with R/R FLT3mutated AML is monotherapy with the sec­ond-gen­er­a­tion type 1
FLT3 inhib­i­tor gilteritinib. This prac­tice was based on results of the
phase 3 Admiral trial, in which gilteritinib was found to be supe­
rior to stan­dard of care che­mo­ther­apy in this set­ting (median OS,
9.3 months vs 5.6 months; haz­ard ratio, 0.64; 95% CI, 0.49-0.83;
P<.001).27 For patients with FLT3-ITD-mutated AML who undergo
allo-HSCT, sorafenib (a type 2 FLT3 inhib­i­tor active against ITDmutated but not TKD-mutated AML) is some­times given off-label
18 | Hematology 2022 | ASH Education Program
as posttransplant main­te­nance ther­apy based on the results of 2
clin­i­cal tri­als, one of which was pla­cebo con­trolled.28,29
Investigational approaches
Intensive che­mo­ther­apy can­di­dates
Several clin­
i­
cal tri­
als are ongo­
ing involv­
ing gilteritinib in
patients with FLT3-mutated AML, many of which com­bine gil­
teritinib with inten­sive che­mo­ther­apy. A phase 1 study com­
bin­ing gilteritinib with 7 + 3 was well tol­er­ated and asso­ci­ated
with a FLT3 muta­tion clear­ance rate of 70%.30 Additional FLT3
inhib­i­tors are also under inves­ti­ga­tion. Quizartinib is a sec­ondgen­er­a­tion type 2 inhib­i­tor that has been com­pared to pla­cebo
in com­bi­na­tion with 7 + 3 in patients with newly diag­nosed, FLT3ITD-mutated AML in the Quantum First trial (NCT02668653).
At a median fol­low-up of 39.2 months, OS was sig­nif­i­cantly
lon­
ger in the quizartinib arm (31.9 months) vs the pla­
cebo
arm (15.1 months; haz­
ard ratio, 0.776; 95% CI, 0.615-0.979;
2-sided P = .0324). CR/CRi rates were 71.6% in the quizartinib
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/15/2021947/15sahasrabudhe.pdf by guest on 09 December 2022
Figure 2. Treatment approach for newly diagnosed TP53-mutated AML.
#
Many clinicians prefer less intensive regimens to reduce toxicity, even in patients who are eligible, given the poor outcomes associ­
ated with intensive chemotherapy. Robust data supporting this practice are lacking.
* Adding venetoclax to the treatment regimen may improve the likelihood of response, but mounting data suggest that this is unlikely
to improve survival.
^ Even though allo-HSCT is the only potentially curative therapy, posttransplant outcomes remain poor. Consequently, the risk/
benefit assessment ends up not favoring allo-HSCT for many patients.
Created with BioRender​.com.
arm vs 64.9% in the pla­cebo arm.31 It is nota­ble that this trial
included patients aged 18 to 75 com­pared to the RATIFY trial,
which had an upper age cut­off of 59. Crenolanib is a sec­ondgen­er­a­tion, type 1 FLT3 inhib­i­tor that has also dem­on­strated
prom­
is­
ing results in early-phase tri­
als.32,33 FLT3 inhib­i­tors
are also being inves­
ti­
gated as main­
te­
nance ther­
apy post
trans­
plant for patients under­
go­
ing allo-HSCT. Posttrans­
plant midostaurin was found to be asso­
ci­
ated with supe­
rior event-free sur­
vival com­
pared to a his­
toric con­
trol in 1
study.34 Another study found that posttransplant midostau­
rin improved relapse-free sur­
vival com­
pared to stan­
dard of
care only when FLT3 phos­
phor­
y­
la­
tion was inhibited to less
than 70% of base­
line.35 Gilteritinib is also being com­
pared
to pla­cebo as main­te­nance ther­apy fol­low­ing allo-HSCT through
the Blood and Marrow Transplant Clinical Trials Network Pro­
tocol 1506 (NCT02997202) and fol­
low­
ing che­
mo­
ther­
apy in
patients who are not under­go­ing allo-HSCT (NCT02927262).
Nonintensive che­mo­ther­apy can­di­dates
For those patients who are not can­di­dates for inten­sive che­mo­
ther­apy, gilteritinib is also being stud­ied in com­bi­na­tion with
HMA ther­apy and other targeted ther­a­pies. The recently com­
pleted phase 3 LACEWING trial com­pared the com­bi­na­tion of
gilteritinib-azacitidine to azacitidine alone in patients with newly
diag­nosed FLT3-mutated AML who were inel­i­gi­ble for inten­sive
che­mo­ther­apy. The com­bi­na­tion arm yielded supe­rior response
rates but not supe­rior OS, and the study there­fore did not meet
its pri­mary end point.36
There is evi­dence that BCL-2 upregulation may play a role in
resis­tance to FLT3 inhi­bi­tion and that adding venetoclax may help
to over­come this.37 A phase 1b study of gilteritinib-­venetoclax
in patients with R/R dis­ease found a favor­able ORR of 90% in
patients with FLT3-mutated dis­ease.38 An ongo­
ing phase 1/2
study (NCT04140487) of gilteritinib-venetoclax-azacitidine found
an ORR of 67% with a median OS of 10.5 months at a median
fol­low-up of 9.9 months in patients with R/R dis­ease. The ORR
was 100% in newly diag­nosed patients inel­i­gi­ble for che­mo­ther­
apy, with median OS not reached at a median fol­low-up of 3.8
months.39 Another study of venetoclax-decitabine-FLT3 inhib­i­tor
of inves­ti­ga­tor’s choice found an ORR of 62% with 2-year OS of
29% in patients with R/R dis­ease and an ORR of 92% with 2-year
OS of 80% in patients over 60 years of age with newly diag­nosed
dis­ease.40
CLINICAL CASE 2 (Con­tin­ued)
The patient presented in this case has AML har­bor­ing an FLT3ITD muta­tion and is an inten­sive induc­tion can­di­date. Again,
it is impor­
tant to deter­
mine the patient’s under­
ly­
ing treat­
ment goals through shared deci­
sion-mak­
ing. A sum­
mary of
our approach and inves­ti­ga­tional strat­e­gies for the treat­ment
of FLT3-mutated AML, both newly diag­nosed and R/R, is pre­
sented in Figure 4.
High-risk AML novel approaches: TP53, KMT2A, FLT3 | 19
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/15/2021947/15sahasrabudhe.pdf by guest on 09 December 2022
Figure 3. Pathogenesis of FLT3-mutated AML and KMT2A-r AML. (A) FLT3-ITD and FLT3-TKD mutations lead to increased downstream
signaling from the FLT3 receptor, which thereby promotes increased cell growth, proliferation, and survival. FLT3 inhibitors block
this downstream signaling. (B) Binding of the KMT2A protein to menin allows the KMT2A/fusion partner protein to facilitate down­
stream transcription of HOX and other developmental proteins that contribute to leukemic pathogenesis. Inhibitors of the meninKMT2A interaction inhibit this transcription. AKT, Akt serine/threonine protein kinases; MAPK, mitogen-activated protein kinases;
PI3K, ­phosphoinositide-3 kinases; STAT5, signal transducer and activator of transcription 5. Created with BioRender​.com.
CLINICAL CASE 3
A 65-year-old woman with­out major comorbidities and nor­mal
per­for­mance sta­tus is diag­nosed with AML har­bor­ing a cyto­ge­
netic trans­lo­ca­tion involv­ing 11q23. She is treated with front­line
venetoclax-azacitidine and achieves CR. However, 4 months
later she is noted to have relapsed AML still har­bor­ing the 11q23
trans­lo­ca­tion.
Background
The KMT2A gene (pre­vi­ously known as MLL1) codes for the
­his­tone-lysine N-methyltransferase 2A pro­tein and is located
on chro­mo­some 11q23. Translocations involv­ing the KMT2A
locus occur in approx­i­ma­tely 3% of AML cases and have a
neg­a­tive impact on prog­no­sis.23,41 Translocations involv­ing
the KMT2A locus result in the expres­sion of fusion pro­teins
that enable an aber­rant tran­scrip­tion pro­gram char­ac­ter­ized
by the overexpression of HOX and other devel­op­men­tal
genes.42,43 The bind­ing of these fusion pro­teins to menin, a
scaf­fold pro­tein, leads to the nuclear local­i­za­tion of the fusion
20 | Hematology 2022 | ASH Education Program
pro­teins, thereby facil­i­tat­ing the aber­rant tran­scrip­tion that is
cen­tral to the path­o­gen­e­sis of KMT2A-rearranged (KMT2A-r)
AML.44,45 The inter­ac­tion of menin with the wild-type KMT2A
pro­tein has also been shown to play a role in the path­o­gen­e­sis
of NPM1-mutated AML.46
Targeting menin inhi­bi­tion in KMT2A-rearranged AML
SNDX-5613 is an oral small-mol­e­cule inhib­i­tor that binds with
high affin­ity to the KMT2A-bind­ing pocket of menin. This inhib­
its the inter­ac­tion between menin and KMT2A, which thereby
inhib­its onco­genic expres­sion and cel­lu­lar pro­lif­er­a­tion. SNDX5613 has dem­
on­
strated activ­
ity against KMT2A-r and NPM1mutated leu­ke­mia in pre­clin­i­cal and xeno­graft mod­els. Augment-1
was the first-in-human phase 1 study involv­ing SNDX-5613 and
included patients with R/R KMT2A-r and NPM1-mutated acute
leu­ke­mias (both acute lym­pho­cytic leu­ke­mia and AML). This
study dem­on­strated prom­is­ing response rates, with a com­pos­ite
CR of 44% and a median dura­tion of response of 5.2 months. The
only dose-lim­it­ing tox­ic­ity was grade 3 QTc pro­lon­ga­tion, which
occurred in 8% of patients, all­of whom were clin­i­cally asymp­
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/15/2021947/15sahasrabudhe.pdf by guest on 09 December 2022
Figure 4. Treatment recommendations for FLT3-Mutated AML. Color code: Blue background with white text: standard of care; green
background with red text: investigational approaches. Aza, azacitidine; ELN, European Leukemia Network; Gilt, gilteritinib; IC, inten­
sive chemotherapy; LDAC, low dose cytarabine; ND, newly diagnosed; R/R, relapsed/refractory; Ven, venetoclax.
*Data exist supporting allo-HSCT for all patients with FLT3-ITD-mutated AML, even if they are classified as favorable risk by the
2017 ELN guidelines based on a low ITD allelic ratio and concurrent NPM1 mutation. The updated 2022 ELN guidelines classify all
FLT3-ITD-mutated AML as intermediate risk.
**Sorafenib maintenance after allo-HSCT is approved per National Comprehensive Cancer Network guidelines but not commonly
used in clinical practice. Sorafenib is only effective against FLT3-ITD mutations, not FLT3-TKD mutations.
Created with BioRender.com.
CLINICAL CASE 3 (Con­t in­u ed)
Patients with newly diag­nosed KMT2A-r AML who are can­di­dates
should undergo allo-HSCT if they achieve remis­sion with front­
line ther­apy. There is no cur­rent stan­dard of care for R/R dis­ease,
but inves­ti­ga­tional inhib­i­tors of the menin-KMT2A inter­ac­tion
rep­re­sent prom­ise in this set­ting.
Conclusions
TP53 muta­tions, FLT3-ITD muta­tions, and KMT2A rearrangements
rep­re­sent high-risk genetic fea­tures in AML, and improved out­comes
are needed in these AML patient sub­sets. Multiple novel agents and
ther­a­peu­tic com­bi­na­tions are being stud­ied in these AML sub­sets
that have the poten­tial to fos­ter these needed improve­ments. This
is an excit­ing time in AML research, with hope for bet­ter out­comes
in patients with high-risk dis­ease mov­ing for­ward.
Conflict-of-inter­est dis­clo­sure
Kieran D. Sahasrabudhe: no com­
pet­
ing finan­
cial inter­
ests to
declare.
Alice S. Mims: sci­en­tific advi­sory com­mit­tee: Syndax Pharmaceu­
ticals, AbbVie, Genentech, BMS, Astellas, Servier Pharmaceuti­
cals, Ryvu Therapeutics; data and safety mon­i­tor­ing com­mit­tee:
Jazz Pharmaceuticals, Daiichi Saynko; senior med­i­cal direc­tor:
Leukemia and Lymphoma Society Beat AML study.
Off-label drug use
Kieran D. Sahasrabudhe: the off-label use of sorafenib is discussed.
Alice S. Mims: the off-label use of sorafenib is discussed.
Correspondence
Alice S. Mims, The Ohio State University, 1800 Cannon Dr, 1120G
Lin­coln Tower, Colum­bus, OH 43210; e-mail: alice​­.mims@osumc​
­.edu.
References
1.
Bowen D, Groves MJ, Burnett AK, et al. TP53 gene muta­tion is fre­quent in
patients with acute mye­loid leu­ke­mia and com­plex kar­yo­type, and is asso­
ci­ated with very poor prog­no­sis. Leukemia. 2009;23(1):203-206.
2. Ok CY, Patel KP, Garcia-Manero G, et al. Mutational pro­fil­ing of ther­apyrelated myelodysplastic syn­dromes and acute mye­loid leu­ke­mia by next
gen­er­a­tion sequenc­ing, a com­par­i­son with de novo dis­eases. Leuk Res.
2015;39(3):348-354.
3. Grossmann V, Schnittger S, Kohlmann A, et al. A novel hier­ar­chi­cal prog­
nos­
tic model of AML solely based on molec­
u­
lar muta­
tions. Blood.
2012;120(15):2963-2972.
4. Poiré X, Labopin M, Maertens J, et al. Allogeneic stem cell trans­plan­ta­tion
in adult patients with acute mye­loid leu­kae­mia and 17p abnor­mal­i­ties in
first com­plete remis­sion: a study from the Acute Leukemia Working Party
(ALWP) of the Euro­pean Society for Blood and Marrow Transplantation
(EBMT). J Hematol Oncol. 2017;10(1):20.
5. Ciurea SO, Chilkulwar A, Saliba RM, et al. Prognostic fac­tors influ­enc­ing
sur­vival after allo­ge­neic trans­plan­ta­tion for AML/MDS patients with TP53
muta­tions. Blood. 2018;131(26):2989-2992.
6. Welch JS, Petti AA, Miller CA, et al. TP53 and decitabine in acute mye­loid
leu­ke­mia and myelodysplastic syn­dromes. N Engl J Med. 2016;375(21):20232036.
7. Bories P, Prade N, Lagarde S, et al. Impact of TP53 muta­
tions in
acute mye­
loid leu­
ke­
mia patients treated with azacitidine. PLoS One.
2020;15(10):e0238795.
8. Short NJ, Kantarjian HM, Loghavi S, et al. Treatment with a 5-day ver­sus
a 10-day sched­ule of decitabine in older patients with newly diag­nosed
acute mye­loid leu­kae­mia: a randomised phase 2 trial. Lancet Haematol.
2019;6(1):e29-e37.
9. DiNardo CD, Jonas BA, Pullarkat V, et al. Azacitidine and venetoclax in pre­
vi­ously untreated acute mye­loid leu­ke­mia. N Engl J Med. 2020;383(7):617629.
10. Pollyea DA, Pratz KW, Wei AH, et al. Outcomes in patients with poor-risk
cyto­ge­net­ics with or with­out TP53 muta­tions treated with venetoclax
com­bined with hypomethylating agents. Blood. 2021;138(suppl 1):224.
11. Kim K, Maiti A, Loghavi S, et al. Outcomes of TP53-mutant acute mye­
loid leu­ke­mia with decitabine and venetoclax. Cancer. 2021;127(20):37723781.
12. Chao MP, Takimoto CH, Feng DD, et al. Therapeutic targeting of the mac­
ro­phage immune check­point CD47 in mye­loid malig­nan­cies. Front Oncol.
2019;9(22 Jan­u­ary):1380.
13. Sallman D, Asch A, Kambhampati S, et al. AML-196: the first-in-class antiCD47 anti­body magrolimab in com­bi­na­tion with azacitidine is well tol­
er­ated and effec­tive in AML patients: phase 1b results. Clin Lymphoma
Myeloma Leuk. 2021;21(suppl 1):s290.
14. Daver N, Konopleva M. Phase I/II study of azacitidine (AZA) with venetoclax
(VEN) and magrolimab (Magro) in patients (Pts) with newly diag­nosed old­
er/unfit or high-risk acute mye­loid leu­ke­mia (AML) and relapsed/refrac­tory
(R/R) AML. Blood. 2021;138(suppl 1):371.
15. Lambert JM, Gorzov P, Veprintsev DB, et al. PRIMA-1 reactivates mutant
p53 by cova­lent bind­ing to the core domain. Cancer Cell. 2009;15(5):376388.
16. Liu DS, Duong CP, Haupt S, et al. Inhibiting the sys­tem xC-/glu­ta­thi­one axis
selec­tively tar­gets can­cers with mutant-p53 accu­mu­la­tion. Nat Commun.
2017;8(27 March):14844.
17. Birsen R, Larrue C, Decroocq J, et al. APR-246 induces early cell death by
ferroptosis in acute mye­loid leu­ke­mia. Haematologica. 2022;107(2):403416.
18. Sallman DA, DeZern AE, Garcia-Manero G, et al. Eprenetapopt (APR-246)
and azacitidine in TP53-mutant myelodysplastic syn­dromes. J Clin Oncol.
2021;39(14):1584-1594.
19. Mishra A, Tamari R, Byrne M, et al. Phase II trial of eprenetapopt (APR-246)
in com­bi­na­tion with azacitidine (AZA) as main­te­nance ther­apy for TP53
mutated acute mye­
loid leu­
ke­
mia (AML) or myelodysplastic syn­
dromes
(MDS) fol­low­ing allo­ge­neic hema­to­poi­etic cell trans­plan­ta­tion (HCT).
Paper presented at: BMT Tandem Meetings; 23-26 April 2022; Salt Lake
City, UT; Abstract 39.
High-risk AML novel approaches: TP53, KMT2A, FLT3 | 21
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/15/2021947/15sahasrabudhe.pdf by guest on 09 December 2022
tom­atic. Other com­mon treat­ment-related adverse events were
nau­sea (22%), vomiting (17%), dif­fer­en­ti­a­tion syn­drome (DS;
15%), and diar­rhea (11%).47 A fol­low-up phase 2 study is ongo­
ing. SNDX-5613 is also being stud­ied in com­bi­na­tion with inten­
sive che­mo­ther­apy in a phase 1 trial in patients with R/R AML or
acute lym­pho­cytic leu­ke­mia har­bor­ing a KMT2A rearrangement
or NPM1 muta­tion (NCT05326516). An all­-oral tri­plet reg­i­men
of SNDX-5613–ASTX727–venetoclax is being stud­ied in a phase
1/2 clin­i­cal trial in patients with R/R AML or mixed phe­no­type
acute leu­ke­mia (NCT05360160). The Beat AML Master Clinical
trial is also study­ing the com­bi­na­tion of SNDX-5613–azacitidine–­
venetoclax in patients older than 60 with newly diag­
nosed
KMT2A-r or NPM1-mutated AML (NCT03013998).
KO-539 and JNJ-75276617 are 2 other novel inhib­i­tors of the
menin-KMT2A inter­ac­tion that are being inves­ti­gated in ­earlyphase clin­
i­
cal tri­
als in patients with R/R AML (NCT04067336,
NCT04811560). DS is an impor­tant adverse event for cli­ni­cians to
be aware of and, if unrec­og­nized, can be fatal. DS was typ­i­cally
low grade in patients treated on Augment-1 but did lead to the
US Food and Drug Administration plac­ing a tem­po­rary hold on
KO-539, which has since been lifted. Signs and symp­toms of DS
include unex­plained fever, weight gain, edema, pleu­ral or peri­
car­dial effu­sions, radio­graphic opac­i­ties, dyspnea, hypo­ten­sion,
renal dys­func­tion, rash, and/or a rap­idly increas­ing white blood
cell count. Treatment involves the prompt ini­ti­a­tion of ste­roids
and fol­low­ing the inves­ti­ga­tor bro­chure guide­lines regard­ing
con­tin­u­a­tion of the inves­ti­ga­tional agent. The path­o­gen­e­sis of
KMT2A-r AML is shown in Figure 3B.
22 | Hematology 2022 | ASH Education Program
main­te­nance ther­apy in acute mye­loid leu­ke­mia with FLT3-ITD. Blood.
2019;133(8):840-851.
35. Maziarz RT, Levis M, Patnaik MM, et al. Midostaurin after allo­ge­neic stem cell
trans­plant in patients with FLT3-inter­nal tan­dem dupli­ca­tion-pos­i­tive acute
mye­loid leu­ke­mia. Bone Marrow Transplant. 2021;56(5):1180-1189.
36. Wang ES, Montesinos P. Phase 3, open-label, ran­dom­ized study of gilter­
itinib and azacitidine vs azacitidine for newly diag­
nosed FLT3-mutated
acute mye­loid leu­ke­mia in patients inel­i­gi­ble for inten­sive induc­tion che­
mo­ther­apy. Blood. 2021;138(suppl 1):700.
37. Yamatani K, Tabe Y, Saito K, et al. Upregulation of Bcl-2 con­fers resis­tance
to FLT3 inhi­bi­tion in FLT3-ITD AML with sec­ond­ary acquired muta­tions.
Blood. 2018;132(suppl 1):3944.
38. Perl AE, Daver NG, Pratz KW, et al. Venetoclax in com­bi­na­tion with gilteri­
tinib in patients with relapsed/refrac­tory acute mye­loid leu­ke­mia: a phase
1b study. Blood. 2019;134(suppl 1):3910.
39. Short NJ, DiNardo CD, Daver N, et al. A trip­let com­bi­na­tion of azacitidine,
venetoclax and gilteritinib for patients with FLT3-mutated acute mye­loid
leu­ke­mia: results from a phase I/II study. Blood. 2021;138(suppl 1):696.
40. Maiti A, DiNardo CD, Daver NG, et al. Triplet ther­apy with venetoclax, FLT3
inhib­i­tor and decitabine for FLT3-mutated acute mye­loid leu­ke­mia. Blood
Cancer J. 2021;11(2):25.
41. Schoch C, Schnittger S, Klaus M, Kern W, Hiddemann W, Haferlach T. AML
with 11q23/MLL abnor­mal­i­ties as defined by the WHO clas­si­fi­ca­tion: inci­
dence, part­ner chro­mo­somes, FAB sub­type, age dis­tri­bu­tion, and prog­
nos­tic impact in an unse­lected series of 1897 cyto­ge­net­i­cally ana­lyzed AML
cases. Blood. 2003;102(7):2395-2402.
42. Meyer C, Burmeister T, Gröger D, et al. The MLL recombinome of acute leu­
ke­mias in 2017. Leukemia. 2018;32(2):273-284.
43. Del Rizzo PA, Trievel RC. Substrate and prod­uct specificities of SET domain
methyltransferases. Epigenetics. 2011;6(9):1059-1067.
44. Yokoyama A, Somervaille TC, Smith KS, Rozenblatt-Rosen O, Meyerson
M, Cleary ML. The menin tumor sup­pres­sor pro­tein is an essen­tial onco­
genic cofac­tor for MLL-asso­ci­ated leu­ke­mo­gen­e­sis. Cell. 2005;123(2):207218.
45. Caslini C, Yang Z, El-Osta M, Milne TA, Slany RK, Hess JL. Interaction of MLL
amino ter­mi­nal sequences with menin is required for trans­for­ma­tion. Cancer Res. 2007;67(15):7275-7283.
46. Kühn MW, Song E, Feng Z, et al. Targeting chro­ma­tin reg­u­la­tors inhib­its
leu­ke­mo­genic gene expres­sion in NPM1 mutant leu­ke­mia. Cancer Discov.
2016;6(10):1166-1181.
47. Stein EM, Aldoss I, DiPersio JF, et al. Safety and effi­cacy of menin inhi­bi­tion
in patients (Pts) with MLL-rearranged and NPM1 mutant acute leu­ke­mia:
a phase (Ph) 1, first-in-human study of SNDX-5613 (AUGMENT 101). Blood.
2021;138(suppl 1):699.
© 2022 by The Amer­i­can Society of Hematology
DOI 10.1182/hema­tol­ogy.2022000325
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/15/2021947/15sahasrabudhe.pdf by guest on 09 December 2022
20. Garcia-Manero G, Goldberg AD, Winer ES, et al. Phase I and expan­sion
study of eprenetapopt (APR-246) in com­bi­na­tion with venetoclax (VEN)
and azacitidine (AZA) in TP53-mutant acute mye­
loid leu­
ke­
mia (AML).
Blood. 2021;138(suppl 1):3409.
21. Fröhling S, Schlenk RF, Breitruck J, et al; AML Study Group Ulm. Acute
mye­loid leu­ke­mia. Prognostic sig­nif­i­cance of acti­vat­ing FLT3 muta­
tions in youn­ger adults (16 to 60 years) with acute mye­loid leu­ke­mia
and nor­mal cyto­ge­net­ics: a study of the AML Study Group Ulm. Blood.
2002;100(13):4372-4380.
22. Mead AJ, Linch DC, Hills RK, Wheatley K, Burnett AK, Gale RE. FLT3 tyro­sine
kinase domain muta­tions are bio­log­i­cally dis­tinct from and have a sig­nif­i­
cantly more favor­able prog­no­sis than FLT3 inter­nal tan­dem dupli­ca­tions in
patients with acute mye­loid leu­ke­mia. Blood. 2007;110(4):1262-1270.
23. Döhner H, Estey E, Grimwade D, et al. Diagnosis and man­age­ment of AML
in adults: 2017 ELN rec­om­men­da­tions from an inter­na­tional expert panel.
Blood. 2017;129(4):424-447.
24. Döhner H, Wei AH, Appelbaum FR, et al. Diagnosis and man­age­ment of
AML in adults: 2022 ELN rec­om­men­da­tions from an inter­na­tional expert
panel. Blood. 2022;140(12):1345-1377.
25. Brunet S, Labopin M, Esteve J, et al. Impact of FLT3 inter­nal tan­dem dupli­
ca­tion on the out­come of related and unre­lated hema­to­poi­etic trans­plan­
ta­tion for adult acute mye­loid leu­ke­mia in first remis­sion: a ret­ro­spec­tive
anal­y­sis. J Clin Oncol. 2012;30(7):735-741.
26. Stone RM, Mandrekar SJ, Sanford BL, et al. Midostaurin plus che­mo­ther­
apy for acute mye­
loid leu­
ke­
mia with a FLT3 muta­
tion. N Engl J Med.
2017;377(5):454-464.
27. Perl AE, Martinelli G, Cortes JE, et al. Gilteritinib or che­mo­ther­apy for relapsed
or refrac­tory FLT3-mutated AML. N Engl J Med. 2019;381(18):1728-1740.
28. Burchert A, Bug G, Fritz LV, et al. Sorafenib main­
te­
nance after allo­
ge­
neic hema­to­poi­etic stem cell trans­plan­ta­tion for acute mye­loid leu­ke­mia
with FLT3–inter­nal tan­dem dupli­ca­tion muta­tion (SORMAIN). J Clin Oncol.
2020;38(26):2993-3002.
29. Xuan L, Wang Y, Huang F, et al. Sorafenib main­te­nance in patients with
FLT3-ITD acute mye­loid leu­kae­mia under­go­ing allo­ge­neic haematopoietic
stem-cell trans­plan­ta­tion: an open-label, multicentre, randomised phase 3
trial. Lancet Oncol. 2020;21(9):1201-1212.
30. Pratz KW, Cherry M, Altman JK, et al. A phase 1 study of gilteritinib in com­
bi­na­tion with induc­tion and con­sol­i­da­tion che­mo­ther­apy in patients with
newly diag­nosed AML: final results. Blood. 2020;136(suppl 1):16-17.
31. Abstract Book for the 27th Congress of the Euro­pean Hematology Associ­
ation. HemaSphere. 2022;6:1-4130.
32. Cortes JE, Kantarjian HM, Kadia TM, et al. Crenolanib besylate, a type I panFLT3 inhib­i­tor, to dem­on­strate clin­i­cal activ­ity in mul­ti­ply relapsed FLT3-ITD
and D835 AML. J Clin Oncol. 2016;34(suppl 15):7008.
33. Goldberg AD, Geyer MB, Kell J, et al. Clinical ben­e­fit of crenolanib, with
or with­out sal­vage che­mo­ther­apy, in mul­ti­ply relapsed, FLT3 mutant AML
patients after prior treat­ment with gilteritinib. Blood. 2020;136(suppl 1):8-9.
34. Schlenk RF, Weber D, Fiedler W, et al; Ger­
man-Austrian AML Study
Group. Midostaurin added to che­mo­ther­apy and con­tin­ued sin­gle-agent
ACUTE MYELOID LEUKEMIA: IMPROVING OUTCOMES IN CHALLENGING SUBSETS
Optimizing outcomes in secondary AML
Abramson Cancer Center, Hospital of University of Pennsylvania, Philadelphia, PA
Acute myeloid leukemia (AML) secondary to antecedent hematologic disorder or prior therapeutics for cancer represent a
diverse group of leukemias often associated with inferior outcomes. Conventional therapy with cytarabine-based chemotherapy has been the mainstay of care for the past 30 years with disappointing overall outcomes. Novel therapies, including liposomal cytarabine/daunorubicin, and venetoclax-based therapies have emerged as options in recent years based
on studies showing improvement in outcomes over standard-of-care therapies. Despite these advances, mutations in TP53
are associated with inferior response to both therapies and represent an area of unmet clinical need. Novel strategies with
immune-targeted therapies such as CD47 monoclonal antibodies appear active in early-phase studies, but randomized
studies have yet to report outcomes leading to approval. Allogeneic transplant remains the only known curative therapy
for many of these cases. Nonetheless, pretransplant high-risk molecular features of secondary AML are associated with
inferior outcome despite transplantation. An optimal approach to secondary AML is yet to be determined.
LEARNING OBJECTIVES
• Secondary acute myeloid leukemia (AML) encompasses both therapy-related myeloid neoplasms and antecedent
myelodysplastic syndrome and/or myeloproliferative neoplasm
• Favorable risk secondary AML is rare but can be managed like de novo favorable risk disease
• Allogeneic stem cell transplant remains the preferred option in first complete remission for intermediate- and
unfavorable-risk disease
• Venetoclax and hypomethylating agent combination may be used as a bridge to transplant or destination therapy
• Emerging therapies may further expand treatment options for this historically older, frailer population with complex genetic changes and TP53 mutations
Introduction
Secondary acute myeloid leukemia (AML) has been historically divided into 2 categories based on known risks for
development of disease from either (1) antecedent myeloid
neoplasm such as myelodysplastic syndrome (MDS) or myeloproliferative neoplasm (MPN) or (2) exposure to ionizing
radiation and/or cytotoxic chemotherapy (Table 1). In population-based cohort series, secondary AML accounts for 20%
to 30% of all AML cases.1,2 The previous World Health Organization (WHO) 2016 classification included the separate categories of AML with myelodysplastic changes (AML-MRC)
and therapy-related myeloid neoplasms (tMNs), which were
the formal categories of secondary AML.3 The WHO classification of tMN was subdivided into therapy-related acute
myeloid leukemia and therapy-related myelodysplastic
syndrome. The fact that both MDS and AML are included in
the single category of “therapy-related myeloid neoplasm”
emphasizes that regardless of precise blast count above or
below 20%, similar biology, pathogenesis, and risk of rapid
progression with poor prognosis guide the need for treatment and transplant evaluation for most patients.
The latest editions of the WHO classification4 and the
International Consensus Classification5 further emphasize
disease biology and genetic features while softening the
boundary between MDS and AML. The WHO classification
now specifically eliminates blast cutoffs for most AML types
with defining genetic alterations but retains a 20% blast
cutoff to delineate MDS from AML in cases lacking such
genetic alterations. AML-MRC has been updated to AML,
myelodysplasia related (AML-MR). AML-MR harbors specific
mutational or cytogenetic abnormalities and can no longer
be defined by morphology alone. Either the presence of
1 or more molecular and cytogenetic abnormalities or a history of MDS/MPN is required for a diagnosis of AML-MR. Further complicating these distinctions, patients with aplastic
anemia or inherited bone marrow failure syndromes may
also develop AML, particularly if dominant clones emerge.
Far more commonly, patients have had prior cytotoxic
Dr Prakash Singh Shekhawat
Optimizing outcomes in secondary AML | 23
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/23/2021835/23matthews.pdf by guest on 09 December 2022
Andrew Matthews and Keith W. Pratz
Table 1. Summary of World Health Organization AML Classifications3
Category
Key fea­tures
De novo AML
• AML with defin­ing genetic abnor­mal­i­ties
Defined bal­anced trans­lo­ca­tion/inver­sions and gene muta­tions (eg, mutated NPM1, biallelic CEBPA,
RUNX1::RUNX1T1). Only AML with BCR::ABL1 fusion and AML with CEBPA muta­tion require ≥20% blast
count in PB or BM.
• AML defined by dif­fer­en­ti­a­tion*
≥20% PB or BM blasts with fur­ther subcategories based on mor­phol­ogy, cyto­chem­is­try, and
immunophenotype
Secondary mye­loid neo­plasms
≥20% PB or BM blasts and prior ther­apy has been excluded with either spe­cific cyto­ge­netic abnor­
mal­i­ties or somatic muta­tions.† This includes patients with known his­tory of MDS or MDS/MPN.
• AML, post–cyto­toxic ther­apy (pre­vi­ously
ther­apy-related mye­loid neo­plasms)
≥20% PB or BM blasts and patient-devel­oped mye­loid neo­plasms fol­low­ing cyto­toxic ther­apy
• Myeloid neo­plasms with germline
pre­dis­po­si­tion
Includes mye­loid neo­plasms with germline path­o­logic/likely path­o­logic muta­tions in CEPBA, DDX41,
TP53, RUNX1, ANKRD26, ETV6, GATA2, SAMD9, SAMD9L, and BLM along with telo­mere biol­ogy dis­or­
ders, RASopathies, Down syn­drome, and bone mar­row fail­ure syn­dromes. Phenotype of AML vs MDS
with germline pre­dis­po­si­tion based on pres­ence or absence of >20% PB or BM blasts, respec­tively.
*AML not oth­er­wise spec­i­fied is no lon­ger rec­og­nized as a cat­e­gory.
†Cytogenetic abnor­mal­i­ties include 5q dele­tion or loss of 5q due to unbal­anced trans­lo­ca­tion; mono­somy 7, 7q dele­tion, or loss of 7q due to unbal­
anced trans­lo­ca­tion; 11q dele­tion; 12p dele­tion or loss of 12p due to unbal­anced trans­lo­ca­tion; mono­somy 13 or 13q dele­tion; 17p dele­tion or loss of
17p due to unbal­anced trans­lo­ca­tion; iso­chro­mo­some 17q; and idic(X)(q13). Molecular somatic muta­tions include SRSF2, SF3B1, U2AF1, ZRSR2, ASXL1,
EZH2, BCOR, and STAG2.
BM, bone mar­row; PB, periph­eral blood.
ther­apy or known MDS/MPN, or they can be clas­si­fied into AMLMR based solely on the pres­
ence of myelodysplasia-related
chro­mo­somal or muta­tional find­ings. Meanwhile, tMN has been
replaced by mye­
loid neo­
plasm post–cyto­
toxic ther­
apy, still
requir­ing a documented his­tory of che­mo­ther­apy treat­ment or
large-field radi­a­tion ther­apy for an unre­lated neo­plasm.
Within both categories of AML-MR and AML, post–cyto­toxic
ther­apy (AML-pCT), sim­i­lar chro­mo­somal and molec­u­lar abnor­
mal­i­ties are found, suggesting a com­mon final path­way. Secondary leu­ke­mia from ante­ced­ent MDS/MPN appears to share a
typ­i­cal evo­lu­tion­ary path­way, with sev­eral genes distinguishing
sec­ond­ary AML from de novo AML with high spec­i­fic­ity.6 Secondary leu­ke­mias after cyto­toxic ther­apy may evolve through a sim­
i­lar MDS evo­lu­tion­ary path­way or acquire TP53 or other driver
muta­tions directly from che­mo­ther­apy, radi­a­tion, or immu­
no­sup­pres­sion. Common cyto­ge­netic and molec­u­lar fea­tures
reflect this var­i­abil­ity, with 50% of patients with AML-pCT hav­ing
poor-risk cyto­ge­net­ics. The most fre­quent molec­u­lar aber­ra­tion
is TP53 (33%), and only 15% of patients with AML-pCT pres­ent
with favor­able-risk fusion genes (RUNX1::RUNX1T1, CBFB::MYH11,
PML::RARA). Age-related increase in clonal hema­to­poi­e­sis7 may
pre­dis­pose patients to develop sec­ond­ary leu­ke­mia as an accre­
tion of muta­tions leads to mye­loid neo­plasms, espe­cially under
the selec­tive pres­sure of che­mo­ther­apy.8
Therapy-related mye­loid neo­plasms may arise after a vari­
ety of treat­ments, with clas­sic exam­ples being alkylating agents
lead­ing to AML-pCT with ante­ced­ent MDS, with abnor­mal­i­ties of
chro­mo­some 5 and/or 7 or topoisomerase II inhib­i­tors lead­ing to
KMT2A gene rearrangements at 11q23. Beyond recur­ring gene or
chro­mo­somal rearrangements, other ther­a­pies can lead to leu­ke­
mia by selecting preexisting clones resis­tant to che­mo­ther­apy,
with recent novel ther­apy exam­ples includ­ing poly (ADP-ribose)
polymerase inhib­i­tors and lenalidomide (see Figure 1).1,9
Outcomes in both AML-MR and AML-pCT appear sim­
i­
lar
across genetic sub­sets. Rare sub­sets of favor­able-risk treat­
24 | Hematology 2022 | ASH Education Program
ment-related dis­ease such as AML with inv16 or t(8;21) and
acute promyelocytic leukemia t(15;17) appear to have dimin­
ished over­all sur­vival com­pared with de novo AML. However,
the impact of confounding from the ante­ced­ent malig­nancy or
advanced age is often not exam­ined in ret­ro­spec­tive stud­ies
com­par­ing out­comes in sec­ond­ary AML to de novo AML. Clinically, diag­no­sis requires inte­gra­tion of clin­i­cal his­tory, mor­pho­
logic changes, cyto­ge­net­ics, and molec­u­lar ana­ly­ses. Each of
these com­po­nents may change treat­ment strat­e­gies and ulti­
mate out­come. Herein we describe the genetic and molec­u­lar
fea­tures asso­ci­ated with the sec­ond­ary AMLs and approaches
and out­comes of con­ven­tional and emerg­ing ther­a­pies.
CLINICAL CASE 1
A 55-year-old woman with a remote his­
tory of Bur­
kitt lym­
phoma presented for rou­tine annual exam­i­na­tion and was found
on com­plete blood count to have pan­cy­to­pe­nia. White blood
cell count was 900/µL, hemo­glo­bin was 7.4 g/dL, and plate­
let count was 101 000/µL. Bone mar­row exam­i­na­tion revealed
28% blasts by mor­phol­ogy, and kar­yo­type was com­plex 43,XX,
del(5)(q11.2q33),-13,add(14) (p11.2),add(17)(p11.2),dic(18;19)(p11;2;
q13.1),psu dic(22;1)(q13;p13)[16]/45,XX,del(5),der(13;14) (q10;q10),
add(17),add(20)(q13.1)[2],46,XX[2]. Molecular stud­
ies revealed
TP53 muta­tion at R248Q with 62% variant allele frequency (VAF).
Performance sta­
tus was excel­
lent, and she had nor­
mal endorgan func­tion. Her lym­phoma his­tory was nota­ble for 30 years
of ongo­ing remis­sion after multiagent che­mo­ther­apy in her early
20 s. She had an unknown quan­tity of anthracycline expo­sure,
and records of the ther­apy were unavail­able. Echocardiogram
at time of diag­no­sis revealed nor­mal ejec­tion frac­tion with­out
regional wall motion abnor­mal­i­ties.
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/23/2021835/23matthews.pdf by guest on 09 December 2022
• AML, myelodysplasia related
Dr Prakash Singh Shekhawat
Optimizing out­comes in sec­ond­ary AML | 25
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/23/2021835/23matthews.pdf by guest on 09 December 2022
Figure 1. Descriptive data by AML type. (A) Incidence of de
novo AML, secondary AML (sAML), and therapy-related AML
(tAML) in Denmark by year of diagnosis. Distribution of previous disease in (B) 203 patients with tAML and (C) 603 patients
with sAML. Originally published in Østgård LSG, Medeiros
BC, Sengeløv H, et al. Epidemiology and clinical significance
of secondary and therapy-related acute myeloid leukemia: a national population-based cohort study. J Clin Oncol.
2015;33(31):3641-3649.1
This patient has mye­loid neo­plasm, post–cyto­toxic ther­apy,
and her blasts >20% put her in the acute mye­loid leu­ke­mia cat­
e­gory. Cohorts of AML-pCT tend to con­tain more women than
men as the most com­mon pri­mary malig­nancy is breast can­
cer followed by non-Hodgkin lym­
phoma.10 Despite the long
latency from treat­ment to AML-pCT pre­sen­ta­tion, her com­plex
kar­yo­type and TP53 muta­tion are com­mon fea­tures of AML-pCT.
Multiple treat­ments are options in this case. Standard-of-care
ther­apy with “7+3” (7 days of cytarabine and 3 days of anthracycline) would be the most com­mon approach in this case,
given a 33% to 55% over­all response rate in sec­ond­ary AML with
inten­sive che­mo­ther­apy.1,11 However, the full molec­u­lar and kar­
yo­typic pat­tern would pre­dict for a low over­all response rate.
Based on kar­yo­typic changes alone, she also meets cri­te­ria for
AML, myelodysplasia related,3,5 and as such, ther­apy with lipo­
so­mal cytarabine and dau­no­ru­bi­cin (CPX-351) could be con­sid­
ered.12 The over­all response rates asso­ci­ated with sec­ond­ary
AML and TP53-mutant leu­
ke­
mia for both 7+3 and lipo­
so­
mal
cytarabine/dau­no­ru­bi­cin are sum­ma­rized in Table 2. The longterm sur­vival with either ther­apy is very lim­ited and only found
in patients who can pro­ceed with allo­ge­neic trans­plan­ta­tion.13,14
The patient’s his­
tory of high-grade lym­
phoma and exten­
sive
remote che­mo­ther­apy expo­sure would also com­pli­cate her tol­
er­ance of high-inten­sity che­mo­ther­apy. With lim­ited knowl­edge
about her anthracycline expo­sure, the choice of a reg­i­men with
anthracyclines becomes prob­lem­atic even with nor­mal ejec­tion
frac­tion and use of cardio­protection with dexrazoxane.
There were lower-inten­sity options offer­ing rea­son­able remis­
sion rates as a bridge to cura­tive intent trans­plant. Azacitidine and
venetoclax (aza/ven) has been asso­ci­ated with com­plete response
(CR) and com­plete remis­sion with incom­plete blood count recov­
ery rate in AML with poor-risk cyto­ge­net­ics and TP53 muta­tions of
41% in a recent review of older adults who were unfit for induc­tion
che­mo­ther­apy in phase 1b and phase 3 stud­ies of aza/ven.15 Given
no pro­spec­tive com­par­a­tive stud­ies and a grow­ing body of ret­ro­
spec­tive stud­ies that show sim­i­lar over­all sur­vival with venetoclax
and hypomethylating agents com­
pared with more tra­
di­
tional
inten­sive che­mo­ther­apy approaches, a patient-cen­tered dis­cus­
sion of puta­tive risks and ben­e­fits is appro­pri­ate.16,17 This patient
had a strong pref­er­ence for out­pa­tient-based ther­apy, and as
such, we proceeded with this ther­apy. She achieved CR after 1
course and com­pleted 5 cycles of aza/ven before pro­ceed­ing to
myeloablative allo­ge­neic stem cell trans­plant 26 months ago with
under 5% VAF TP53 muta­tion and nor­mal kar­yo­type at the time of
trans­plant. She has ongo­ing remis­sion and nor­mal blood counts
with­out active graft-vs-host dis­ease.
Data on trans­plant in this sec­ond­ary AML sub­set reflect out­
comes predicted by pretransplant molec­u­lar and cyto­ge­netic
risks in de novo AML. Several stud­
ies sup­
port myeloablative
con­di­tion­ing over reduced inten­sity con­di­tion­ing in improved
relapse-free sur­
vival and equal or supe­
rior over­
all sur­
vival,
although his­tor­i­cally for TP53 muta­tion, the ben­e­fit has been less
clear, espe­cially in MDS.18 In addi­tion, the pres­ence of mea­sur­able
resid­ual dis­ease (MRD), defined as detec­tion of any of a sub­set
of 13 com­monly mutated genes on ultra-deep error-corrected
sequenc­
ing of preconditioning blood, appears to con­
fer the
highest risk of relapse on patients who receive a nonmyeloablative trans­plant.19 It is unknown if MRD-neg­a­tive patients may forgo
Table 2. Representative efficacy outcomes in AML subsets
Regimen
Response rates
Overall sur­vival
All groups
TP53
All groups
TP53
7+3 (cytarabine and anthracycline)2,11,12,37
CR: 35%-71%
CR/CRi: 40%-71%
sAML sub­set:
CR: 26%-52%
CR/CRi: 33%-55%
CR: 30%-34%
CR/CRi: 40%
sAML sub­set:
5-10 mo
5-6 mo
CPX-351 (lipo­so­mal cytarabine
and dau­no­ru­bi­cin)12,16,37
sAML sub­set:
CR: 7%-12%
CR/CRi: 45%-48%
CR: 29%
CR/CRi: 29%
sAML sub­set:
10-13 mo
4-6 mo
Azacitidine and venetoclax16,17,21
CR: 40%
CR/CRi: 65%
sAML subset:
CR/CRi: 60%
CR: NA
CR/CRi: 50%-55%
11-16 mo
sAML sub­set:
11-16 mo
5-7 mo
Azacitidine or decitabine
monotherapy21,38
CR: 13%-24%
CR/CRi: 18%-27%
sAML sub­set:
CR/CRi: 25%
CR: 24%-40%
CR/CRi: 0%-40%
6-11 mo
sAML sub­set:
7-8 mo
2-7 mo
Low-dose cytarabine (LoDAC)38-40
CR: 3%-24%
CR/CRi: 5%-34%
sAML sub­set:
CR: 0%
CR: 0%-11%
CR/CRi: 0%-22%
4-5 mo
sAML sub­set:
4 mo
2-3 mo
Glasdegib and LoDAC41,42
CR: 18%
CR/CRi: 24%
sAML sub­set:
CR: 24%
CR: 24%*
CR/CRi: NA
8-9 mo
sAML sub­set:
9 mo
5-6 mo*
CD47/SIRPα inhib­i­tors with HMA
(eg, magrolimab)35
CR: 44%
CR/CRi: 56%
CR: 48%
CR/CRi: 67%
18.9 mo
12.9 mo
TIM-3 inhib­i­tors with HMA
(eg, sabatolimab)43
NA
CR: 25%
CR/CRi: 30%
NA
NA
Bispecific anti­body ther­a­pies
(eg, flotetuzumab) in R/R AML44
NA
CR: 27%
CR/CRi: 40%
NA
4.5 mo
Intensive reg­i­mens
Experimental reg­i­mens
Representative effi­cacy out­comes in sub­sets of acute mye­loid leu­ke­mia with myelodysplasia-related changes, ther­apy-related mye­loid neo­plasm,
and TP53. Early clin­i­cal trial data are presented for rep­re­sen­ta­tive exper­i­men­tal reg­i­mens with listed agents.
*Poor cytogenic risk group reported, no sep­a­rate TP53 mutated sub­set data avail­­able.
CRi, com­plete remis­sion with incom­plete count recov­ery; HMA, hypomethylating agent; NA, not available; R/R, relapsed refrac­tory dis­ease; sAML,
sec­ond­ary AML.
myeloablative reg­i­mens.20 In this case, the achieve­ment of remis­
sion allowed for trans­plan­ta­tion to pro­ceed. Despite the mod­
est response rate published with aza/ven, the dura­bil­ity of these
responses is often lim­ited to less than 6 months, high­light­ing the
need for more effec­tive and dura­ble ther­apy in this sub­set.15,21 Our
cur­rent approach favors stem cell trans­plant when pos­si­ble, but
for inter­me­di­ate- or adverse-risk patients who are inel­i­gi­ble or
bor­der­line can­di­dates for trans­plant, novel ther­a­pies are needed.
CLINICAL CASE 2
A 52-year-old woman with a his­tory of tri­ple-neg­a­tive breast
can­cer fol­low­ing adju­vant radiation therapy along with doxo­
ru­bi­cin, cyclo­phos­pha­mide, and docetaxel was found to have
40% blasts on rou­tine blood work 12 months after com­ple­tion
of che­mo­ther­apy. Bone mar­row biopsy spec­i­men con­firmed
26 | Hematology 2022 | ASH Education Program
AML with 73% blasts, and chro­mo­some anal­y­sis revealed 48,XX,
+8,+13,t(16;16) (p13.1;q22)[20]. Molecular stud­
ies revealed a
NRAS G13D muta­tion with a VAF of 38%. Her life­time anthracycline dose was 240 mg/m2. She has no evi­dence of breast
can­cer at the time of diag­no­sis of AML.
Acute leu­ke­mia after expo­sure to ion­iz­ing radi­a­tion or che­mo­
ther­apy rarely pres­ents with oth­er­wise favor­able genetic mark­ers
such as fusions of RUNX1-RUNXT1, CBFB-MYH11, and PML-RARA in
~15% of cases.10 In the case of acute promyelocytic leukemia after
che­mo­ther­apy, the out­comes appear as favor­able as the out­
comes in patients who had never been exposed to che­mo­ther­
apy.22 On occa­sion, core bind­ing fac­tor AML is found in patients
with prior che­mo­ther­apy expo­sure, and review of the out­comes
of these patients sug­gests stan­dard-of-care ther­a­pies are asso­
ci­ated with shorter over­all sur­vival than de novo cases. That said,
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/23/2021835/23matthews.pdf by guest on 09 December 2022
Nonintensive reg­i­mens
CLINICAL CASE 3
A 69-year-old woman presented for short­ness of breath and
was found to have pan­cy­to­pe­nia. She had no prior diag­no­sis of
an ante­ced­ent MDS or MPN and no his­tory of cyto­toxic ther­a­
pies. Bone mar­row biopsy spec­i­men revealed AML, myelodysplasia related. Blasts were enu­mer­ated at 68%, and dys­pla­sia
was greater than 50% in 2 cell lines. Karyotype was nor­mal.
Molecular test­ing revealed NPM1 (46% VAF), IDH1 R132H (48%
VAF), SRSF2 (45% VAF), and FLT3-ITD (4.5% VAF). She is fit for
induc­tion che­mo­ther­apy.
AML-MR falls under a cat­e­gory of AML included in the sec­
ond­ary sub­set. This case high­lights a com­mon sce­nario where
a patient pres­ents with­out a known ante­ced­ent dis­or­der and
is still man­
aged as hav­
ing sec­
ond­
ary AML. In this case, the
founder muta­tion in the IDH1 gene at R132H likely predisposed
the patient to the dys­plas­tic fea­tures seen in this pre­sen­ta­tion.
Therapeutic options include con­ven­tional ther­apy with 7+3 or
CPX-351, which has been approved for patients with AML-MR.12
In this case, before know­ing the results of the NPM1 and FLT3
test­ing, the patient ini­ti­ated ther­apy with lipo­so­mal cytarabine/
dau­no­ru­bi­cin. Upon deter­mi­na­tion of the low-level FLT3-ITD
muta­
tion, midostaurin was added on days 8 to 21 based on
improved over­all sur­vival with stan­dard 7+3 induc­tion ther­apy
plus midostaurin in the RATIFY trial.26 She was found to have
MRD-neg­a­tive remis­sion at the com­ple­tion of induc­tion via
assess­ment of NPM1 with quan­ti­ta­tive RT-PCR and proceeded
to nonmyeloablative stem cell trans­plant given her age over
65 years.
This case high­lights some com­plex­ity in the diag­nos­tics and
char­ac­ter­iza­tions of AML with the evo­lu­tion of novel ther­a­pies
and evolved risk cat­e­go­ri­za­tions to include molec­u­lar changes.
By the WHO 2016 clas­si­fi­ca­tion sys­tem, the pres­ence of dys­pla­
sia alone would have been suf­fi­cient for a diag­no­sis of AML-MRC,
but by the 2022 clas­si­fi­ca­tion sys­tem cri­te­ria, the pres­ence of
SRSF2 or other molec­u­lar or cyto­ge­netic changes is nec­es­sary
for AML-MR. By the European LeukemiaNet 2017 criteria, she
would fall into favorable-risk disease.27 In youn­ger patients, over­
all sur­vival (OS) is sim­i­lar in those who receive an allo­genic stem
cell trans­plant in first CR and those who do not. It is unclear if this
find­ing can be extrap­o­lated to older patients.28 While favor­able
via this method, the median OS in the phase 3 study of CPX351 was only 9.7 months.12 As such, our stan­dard prac­tice is to
con­sol­i­date remis­sions achieved in patients given CPX-351 who
are trans­plant eli­gi­ble with an allo­ge­neic stem cell trans­plant. In
the registrational study, the out­comes in patients who went to
trans­plant after CPX-351 appear to be improved over those who
received 7+3, although MRD assess­ments were not avail­­able.14
An inter­est­ing subanalysis of these OS data suggested lower
trans­plant-related (nonrelapse) mor­tal­ity in the CPX-351 arm.29
Specific to the NPM1 sub­set, quan­ti­ta­tive RT-PCR mon­i­tor­ing
of blood or mar­row after induc­tion can iden­tify patients likely
to relapse and those with inter­me­di­ate-risk dis­ease who would
ben­e­fit from trans­plant, mon­i­tor­ing of blood once in remis­sion
can iden­tify favor­able-risk patients likely to relapse, and mon­
i­tor­ing of blood or bone mar­row before and after trans­plant
can help prog­nos­ti­cate risk of relapse.30 In older adults in whom
improve­ments from favor­able-risk dis­ease appear less,31 stem cell
trans­plant still appears to be advan­ta­geous.32 This patient was fit
for inten­sive ther­apy, but venetoclax and azacitidine would be
any option for many other 69-year-old patients. We would have
favored that approach over, say, ivosidenib and azacitidine given
the higher response rate and faster time to response.
CLINICAL CASE 4
A 77-year-old man presented for eval­u­a­tion of diz­zi­ness and
short­ness of breath and was found to have a white blood cell
count of 7480/µL with 17% periph­eral blood blasts, hemo­glo­
bin of 6.1 g/dL, and plate­let count of 70 000/µL. Bone mar­row
exam­i­na­tion revealed 48% blasts and chro­mo­some anal­y­sis
revealed a com­plex kar­yo­type with evi­dence of mono­somy 5,
7, and 17. Molecular stud­ies revealed muta­tion in TP53 at R273
with a VAF of 80%. His per­for­mance sta­tus prior to this pre­sen­
ta­tion was excel­lent for his age.
AML with com­plex kar­yo­type and TP53 muta­tions appears to
have the worst out­comes of the sec­ond­ary AMLs. We esti­mate
the like­li­hood of achiev­ing remis­sion with stan­dard 7+3 and CPX351 to be about 30% in these cases (Table 2). This patient may
be con­sid­ered inel­i­gi­ble for stan­dard induc­tion by age alone,
but the bio­logic fea­tures of his dis­ease make a stron­ger case
for alter­na­tive ther­a­pies. In the phase 3 study of azacitidine and
Dr Prakash Singh Shekhawat
Optimizing out­comes in sec­ond­ary AML | 27
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/23/2021835/23matthews.pdf by guest on 09 December 2022
over­
all out­
comes appear deter­
mined by addi­
tional adverse
genetic fea­tures more so than by prior expo­sure to cyto­toxic
ther­a­pies.23 A ret­ro­spec­tive col­lec­tion of 69 patients with treat­
ment-related core bind­
ing fac­
tor leu­
ke­
mia revealed patients
with sec­ond­ary cyto­ge­netic abnor­mal­i­ties had lon­ger over­all
sur­vival than those with­out abnor­mal­i­ties. Presence of tri­somy 8
and tri­somy 22 and loss of X or Y chro­mo­some were asso­ci­ated
with improved sur­vival in both the post–cyto­toxic ther­apy and
de novo sub­sets.23 In these sit­u­a­tions, we advo­cate stan­dardof-care ther­a­pies with cytarabine-containing induc­tion reg­i­mens
plus gemtuzumab ozogamicin.24 This high­lights a prin­ci­ple that
the key fea­ture of sec­ond­ary AML is the spe­cific genetic changes
pres­ent rather than prior ther­apy or ante­ced­ent MDS or MPN.
In these cases, mon­i­tor­ing MRD via high-sen­si­tiv­ity poly­mer­
ase chain reac­tion (PCR)–based test­ing of the blood should be
performed to doc­u­ment an MRD-neg­a­tive remis­sion for ther­apy
to con­tinue with cura­tive intent con­sol­i­da­tion. RUNX1::RUNX1T1
or CBFB::MYH11 fusion genes cor­
re­
spond­
ing to t(8;21) or
inv(16/t(16;16) changes can be mon­i­tored by blood-level realtime quan­ti­ta­tive PCR with high sen­si­tiv­ity (10–4 to 10–6), allowing poten­tial early inter­ven­tion.25 For these assays to be most
reli­able, assess­ment via molec­u­lar assays at the time of diag­no­
sis is crit­i­cal to estab­lish the abil­ity to detect fusions that occa­
sion­ally can be missed if noncanonical gene fusions are pres­ent.
Even in the sec­
ond­
ary leu­
ke­
mia set­
ting, patients with oth­
er­
wise favor­able leu­ke­mias can have rea­son­able out­comes with
non-trans­plant-based ther­apy. Measurement of RUNX1-RUNX1T1
MRD via reverse tran­scrip­tion (RT)–PCR after course 1 of 7+3 with
gemtuzumab and course 2 of high-dose cytarabine revealed no
detect­able evi­dence of dis­ease, and the patient has com­pleted
4 courses of con­sol­i­da­tion with­out trans­plant in first CR.
Conclusions
Secondary AML rep­re­sents a het­er­og­e­nous group of patients
with fea­tures often overlapping de novo cases with sim­i­lar genet­
ics. Management and risks are often com­pa­ra­ble to de novo dis­
ease with sim­i­lar genetic fea­tures. Secondary AML encompasses
a wide spec­trum of molec­u­lar fea­tures. The his­tor­i­cally poor out­
comes in sec­ond­ary AML reflect the higher pro­por­tion of neg­
a­tive fea­tures like poor risk cyto­ge­net­ics and TP53 muta­tion.
Furthermore, patients are fre­quently older with com­plex med­
i­cal his­to­ries, lim­it­ing treat­ment options. The com­plete pro­fil­ing
of his­to­pa­thol­ogy and genetic driv­ers is crit­i­cal with the evo­lu­
tion of alter­na­tives to stan­dard 7+3 ther­apy. Patients with more
favor­able under­ly­ing cyto­ge­netic and molec­u­lar fea­tures may do
well with ther­apy, espe­cially those who are can­di­dates for allo­
ge­neic stem cell trans­plan­ta­tion. Overwhelmingly, patients have
dis­
ease with poor-risk fea­
tures. Only rarely do patients have
AML-pCT that is truly low risk. Maintenance ther­apy or trans­plan­
ta­tion is appro­pri­ate in most cases. Venetoclax-based reg­i­mens
have expanded treat­ment options for frailer, older patients with
sta­
ble comorbidities. Emerging treat­
ments seek to open the
ther­a­peu­tic win­dow by targeting spe­cific genetic vulnerabilities
or alter­ing the micro­en­vi­ron­ment. Current treat­ment deci­sions
28 | Hematology 2022 | ASH Education Program
require a care­ful inte­gra­tion of clin­i­cal his­tory, genetic fea­tures of
dis­ease, and eval­u­a­tion for allo­ge­neic trans­plant.
Conflict-of-inter­est dis­clo­sure
Andrew Mat­thews: no com­pet­ing finan­cial inter­ests to declare.
Keith W. Pratz: research funding from AbbVie, Agios, Daiichi Sankyo, and Millennium; advi­sory board mem­ber for AbbVie, Astellas, Bos­ton BioMedical, BMS, Celgene, Novartis, Jazz Pharmaceuticals, and Servier.
Off-label drug use
Midostaurin is approved in combination with standard cytarabine and daunorubicin induction and cytarabine consolidation.
Case discussion gives an example of a patient starting CPX-351
with addition of midostaurin after FLT3 mutation resulted.
Correspondence
Keith W. Pratz, University of Pennsylvania, PCAM South 12155, 3400 Civic Center Boulevard, Philadelphia, PA 19104, USA;
e-mail: keith​­.pratz@pennmedicine​­.upenn​­.edu.
References
1.
Granfeldt Østgård LS, Medeiros BC, Sengeløv H, et al. Epidemiology
and clin­
i­
cal sig­
nif­i­
cance of sec­
ond­
ary and ther­
apy-related acute mye­
loid leu­
ke­
mia: a national pop­
u­
la­
tion-based cohort study. J Clin Oncol.
2015;33(31):3641-3649.
2. Martínez-Cuadrón D, Megías-Vericat JE, Serrano J, et al. Treatment pat­terns
and out­comes of 2310 patients with sec­ond­ary acute mye­loid leu­ke­mia: a
PETHEMA reg­is­try study. Blood Adv. 2022;6(4):1278-1295.
3. Arber DA, Orazi A, Hasserjian R, et al. The 2016 revi­sion to the World Health
Organization clas­si­fi­ca­tion of mye­loid neo­plasms and acute leu­ke­mia.
Blood. 2016;127(20):2391-2405.
4. Khoury JD, Solary E, Abla O, et al. The 5th edi­tion of the World Health Organization clas­si­fi­ca­tion of haematolymphoid tumours: mye­loid and his­tio­
cytic/den­dritic neo­plasms. Leukemia. 2022;36(7):1703-1719.
5. Arber DA, Orazi A, Hasserjian RP, et al. International Consensus Classification of mye­loid neo­plasms and acute leu­ke­mia: inte­grat­ing mor­pho­log­i­cal,
clin­i­cal, and geno­mic data. Blood. 2022;140(11):1200-1228.
6. Lindsley RC, Mar BG, Mazzola E, et al. Acute mye­loid leu­ke­mia ontog­eny is
defined by dis­tinct somatic muta­tions. Blood. 2015;125(9):1367-1376.
7. Zink F, Stacey SN, Norddahl GL, et al. Clonal hema­to­poi­e­sis, with and
with­
out can­
di­
date driver muta­
tions, is com­
mon in the elderly. Blood.
2017;130(6):742-752.
8. Bolton KL, Ptashkin RN, Gao T, et al. Cancer ther­apy shapes the fit­ness
land­scape of clonal hema­to­poi­e­sis. Nat Genet. 2020;52(11):1219-1226.
9. McNerney ME, Godley LA, Le Beau MM. Therapy-related mye­
loid neo­
plasms: when genet­ics and envi­ron­ment col­lide. Nat Rev Cancer.
2017;17(9):513-527.
10. Kayser S, Döhner K, Krauter J, et al; Ger­man-Austrian AMLSG. The impact of
ther­apy-related acute mye­loid leu­ke­mia (AML) on out­come in 2853 adult
patients with newly diag­nosed AML. Blood. 2011;117(7):2137-2145.
11. Löwenberg B, Ossenkoppele GJ, van Putten W, et al; Dutch-Bel­gian Cooperative Trial Group for Hemato-Oncology (HOVON); Ger­man AML Study
Group (AMLSG); Swiss Group for Clinical Cancer Research (SAKK) Collaborative Group. High-dose dau­no­ru­bi­cin in older patients with acute mye­loid
leu­ke­mia. N Engl J Med. 2009;361(13):1235-1248.
12. Lancet JE, Uy GL, Cortes JE, et al. CPX-351 (cytarabine and dau­no­ru­bi­cin)
lipo­some for injec­tion ver­sus con­ven­tional cytarabine plus dau­no­ru­bi­cin in
older patients with newly diag­nosed sec­ond­ary acute mye­loid leu­ke­mia. J
Clin Oncol. 2018;36(26):2684-2692.
13. Nilsson C, Hulegårdh E, Garelius H, et al. Secondary acute mye­loid leu­ke­
mia and the role of allo­ge­neic stem cell trans­plan­ta­tion in a pop­u­la­tionbased set­ting. Biol Blood Marrow Transplant. 2019;25(9):1770-1778.
14. Lancet JE, Uy GL, Newell LF, et al. CPX-351 ver­sus 7+3 cytarabine and
dau­no­ru­bi­cin che­mo­ther­apy in older adults with newly diag­nosed
high-risk or sec­ond­ary acute mye­loid leu­kae­mia: 5-year results of a randomised, open-label, multicentre, phase 3 trial. Lancet Haematol.
2021;8(7):e481-e491.
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/23/2021835/23matthews.pdf by guest on 09 December 2022
venetoclax, the remis­sion rates were mod­est (55%), but the OS
was short (6.5 months).21 There are sev­eral novel strat­e­gies under
devel­op­ment spe­cif­i­cally in the TP53 sub­set, includ­ing TP53 acti­
vat­ing ther­apy in the form of eprenetapopt (APR-246). In early-phase stud­ies, the reported response rates were high33 but,
in sub­se­quent ran­dom­ized stud­ies com­pared with azacitidine
alone, failed to reach a pri­mary end point of improve­ment in CR
rates (Aprea press release 12/2020). The more recent strat­e­gies
involve the use of immune tar­get acti­va­tion. There are sev­eral
emerg­ing immu­no­ther­apy strat­e­gies, includ­ing immune check­
point block­ade (such as Tim-3), bispecific anti­body ther­a­pies
(such as flotetuzumab, an inves­
ti­
ga­
tional bispecific anti­
bodybased mol­e­cule to CD3ε and CD123), and the most advanced
can­di­date: mono­clo­nal antibodies to CD47 (such as magrolimab).
The expres­sion of CD47 on mye­loid cells acts to block immune
response to malig­nant cells directly at the site of phago­cy­to­
sis. Interestingly, CD47 mes­sen­ger RNA expres­sion varies across
cytogenic and molec­u­lar sub­groups, with lower expres­sion in
cases with t(8;21), a favor­
able-risk trans­
lo­
ca­
tion, and higher
expres­sion in cases har­bor­ing FLT3-ITD muta­tions. High CD47
expres­sion has been shown to be an inde­pen­dent prog­nos­tic
fac­tor for poor OS in AML patient cohorts.34 The mono­clo­nal anti­
body magrolimab has been devel­oped to block the CD47 sig­
nal­ing and in early-phase devel­op­ment shows prom­is­ing activ­ity
against TP53 AML, with an over­all response rate of 71% (48% CR
and 19% CR with incom­plete count recov­ery, 5% mor­pho­logic
leu­ke­mia-free state) in 21 patients.35 The defin­i­tive ran­dom­ized
phase 3 stud­ies in AML and MDS are enroll­ing. In addi­tion, trip­let
ther­apy with venetoclax and a hypomethylating agent has been
reported in abstract form, with a 64% CR rate in TP53-mutant
dis­ease. Fifty-five per­cent of these CRs were MRD neg­a­tive by
flow cytom­e­try, suggesting deep remis­sions in a sub­set of the
respond­ers.36 In this patient, we elected to enroll him on a clin­
i­cal trial exam­in­ing the effi­cacy of a CD47 mono­clo­nal anti­body
with the knowl­edge that trans­plant in this age group is gen­er­ally
not an option.
molec­u­lar alter­ations in adults with acute mye­loid leu­ke­mia. J Clin Oncol.
2012;30(36):4515-4523.
32. Koreth J, Schlenk R, Kopecky KJ, et al. Allogeneic stem cell trans­plan­ta­tion
for acute mye­loid leu­ke­mia in first com­plete remis­sion: sys­tem­atic review
and meta-anal­y­sis of pro­spec­tive clin­i­cal tri­als. JAMA. 2009;301(22):23492361.
33. Sallman DA, DeZern AE, Garcia-Manero G, et al. Eprenetapopt (APR-246)
and azacitidine in TP53-mutant myelodysplastic syn­dromes. J Clin Oncol.
2021;39(14):1584-1594.
34. Majeti R, Chao MP, Alizadeh AA, et al. CD47 is an adverse prog­nos­tic fac­tor
and ther­a­peu­tic anti­body tar­get on human acute mye­loid leu­ke­mia stem
cells. Cell. 2009;138(2):286-299.
35. Sallman D, Asch A, Kambhampati S, et al. AML-196: the first-in-class antiCD47 anti­body magrolimab in com­bi­na­tion with azacitidine is well tol­
er­ated and effec­tive in AML patients: phase 1b results. Clin Lymphoma
Myeloma Leuk. 2021;21:S290.
36. Daver N, Konopleva M, Maiti A, et al. Phase I/II study of azacitidine (AZA)
with venetoclax (VEN) and magrolimab (magro) in patients (pts) with
newly diag­
nosed older/unfit or high-risk acute mye­
loid leu­
ke­
mia (AML)
and relapsed/refrac­tory (R/R) AML [Abstract #371]. Paper presented at:
2021 Amer­i­can Society of Hematology Annual Meeting; 12 December 2021;
Atlanta, GA.
37. Lindsley RC, Gib­son CJ, Murdock HM, et al. Genetic char­ac­ter­is­tics and
out­comes by muta­tion sta­tus in a phase 3 study of CPX-351 ver­sus 7+3
in older adults with newly diag­nosed, high-risk/sec­ond­ary acute mye­loid
leu­ke­mia (AML). Blood. 2019;134(suppl 1):15.
38. Boddu P, Kantarjian H, Ravandi F, et al. Outcomes with lower inten­
sity ther­apy in TP53-mutated acute mye­loid leu­ke­mia. Leuk Lymphoma.
2018;59(9):2238-2241.
39. Wei AH, Montesinos P, Ivanov V, et al. Venetoclax plus LDAC for newly diag­
nosed AML inel­i­gi­ble for inten­sive che­mo­ther­apy: a phase 3 ran­dom­ized
pla­cebo-con­trolled trial. Blood. 2020;135(24):2137-2145.
40. Stone A, Zukerman T, Flaishon L, Yakar RB, Rowe JM. Efficacy out­comes in
the treat­ment of older or med­i­cally unfit patients with acute mye­loid leu­
kae­mia: a sys­tem­atic review and meta-anal­y­sis. Leuk Res. 2019;82:36-42.
41. Heuser M, Smith BD, Fiedler W, et al. Clinical ben­e­fit of glasdegib plus lowdose cytarabine in patients with de novo and sec­ond­ary acute mye­loid
leu­ke­mia: long-term anal­y­sis of a phase II ran­dom­ized trial. Ann Hematol.
2021;100(5):1181-1194.
42. Cortes JE, Heidel FH, Hellmann A, et al. Randomized com­par­i­son of low
dose cytarabine with or with­out glasdegib in patients with newly diag­
nosed acute mye­
loid leu­
ke­
mia or high-risk myelodysplastic syn­
drome.
Leukemia. 2019;33(2):379-389.
43. Brunner AM, Esteve J, Porkka K, et al. Efficacy and safety of sabatolimab (MBG453) in com­bi­na­tion with hypomethylating agents (HMAs) in
patients with acute mye­loid leu­ke­mia (AML) and high-risk myelodysplastic syn­drome (HR-MDS): updated results from a phase 1b study. Blood.
2020;136(suppl 1):1-2.
44. Vadakekolathu J, Lai C, Reeder S, et al. TP53 abnor­mal­i­ties cor­re­late with
immune infil­tra­tion and asso­ci­ate with response to flotetuzumab immu­no­
ther­apy in AML. Blood Adv. 2020;4(20):5011-5024.
© 2022 by The Amer­i­can Society of Hematology
DOI 10.1182/hema­tol­ogy.2022000324
Dr Prakash Singh Shekhawat
Optimizing out­comes in sec­ond­ary AML | 29
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/23/2021835/23matthews.pdf by guest on 09 December 2022
15. Pollyea DA, Pratz KW, Wei AH, et al. Outcomes in patients with poor-risk
cytogenetics with or without TP53 mutations treated with venetoclax and
azacitidine. Clin Cancer Res. 2022;25;CCR-22-1183.
16. Mat­thews AH, Perl AE, Luger SM, et al. Real-world effec­tive­ness of CPX351 vs venetoclax and azacitidine in acute mye­loid leu­ke­mia. Blood Adv.
2022;6(13):3997-4005.
17. Cherry EM, Abbott D, Amaya M, et al. Venetoclax and azacitidine com­pared
with induc­tion che­mo­ther­apy for newly diag­nosed patients with acute
mye­loid leu­ke­mia. Blood Adv. 2021;5(24):5565-5573.
18. Sengsayadeth S, Gatwood KS, Boumendil A, et al. Conditioning inten­sity
in sec­ond­ary AML with prior myelodysplastic syn­drome/mye­lo­pro­lif­er­a­tive
dis­or­ders: an EBMT ALWP study. Blood Adv. 2018;2(16):2127-2135.
19. Hourigan CS, Dillon LW, Gui G, et al. Impact of con­di­tion­ing inten­sity of
allo­ge­neic trans­plan­ta­tion for acute mye­loid leu­ke­mia with geno­mic evi­
dence of resid­ual dis­ease. J Clin Oncol. 2020;38(12):1273-1283.
20. Gilleece MH, Labopin M, Yakoub-Agha I, et al. Measurable resid­ual dis­
ease, con­di­tion­ing reg­i­men inten­sity, and age pre­dict out­come of allo­
ge­neic hema­to­poi­etic cell trans­plan­ta­tion for acute mye­loid leu­ke­mia in
first remis­sion: a reg­is­try anal­y­sis of 2292 patients by the Acute Leukemia
Working Party Euro­pean Society of Blood and Marrow Transplantation. Am
J Hematol. 2018;93(9):1142-1152.
21. DiNardo CD, Jonas BA, Pullarkat V, et al. Azacitidine and venetoclax in pre­
vi­ously untreated acute mye­loid leu­ke­mia. N Engl J Med. 2020;383(7):617629.
22. Duffield AS, Aoki J, Levis M, et al. Clinical and path­o­logic fea­tures of sec­
ond­ary acute promyelocytic leu­ke­mia. Am J Clin Pathol. 2012;137(3):395402.
23. Rogers HJ, Wang X, Xie Y, et al. Comparison of ther­apy-related and de
novo core bind­ing fac­tor acute mye­loid leu­ke­mia: a bone mar­row pathol­
ogy group study. Am J Hematol. 2020;95(7):799-808.
24. Hills RK, Castaigne S, Appelbaum FR, et al. Addition of gemtuzumab ozogamicin to induc­tion che­mo­ther­apy in adult patients with acute mye­loid
leu­kae­mia: a meta-anal­y­sis of indi­vid­ual patient data from randomised con­
trolled tri­als. Lancet Oncol. 2014;15(9):986-996.
25. Willekens C, Blanchet O, Renneville A, et al; French AML Intergroup. Prospective long-term min­i­mal resid­ual dis­ease mon­i­tor­ing using RQ-PCR in
RUNX1-RUNX1T1-pos­
i­
tive acute mye­
loid leu­
ke­
mia: results of the French
CBF-2006 trial. Haematologica. 2016;101(3):328-335.
26. Stone RM, Mandrekar SJ, Sanford BL, et al. Midostaurin plus che­mo­ther­
apy for acute mye­
loid leu­
ke­
mia with a FLT3 muta­
tion. N Engl J Med.
2017;377(5):454-464.
27. Döhner H, Estey E, Grimwade D, et al. Diagnosis and man­age­ment of AML
in adults: 2017 ELN rec­om­men­da­tions from an inter­na­tional expert panel.
Blood. 2017;129(4):424-447.
28. Schlenk RF, Döhner K, Krauter J, et al; Ger­man-Austrian Acute Myeloid Leukemia Study Group. Mutations and treat­ment out­come in cyto­ge­net­i­cally
nor­mal acute mye­loid leu­ke­mia. N Engl J Med. 2008;358(18):1909-1918.
29. Uy GL, Newell LF, Lin TL, et al. Transplant out­comes after CPX-351 vs 7+3 in
older adults with newly diag­nosed high-risk and/or sec­ond­ary AML. Blood
Adv. 2022;6(17):4989-4993.
30. Falini B, Sciabolacci S, Falini L, Brunetti L, Martelli MP. Diagnostic and ther­
a­
peu­
tic pit­
falls in NPM1-mutated AML: notes from the field. Leukemia.
2021;35(11):3113-3126.
31. Mrózek K, Marcucci G, Nicolet D, et al. Prognostic sig­nif­i­cance of the Euro­
pean LeukemiaNet stan­
dard­
ized sys­
tem for reporting cyto­
ge­
netic and
ACUTE MYELOID LEUKEMIA: IMPROVING OUTCOMES IN CHALLENGING SUBSETS
EVIDENCE-BASED MINIREVIEW
Brandon J. Aubrey and Andrew Brunner
Harvard Medical School and Massachusetts General Hospital, Boston, MA
LEARNING OBJECTIVES
• Understand the role of panel­based molecular testing at the time of diagnosis in AML
• Consider the characteristics and limitations of serial panel­based molecular testing in AML
CLINICAL CASE
A 52­year­old woman is diagnosed with acute myeloid
leukemia (AML). At diagnosis, pathology demonstrated
myelomonocytic morphology with a normal karyotype.
The patient underwent standard induction chemotherapy
with cytarabine and daunorubicin, achieving a complete
remission, and is currently having consolidation therapy
with high­dose cytarabine without an initial plan for allo­
geneic hematopoietic stem cell transplantation. What is
the utility of broad panel­based molecular testing at dif­
ferent time points during the management of this patient?
Introduction
One of the great advances in understanding the pathobi­
ology of AML has been through the discovery and charac­
terization of recurrent somatic gene mutations. This is a
rapidly evolving area where clinical decision­making may
now capitalize on knowledge of a patient’s mutational pro­
file for diagnostic subtyping, prognostication, and choice
of therapy, carried out in conjunction with standard histo­
pathologic and cytogenetic studies at the time of diag­
nosis. Somatic mutations can be identified using several
techniques, including targeted sequencing of a single gene,
RNA­ or DNA­based quantitative polymerase chain reaction
(qPCR) assays, panel­based next­generation sequencing
(NGS), and whole­exome or whole­genome sequencing,
each of which exhibits varying sensitivity and specificity
for the detection of gene mutations. Application of these
technologies has revealed the genomic complexity of AML,
with molecular heterogeneity between patients that may
evolve over time and in response to treatment. A strong
evidence base supports the testing of a small number of
30 | Hematology 2022 | ASH Education Program
specific gene mutations in routine clinical practice, includ­
ing NPM1, FLT3, and CEBPA, as they may influence diagnostic
subtyping and provide clear prognostic information in
patients with a normal karyotype on cytogenetic testing.1
Improved technology and declining cost for gene sequenc­
ing has provided more widely available panel­based gene
testing for patients with AML, providing accessible testing
for a larger number of genes. NGS­based gene panels used
in AML have grown to incorporate an ever­expanding set
of genes that typically include FLT3, NPM1, CEBPA, RUNX1,
isocitrate dehydrogenase 1 and 2 (IDH1/2), TET2, TP53, KIT,
WT1, and ASXL1, among other recurring gene mutations
found in myeloid malignancy (Table 1). While NGS­based
gene panels are able to detect a larger number of gene
mutations and comutations in a single patient, it can be
challenging to interpret and clinically use all of these data
points. In this evidence­based minireview, we assess the
decision to test a broad NGS panel, rather than individual
genes, in AML and how this decision may affect the diag­
nosis, prognosis, and/or treatment selection for a patient.
What role does molecular testing play
in establishing a diagnosis of AML?
AML diagnosis is primarily based on morphology and cyto­
genetics but increasingly takes molecular features into
consideration. Some AML subtypes now incorporate spe­
cific gene mutations into the diagnostic criteria, including
NPM1 and CEBPA.1­3 This is based on the observation that
these mutations are associated with distinct clinicopatho­
logic features that are reflected in patient outcomes. Nota­
bly, AML may now be diagnosed by the presence of NPM1
mutation irrespective of bone marrow myeloblast per­
centage, analogous to the finding of specific cytogenetic
abnormalities, such as t(8;21) and inv(16). An emerging
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/30/2021776/30aubrey.pdf by guest on 09 December 2022
Clinical utilization of panel-based molecular
testing for patients with AML
Table 1. Clinical util­ity of somatic gene muta­tion test­ing in AML and molec­u­lar fea­tures that may guide treat­ment strat­egy and
can result in a ben­e­fit to patient sur­vival
Gene
Clinical util­ity
NPM1
Diagnostic, good prog­no­sis in iso­la­tion, MRD mon­i­tor­ing
ITD con­fers poorer prog­no­sis, ther­a­peu­tic tar­get for ITD and TKD
Role in diag­no­sis, famil­ial risk, good prog­no­sis with in-frame bZIP
muta­tion
IDH1/2
Therapeutic tar­gets, pro­posed role in diag­no­sis, clonal hema­to­poi­e­sis
TP53
Very poor prog­no­sis, che­mo­ther­apy refrac­tory, may favor
hypomethylating agent-based treat­ment (azacitidine/decitabine) or
clin­i­cal trial
RUNX1
Poorer prog­no­sis, famil­ial risk
KIT
Poorer prog­no­sis in CBF AML, poten­tial ther­a­peu­tic tar­get
TET2
Associated with clonal hema­to­poi­e­sis
EZH2
Associated with sec­ond­ary AML
STAG2
Associated with sec­ond­ary AML
SRSF2
Associated with sec­ond­ary AML
BCOR
Associated with sec­ond­ary AML
SF3B1
Associated with sec­ond­ary AML
U2AF1
Associated with sec­ond­ary AML
ASXL1
Poor prog­no­sis, sec­ond­ary AML, clonal hema­to­poi­e­sis
DNMT3A
Associated with clonal hema­to­poi­e­sis
PTPN11
Associated with clonal hema­to­poi­e­sis, poorer prog­no­sis
Feature
Therapeutic con­fer­ring sur­vival advan­tage
FLT3
(ITD and TKD) midostaurin with induc­tion ther­apy, gilteritinib at relapse
IDH1⁺
Ivosidinib + azacitidine
sAML*
Cytarabine/dau­no­ru­bi­cin lipo­so­mal (Vyxeos; CPX-351)
bZIP, basic leu­cine zip­per region; CBF, core-bind­ing fac­tor; ITD, inter­nal tan­dem dupli­ca­tion; MRD, mea­sur­able resid­ual dis­ease; TKD, tyro­sine kinase
domain muta­tion.
⁺In the phase 3 study com­par­ing azacytidine and venetoclax to azacytidine monotherapy, patients with IDH1 and IDH2 muta­tions appeared to have
improved responses in sub­group anal­y­sis.
*sAML may be suggested by the find­ing of high muta­tion bur­den and/or EZH2, STAG2, SRSF2, BCOR, RUNX1, SF3B1, U2AF1, and ASXL1, but diag­no­sis
requires cor­re­la­tion with clin­i­cal his­tory, mor­phol­ogy, and cyto­ge­netic find­ings. Vyxeos has shown ben­e­fit to patients with sAML.
area is the under­stand­ing that com­pre­hen­sive molec­u­lar pro­
files may bet­ter char­ac­ter­ize the ori­gin of a given AML sub­type;
for instance, the detec­tion of gene muta­tions in SRSF2, SF3B1,
U2AF1, ZRSR2, ASXL1, EZH2, BCOR, or STAG2 are asso­ci­ated with
>95% spec­i­fic­ity for sec­ond­ary AML (sAML).4 However, to date,
the for­mal diag­no­sis of sAML uses the cor­re­la­tion of such molec­
u­lar fea­tures with the clin­i­cal his­tory, mor­phol­ogy, and cyto­ge­
netic find­ings. In the inves­ti­ga­tional set­ting, detailed molec­u­lar
pro­fil­ing of patients with AML has been shown to effec­
tively
sub­clas­sify patients into diverse, clin­i­cally rel­e­vant sub­groups
and pre­dicts the exis­tence of leu­ke­mia sub­types not cur­rently
known within the existing clas­si­fi­ca­tion sys­tem,5 suggesting the
strong poten­tial for future clin­i­cal use of in-depth molec­u­lar dis­
ease clas­si­fi­ca­tion. At this time, it is impor­tant to rec­og­nize the
expanding num­ber of muta­tions nec­es­sary for a com­plete AML
clas­si­fi­ca­tion; although sin­gle-gene muta­tion test­ing can be
done for a min­i­mal set of genes, panel sequenc­ing increas­ingly
has an advan­tage in this domain.
What role does molec­u­lar test­ing play in predicting
a patient’s prog­no­sis?
FLT3, NPM1, and CEBPA were iden­ti­fied early as hav­ing a role
in risk strat­
i­
fi­
ca­
tion of patients with AML who have nor­
mal
­cyto­ge­net­ics,6 effec­tively iden­ti­fy­ing patient risk of relapse and
sur­vival after che­mo­ther­apy.1,2 AML with an iso­lated NPM1 muta­
tion pre­dicts for more favor­able-risk dis­ease with good response
to stan­dard che­mo­ther­apy.7 Similarly, in-frame muta­tions within
the basic leu­
cine zip­
per region of the CEBPA gene iden­
tify
patients with favor­able-risk dis­ease.8,9 In con­trast, the pres­ence
of FLT3 gene inter­nal tan­dem dupli­ca­tion pre­dicts a higher risk
of relapse. Recent data sug­gest that broader muta­tion pro­files—
often facil­i­tated through panel-based gene test­ing approaches—
can improve upon risk strat­i­fi­ca­tion and dis­ease prog­no­sis
(Table 1). For exam­ple, TP53 muta­tions con­fer a dis­tinctly poor
prog­no­sis, have a high risk for fail­ure with stan­dard che­mo­ther­
apy, and may prompt alter­na­tive treat­ment approaches such
as those with a hypomethylating agent back­bone.10,11 RUNX1,12
Dr Prakash Singh Shekhawat
Molecular test­ing for patients with AML | 31
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/30/2021776/30aubrey.pdf by guest on 09 December 2022
FLT3
CEBPA
ASXL1,13 and other sec­ond­ary-like muta­tions may also indi­cate
higher risk of patient relapse or ther­apy resis­tance. In addi­tion,
high muta­tional bur­den or the detec­tion of cer­tain muta­tions
(Table 1) may sug­gest a diag­no­sis of sAML, which has clin­i­cal
impor­tance, in that these gene muta­tions iden­tify a group of
patients with very poor prog­no­sis, and increas­ingly guides the
choice of induc­tion che­mo­ther­apy and con­sol­i­da­tion strat­egy.
Can gene panel test­ing affect treat­ment selec­tion
for AML dif­fer­ently than test­ing for sin­gle muta­tions?
Where is molec­u­lar test­ing likely to play an impact after
ini­tial AML diag­no­sis?
Molecular test­ing has a grow­ing role fol­low­ing a patient’s ini­
tial diag­no­sis dur­ing treat­ment. There is a strong inter­est and
active research to opti­mize molec­u­lar mon­i­tor­ing of mea­sur­able
resid­ual dis­ease to guide postremission treat­ment strat­e­gies for
AML.18 In the set­ting of NPM1-mutant AML, spe­cific detec­tion of
the mutant NPM1 tran­script by qPCR is a pow­er­ful indi­ca­tor of
relapse risk that can be used to risk-adapt patient treat­ment.19 A
lim­i­ta­tion to the uni­ver­sal appli­ca­tion of such tech­niques is the
genetic het­ero­ge­ne­ity of AML; NGS-based gene panel assays are
being eval­u­ated as a way to mon­i­tor dis­ease over time and are
bet­ter ­able to track clonal het­ero­ge­ne­ity com­pared with genespe­cific qPCR assays. This is rel­e­vant as most patients with AML
will not have a muta­tion as easy to track as NPM1. If such track­
ing is val­i­dated, NGS pan­els have an advan­tage com­pared with
qPCR assays because they assess a larger selec­tion of genes, and
most patients will have at least 1 genetic alter­ation detected by
the panel assay that can be mon­i­tored over time.5 Furthermore,
broad NGS-based gene panel test­ing enables detec­tion of the
emer­gence of addi­tional molec­u­lar abnor­mal­i­ties over time that
may not have been pres­ent at the time of diag­no­sis. NGS-based
mea­sur­able resid­ual dis­ease mon­i­tor­ing faces sev­eral lim­i­ta­tions,
includ­ing rel­a­tively lower sen­si­tiv­ity com­pared with ded­i­cated
qPCR assays, includ­ing a lack of stan­dard­i­za­tion, and sig­nif­i­cant
chal­
lenges related to distinguishing what rep­
re­
sents resid­
ual
leu­ke­mia from what may be back­ground clonal hema­to­poi­e­sis
(clonal hematopoiesis of indeterminate potential or clonal cyto­
penias of unknown significance).20 At this time, inter­
val NGS
32 | Hematology 2022 | ASH Education Program
What is the role of molec­u­lar test­ing in patients with AML
after relapse?
Finally, AML relapse after che­mo­ther­apy, like at diag­no­sis, rep­re­
sents a time point to reassess dis­ease in order to develop a sal­
vage strat­egy. In con­trast to mon­i­tor­ing in remis­sion, at relapse,
there is more empha­sis on imme­di­ately action­able muta­tions.
These include muta­
tions in FLT3, IDH1, and IDH2. Whether a
broad NGS-based panel offers sig­nif­i­cant advan­tage over tar­
geted sequenc­ing at relapse is less clear, how­ever. Particularly
for late relapses, NGS can help to clar­
ify the rela­
tion of the
relapsed leu­ke­mia to the ini­tial diag­no­sis. Since most dis­ease at
relapse has a uni­ver­sal poor prog­no­sis, risk assess­ment has less
imme­di­ate util­ity. It is none­the­less impor­tant to assess for FLT3
muta­tions, which can emerge dur­ing treat­ment and for which
sev­eral targeted ther­a­pies are avail­­able. Mutations in IDH1 or
IDH2 appear more sta­ble over time but can be sub­ject to sub­
clonal evo­lu­tion that may evade detec­tion at low lev­els early in
the dis­ease course. An increas­ing num­ber of targeted ther­a­pies
in clin­i­cal devel­op­ment also require the pres­ence of a mutated
gene tar­get; given the impor­tance of clin­i­cal tri­als as an option
for patients with relapsed or refrac­tory AML, such test­ing can be
infor­ma­tive.
Conclusions
Gene panel test­ing, now com­monly performed by NGS-based
tech­nol­ogy, is an increas­ingly avail­­able tool for patients with
AML. We rec­om­mend all­patients with AML have broad gene
panel test­ing at diag­no­sis, which includes the most com­mon
recur­rently mutated genes in this dis­ease where test­ing may
influ­ence diag­no­sis, prog­no­sis, and treat­ment selec­tion. At this
time, there is an evolv­ing evi­dence base for serial NGS-based
broad gene panel test­ing dur­ing treat­ment in rou­tine clin­i­cal
prac­tice. We await robust evi­dence to sup­port the rou­tine use
of this out­side of a clin­i­cal trial, although it may be valu­able in
spe­cific clin­i­cal sce­nar­ios such as prior to allo­ge­neic trans­plant.
Many open ques­tions remain, includ­ing opti­mal tim­ing for serial
panel muta­tion assess­ment, thresh­olds for detec­tion in NGSbased meth­ods that bal­ance sen­si­tiv­ity and spec­i­fic­ity, and bet­
ter defin­ing which gene muta­tions, and com­bi­na­tions thereof,
are most pre­dic­tive of patient out­comes. Finally, a major bar­rier
remains regard­ing the acces­si­bil­ity, cost, and stan­dard­i­za­tion
of these tech­niques to make them more acces­si­ble in rou­tine
­clin­i­cal prac­tice.
Recommendations for NGS-based broad molec­u­lar
test­ing in AML
1. Broad NGS-based gene panel test­ing should be performed at
the time of AML diag­no­sis. (Strong rec­om­men­da­tion, mod­er­
ate cer­tainty of the evi­dence)
2. At AML relapse, patients should have molec­u­lar test­ing repeat­
ed, although it is less clear whether NGS-based panel test­ing
has an advan­tage over test­ing sin­gle action­able genes at this
time. (Strong rec­om­men­da­tion for test­ing in gen­eral, low cer­
tainty of evi­dence between test­ing modal­i­ties)
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/30/2021776/30aubrey.pdf by guest on 09 December 2022
Several muta­tion pro­files affect the ini­tial treat­ment strat­egy. The
pres­ence of FLT3 inter­nal tan­dem dupli­ca­tion or FLT3 muta­tions in
the tyro­sine kinase domain, for instance, informs the use of drug
ther­a­pies that tar­get the FLT3 kinase. Incorporating FLT3 inhib­i­
tors with stan­dard che­mo­ther­apy pro­vi­des a sur­vival advan­tage
for patients with this AML sub­type.14 More chal­leng­ing per­haps is
the entity of sAML, which may be more variably defined but for
which CPX-351 (Vyxeos; Jazz Pharmaceuticals) may be advan­ta­
geous to stan­dard induc­tion ther­apy.15 A grow­ing num­ber of avail­­
able targeted ther­a­pies may be bet­ter suited to other genetic
alter­ations. In patients with IDH1/2 muta­tions, small-mol­e­cule
inhib­
i­
tors of IDH1 or IDH2, com­
bined with azacitidine, can
improve out­comes in select patients,16 although these patients
also have high response rates to other reg­i­mens such as azac­
itidine in com­bi­na­tion with venetoclax.17 There is also evi­dence
that TP53-mutated AML—and per­haps other sec­ond­ary-like
muta­tions—may be equally if not more respon­sive to alter­na­tive
ther­a­pies to induc­tion, such as hypomethylating agent-based
ther­a­pies. With the expanding avail­abil­ity of targeted drug ther­
a­pies, the value of panel-based gene test­ing is likely to expand.
test­
ing in remis­
sion remains under inves­
ti­
ga­
tion but may be
most valu­able at the time of clear treat­ment deci­sions where
esti­ma­tion of patient risk is clin­i­cally impor­tant, such as with allo­
ge­neic trans­plant.21
3. At this time, rou­tine serial NGS-based gene panel test­ing for
patients with AML in remis­
sion remains inves­
ti­
ga­
tional, al­
though may have value in spe­cific sce­nar­ios. (Low cer­tainty
of evi­dence)
Conflict-of-inter­est dis­clo­sure
Off-label drug use
Brandon J. Aubrey: nothing to disclose.
Andrew Brunner: nothing to disclose.
Correspondence
Andrew Brunner, Harvard Medical School and Mas­
sa­
chu­
setts
General Hospital, Zero Emerson Place, Suite 118, Bos­
ton, MA
02114, USA; e-mail: abrunner@mgh​­.harvard​­.edu
References
1.
2.
3.
4.
5.
6.
7.
Khoury JD, Solary E, Abla O, et al. The 5th edi­tion of the World Health Orga­
nization Classification of Haematolymphoid Tumours: mye­loid and Histio­
cytic/Dendritic Neoplasms. Leukemia. 2022;36(7):1703-1719.
Döhner H, Estey E, Grimwade D, et al. Diagnosis and man­age­ment of AML
in adults: 2017 ELN rec­om­men­da­tions from an inter­na­tional expert panel.
Blood. 2017;129(4):424-447.
Arber DA, Orazi A, Hasserjian RP, et al. International con­sen­sus clas­si­fi­ca­
tion of mye­loid neo­plasms and acute leu­ke­mia: inte­grat­ing mor­pho­log­i­cal,
clin­i­cal, and geno­mic data. Blood. 2022;140(11):1200-1228.
Lindsley RC, Mar BG, Mazzola E, et al. Acute mye­loid leu­ke­mia ontog­eny is
defined by dis­tinct somatic muta­tions. Blood. 2015;125(9):1367-1376.
Papaemmanuil E, Gerstung M, Bullinger L, et al. Genomic clas­si­fi­ca­tion and
prog­no­sis in acute mye­loid leu­ke­mia. N Engl J Med. 2016;374(23):22092221.
Schlenk RF, Döhner K, Krauter J, et al; Ger­man-Austrian Acute Myeloid Leu­
kemia Study Group. Mutations and treat­ment out­come in cyto­ge­net­i­cally
nor­mal acute mye­loid leu­ke­mia. N Engl J Med. 2008;358(18):1909-1918.
Falini B, Mecucci C, Tiacci E, et al; GIMEMA Acute Leukemia Working Party.
Cytoplasmic nucleophosmin in acute mye­log­e­nous leu­ke­mia with a nor­mal
kar­yo­type. N Engl J Med. 2005;352(3):254-266.
© 2022 by The Amer­i­can Society of Hematology
DOI 10.1182/hema­tol­ogy.2022000417
Dr Prakash Singh Shekhawat
Molecular test­ing for patients with AML | 33
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/30/2021776/30aubrey.pdf by guest on 09 December 2022
Brandon J. Aubrey is a for­mer employee of the Walter and Eliza
Hall Institute, Melbourne, Australia, and has received pay­ments
relat­
ing to roy­
al­
ties and mile­
stone pay­
ments for the BCL-2
antag­o­nist venetoclax/ABT-199.
Andrew Brunner is a con­
sul­
tant for Acceleron, Agios,
BMS/Celgene, CTI Biopharma, Gilead, Keros Therapeutics, No­
vartis, Taiho, and Takeda.
8. Taube F, Georgi JA, Kramer M, et al; Study Alliance Leukemia (SAL). CEBPA
muta­tions in 4708 patients with acute mye­loid leu­ke­mia: dif­fer­en­tial impact
of bZIP and TAD muta­tions on out­come. Blood. 2022;139(1):87-103.
9. Wakita S, Sakaguchi M, Oh I, et al. Prognostic impact of CEBPA bZIP domain
muta­tion in acute mye­loid leu­ke­mia. Blood Adv. 2022;6(1):238-247.
10. Welch JS, Petti AA, Miller CA, et al. TP53 and decitabine in acute mye­loid
leu­ke­mia and myelodysplastic syn­dromes. N Engl J Med. 2016;375(21):20232036.
11. Grob T, Al Hinai ASA, Sanders MA, et al. Molecular char­
ac­
ter­
iza­
tion of
mutant TP53 acute mye­loid leu­ke­mia and high-risk myelodysplastic syn­
drome. Blood. 2022;139(15):2347-2354.
12. Mendler JH, Maharry K, Radmacher MD, et al. RUNX1 muta­tions are asso­ci­
ated with poor out­come in youn­ger and older patients with cyto­ge­net­i­
cally nor­mal acute mye­loid leu­ke­mia and with dis­tinct gene and microRNA
expres­sion sig­na­tures. J Clin Oncol. 2012;30(25):3109-3118.
13. Schnittger S, Eder C, Jeromin S, et al. ASXL1 exon 12 muta­tions are fre­quent
in AML with inter­me­di­ate risk kar­yo­type and are inde­pen­dently asso­ci­ated
with an adverse out­come. Leukemia. 2013;27(1):82-91.
14. Stone RM, Mandrekar SJ, Sanford BL, et al. Midostaurin plus che­mo­ther­
apy for acute mye­
loid leu­
ke­
mia with a FLT3 muta­
tion. N Engl J Med.
2017;377(5):454-464.
15. Lancet JE, Uy GL, Cortes JE, et al. CPX-351 (cytarabine and dau­no­ru­bi­cin)
lipo­some for injec­tion ver­sus con­ven­tional cytarabine plus dau­no­ru­bi­cin in
older patients with newly diag­nosed sec­ond­ary acute mye­loid leu­ke­mia. J
Clin Oncol. 2018;36(26):2684-2692.
16. Montesinos P, Recher C, Vives S, et al. Ivosidenib and azacitidine in
IDH1-mutated acute mye­loid leu­ke­mia. N Engl J Med. 2022;386(16):15191531.
17. DiNardo CD, Jonas BA, Pullarkat V, et al. Azacitidine and venetoclax in pre­
vi­ously untreated acute mye­loid leu­ke­mia. N Engl J Med. 2020;383(7):617629.
18. Ghannam J, Dillon LW, Hourigan CS. Next-gen­er­a­tion sequenc­ing for mea­
sur­able resid­ual dis­ease detec­tion in acute mye­loid leu­kae­mia. Br J Haematol. 2020;188(1):77-85.
19. Ivey A, Hills RK, Simpson MA, et al; UK National Cancer Research Institute
AML Working Group. Assessment of min­i­mal resid­ual dis­ease in stan­dardrisk AML. N Engl J Med. 2016;374(5):422-433.
20. Pløen GG, Nederby L, Guldberg P, et al. Persistence of DNMT3A muta­
tions at long-term remis­sion in adult patients with AML. Br J Haematol.
2014;167(4):478-486.
21. Getta BM, Devlin SM, Levine RL, et al. Multicolor flow cytom­e­try and multi­
gene next-gen­er­a­tion sequenc­ing are com­ple­men­tary and highly pre­dic­
tive for relapse in acute mye­loid leu­ke­mia after allo­ge­neic trans­plan­ta­tion.
Biol Blood Marrow Transplant. 2017;23(7):1064-1071.
ANXIETY PROVOKING CONSULTATIONS: MAST CELLS AND EOSINOPHILS
Jason Gotlib
Division of Hematology, Stanford Cancer Institute/Stanford University School of Medicine, Stanford, CA
The historically poor prognosis of patients with advanced systemic mastocytosis (AdvSM) and primary eosinophilic neoplasms has shifted to increasingly favorable outcomes with the discovery of druggable targets. The multikinase/KIT
inhibitor midostaurin and the highly selective KIT D816V inhibitor avapritinib can elicit marked improvements in measures of mast cell (MC) burden as well as reversion of MC-mediated organ damage (C-findings) and disease symptoms.
With avapritinib, the achievement of molecular remission of KIT D816V and improved survival compared with historical
therapy suggests a potential to affect disease natural history. BLU-263 and bezuclastinib are KIT D816V inhibitors currently being tested in trials of AdvSM. In the new World Health Organization and International Consensus Classifications,
the category of “myeloid/lymphoid neoplasms with eosinophilia and tyrosine kinase (TK) gene fusions” is inclusive of
rearrangements involving PDGFRA, PDGFRB, FGFR1, JAK2, FLT3, and ETV6::ABL1. While the successful outcomes with imatinib in FIP1L1::PDGFRA-positive cases and PDGFRB-rearranged neoplasms have become the “poster children” of these
disorders, the responses of the other TK-driven neoplasms to small-molecule inhibitors are more variable. The selective
FGFR inhibitor pemigatinib, approved in August 2022, is a promising therapy in aggressive FGFR1-driven diseases and
highlights the role of such agents in bridging patients to allogeneic transplantation. This review summarizes the data for
these approved and investigational agents and discusses open questions and future priorities regarding the management of these rare diseases.
LEARNING OBJECTIVES
• Summarize the efficacy and safety data for targeted therapy of advanced systemic mastocytosis and primary
eosinophilic neoplasms
• Discuss the impact of these novel agents on current treatment algorithms and how they shape future management priorities
CLINICAL CASE
A 61-year-old man with systemic mastocytosis with chronic
myelomonocytic leukemia (SM-CMML) presented with
severe fatigue and a 30-pound weight loss over the prior
6 months. He did not respond to PEG-interferon-alfa-2a. For
several months, his paracenteses requirement increased
to twice weekly for a volume of 10 to 15 L. Examination
was notable for temporal wasting and marked atrophy of
muscles in the bilateral supraclavicular regions. Imaging
findings of hepatosplenomegaly were not palpable on
examination due to his marked ascites. Laboratory studies revealed a white blood cell count of 17.8 × 109/L (51%
monocytes), hemoglobin of 11.9 g/dL, platelet count of
34 | Hematology 2022 | ASH Education Program
106 × 109/L, albumin of 2.3 g/dL (normal, 3.5-5.0 g/dL), and
serum tryptase level of 416 ng/mL (normal, <11.5 ng/mL).
The bone marrow biopsy specimen revealed 30% mast
cells, increased monocytes, and 8% blasts/blast equivalents; karyotype revealed del(13q). Next-generation
sequencing revealed KIT D816V and TET2 Q243X mutations (variant allele frequencies of 11% and 44%, respectively). The patient was referred for management.
Advanced systemic mastocytosis
Systemic mastocytosis (SM), defined by established
World Health Organization (WHO) criteria, is divided into
nonadvanced forms (indolent SM [ISM], bone marrow
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/34/2021847/34gotlib.pdf by guest on 09 December 2022
Available and emerging therapies for bona fide
advanced systemic mastocytosis and primary
eosinophilic neoplasms
from both the MC and AHN com­part­ments.14,15,27,28 The expe­ri­
ence with midostaurin and avapritinib has val­i­dated this clin­i­cal
strat­egy but has also revealed chal­lenges of KIT inhib­i­tor monotherapy. In par­tic­u­lar, patients with SM-AHN and/or cases with a
com­plex molec­u­lar land­scape are vul­ner­a­ble to AHN pro­gres­sion
or trans­for­ma­tion to sec­ond­ary AML.4,5,22-24
Midostaurin
Midostaurin (Novartis) is a multikinase/KIT inhib­i­tor that tar­
gets not only D816V-mutated KIT but also wild-type (WT) KIT,
PDGFRα/β, VEGFR2, and FLT3.29 Based on an encour­
ag­
ing
­par­tial response to midostaurin in a patient with SM-MDS/MPNunclas­si­fi­able30 and an over­all response rate (ORR) of 69% in
an inves­ti­ga­tor-ini­ti­ated trial,31 the drug was eval­
u­
ated in a
global registrational, phase 2, sin­gle-am, open-label study of
89 evaluable patients with AdvSM with ≥1 C-find­ings. The ORR
was 60% (45% major responses) by (mod­
i­
fied) Valent and
Cheson cri­te­ria.32 The Euro­pean Medicines Agency (EMA) and
FDA performed post hoc ana­ly­ses of the trial using International
Working Group–Myeloproliferative Neoplasms Research and
Treatment-Euro­
pean Competence Network on Mastocytosis
(IWG) response cri­te­ria. The EMA and FDA iden­ti­fied IWG ORRs
of 28% and 17%, respec­
tively, the dif­
fer­
ence reflecting the
FDA’s deci­sion not to include the cat­e­gory of “clin­i­cal improve­
ment” in their cal­cu­la­tion of response rate.33,34 Midostaurin elicited a median best per­cent reduc­tion of serum tryptase level
and BM MC bur­den by −58% and −59%, respec­tively; reduced
spenomegaly; and elicited a sig­nif­i­cant reduc­tion in symp­toms
(using the Memorial Symptom Assessment Scale) except for
nau­sea and vomiting, which are com­mon midostaurin-related
adverse events.32,35 The median pro­gres­sion-free sur­vival was
14.1 months, and the over­all sur­vival (OS) was 28.7 months.32
These data led to midostaurin’s approval for advanced SM in
2017 by the FDA and EMA. Additional real-world cohorts of
AdvSM patients have cor­
rob­
o­
rated midostaurin’s activ­
ity,
includ­ing an anal­y­sis that showed that a ≥25% reduc­tion in KIT
D816V RNA expressed allele bur­den was the stron­gest var­i­able
asso­ci­ated with prolonged sur­vival.36-39 A ret­ro­spec­tive reg­is­
try anal­y­sis also con­firmed the supe­rior sur­vival of midostaurin
com­pared with cladribine.40
Avapritinib
Avapritinib (BLU-265; Blueprint Medicines) was designed as a
highly selec­tive, type 1 inhib­i­tor of D816V-mutated KIT.41 It exhib­
its a 10-fold lower 50% inhib­i­tory con­cen­tra­tion (IC50) ­com­pared
with midostaurin in an assay of KIT D816V kinase activ­ity (0.27
vs 2.9nM). Avapritinib’s restricted tar­
get pro­
file also includes
plate­
let-derived growth fac­
tor recep­
tor alpha (PDGFRα) with
neg­
­
li­
gi­
ble activ­
ity against WT KIT; the drug received FDA
approval in 2020 for adults with an unresectable or met­a­static
gas­tro­in­tes­ti­nal stro­mal tumor har­bor­ing a PDGFRA exon 18
muta­tion, includ­ing the D842V muta­tion.42
Avapritinib’s FDA approval in 202143 as first-line ther­apy for
adults with AdvSM (recommended plate­
let count ≥50×109/L)
was based on the phase 1 EXPLORER study44 and an interim
anal­y­sis of the phase 2 PATHFINDER study.45 The centrally adju­
di­cated EXPLORER study consisted of a dose esca­la­tion phase
eval­u­at­ing doses of 30 to 400 mg daily in patients with AdvSM
with ≥1 eli­gi­ble organ dam­age find­ing by IWG cri­te­ria. The sub­
se­quent dose expan­sion phase eval­u­ated 2 dos­ing cohorts of
200 mg and 300 mg daily. Ultimately, 200 mg daily was cho­
Therapies for advanced SM & eosin­o­philic neo­plasms | 35
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/34/2021847/34gotlib.pdf by guest on 09 December 2022
mastocytosis, smol­
der­
ing SM) and advanced SM var­
i­
ants
(aggres­
sive SM [ASM], SM with an asso­
ci­
ated hema­
to­
logic
neo­plasm [SM-AHN], and mast cell leu­ke­mia [MCL]).1-3 ASM is
defined by ≥1 C-find­ings (neo­plas­tic mast cell [MC]–related
organ dam­age). Diagnostic cri­te­ria for SM-AHN include the
pres­ence of SM and an asso­ci­ated (mye­loid) neo­plasm (most
com­monly CMML),4 and MCL is his­to­path­o­log­i­cally defined by
≥20% MCs on a bone mar­row (BM) aspi­rate.1-4 Advanced SM
(AdvSM) var­i­ants carry a poor prog­no­sis (typ­i­cally less than
4 years),5-9 and the worst out­comes are observed in patients
with MCL (eg, <6 months to ~2 years).5,10-13 In both the International Consensus Classification (ICC) and WHO fifth edi­tion,1,2
mod­i­fi­ca­tions to cur­rent minor diag­nos­tic cri­te­ria include
addi­tion of CD30 as an immu­no­his­to­chem­i­cal marker of SM,
and an acti­vat­ing muta­tion in KIT besides D816V may also
qual­ify. In the ICC,1 “AHN” is changed to “AMN” (asso­ci­ated
mye­loid neo­plasm), and a core biopsy spec­i­men may be used
to diag­nose MCL if the aspi­rate is a dry tap. While bone mar­
row mastocytosis is maintained as a sub­type of ISM in the ICC,1
it is a dis­tinct sub­type from ISM in the WHO clas­si­fi­ca­tion.2
Using high-sen­si­tiv­ity assays such as dig­i­tal drop­let poly­mer­
ase chain reac­tion (PCR) and allele-spe­cific PCR, KIT D816V can
be detected in ~95% of patients with AdvSM.14,15 Compared with
nonadvanced SM, AdvSM var­i­ants are more fre­quently char­ac­
ter­ized by multilineage involve­ment by KIT D816V and a com­
plex, multimutated genetic land­
scape, best exem­
pli­
fied by
SM-AHN.16-18 SRSF2, ASXL1, and RUNX1 (S/A/R panel) are high-risk
muta­tions,19,20 but other mye­loid muta­tions are com­monly iden­
ti­fied (eg, TET2, DNMT3A, JAK2, NRAS, CBL, EZH2),15 some of
which have been incor­po­rated into prog­nos­tic scor­ing sys­tems
based on clin­i­cal, lab­o­ra­tory, and molec­u­lar fea­tures.6-9
Historically, ther­apy of AdvSM has consisted of off-label use
of PEG-inter­feron (IFN) α±cor­ti­co­ste­roids or cladribine. Studies
eval­u­at­ing these agents have often consisted of an admix­ture of
patients with nonadvanced and AdvSM with over­all responses in
the range of 30% to 50% using het­er­og­e­nous response cri­te­ria and
var­i­able reporting of bio­mark­ers of response, includ­ing changes
in BM MC bur­
den and serum tryptase level.21-24 Imatinib was
approved by the US Food and Drug Administration (FDA) in 2005
for patients with ASM with­out KIT D816V or unknown muta­tion
sta­tus, an exceed­ingly rare patient pop­u­la­tion. While exon 17 KIT
D816V is an imatinib-resis­tant muta­tion, the drug can be effec­tive
in patients with SM with exon 8 to 11 muta­tions/juxtamembrane
var­i­ants such as F522C, which is asso­ci­ated with a well-dif­fer­en­ti­
ated SM phe­no­type.25 Acute mye­loid leu­ke­mia (AML)–type induc­
tion che­mo­ther­apy has been used in patients with kinet­i­cally
aggres­sive or refrac­tory/relapsed dis­ease to the afore­men­tioned
ther­
a­
pies. For patients with SM-AHN, in whom AHN-directed
ther­apy is required, hydroxy­urea has been used to con­trol leu­
ko­cy­to­sis and spleno­meg­aly. Hypomethylating agents are com­
monly employed when the AHN is a higher-risk myelodysplastic
syn­drome (MDS) or MDS/mye­lo­pro­lif­er­a­tive neo­plasm (MPN).
A ret­ro­spec­tive anal­y­sis of allo­ge­neic hema­to­poi­etic stem cell
trans­plan­ta­tion (HSCT) revealed a 3-year over­all sur­vival of 57%
among 57 patients, with a diag­no­sis of MCL and reduced inten­
sity (vs myeloablative) con­di­tion­ing being iden­ti­fied as adverse
prog­nos­tic fac­tors.26 These data were derived before the use of
KIT inhib­i­tors; there­fore, the role of trans­plant in the mod­ern KIT
inhib­i­tor era is not well defined.
The ratio­nale for KIT inhi­bi­tion in AdvSM is based on the high
fre­quency of the KIT D816V muta­tion and its pres­ence in cells
Table 1. Response rates to avapritinib in the EXPLORER and PATHFINDER stud­ies
Characteristic
Phase 1 EXPLORER
Phase 2 Interim PATHFINDER
Response evaluable, n
53
32
ASM
3
2
SM-AHN
37
26
MCL
13
4
75
75
CR
15
0
CRh
21
19
PR
34
31
CI
6
25
SD
23
13
PD
0
3
NE
2
9
ASM
100
100
SM-AHN
76
81
MCL
69
25
Yes
69
74
No
86
78
Prior midostaurin
59
82
Midostaurin naive
83
67
Yes
74
71
No
77
80
AdvSM sub­type, n
ORR by mod­i­fied IWG cri­te­ria, % (CR + CRh + PR + CI)
Best response, %
ORR by any prior ther­apy, %
ORR by midostaurin his­tory, %
ORR by S/A/R comutation sta­tus, %
CI, clin­i­cal improve­ment; CRh, com­plete remis­sion with par­tial hema­to­logic recov­ery; NE, not evaluable; PD, pro­gres­sive dis­ease; S/A/R, SRSF2,
ASXL1, RUNX1; SD, sta­ble dis­ease.
sen as the recommended phase 2 dose based on a com­pos­ite
anal­y­sis of safety/tol­er­a­bil­ity, phar­ma­co­ki­net­ics, effi­cacy, and
bio­mark­ers of response, includ­ing dynamic changes in BM MC
bur­den and serum tryptase lev­els. Table 1 shows the over­all rates
of response by mod­i­fied IWG cri­te­ria, AdvSM var­i­ant, prior ther­
apy and midostaurin expo­sure, and S/A/R muta­tion sta­tus in the
EXPLORER study (n=69 evaluable patients)44 and from an interim
anal­
y­
sis of PATHFINDER (n=32 evaluable patients).45 Figure 1
high­lights the marked decreases in mea­sures of MC dis­ease from
the EXPLORER study, includ­ing BM MC bur­den, serum tryptase
level, and KIT D816V var­i­ant allele frac­tion from BM using dig­i­
tal drop­let PCR with a limit of detec­tion of 0.17%. Importantly,
a com­
plete molec­
u­
lar remis­
sion (CMR) was achieved in 30%
of patients, which rep­re­sents a new response bench­mark for
AdvSM. Reversion of MC skin lesions and sig­nif­i­cant improve­
ment in symp­toms using the AdvSM Symptom Assessment Form
were also observed.44-46
An anal­y­sis of pooled out­comes in 53 patients treated with
an avapritinib starting dose of ≤200 mg daily revealed an ORR
36 | Hematology 2022 | ASH Education Program
of 72% (com­plete response [CR]/CR with par­tial hema­to­logic
recov­ery [28%] + par­tial response [PR] [28%] + clin­i­cal improve­
ment [15%]).47 The ORR among the 31 patients who had received
prior ther­apy was 71%, includ­ing a CR/CR with par­tial hema­to­
logic recov­ery rate of 19%.48 Overall sur­vival at 12 and 24 months
was 80% and 65%, respec­tively, and the median OS sur­vival was
not reached in this pre­vi­ously treated pop­u­la­tion with a median
fol­low-up of 17.7 months.48 In 2022, the EMA granted approval
for avapritinib in patients with AdvSM exposed to at least 1 prior
sys­temic ther­apy.49
Long-term out­comes
The esti­
mated pro­
gres­
sion-free sur­
vival rates were 84% at
12 months and 63% at 24 months in the EXPLORER response-­
evaluable pop­u­la­tion (n = 53).44 During a median fol­low-up of
23 months, 14 (20%) exhibited dis­ease pro­gres­sion, includ­ing
6 patients (9%) with trans­for­ma­tion to sec­ond­ary AML. There
was no con­sis­tent pat­tern of base­line or on-treat­ment mye­loid
muta­tions, changes in the var­i­ant allele fre­quency of spe­cific
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/34/2021847/34gotlib.pdf by guest on 09 December 2022
ORR by sub­type, %
genes, or resis­tance muta­tions in KIT in patients with or with­
out clin­i­cal pro­gres­sion.50
In the EXPLORER over­
all AdvSM safety pop­
u­
la­
tion, the
median OS was not reached with a median fol­low-up dura­tion
of 23 months.44 Patients with S/A/R muta­tions or a base­line
muta­
tion-adjusted risk score ≥2 exhibited shorter sur­
vival.44
Among all­patients with AdvSM, the esti­mated 24-month OS
rate was 76% and 100%, 67%, and 92% for ASM, SM-AHN, and
MCL sub­types, respec­tively.44 The com­par­a­tive sur­vival rates
from the global midostaurin trial were 53% for all­patients
with AdvSM and 86%, 49%, and 26% for patients with ASM,
SM-AHN, and MCL, respec­tively.32 The esti­mated median OS
for all­patients with AdvSM was 46.9 months in a pooled anal­y­
sis of patients from EXPLORER and PATHFINDER ini­ti­ated with
avapritinib ≤200 mg daily.47 Compared with a his­tor­i­cal cohort
of patients with AdvSM treated with best avail­­able ther­apy,
avapritinib-treated patients exhibited sig­
nif­i­
cantly improved
sur­vival (adjusted haz­ard ratio, 0.48; 95% con­fi­dence inter­val,
0.29-0.79; P = .004), lon­ger dura­tion of treat­ment (23.8 vs 5.4
months; P < .001), and a 60% greater mean dif­fer­ence in the
per­cent max­i­mum decrease in lev­els of serum tryptase.51 An
indi­rect treat­ment com­par­i­son found that avapritinib improved
sur­vival com­pared with midostaurin.52
eral/periorbital edema (81%/4%), diar­rhea (34%/<1%), ­nau­sea
(31%/3%), fatigue/asthe­
nia (28%/7%), and cog­
ni­
tive effects
(25%/2%).41 Hematologic AEs consisted of neutropenia
(17%/16%), ane­mia (44%/27%), and the grouped terms throm­bo­
cy­to­pe­nia/plate­let count decreased (50%/30%). In EXPLORER,
intracranial bleed­
ing (ICB; eg, intraparenchymal hem­
or­
rhage,
sub­dural hema­toma) emerged as an AE of spe­cial inter­est that
occurred in 9 (13%) patients, with 7 of these cases devel­op­ing
in the set­ting of throm­bo­cy­to­pe­nia (plate­let count <50×109/L).
Asymptomatic cases (n=5, grade 1) were detected by prespecified ­pro­to­col mag­netic res­o­nance imag­ing of the brain; 2 events
were grade 2, and one each was grade 3 and grade 5 (asso­ci­
ated with head trauma). Mitigation pro­ce­dures (eg, exclu­sion of
patients with a starting plate­let count <50×109/L, dose hold and
reduc­tion for emer­gent throm­bo­cy­to­pe­nia reaching this thresh­
old, plate­let trans­fu­sions, increased blood count sur­veil­lance) led
to a decrease in ICB in the interim PATHFINDER anal­y­sis, with only
1 patient (1.6%) expe­ri­enc­ing a grade 2 sub­dural hema­toma in
the set­ting of pro­gres­sive throm­bo­cy­to­pe­nia.45 The recommended fre­quency of plate­let count mon­i­tor­ing is detailed in the
avapritinib pre­scrib­ing insert.
Avapritinib: adverse events
BLU-263 (Blueprint Medicines) has com­pa­ra­ble selec­tiv­ity and
potency to avapritinib but lim­ited brain pen­e­tra­tion poten­tial,
which may mit­i­gate cog­ni­tive changes and ICB. It is cur­rently
being eval­u­ated in the phase 2/3 HARBOR study (NCT04910685)
In the pooled out­comes from the over­all safety pop­u­la­tion of 131
avapritnib-treated patients,47 the most com­mon ­nonhematologic
adverse events (AEs) (all­grades %/grade ≥3%) were periph­
Emerging KIT inhib­i­tors
BLU-263 and bezuclastinib
Therapies for advanced SM & eosin­o­philic neo­plasms | 37
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/34/2021847/34gotlib.pdf by guest on 09 December 2022
Figure 1. Phase 1 EXPLORER Study: Avapritinib effects on measures of mast cell burden. Waterfall plots are shown for bone marrow
mast cell burden, serum tryptase level, spleen volume, and KIT D816V variant allele fraction. A, other antineoplastic therapy; M, prior
midostaurin exposure; MAF, mutant allele fraction; VAF, variant allele fraction.
AdvSM and KIT inhib­i­tors: open ques­tions and future
direc­tions
Avapritinib, midostaurin, or enroll­ment in clin­i­cal tri­als should
be con­sid­ered the first-line treat­ment option for AdvSM. While
avapritinib is not rec­om­mend for patients with plate­let count
<50 × 109/L, it may oth­er­wise merit pre­ferred sta­tus. The drug
can gen­er­ate molec­u­lar remis­sions of KIT D816V, and recent
data indi­cate more favor­able long-term out­comes com­pared
with his­tor­i­cal treat­ments, includ­ing midostaurin. The major
chal­lenge of these KIT inhib­i­tors is the AdvSM var­i­ant SM-AHN,
in which prog­no­sis is usu­ally driven by the AHN, and the pres­
ence of a com­plex muta­tional land­scape beyond KIT D816V
may pro­mote pro­gres­sion and resis­tance. In this regard, how to
sequence KIT inhi­bi­tion with AHN-targeted ther­apy is a major
focus of clin­i­cal trial devel­op­ment in AdvSM. Table 2 high­lights
some of the major chal­lenges and open ques­tions that have
emerged in the era of KIT inhib­i­tors.
CLINICAL CASE (Con­tin­ued)
The patient was started on avapritinib 200 mg daily. After 4
months, he became paracentesis inde­pen­dent. By 6 months, the
albu­min had increased to 4.8 g/dL, and the patient had gained
30 pounds and no lon­ger exhibited mus­cle wast­ing. In addi­tion,
the serum tryptase level nor­mal­ized to 10.1 ng/mL and the BM
biopsy spec­i­men showed no mast cell aggre­gates. He exhibited
a 38% improve­ment in total symp­tom score on the AdvSM Symptom Assessment Form. Due to per­sis­tence of periorbital edema
and achieve­ment of a com­plete remis­sion, the avapritinib dose
was decreased to 100 mg daily. After 3 years of fol­low-up, he
main­tains a com­plete remis­sion; the BM shows no evi­dence of
MC aggre­gates and blasts are not increased. The serum tryptase
level remains nor­mal. The del(13q) abnor­mal­ity and TET2 muta­
tion per­sist, but the KIT D816V muta­tion is no lon­ger detect­able.
Primary eosin­o­philic neo­plasms
The WHO major cat­e­gory of “mye­loid/lym­phoid neo­plasms with
eosin­o­philia and rearrangement of PDGFRA, PDGFRB, or FGFR1 or
with PCM1-JAK2” has had its title changed to “mye­loid/lym­phoid
neo­plasms with eosin­o­philia and tyro­sine kinase gene fusions” in
both the fifth edi­tion of the WHO clas­si­fi­ca­tion of hematolymphoid
tumors and the International Consensus Classification (ICC).1,2 The
mod­i­fi­ca­tion in nomen­cla­ture reflects the com­mon molec­u­lar
theme of rearranged, con­sti­tu­tively acti­vated tyro­sine kinases.
These TK gene fusions (and asso­ci­ated chro­mo­somal breakpoints)
involve PDGFRA (the most com­
mon fusion, FIP1L1::PDGRA, is
cryp­tic; oth­er­wise, breakpoint 4q12), PDGFRB (5q31~q33), FGFR1
(8p11), JAK2 (9p24), FLT3 (13q12), and ETV6::ABL1. ETV6::ABL1 presenting as Philadelphia chromosome-like acute lymphoblastic
leukemia (which should be dis­tin­guished from this cat­e­gory of TK
gene fusions) can also rarely pres­ent as de novo T-cell acute lymphoblastic leukemia or a mye­loid neo­plasm.55,56
Table 2. Major ques­tions and future chal­lenges with KIT inhib­i­tors in AdvSM
• What are the most pre­dic­tive bio­mark­ers of prolonged response and sur­vival?
• Will response cri­te­ria based on pure path­o­logic cri­te­ria (eg, BM MC bur­den, serum tryptase level, blood counts), with­out confounding by
lin­ger­ing C-find­ings, bet­ter pre­dict long-term end points such as over­all sur­vival?
• How do we define and har­mo­nize molec­u­lar eval­u­a­tion of MRD?
• Is time-lim­ited treat­ment with KIT inhib­i­tors fea­si­ble in patients who achieve a min­i­mum dura­tion of com­plete molec­u­lar remis­sion of KIT D816V?
• What is the clin­i­cal impact of KIT inhib­i­tors on the nat­u­ral his­tory of the AHN com­po­nent?
• Does the pres­ence of an AHN or multimutated molec­u­lar pro­file abro­gate the poten­tial ben­e­fits of molec­u­lar remis­sion of KIT D816V or neg­a­tive
MRD?
• How do we sequence KIT inhib­i­tors and AHN-directed ther­apy in patients with SM-AHN?
• What is the opti­mal role and tim­ing of allo­ge­neic HSCT in the era of KIT inhib­i­tors and who are most appro­pri­ate can­di­dates for trans­plant?
• How do we use KIT inhib­i­tors as a cytoreduction strat­egy before transplant?
• Should KIT inhib­i­tors be con­tin­ued after trans­plant only in patients with detect­able KIT D816V?
• Will KIT inhib­i­tors with less CNS pen­e­tra­tion (eg, BLU-263, bezuclastinib) per­mit their use with ther­apy for high-risk AHN (eg, HMA±venetoclax) or
with inten­sive induc­tion che­mo­ther­apy in KIT D816V-pos­i­tive AML?
• Is there a role for cladribine + selec­tive KIT inhib­i­tors in patients with AdvSM with refrac­tory/relapsed or rap­idly pro­gres­sive dis­ease?
CNS, cen­tral ner­vous sys­tem; HMA, hypomethylating agent; MRD, measurable resid­ual dis­ease.
38 | Hematology 2022 | ASH Education Program
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/34/2021847/34gotlib.pdf by guest on 09 December 2022
of patients with ISM.53 Substantial clin­i­cal inter­est exists in eval­
u­at­ing BLU-263 in patients with AdvSM as monotherapy or as
sequenced ther­apy with AHN-directed agents given the con­cern
about the poten­tial for throm­bo­cy­to­pe­nia-emer­gent ICB in this
patient pop­u­la­tion with avapritinib.
Bezuclastinib (CGT9486; Cogent Biosciences) is a type I TK
inhib­i­tor with activ­ity against muta­tions in KIT exons 9, 11, 17, and
18, includ­ing D816V.54 Its tar­get pro­file avoids other kinases such
as WT KIT, PDGFRα, PDGFRβ, wild-type KIT, VEGFR2, and CSF1R
and exhib­its lim­ited blood-brain bar­rier pen­e­tra­tion and no cen­
tral ner­vous sys­tem toxicities in pre­clin­i­cal stud­ies. The drug is
cur­rently being eval­u­ated in a phase 2, open-label clin­i­cal trial
(APEX; NCT04996875) of AdvSM in which patients are first ran­
dom­ized to 1 of 4 doses of the drug in a dose opti­mi­za­tion stage,
followed by dose expan­sion at the recommended phase 2 dose.
A phase 2/3 study of bezuclastinib in ISM/smol­der­ing SM has
also been ini­ti­ated (NCT05186753).
Table 3. Tyrosine kinase fusion genes and poten­tial targeted ther­a­pies in mye­loid/lym­phoid neo­plasms with eosin­o­philia
Most com­mon part­ner
fusion gene
Other part­ner fusion genes and chro­mo­some
breakpoint
Tyrosine kinase inhib­i­tors with
established or poten­tial activ­ity
PDGFRA (4q12)
FIP1L1 (4q12)
BCR (22q11)
KIF5B (10p11)
TNKS2 (10q23)
CDK5RAP2 (9q33)
ETV6 (12p13)
STRN (2p24)
FOXP1 (3p14)
Imatinib
PDGFRB (5q31~33)
ETV6 (12p13)
SPTBN1 (2p16)
PDE4DIP (1q22)
WDR48 (3p22)
GOLGB1 (3q12)
DIAPH1 (5q31)
KANK1 (9p24)
CEP85L (6q22)
GIT2 (12q24)
HIP1 (7q11)
NIN (14q24)
ERC1 (12p13)
DTD1 (20p11)
MYO18A (17q11)
NDE1 (16p13).
CPSF6 (12q15)
CCDC88C (14q32)
TPM3 (1q21)
SPDR (2q32)
GOLGA4 (3p22)
PRKG2 (4q21)
TNIP1 ((5q33)
SART3 ((12q23)
CCDC6 (10q21)
NDEL1 (17p13)
GPIAP1 (11p13)
SPECC1 (17p11)
TRIP11 (14q32)
RABEP1 (17p13)
MPRIP (17p11)
TP53BP1 (15q22)
BIN2 (12q13)
Imatinib
FGFR1 (8p11)
ZMYM2 (13q12)
FGFR1OP (6q27)
LRRFIP1 (2q37)
SQSTM1 (5q35)
TRIM24 (7q34)
HERV-K (19q13)
BCR (22q11)
CPSF6 (12q15)
CNTRL (9q33)
RANBP2 (2q13)
CUX1 (7q22)
TPR1 (1q25)
FGFR1OP2 (12p11)
MYO18A (17q11)
TFG (3q12)
Pemigatinib
Futibatinib
Midostaurin
Ponatinib
JAK2 (9p24)
PCM1 (8p21)
ETV6 (12p13)
BCR (22q11)
FLT3 (13q12)
ETV6 (12p13)
SPTBN1 (2p16)
TRIP11 (14q32)
LYN (8q12)
ABL1 (9q34)
ETV6 (12p13)
Ruxolitinib
Fedratinib
Pacritinib
Momelotinib
GOLGB1 (3q12)
NTRK3 (15q25)
SYK (9q22)
Gilteritinib
Midostaurin
Sorafenib
Sunitinib
Dasatinib
Nilotinib
Imatinib
Bosutinib
Ponatinib
Asciminib
Eosinophilia is not an invari­able fea­ture of these neo­plasms
but can serve as a use­ful pre-diagnostic check­point to think
about these dis­ease enti­ties. At the time of writ­ing, there have
been 8 PDGFRA part­ner fusion genes char­ac­ter­ized, over 30
for PDGFRB, 16 for FGFR1, 3 for JAK2, and 7 for FLT3 (Table 3).55-57
In some cases besides the cyto­ge­net­i­cally invis­i­ble FIP1L1::
PDGFRA, stan­dard cyto­ge­net­ics and/or fluo­res­cence in situ
hybrid­iza­tion (FISH) may not be a
­ ble to iden­tify the TK gene
fusion. Integrated geno­mic ana­ly­ses, includ­ing chro­mo­
somal microarrays, whole-genome sequenc­ing, and/or RNA
sequenc­ing, may be nec­es­sary to uncover small dele­tions or
inver­sions resulting in cryp­tic fusion genes.58
The clin­i­cal pre­sen­ta­tion of myeloid/lymphoid neoplasms
with eosinophilia (MLN-eo) with TK gene fusions can be highly
com­plex. One should deter­mine whether the dis­ease involves
the BM and periph­eral blood (PB) only, extramedullary dis­ease
(EMD) only, or both BM/PB and EMD. If clin­i­cally suspected,
imag­ing (pref­er­a­bly pos­i­tron emis­sion tomog­ra­phy/com­puted
tomog­ra­phy) should be under­taken to iden­tify EMD so it can
be seri­ally followed dur­ing treat­ment. Second, the pres­ence
of chronic phase (CP) or blast phase (BP) dis­ease should be
char­ac­ter­ized in the BM/PB. The pres­ence of EMD rep­re­sents
a BP com­po­nent. Last, the dis­ease lin­e­ages of the BM/PB and
EMD com­po­nents should be char­ac­ter­ized. The most com­mon
CP pre­sen­ta­tions of MLN include mye­loid neo­plasms: MPN or
MDS/MPN (CMML or MDS/MPN-unclas­si­fi­able) with or with­out
eosin­o­philia. The BM may exhibit an atyp­i­cal (usu­ally inter­sti­
tial) infil­trate of MCs in the absence of the KIT D816V muta­tion.
BP dis­ease in the BM/PB and/or EMD may pres­ent as AML,
B- or T-cell leu­ke­mia/lym­pho­blas­tic lym­phoma, or a mixedphe­no­type acute leu­ke­mia. Importantly, the dis­ease lin­e­age
in the BM/PB can be dif­fer­ent from the EMD; there­fore, biopsy
of the EMD for lin­e­age ascer­tain­ment may be help­ful but not
always pos­si­ble. In some cases, we have seen tan­dem involve­
ment of the BM/PB by both a chronic mye­loid neo­plasm and
B- or T-cell acute lymphoblastic leukemia. If patients have
Therapies for advanced SM & eosin­o­philic neo­plasms | 39
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/34/2021847/34gotlib.pdf by guest on 09 December 2022
Tyrosine kinase gene
BM/PB: Chronic Phase*
with EMD
BM/PB: Blast Phase
No EMD
BM/PB: Blast Phase
with EMD
EMD Only
Treated, no morphologic
or radiologic evidence
of disease;
+ Cytogenetic/FISH, or
molecular evidence of
FGFR1 rearrangement
Figure 2. Clinical presentations of MLN with eosinophilia and tyrosine kinase gene fusions. *BM/PB can present with a concurrent
chronic myeloid neoplasm as well as acute lymphoblastic leukemia/lymphoma (B- or T-cell lineage) or mixed-phenotype acute
leukemia.
received prior treat­ment, only cyto­ge­netic, FISH, or molec­u­lar
evi­dence of a TK fusion gene may be evi­dent as a marker to
fol­low dur­ing sub­se­quent ther­apy. The het­ero­ge­neous pre­
sen­ta­tions of MLN with TK fusion genes are shown in Figure 2.
Their het­ero­ge­ne­ity may reflect the TK fusion gene, its part­
ner, or the pres­ence of addi­tional cyto­ge­netic or molec­u­lar
abnor­mal­i­ties (eg, RUNX1 muta­
tion in patients with FGFR1
fusions).
Imatinib in PDGFRA and PDGFRB TK gene fusions
FIP1L1::PDGFRA is the pro­to­typic fusion TK gene and is almost
always asso­ci­ated with eosin­o­philia. An increase in the vita­
min B12 and/or serum tryptase level often accompanies the
find­ing of atyp­i­cal, inter­sti­tial MCs in the BM.59-61 Most patients
are male who pres­ent with a chronic mye­loid neo­plasm with
eosin­o­philia; how­ever, rare cases of AML or T-cell ALL, as well
as iso­lated or con­cur­rent EMD, have been reported.62 FIP1L1::
PDGFRA results from a sub­mi­cro­scopic dele­tion of 800 kb on
chro­mo­some 4q12 and is not detected by stan­dard karyotyping.59 FISH for the CHIC2 dele­tion63 or reverse ­tran­scrip­tion
PCR is used to iden­tify the fusion and for serial mon­i­tor­ing;
false-neg­
a­
tive cases using FISH have been reported and
there­fore PCR should be employed in such cases with high
clin­i­cal sus­pi­cion.64
Studies have con­firmed the deep and dura­ble responses of
FIP1L1::PDGFRA-pos­i­tive dis­ease to imatinib at starting doses
of 100 mg daily.65-68 In patients with known or pos­si­ble car­
diac involve­ment with eosin­o­phils, con­com­i­tant ini­ti­a­tion of
cor­ti­co­ste­roids (eg, pred­ni­sone 1 mg/kg for 7-10 days) with
imatinib is recommended to mit­i­gate poten­tial com­pli­ca­tions
from heart fail­ure/car­dio­genic shock. Complete ­hema­to­logic
40 | Hematology 2022 | ASH Education Program
remis­sions are achiev­able in >95% of patients and can achieved
within sev­eral weeks; com­plete FISH and/or PCR responses
are typ­i­cally observed within 3 to 6 months. In patients achiev­
ing FISH or PCR neg­a­tiv­ity, main­te­nance dos­ing of 100 mg
3 times a week or weekly can main­
tain deep responses.69
Similar to chronic myeloid leukemia, increas­
ing expe­ri­ence
is accu­mu­lat­ing with imatinib dis­con­tin­u­a­tion and the con­
cept of treat­ment-free remis­sion in FIP1L1::PDGFRA-­pos­i­tive
cases. In case series, hema­to­logic and/or molec­u­lar relapse
has been documented in <6 months, but in some cases,
molec­u­lar remis­sions have been maintained for more than 5
years.70-73 In a ret­ro­spec­tive French series, the relapse-free sur­
vival rate of patients who discontinued imatinib was 57% after
a median fol­low-up of 58 months (range, 25-100 months).74
Relapses occurred after a median of 10 months (range, 4-23),
consisting of hema­to­logic (n = 18, 90%) or iso­lated molec­u­lar
relapses (n = 2, 10%). Imatinib was suc­cess­fully resumed in 17
of 20 (85%) relapsed patients, with com­plete hema­to­logic
remis­sion within 1 month in all­cases and molec­u­lar responses
within 12 months in all­tested patients.74 In mul­ti­var­i­able anal­
y­sis, dura­tion of prior imatinib ther­apy treat­ment was a sta­
tis­ti­cally sig­nif­i­cant fac­tor for relapse while time to imatinib
ini­ti­a­tion showed a trend for sig­nif­i­cance. After a mean fol­lowup of 80 months, the 1-, 5-, and 10-year over­all sur­vival rates
of imatinib-treated patients in this cohort were 99%, 95%, and
84%, respec­tively.74
In another study, 12 patients with FIP1L1::PDGFRA who
achieved a CMR were followed after imatinib dis­con­tin­u­a­tion.75
Median time of treat­ment and median time of CMR before imatinib dis­con­tin­u­a­tion were 80 months (range, 43-175 months)
and 66 months (range, 37-174 months), respec­tively. A molec­u­lar
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/34/2021847/34gotlib.pdf by guest on 09 December 2022
BM/PB: Chronic Phase*
No EMD
JAK2, FLT3, and ETV6::ABL1 TK gene fusions
In sev­eral case series, patients have been treated with ruxolitinib
for JAK2 gene fusions or with FLT3 inhib­i­tors (midostaurin, sunitinib, sorafenib, gilteritinib) for gene fusions involv­ing FLT3.84-86
Hematologic and cyto­ge­netic remis­sions have been described
with JAK2 or FLT3 inhib­i­tors in both TK-driven neo­plasms, but
they tend not to be dura­ble.79-85 Since JAK2- and FLT3-targeting
TKIs in these dis­eases have gen­er­ally not pro­vided suf­fi­cient dis­
ease con­trol, the intent should be to use them as a cytoreductive bridge to HSCT. This was recently dem­on­strated in an infant
with ETV6::FLT3 MLN who achieved a mor­pho­logic, immunophenotypic, and cyto­ge­netic remis­sion with gilteritinib before
pro­ceed­ing to trans­plant.87 In patients with CP pre­sen­ta­tions
of ETV6::ABL1-pos­i­tive dis­ease, the lim­ited expe­ri­ence to date
sug­gests that the sec­ond-gen­er­a­tion TKIs nilotinib and dasatinib
may elicit more dura­ble com­plete remis­sions than imatinib.86
FGFR1 inhib­i­tors in MLN with FGFR1 gene fusions
Patients with FGFR1-rearranged MLN-eo exhibit an aggres­
sive dis­ease course.88,89 Individuals with chronic phase dis­ease
exhibit a cumu­la­tive rate of trans­for­ma­tion to blast phase at
12 months approaching 50%, and 1-year OS of patients presenting
with blast phase dis­ease is 30%.89 Hydroxyurea and multikinase
inhib­
i­
tors with anti-FGFR1 activ­
ity (eg, ponatinib, midostaurin) in CP and lin­e­age-spe­cific induc­tion ­che­mo­ther­apy ± TKI in
BP can lead to par­tial response (PR) or short-lived com­plete
response (CR), but cyto­ge­netic responses (CyR) are rare.90-92
Allogeneic HSCT is the only treat­ment modal­ity that has been
shown to pro­duce long-term remis­sions in MLN with a FGFR1
fusion gene.93
Futibatinib is a selec­tive inhib­i­tor of FGFR1-4 that was eval­
u­ated in a patient with a PCM1::FGFR1 fusion.94 Futibatinib at a
dose of 20 mg daily (dose reduced to 16 mg daily after 3 months
for a grade 2 bul­lous rash) pro­duced a dura­ble, com­plete hema­
to­logic and cyto­ge­netic remis­sion that was ongo­ing after >18
months of ther­apy. Futibatinib is cur­rently being eval­u­ated in a
phase 2 study of MLN as well as solid tumors (NCT04189445).
The most mature data are avail­­
able for pemigatinib
(INCB054828), a selec­tive inhib­i­tor of FGFR1-3. It has received
reg­u­la­tory approval for adult patients with pre­vi­ously treated,
unresectable, locally advanced, or met­
a­
static cholangiocarcinoma with an FGFR2 fusion or other rearrangements.95 The
ongo­ing FIGHT-203 study eval­u­at­ing pemigatinib is a phase 2,
mul­ti­cen­ter trial enroll­ing adults with MLN-eo with a FGFR1 fusion
TK gene. Initially, patients had to have received ≥1 prior ther­
apy with a starting dose of pemigatinib 13.5 mg daily in 3-week
cycles (2 weeks on, 1 week off). With amend­ments, patients
with­out prior ther­apy were also eli­gi­ble and the starting dose
was mod­i­fied to 13.5 mg daily on a con­tin­u­ous sched­ule. The pri­
mary end point is CR rate; sec­ond­ary end points include over­
all response (ORR [CR + PR]), com­plete CyR or par­tial CyR, and
safety.96 All pri­mary and sec­ond­ary end points were inves­ti­ga­
tor assessed and also adju­di­cated ret­ro­spec­tively by a Central
Review Committee (CRC) with CRC-defined cri­te­ria based on
local lab and radio­logic results and cen­tral review of his­to­pa­thol­
ogy and stan­dard karyotyping/FISH.
Thirty-four patients were enrolled and treated (1 sub­ject with­
out an FGFR1 rearrangement was excluded from the effi­cacy
anal­y­sis). The aver­age num­ber of prior ther­a­pies was 1.6 (range,
0-6); 3 patients had under­gone prior trans­plant and 5 patients
were treat­
ment naive.96 The lon­
gest dura­
tion of pemigatinib
expo­sure was 192 weeks with a median dos­ing dura­tion of 29
weeks. Patients com­pleted a median of 10.0 treat­ment cycles
(range, 2-65). At data cut­off (Decem­ber 31, 2020), treat­ment was
ongo­ing in 18 patients (53%); rea­sons for treat­ment dis­con­tin­
u­a­tion (n, %) included bridg­ing to trans­plant (n=6, 18%), pro­
gres­sive dis­ease (n=5, 15%), adverse event (n=3, 9%), phy­si­cian
deci­sion (n=1, 3%), and patient deci­sion (n=1, 3%).96
The base­
line fea­
tures of the 33 patients with FGFR1-rearranged MLN have been pre­vi­ously reported.96 Two patients had
treated MLN with FGFR1 with a per­sis­tent cyto­ge­netic abnor­mal­
ity only (no mor­pho­logic evi­dence of dis­ease), and the remaining patients had CP (n=18) or BP (n=13) dis­ease. Among these 31
patients, CR rates per inves­ti­ga­tor and CRC assess­ments were
64.5% and 77.4%, respec­tively; among the 33 patients evaluable
for CyR, com­plete CyR rates were 72.7% and 75.8%, respec­tively
(Figure 3).
The most com­
mon treat­
ment-emer­
gent adverse events
(TEAEs) were hyperphosphatemia (68%), alo­
pe­
cia (59%),
diar­rhea (50%), sto­ma­ti­tis (44%), and ane­mia (35%). Grade
Therapies for advanced SM & eosin­o­philic neo­plasms | 41
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/34/2021847/34gotlib.pdf by guest on 09 December 2022
relapse was observed in 4 patients between 10 and 24 months. A
sec­ond CMR was achieved in 3 patients after 3, 4, and 21 months.
Eight patients (62%) are in ongo­ing CMR (median, 17 months;
range, 3-71 months). Molecular relapse-free sur­vival was 91% at 12
months and 65% at 24 months.75 No sig­nif­i­cant dif­fer­ences were
iden­ti­fied between patients with and with­out molec­u­lar relapse
regard­ing dose and dura­tion of imatinib ther­apy or prior dura­
tion of CMR. Further pro­spec­tive stud­ies are needed to bet­ter
define pre­dic­tive fac­tors for treat­ment-free remis­sion in these
patients.
For patients with fusion tyro­sine kinases involv­ing PDGFRB, starting doses of imatinib 100 to 400 mg daily are recommended.55-57 A
main­te­nance dose of 100 mg daily can be con­sid­ered in patients
achiev­ing hema­to­logic and molec­u­lar remis­sions fol­low­ing an
induc­tion dose of 400 mg daily. Long-term treat­ment out­comes
with imatinib are sim­i­lar to the expe­ri­ence with FIP1L1::PDGFRA.76
Long-term fol­low-up (median, 10.2 years) of PDGFRB-rearranged
patients treated with imatinib for a median dura­tion of 6.6 years
showed a 96% response rate and a 10-year over­all sur­vival rate
of 90%.77 None of the patients achiev­ing a com­plete cyto­ge­netic
(n=13) or molec­u­lar (n=8) remis­sion lost their response or exhibited pro­gres­sion to blast phase dis­ease.77
Imatinib monotherapy can be effec­tive in de novo pre­sen­ta­
tions of blast phase dis­ease with PDGFRA or PDGFRB fusion TK
genes. In a case series of 17 patients in blast phase and/or with
EMD, 15 patients treated with imatinib only achieved dura­ble com­
plete hema­to­logic and molec­u­lar remis­sions.78 Two of 17 patients
(12%) died after a median fol­low-up of 65 months. While these
data are encour­ag­ing, an alter­nate approach may be to com­bine
lin­e­age-spe­cific induc­tion che­mo­ther­apy + imatinib with con­sid­
er­ation of allografting patients in first complete remission.
Primary resis­tance is not a clin­i­cal con­cern. Although uncom­
mon, sec­ond­ary resis­tance almost always defaults to 2 canon­i­
cal muta­tions: PDGFRA T674I or PDGFRA D842V. Although these
muta­tions have exhibited var­i­able sen­si­tiv­ity to tyrosine kinase
inhibitors (TKIs) in vitro, clin­i­cal responses have been underwhelming.79-83 However, avapritinib’s activ­ity against PDGFRA
D842V in patients with GIST42 augurs prom­ise for sim­i­lar ben­e­fit
in the con­text of FIP1L1::PDGFRA.
≥3 TEAEs in ≥10% of patients were ane­mia (18%) and pain
in extrem­ity and sto­ma­ti­tis (both 12%). Dose mod­i­fi­ca­tions
due to any TEAE included dose inter­
rup­
tion (65%), reduc­
tion (59%), and dis­con­tin­u­a­tion (12%: due to car­diac fail­ure,
­mul­ti­ple-organ dys­func­tion, increased alka­line phos­pha­tase,
and calciphylaxis).96
While Kaplan-Meier median dura­tions of CR and ORR had
not been reached in the over­all effi­cacy pop­u­la­tion, clin­i­cal
and cyto­ge­netic responses in BP dis­ease were less fre­quent
and less dura­ble than in patients with CP dis­ease. The most
fre­quent rea­son for treat­ment dis­con­tin­u­a­tion was bridg­ing
to trans­plant, which was achieved in 23% of BP patients.96
Taken together, these data indi­cate that pemigatinib can pro­
duce dura­ble and high rates of com­plete clin­i­cal and cyto­
ge­netic responses, most of whom had progressed on prior
ther­apy. Pemigatinib may be a use­ful option for patients inel­
i­gi­ble for trans­plant or may facil­i­tate bridg­ing to trans­plant
in eli­gi­ble patients. Pemigatinib was approved by the FDA on
August 26, 2022, for relapsed or refrac­tory MLN with FGFR1
rearrangement.
Chronic eosin­o­philic leu­ke­mia
In both the new ICC and WHO clas­
si­
fi­
ca­
tions,1,2 diag­nos­
tic cri­
te­
ria for chronic eosin­
o­
philic leu­
ke­
mia (CEL) require
both abnor­mal BM mor­phol­ogy (eg, dys­plas­tic mega­kar­yo­
cytes with or with­out dys­plas­tic fea­tures in other lin­e­ages
or increased blasts ≥5% in the bone mar­row and/or ≥2% in
the periph­eral blood) as well as dem­on­stra­tion of a clonal
cyto­ge­netic abnor­mal­ity and/or somatic muta­tion(s). In the
absence of a clonal cyto­ge­netic abnor­mal­ity and/or somatic
muta­
tion(s) or increased blasts, the afore­
men­
tioned BM
find­ings are suf­fi­cient in the pres­ence of per­sis­tent eosin­o­
philia, for which other causes have been excluded. The focus
42 | Hematology 2022 | ASH Education Program
on ­abnor­mal BM his­to­pa­thol­ogy97 is intended to dis­tin­guish
CEL from idi­o­pathic hypereosinophilic syn­drome and hypereosinophilia of unknown sig­nif­i­cance. The ICC also requires
per­sis­tent PB hypereosinophilia (eosin­o­phil count ≥1.5 × 109/L
and ­eosin­o­phils ≥10% of white blood cells)1 while the WHO
specifies the hypereosinophilia should endure for at least
4 weeks from the prior require­ment of 6 months.2 The WHO
also removed the CEL dis­ease qual­i­fier “NOS.”2 Notably, CEL
is still included among the BCR::ABL1-neg­a­tive MPNs, not
MDS/MPNs, despite dys­plas­tic mar­row find­ings being a core
his­to­path­o­logic fea­ture of these neo­plasms.
CEL lacks recur­rent cyto­ge­netic and/or molec­u­lar genetic
abnor­mal­i­ties such as BCR::ABL1 or TK gene fusions asso­
ci­
ated with MLN-eo. Case reports and series have anno­
tated
non­spe­cific abnor­mal­i­ties, includ­ing del(13q), del(20q), tri­
somy 8, mono­somy 7, and com­plex kar­yo­types.98 In some CEL
cases, next-gen­er­a­tion sequenc­ing (NGS) pan­els have uncov­
ered gene muta­
tions that are pro­
mis­
cu­
ous among mye­
loid
neo­plasms (eg, DNMT3A, ASXL1, TET2, EZH2, SETBP1, CBL,
SF3B1, among oth­ers).99 It may not always be pos­si­ble to dis­
tin­guish whether some var­i­ants are driv­ers of eosin­o­philia or
clonal hematopoiesis of indeterminate potential muta­
tions.
Rarely, NGS may uncover poten­
tial druggable tar­
gets or
path­ways, such as the JAK-STAT sig­nal­ing axis. In 1 study, the
STAT5B N642H muta­tion was iden­ti­fied in 27 of 1715 (1.6%) cases
referred for eosin­o­philia.100 Somatic inser­tion/dele­tion muta­
tions (p.L583_A586delinsS) within exon 13 of the pseudokinase
domain of JAK2 have also been described in patients with a
diag­no­sis of both CEL and poly­cy­the­mia vera,101 and a somatic
acti­vat­ing JAK1 R629_S632delinsSA muta­tion was iden­ti­fied in
another patient with long-stand­ing CEL.102
CEL car­ries a poor prog­no­sis, with a median OS in the range
of 1 to 2 years in 2 case series.98,103 No con­sen­sus stan­dard
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/34/2021847/34gotlib.pdf by guest on 09 December 2022
Figure 3. FIGHT-203 Trial: Rates of complete clinical and cytogenetic response with pemigatinib. Responses to pemigatinib were
adjudicated by local investigators and according to a central review committee. Central FISH was given priority over local cytogenetic results during CRC adjudications. aIn the 2 patients with treated MLN with no morphologic evidence of disease but persistent
cytogenetic abnormality, cytogenetic responses were adjudicated, but clinical responses were not.
Table 4. Open ques­tions and future pri­or­i­ties with novel agents in pri­mary eosin­o­philic neo­plasms
• Need for response cri­te­ria for MLN with eosin­o­philia and tyro­sine kinase gene fusions
• Increasing diag­nos­tic rec­og­ni­tion of MLN with eosin­o­philia and tyro­sine kinase gene fusions given the avail­abil­ity of targeted ther­a­pies
• How do we incor­po­rate stan­dard cyto­ge­netic/FISH and molec­u­lar mon­i­tor­ing of tyro­sine kinase fusion genes into clin­i­cal deci­sion-mak­ing?
• Harmonization of molec­u­lar mon­i­tor­ing tech­niques for tyro­sine kinase fusion genes
• What is the opti­mal induc­tion and main­te­nance dos­ing of tyro­sine kinase inhib­i­tors in spe­cific MLN with eosin­o­philia and tyro­sine kinase gene
fusions?
• What are the best pre­dic­tors of treat­ment-free remis­sion in PDGFRA- and PDGFRB-rearranged MLN treated with imatinib?
• In FGFR1-rearranged MLN, is pemigatinib monotherapy suf­fi­cient for chronic phase dis­ease, and what is the role of lin­e­age-spe­cific induc­tion
che­mo­ther­apy±pemigatinib in blast phase dis­ease?
• What is the role of pemigatinib and other tyro­sine kinase inhib­i­tors posttransplant to min­i­mize dis­ease relapse?
• Unmet ther­a­peu­tic need in CEL: eval­u­ate HMAs and other ther­a­pies (eg, venetoclax) in the con­text of clin­i­cal tri­als
­front­line treat­ment exists for CEL. Corticosteroids, hydroxy­urea,
PEG-IFN-α, and imatinib can improve leu­ko­cy­to­sis and hypereosinophilia, but responses tend not to be dura­ble. PEG-IFN-α
can elicit hema­to­logic and cyto­ge­netic responses, as well as
rever­sion of end-organ dam­age, includ­ing in patients who are
relapsed/refrac­tory to cor­ti­co­ste­roids and hydroxy­urea.55,56 In the
absence of a tyro­sine kinase tar­get, hema­to­logic improve­ment
with empiric use of imatinib may reflect non­spe­cific myelosuppression, and responses tend to be short-lived. Anti–inter­leu­kin
5 and anti–inter­leu­kin 5 recep­tor antibodies (mepolizumab and
benralizumab, respec­tively), which dem­on­strate activ­ity in idi­
o­pathic hypereosinophilic syn­drome and lym­pho­cyte var­i­ant
hypereosinophilia, are gen­er­ally not active in pri­mary eosin­o­
philic neo­plasms.104,105 As BM dys­pla­sia is a prominent fea­ture of
bona fide CEL, the role of hypomethylating agents should be
fur­ther explored in these cases.
Novel agents in pri­mary eosin­o­philic neo­plasms:
open ques­tions and future pri­or­i­ties
Imatinib has reversed the once poor sur­vival of patients with
PDGFRA and PDGFRB fusions into “good-risk” dis­eases. Despite
the avail­abil­ity of TK inhib­i­tors in MLN with rearranged FGFR1,
JAK2, FLT3, and ABL1, responses to targeted agents are usu­ally
par­tial and less dura­ble. Their nat­u­ral his­to­ries gen­er­ally remain
poor and should be dis­tin­guished from the very favor­able expe­ri­
ence with imatinib in PDGFRA- and PDGFRB-rearranged dis­ease.
The use of pemigatinib as a bridg­
ing agent to trans­
plan­
ta­
tion in MLN with FGFR1 fusions exemplifies the approach that
will need to be con­sid­ered more broadly in MLN, espe­cially in
the BP of dis­ease, where TKI response is often short-lived. For
patients with CEL, the results of NGS may pro­vide druggable
tar­gets in rare instances, but oth­er­wise a treat­ment back­bone
with hypomethylating agents or PEG-IFN-α or enroll­ment in clin­
i­cal tri­als of novel agents is recommended with an eye toward
trans­plan­ta­tion when fea­si­ble. Table 4 high­lights out­stand­ing
ques­
tions and future pri­
or­
i­
ties for the treat­
ment of pri­
mary
eosin­o­philic ­neo­plasms.
Acknowledgments
Dr. Gotlib expresses grat­i­tude to the Charles and Ann Johnson
Foundation for its sup­port of MPN Research at Stanford, as well the
Stanford Hematology Division, US and inter­na­tional SM inves­ti­ga­
tors, and patients with SM and their fam­i­lies. Dr. Gotlib expresses
spe­cial acknowl­edg­ment of mem­bers of the Stanford mastocytosis clin­i­cal research col­lab­o­ra­tive: Dr. William Shomali, Dr. Asiri
Ediriwickrema, Cheryl Langford, Justin Abuel, Cecelia Perkins,
Lenn Fechter, Annie Jinsuh Jung, Denise DeVore, Leslie Hwang,
Robert Jones, Parveen Abidi, and Bhuvana Ramachandran.
Conflict-of-inter­est dis­clo­sure
Research funding: Incyte, Novartis, Kartos, Blueprint Medicines,
Cogent Biosciences, Abbvie, Celgene, BMS, Protagonist Therapeutics. Advisory boards/con­sul­ting/hon­o­raria: Incyte, ­Novartis,
Kartos, Blueprint Medicines, Cogent Biosciences, Abbvie, Protagonist Therapeutics, PharmaEssentia, Imago Biosciences.
Off-label drug use
Bezuclastinib and BLU-263 are under inves­ti­ga­tion for indo­lent
and advanced SM. Futibatinib is under clinical investigation for
FGFR1-rearranged MLN. Off-label use: ruxolitinib, fedratinib, pacritinib, and momelotinib for JAK2-rearranged MLN; midostaurin
and ponatinib for FGFR1-rearranged MLN; ­midostaurin, sorafenib,
sunitinib, and gilteritinib for FLT3-rearranged MLN; and imatinib, dasatinib, nilotinib, bosutinib, ponatinib, and asciminib for
ETV6::ABL1-rearrranged MLN.
Correspondence
Jason Gotlib, Stanford Cancer Institute, 875 Blake Wilbur Drive,
Room 2324, Stanford, CA 94305-6555, USA; e-mail: jason​­.gotlib@
stanford​­.edu.
References
1.
Arber DA, Orazi, Hasserjian RP, et al. International con­sen­sus clas­si­fi­ca­tion of
mye­loid neo­plasms and acute leu­ke­mia: inte­grat­ing mor­pho­log­i­cal, clin­i­cal
and geno­mic data [published online ahead of print 29 June 2022]. Blood.
2. Khoury JD, Solary E, Abla O, et al. The 5th edi­tion of the World Health Organization clas­si­fi­ca­tion of haematolymphoid tumours: mye­loid and his­tio­
cytic/den­dritic neo­plasms. Leukemia. 2022;36(7):1703-1719.
3. Valent P, Akin C, Hartmann K, et al. Updated diag­nos­tic cri­te­ria and clas­
si­fi­ca­tion of mast cell dis­or­ders: a con­sen­sus pro­posal. Hemasphere.
2021;5(11):e646.
4. Pardanani A, Lim K-H, Lasho T-L, et al. Prognostically rel­e­vant break­down
of 123 patients with sys­temic mastocytosis asso­ci­ated with other mye­loid
malig­nan­cies. Blood. 2009;114(18):3769-3772.
Therapies for advanced SM & eosin­o­philic neo­plasms | 43
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/34/2021847/34gotlib.pdf by guest on 09 December 2022
• How do we incor­po­rate small-mol­e­cule inhib­i­tors of MLN tyro­sine kinase gene fusions as a cytoreduction strat­egy before transplant?
44 | Hematology 2022 | ASH Education Program
30. Gotlib J, Berubé C, Growney JD, et al. Activity of the tyro­sine kinase inhib­
i­tor PKC412 in a patient with mast cell leu­ke­mia with the D816V KIT muta­
tion. Blood. 2005;106(8):2865-2870.
31. DeAngelo DJ, George TI, Linder A, et al. Efficacy and safety of midostaurin
in patients with advanced sys­temic mastocytosis: 10-year median fol­lowup of a phase II trial. Leukemia. 2018;32(2):470-478.
32. Gotlib J, Kluin-Nelemans HC, George TI, et al. Efficacy and safety of midostaurin in advanced sys­temic mastocytosis. N Engl J Med. 2016;374(26):
2530-2541.
33. US Food and Drug Administration. FDA med­
i­
cal review. https:​­
/​­
/www​
­.accessdata​­.fda​­.gov​­/drugsatfda_docs​­/nda​­/2017​­/207997Orig1Orig2s
000MedR​­.pdf. Accessed 18 July 2022.
34. EMA. Summary of prod­uct char­ac­ter­is­tics. https:​­/​­/www​­.ema​­.europa​­.eu​­
/en​­/documents​­/product​­-information​­/rydapt​­-epar​­-product​­-information
_en​­.pdfformidostaurin. Accessed 18 July 2022.
35. Hartmann K, Gotlib J, Akin C, et al. Midostaurin improves qual­ity of life and
medi­a­tor-related symp­toms in advanced sys­temic mastocytosis. J Allergy
Clin Immunol. 2020;146(2):356-366.e4366e4.
36. Chandesris M-O, Damaj G, Canioni D, et al; CEREMAST Study Group. Midostaurin in advanced sys­temic mastocytosis. N Engl J Med. 2016;374(26):26052607.
37. Jawhar M, Schwaab J, Naumann N, et al. Response and pro­gres­sion on
midostaurin in advanced sys­
temic mastocytosis: KIT D816V and other
molec­u­lar mark­ers. Blood. 2017;130(2):137-145.
38. Szudy-Szczyrek A, Bachanek-Mitura O, Gromek T, et al. Real-world effi­cacy
of midostaurin in aggres­sive sys­temic mastocytosis. J Clin Med. 2021;10(5).
39. Singh A, Al-Kali A, Begna KH, et al. Midostaurin ther­apy for advanced sys­
temic mastocytosis: Mayo Clinic expe­ri­ence in 33 con­sec­u­tive cases. Am J
Hematol. 2022;97(5):630-637.
40. Lübke J, Schwaab J, Naumann N, et al. Superior effi­cacy of midostaurin over
cladribine in advanced sys­temic mastocytosis: a reg­is­try-based anal­y­sis.
J Clin Oncol. 2022;40(16):1783-1794.
41. Evans EK, Gardino AK, Kim JL, et al. A pre­ci­sion ther­apy against can­cers
driven by KIT/PDGFRA muta­tions. Sci Transl Med. 2017;9(414):eaao1690.
42. Jones RL, Serrano C, von Mehren M, et al. Avapritinib in unresectable or
met­a­static PDGFRA D842V-mutant gas­tro­in­tes­ti­nal stro­mal tumours: longterm effi­cacy and safety data from the NAVIGATOR phase I trial. Eur J Cancer. 2021;145:132-142.
43. US Food and Drug Administration. FDA med­
i­
cal review. https:​­
/​­
/www​
­.accessdata​­.fda​­.gov​­/drugsatfda_docs​­/label​­/2021​­/212608s007lbl​­.pdf.
Accessed 18 July 2022.
44. DeAngelo DJ, Radia DH, George TI, et al. Safety and effi­cacy of avapritinib
in advanced sys­temic mastocytosis: the phase 1 EXPLORER trial. Nat Med.
2021;27(12):2183-2191.
45. Gotlib J, Reiter A, Radia DH, et al. Efficacy and safety of avapritinib
in advanced sys­
temic mastocytosis: interim anal­
y­
sis of the phase 2
PATHFINDER trial. Nat Med. 2021;27(12):2192-2199.
46. Taylor F, Li X, Yip C, et al. Psychometric eval­u­a­tion of the Advanced Systemic Mastocytosis Symptom Assessment Form (AdvSM-SAF). Leuk Res.
2021;108:106606.
47. Radia D, DeAngelo DJ, Deininger DW, et al. Efficacy and safety of ≤200 mg
avapritinib in patients with advanced sys­
temic mastocytosis: pooled
results from the phase 1 EXPLORER and interim phase 2 PATHFINDER stud­
ies [ab­stract]. Clin Lymphoma Myeloma Leuk. 2021;21(suppl 1):S367-S368.
48. Reiter A, Schwaab J, DeAngelo DJ, et al. Efficacy and safety of avapritinib
in pre­vi­ously treated patients with advanced sys­temic mastocytosis [published online ahead of print 31 May 2022]. Blood Adv.
49. EMA. Summary of prod­uct char­ac­ter­is­tics. https:​­/​­/www​­.ema​­.europa​­.eu​­
/en​­/documents​­/product​­-information​­/ayvakyt​­-epar​­-product​­-information
_en​­.pdf. Accessed 18 July 2022.
50. Deininger MW, DeAngelo DJ, Radia DH. Effective con­
trol of advanced
sys­
temic mastocytosis with avapritinib: muta­
tional anal­
y­
sis from the
EXPLORER clin­i­cal study [ab­stract]. Blood. 2021;138(suppl 1): 318.
51. Reiter A, Gotlib J, Álvarez-Twose I, et al. Efficacy of avapritinib ver­sus best
avail­­able ther­apy in the treat­ment of advanced sys­temic mastocytosis.
Leukemia. 2022;36(8):2108-2120.
52. Pilkington H, Smith S, Roskell N, Iannazzo S. Indirect treat­ment com­par­i­
sons of avapritinib ver­sus midostaurin for patients with advanced sys­temic
mastocytosis. Future Oncol. 2022;18:1583-1594. doi: 10.2217/fon-2021-1509.
53. Castells M, Bhavsar V, He K, Scherber R, Akin C. HARBOR, a phase 2/3 study
of BLU-263 in patients with indo­lent sys­temic mastocytosis and mono­clo­
nal mast cell acti­va­tion syn­drome [ab­stract]. Paper presented at: Euro­pean
Hematology Association Congress; 10 June 2022; Vienna, Austria.
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/34/2021847/34gotlib.pdf by guest on 09 December 2022
5. Lim K-H, Tefferi A, Lasho T-L, et al. Systemic mastocytosis in 342 con­sec­u­tive
adults: sur­vival stud­ies and prog­nos­tic fac­tors. Blood. 2009;113(23):57275736.
6. Sperr WR, Kundi M, Alvarez-Twose I, et al. International Prognostic Scoring
System for Mastocytosis (IPSM): a ret­ro­spec­tive cohort study. Lancet Haematol. 2019;6(12):e638-e649.
7. Pardanani A, Shah S, Mannelli F, et al. Mayo alli­ance prog­nos­tic sys­tem for
mastocytosis: clin­i­cal and hybrid clin­i­cal-molec­u­lar mod­els. Blood Adv.
2018;2(21):2964-2972.
8. Jawhar M, Schwaab J, Álvarez-Twose I, et al. MARS: muta­tion-adjusted risk
score for advanced sys­temic mastocytosis. J Clin Oncol. 2019;37(31):28462856.
9. Muñoz-González JI, Álvarez-Twose I, Jara-Acevedo M, et al. Proposed global
prog­nos­tic score for sys­temic mastocytosis: a ret­ro­spec­tive prog­nos­tic
mod­el­ling study. Lancet Haematol. 2021;8(3):e194-e204.
10. Georgin-Lavialle S, Lhermitte L, Dubreuil P, Chandesris M-O, Hermine O,
Damaj G. Mast cell leu­ke­mia. Blood. 2013;121(8):1285-1295.
11. Jawhar M, Schwaab J, Meggendorfer M, et al. The clin­i­cal and molec­u­lar
diver­sity of mast cell leu­ke­mia with or with­out asso­ci­ated hema­to­logic
neo­plasm. Haematologica. 2017;102(6):1035-1043.
12. Jain P, Wang S, Patel KP, et al. Mast cell leu­ke­mia (MCL): clinico-path­o­logic
and molec­u­lar fea­tures and sur­vival out­come. Leuk Res. 2017;59:105-109.
13. Kennedy VE, Perkins C, Reiter A, et al. Mast cell leu­ke­mia: clin­i­cal and
molec­u­lar fea­tures and sur­vival out­comes of patients in the ECNM Registry
[published online ahead of print 12 Sep­tem­ber 2022]. Blood Adv.
14. Arock M, Sotlar K, Akin C, et al. KIT muta­tion anal­y­sis in mast cell neo­
plasms: rec­
om­
men­
da­
tions of the Euro­
pean Competence Network on
Mastocytosis. Leukemia. 2015;29(6):1223-1232.
15. Kristensen T, Vestergaard H, Bindslev-Jensen C, Møller MB, Broesby- Olsen
S; Mastocytosis Centre, Odense University Hospital (MastOUH). Sensitive
KIT D816V muta­tion anal­y­sis of blood as a diag­nos­tic test in mastocytosis.
Am J Hematol. 2014;89(5):493-498.
16. Schwaab J, Schnittger S, Sotlar K, et al. Comprehensive muta­tional pro­fil­
ing in advanced sys­temic mastocytosis. Blood. 2013;122(14):2460-2466.
17. Garcia-Montero AC, Jara-Acevedo M, Teodosio C, et al. KIT muta­tion in
mast cells and other bone mar­row hema­to­poi­etic cell lin­e­ages in sys­temic
mast cell dis­or­ders: a pro­spec­tive study of the Span­ish Network on Mastocytosis (REMA) in a series of 113 patients. Blood. 2006;108(7):2366-2372.
18. Jawhar M, Schwaab J, Schnittger S, et al. Molecular pro­fil­ing of mye­loid
pro­gen­i­tor cells in multi-mutated advanced sys­temic mastocytosis identifies KIT D816V as a dis­tinct and late event. Leukemia. 2015;29(5):1115-1122.
19. Jawhar M, Schwaab J, Schnittger S, et al. Additional muta­tions in SRSF2,
ASXL1 and/or RUNX1 iden­tify a high-risk group of patients with KIT D816V(+)
advanced sys­temic mastocytosis. Leukemia. 2016;30(1):136-143.
20. Jawhar M, Schwaab J, Hausmann D, et al. Splenomegaly, ele­vated alka­line
phos­pha­tase and muta­tions in the SRSF2/ASXL1/RUNX1 gene panel are
strong adverse prog­nos­tic mark­ers in patients with sys­temic mastocytosis.
Leukemia. 2016;30(12):2342-2350.
21. Lim KH, Pardanani A, Butterfield JH, Li C-Y, Tefferi A. Cytoreductive ther­apy
in 108 adults with sys­temic mastocytosis: out­come anal­y­sis and response
pre­dic­tion dur­ing treat­ment with inter­feron-alpha, hydroxy­urea, imatinib
mesylate or 2-chlorodeoxyadenosine. Am J Hematol. 2009;84(12):790-794.
22. Reiter A, George TI, Gotlib J. New devel­
op­
ments in diag­
no­
sis, prog­
nos­
ti­
ca­
tion, and treat­
ment of advanced sys­
temic mastocytosis. Blood.
2020;135(16):1365-1376.
23. Shomali W, Gotlib J. The new tool “KIT” in advanced sys­temic mastocytosis. Hematology Am Soc Hematol Educ Program. 2018;2018(1):127-136.
24. Pardanani A. Systemic mastocytosis in adults: 2021 update on diag­no­sis,
risk strat­i­fi­ca­tion and man­age­ment. Am J Hematol. 2021;96(4):508-525.
25. Akin C, Fumo G, Yavuz AS, Lipsky PE, Neckers L, Metcalfe DD. A novel form
of mastocytosis asso­
ci­
ated with a trans­
mem­
brane c-kit muta­
tion and
response to imatinib. Blood. 2004;103(8):3222-3225.
26. Ustun C, Reiter A, Scott BL, et al. Hematopoietic stem-cell trans­plan­ta­tion
for advanced sys­temic mastocytosis. J Clin Oncol. 2014;32(29):3264-3274.
27. Sotlar K, Marafioti T, Griesser H, et al. Detection of c-kit muta­tion Asp 816 to
Val in microdissected bone mar­row infil­trates in a case of sys­temic mastocytosis asso­ci­ated with chronic myelomonocytic leu­kae­mia. Mol Pathol.
2000;53(4):188-193.
28. Sotlar K, Colak S, Bache A, et al. Variable pres­ence of KITD816V in clonal
haematological non-mast cell lin­e­age dis­eases asso­ci­ated with sys­temic
mastocytosis (SM-AHNMD). J Pathol. 2010;220(5):586-595.
29. Valent P, Akin C, Hartmann K, et al. Midostaurin: a magic bul­let that blocks
mast cell expan­sion and acti­va­tion. Ann Oncol. 2017;28(10):2367-2376.
74. Rohmer J, Couteau-Chardon A, Trichereau J, et al; CEREO and GBMHM
col­lab­o­ra­tors. Epidemiology, clin­i­cal pic­ture and long-term out­comes of
FIP1L1-PDGFRA-pos­i­tive mye­loid neo­plasm with eosin­o­philia: data from 151
patients. Am J Hematol. 2020;95(11):1314-1323.
75. Metzgeroth G, Schwaab J, Naumann N, et al. Treatment-free remis­sion in
FIP1L1-PDGFRA-pos­i­tive mye­loid/lym­phoid neo­plasms with eosin­o­philia
after imatinib dis­con­tin­u­a­tion. Blood Adv. 2020;4(3):440-443.
76. David M, Cross NC, Burgstaller S, et al. Durable responses to imatinib in
patients with PDGFRB fusion gene-pos­i­tive and BCR-ABL-neg­a­tive chronic
mye­lo­pro­lif­er­a­tive dis­or­ders. Blood. 2007;109(1):61-64.
77. Cheah CY, Burbury K, Apperley JF, et al. Patients with mye­loid malig­nan­cies
bear­ing PDGFRB fusion genes achieve dura­ble long-term remis­sions with
imatinib. Blood. 2014;123(23):3574-3577.
78. Jawhar M, Naumann N, Schwaab J, et al. Imatinib in mye­loid/lym­phoid
neo­plasms with eosin­o­philia and rearrangement of PDGFRB in chronic or
blast phase. Ann Hematol. 2017;96(9):1463-1470.
79. Ohnishi H, Kandabashi K, Maeda Y, Kawamura M, Watanabe T. Chronic
eosin­o­philic leu­kae­mia with FIP1L1-PDGFRA fusion and T6741 muta­tion that
evolved from Lang­er­hans cell histiocytosis with eosin­o­philia after che­mo­
ther­apy. Br J Haematol. 2006;134(5):547-549.
80. Lierman E, Michaux L, Beullens E, et al. FIP1L1-PDGFRalpha D842V, a novel
panresistant mutant, emerg­
ing after treat­
ment of FIP1L1-PDGFRalpha
T674I eosin­o­philic leu­ke­mia with sin­gle agent sorafenib. Leukemia.
2009;23(5):845-851.
81. Lierman E, Folens C, Stover EH, et al. Sorafenib is a potent inhib­i­tor of
FIP1L1-PDGFRalpha and the imatinib-resis­
tant FIP1L1-PDGFRalpha T674I
mutant. Blood. 2006;108(4):1374-1376.
82. von Bubnoff N, Gorantla SP, Thöne S, Peschel C, Duyster J. The FIP1L1-PDGFRA T674I muta­
tion can be inhibited by the tyro­
sine kinase inhib­
i­
tor
AMN107 (nilotinib). Blood. 2006;107(12):4970-4971; author reply 4972.
83. Metzgeroth G, Erben P, Martin H, et al. Limited clin­i­cal activ­ity of nilotinib
and sorafenib in FIP1L1-PDGFRA pos­i­tive chronic eosin­o­philic leu­ke­mia with
imatinib-resis­tant T674I muta­tion. Leukemia. 2012;26(1):162-164.
84. Rumi E, Milosevic JD, Sellesag D, et al. Efficacy of ruxolitinib in myeloid neoplasms with PCM1-JAK2 fusion gene. Ann Hematol. 2015;94(11):1927-1928.
85. Tang G, Tam W, Short NJ, et al. Myeloid/lymphoid neoplasms with FLT3
rearrangement. Mod Pathol. 2021;34(9):1673-1685.
86. Schwaab J, Naumann N, Luebke J, et al. Response to tyro­sine kinase inhib­
i­
tors in mye­
loid neo­
plasms asso­
ci­
ated with PCM1-JAK2, BCR-JAK2 and
ETV6-ABL1 fusion genes. Am J Hematol. 2020;95(7):824-833.
87. Spitzer B, Dela Cruz FS, Ibanez Sanchez GD, et al. ETV6-FLT3-pos­
i­
tive
mye­loid/lym­phoid neo­plasm with eosin­o­philia presenting in an infant: an
entity dis­tinct from JMML. Blood Adv. 2021;5(7):1899-1902.
88. Macdonald D, Aguiar RC, Mason PJ, Goldman JM, Cross NC. A new mye­lo­
pro­lif­er­a­tive dis­or­der asso­ci­ated with chro­mo­somal trans­lo­ca­tions involv­
ing 8p11: a review. Leukemia. 1995;9(10):1628-1630.
89. Umino K, Fujiwara S-I, Ikeda T, et al. Clinical out­comes of mye­loid/lym­
phoid neo­plasms with fibro­blast growth fac­tor recep­tor-1 (FGFR1) rearrangement. Hematology. 2018;23(8):470-477.
90. Chen J, Deangelo DJ, Kutok JL, et al. PKC412 inhib­
its the zinc fin­
ger
198-fibro­blast growth fac­tor recep­tor 1 fusion tyro­sine kinase and is active
in treat­ment of stem cell mye­lo­pro­lif­er­a­tive dis­or­der. Proc Natl Acad Sci
U S A. 2004;101(40):14479-14484.
91. Khodadoust MS, Luo B, Medeiros BC, et al. Clinical activ­ity of ponatinib in
a patient with FGFR1-rearranged mixed-phe­no­type acute leu­ke­mia. Leukemia. 2016;30(4):947-950.
92. Kreil S, Ades L, Bommer M, et al. Limited effi­cacy of ponatinib in mye­lo­pro­
lif­er­a­tive neo­plasms asso­ci­ated with FGFR1 fusion genes [ab­stract]. Blood.
2015;126:Abstract 2812.
93. Hernández-Boluda J-C, Pereira A, Zinger N, et al. Allogeneic hema­to­poi­
etic cell trans­plan­ta­tion in patients with mye­loid/lym­phoid neo­plasm with
FGFR1-rearrangement: a study of the Chronic Malignancies Working Party
of EBMT. Bone Marrow Transplant. 2022;57(3):416-422.
94. Kasbekar M, Nardi V, Dal Cin P, et al. Targeted FGFR inhi­bi­tion results in a
dura­ble remis­sion in an FGFR1-driven mye­loid neo­plasm with eosin­o­philia.
Blood Adv. 2020;4(13):3136-3140.
95. Merz V, Zecchetto C, Melisi D. Pemigatinib, a potent inhib­i­tor of FGFRs for
the treat­ment of cholangiocarcinoma. Future Oncol. 2021;17(4):389-402.
96. Gotlib J, Kiladjian J-J, Vannucchi A, et al. A phase 2 study of pemigatinib
(FIGHT-203; INCB054828) in patients with mye­loid/lym­phoid neo­plasms
(MLNs) with fibro­blast growth fac­tor recep­tor 1 (FGFR1) rearrangement
(MLNFGFR1) [ab­stract]. Blood. 2021;138(suppl 1):385.
Therapies for advanced SM & eosin­o­philic neo­plasms | 45
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/34/2021847/34gotlib.pdf by guest on 09 December 2022
54. Gotlib J, George TI, Deininger MW, et al. A phase 2 study of bezuclastinib
(CGT9486), an oral, selec­
tive, and potent KIT D816V inhib­
i­
tor, in adult
patients with advanced sys­
temic mastocytosis. Blood. 2021:138(suppl
1):3636.
55. Reiter A, Gotlib J. Myeloid neo­
plasms with eosin­
o­
philia. Blood.
2017;129(6):704-714.
56. Shomali W, Gotlib J. World Health Organization-defined eosin­o­philic dis­or­
ders: 2022 update on diag­no­sis, risk strat­i­fi­ca­tion, and man­age­ment. Am J
Hematol. 2022;97(1):129-148.
57. NCCN Clinical Practice Guidelines in Oncology. Myeloid/lym­phoid neo­
plasms with eosin­o­philia and tyro­sine kinase fusion genes. Version 1.2022.
https:​­ /​­/www​­ . nccn​­ .org​­ /professionals​­ /physician_gls​­ /pdf​­ /mlne​­ . pdf.
Accessed 18 July 2022.
58. Snider JS, Znoyko I, Lindsey KG, et al. Integrated geno­mic anal­y­sis using
chro­mo­somal microarray, fluo­res­cence in situ hybrid­iza­tion and mate pair
ana­ly­ses: char­ac­ter­iza­tion of a cryp­tic t(9;22)(p24.1;q11.2)/BCR-JAK2 in
mye­loid/lym­phoid neo­plasm with eosin­o­philia. Cancer Genet. 2020;246247:44-47.
59. Cools J, DeAngelo DJ, Gotlib J, et al. A tyro­sine kinase cre­ated by fusion
of the PDGFRA and FIP1L1 genes as a ther­a­peu­tic tar­get of imatinib in
idi­o­pathic hypereosinophilic syn­drome. N Engl J Med. 2003;348(13):
1201-1214.
60. Klion AD, Noel P, Akin C, et al. Elevated serum tryptase lev­els iden­tify a
sub­set of patients with a mye­lo­pro­lif­er­a­tive var­i­ant of idi­o­pathic hypereosinophilic syn­drome asso­ci­ated with tis­sue fibro­sis, poor prog­no­sis, and
imatinib respon­sive­ness. Blood. 2003;101(12):4660-4666.
61. Pardanani A, Brockman SR, Paternoster SF, et al. FIP1L1-PDGFRA fusion:
prev­a­lence and clin­i­co­path­o­logic cor­re­lates in 89 con­sec­u­tive patients
with mod­er­ate to severe eosin­o­philia. Blood. 2004;104(10):3038-3045.
62. Metzgeroth G, Walz C, Score J, et al. Recurrent find­ing of the FIP1L1-PDGFRA fusion gene in eosin­o­philia-asso­ci­ated acute mye­loid leu­ke­mia and
lym­pho­blas­tic T-cell lym­phoma. Leukemia. 2007;21(6):1183-1188.
63. Pardanani A, Ketterling RP, Brockman SR, et al. CHIC2 dele­tion, a sur­ro­gate
for FIP1L1-PDGFRA fusion, occurs in sys­temic mastocytosis asso­ci­ated with
eosin­o­philia and pre­dicts response to imatinib mesylate ther­apy. Blood.
2003;102(9):3093-3096.
64. Olsson-Arvidsson L, Norberg A, Sjögren H, Johansson B. Frequent falseneg­
a­
tive FIP1L1-PDGFRA FISH ana­
ly­
ses of bone mar­
row sam­
ples from
clonal eosin­o­philia at diag­no­sis. Br J Haematol. 2020;188(5):e76-e79.
65. Klion AD, Robyn J, Akin C, et al. Molecular remis­sion and rever­sal of mye­
lo­
fi­
bro­
sis in response to imatinib mesylate treat­
ment in patients with
the mye­lo­pro­lif­er­a­tive var­i­ant of hypereosinophilic syn­drome. Blood.
2004;103(2):473-478.
66. Baccarani M, Cilloni D, Rondoni M, et al. The effi­cacy of imatinib mesylate
in patients with FIP1L1-PDGFRalpha-pos­i­tive hypereosinophilic syn­drome.
Results of a mul­ti­cen­ter pro­spec­tive study. Haematologica. 2007;92(9):
1173-1179.
67. Jovanovic JV, Score J, Waghorn K, et al. Low-dose imatinib mesylate leads
to rapid induc­tion of major molec­u­lar responses and achieve­ment of com­
plete molec­u­lar remis­sion in FIP1L1-PDGFRA-pos­i­tive chronic eosin­o­philic
leu­ke­mia. Blood. 2007;109(11):4635-4640.
68. Gotlib J, Cools J. Five years since the dis­cov­ery of FIP1L1-PDGFRA: what we
have learned about the fusion and other molec­u­larly defined eosin­o­phil­ias.
Leukemia. 2008;22(11):1999-2010.
69. Helbig G, Stella-Hołowiecka B, Majewski M, et al. A sin­gle weekly dose of
imatinib is suf­fi­cient to induce and main­tain remis­sion of chronic eosin­
o­philic leu­kae­mia in FIP1L1-PDGFRA-expressing patients. Br J Haematol.
2008;141(2):200-204.
70. Pardanani A, Ketterling RP, Li C-Y, et al. FIP1L1-PDGFRA in eosin­
o­
philic
dis­or­ders: prev­a­lence in rou­tine clin­i­cal prac­tice, long-term expe­ri­ence
with imatinib ther­apy, and a crit­i­cal review of the lit­er­a­ture. Leuk Res.
2006;30(8):965-970.
71. Klion AD, Robyn J, Maric I, et al. Relapse fol­
low­
ing dis­
con­
tin­
u­
a­
tion of
imatinib mesylate ther­
­
apy for FIP1L1/PDGFRA-pos­
i­
tive chronic eosin­
o­
philic leu­ke­mia: impli­ca­tions for opti­mal dos­ing. Blood. 2007;110(10):35523556.
72. Pardanani A, D’Souza A, Knudson RA, Hanson CA, Ketterling RP, Tefferi A.
Long-term fol­low-up of FIP1L1-PDGFRA-mutated patients with eosin­o­philia:
sur­vival and clin­i­cal out­come. Leukemia. 2012;26(11):2439-2441.
73. Helbig G, Kyrcz-Krzemień S. Cessation of imatinib mesylate may lead to
sustained hema­to­logic and molec­u­lar remis­sion in FIP1L1-PDGFRA-mutated
hypereosinophilic syn­drome. Am J Hematol. 2014;89(1):115.
46 | Hematology 2022 | ASH Education Program
102. Shomali W, Damnernsawad A, Theparee T, et al. A novel acti­vat­ing JAK1
muta­tion in chronic eosin­o­philic leu­ke­mia. Blood Adv. 2021;5(18):35813586.
103. Helbig G, Soja A, Bartkowska-Chrobok A, Kyrcz-Krzemień S. Chronic
eosin­o­philic leu­ke­mia-not oth­er­wise spec­i­fied has a poor prog­no­sis with
unre­spon­sive­ness to con­ven­tional treat­ment and high risk of acute trans­
for­ma­tion. Am J Hematol. 2012;87(6):643-645.
104. Kuang FL, Fay MP, Ware J, et al. Long-term clin­i­cal out­comes of high-dose
mepolizumab treat­ment for hypereosinophilic syn­drome. J Allergy Clin
Immunol Pract. 2018;6(5):1518-1527.e51527e5.
105. Kuang FL, Legrand F, Makiya M, et al. Benralizumab for PDGFRA-neg­a­tive
hypereosinophilic syn­drome. N Engl J Med. 2019;380(14):1336-1346.
© 2022 by The Amer­i­can Society of Hematology
DOI 10.1182/hema­tol­ogy.2022000368
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/34/2021847/34gotlib.pdf by guest on 09 December 2022
97. Wang SA, Hasserjian RP, Tam W, et al. Bone mar­row mor­phol­ogy is a
strong dis­crim­i­na­tor between chronic eosin­o­philic leu­ke­mia, not oth­er­
wise spec­i­fied and reac­tive idi­o­pathic hypereosinophilic syn­drome. Haematologica. 2017;102:1352-1360.
98. Morsia E, Reichard K, Pardanani A, Tefferi A, Gangat N. WHO defined chronic
eosin­o­philic leu­ke­mia, not oth­er­wise spec­i­fied (CEL, NOS): a con­tem­po­
rary series from the Mayo Clinic. Am J Hematol. 2020;95(7):E172-E174.
99. Wang SA, Tam W, Tsai AG, et al. Targeted next-gen­er­a­tion sequenc­ing
identifies a sub­set of idi­o­pathic hypereosinophilic syn­drome with fea­
tures sim­i­lar to chronic eosin­o­philic leu­ke­mia, not oth­er­wise spec­i­fied.
Mod Pathol. 2016;29(8):854-864.
100. Cross NCP, Hoade Y, Tapper WJ, et al. Recurrent acti­
vat­
ing STAT5B
N642H muta­tion in mye­loid neo­plasms with eosin­o­philia. Leukemia.
2019;33(2):415-425.
101. Patel AB, Franzini A, Leroy E, et al. JAK2 ex13InDel drives onco­genic trans­
for­ma­tion and is asso­ci­ated with chronic eosin­o­philic leu­ke­mia and poly­
cy­the­mia vera. Blood. 2019;134(26):2388-2398.
ANXIETY PROVOKING CONSULTATIONS: MAST CELLS AND EOSINOPHILS
Approach to the patient with suspected
hypereosinophilic syndrome
Human Eosinophil Section, Laboratory of Parasitic Diseases, National Institute of Allergy and Infectious Diseases, National Institutes of Health,
Bethesda, MD
Hypereosinophilic syndromes (HES) are a heterogenous group of rare disorders with clinical manifestations ranging from
fatigue to life-threatening endomyocardial fibrosis and thromboembolic events. Given the broad differential diagnosis
of HES, a comprehensive approach is needed to identify potential secondary (treatable) causes and define end-organ
manifestations. Classification by clinical HES subtype is also useful in terms of assessing prognosis and guiding therapy.
Corticosteroids remain the mainstay of initial therapy in the setting of acute, life-threatening PDGFR mutation-negative
HES. Whereas the recent availability of eosinophil-targeted therapies with extraordinary efficacy and little apparent toxicity is changing the treatment paradigm, especially for idiopathic HES and overlap syndromes, questions remain unanswered regarding the choice of agent, impact of combination therapies, and long-term effects of eosinophil depletion.
This review provides a case-based discussion of the differential diagnosis of HES, including the classification by clinical
HES subtype. Treatment options are reviewed, including novel eosinophil-targeted agents recently approved for the
treatment of HES and/or other eosinophil-associated disorders. Primary (myeloid) disorders associated with hypereosinophilia are not be addressed in depth in this review.
LEARNING OBJECTIVES
• To describe the heterogeneity of clinical presentations of hypereosinophilia
• To discuss the approach to targeted therapy of hypereosinophilic syndrome
CLINICAL CASE
A 38­year­old previously healthy woman presented with
intermittent but intense pruritus without rash. The pruritus
initially involved only her ankles but subsequently spread
up her legs with accompanying angioedema of the thighs
and buttocks that prevented her from being able to wear
pants. Antihistamines were ineffective, and a short course
of solumedrol prescribed for sinusitis did not relieve the
pruritus or swelling. A dermatologist prescribed oral
and topical antibiotics for presumed folliculitis without
improvement. One year after the onset of the symptoms,
a complete blood count was performed and revealed ane­
mia, thrombocytopenia (platelets 34 000), and eosinophilia
(14.0×109/L). She was referred to hematology.
Differential diagnosis and initial evaluation
Eosinophilia, defined as an absolute eosinophil count
(AEC) >0.45 × 109/L, is quite common, occurring in 1% to
2% of the general population.1 In contrast, hypereosino­
philia (HE; AEC ≥1.5 × 109/L) is extremely rare, with an esti­
mated incidence of 0.315 to 6.3 per 100 000 in the United
States.2 The potential etiologies of eosinophilia (including
HE) are varied and include allergic, infectious, neoplas­
tic, genetic, and immune disorders (Table 1). Moreover,
clinical symptoms are extremely heterogeneous. Derma­
tologic, pulmonary, and gastrointestinal manifestations
are most frequently reported, but any organ system can
be affected, and progression can occur over time with­
out effective therapy.3 A careful and complete history and
physical examination, including prior complete blood
counts (if available), medication and travel history, assess­
ment of cancer risk factors, and family history, is essential
to narrow the differential. Initial laboratory and diagnos­
tic testing should include complete blood count with
differential, routine chemistries, serum immunoglobulin
levels, B12 and tryptase, and assessment of lymphocyte
clonality and phenotype. If there is a possible history of
Strongyloides exposure, no matter how remote, serologic
testing should be performed and/or empiric ivermectin
(150 µg/kg × 1 dose) administered to prevent potentially
Hypereosinophilia and hypereosinophilic syndromes | 47
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/47/2021769/47klion.pdf by guest on 09 December 2022
Amy D. Klion
Table 1. Disorders asso­ci­ated with marked eosin­o­philia
Examples
Comments
Asthma; atopic der­ma­ti­tis; chronic
rhinosinusitis
Typically cause mild to mod­er­ate eosin­o­philia
Drug hyper­sen­si­tiv­ity
Varied; drug rash, eosin­o­philia, and sys­temic
symp­toms; eosin­o­philia-myal­gia syn­drome
Can occur with any drug or sup­ple­ment;
man­i­fes­ta­tions range from asymp­tom­atic
eosin­o­philia to life-threat­en­ing com­pli­ca­tions40
Infection and infes­ta­tion
Helminth infec­tion (espe­cially those with a
tis­sue inva­sive phase)
Fungal infec­tions
Viral infec­tion (HIV, COVID-19)
Ectoparasite infes­ta­tion
Protozoal infec­tion (lim­ited to Sarcocystis and
Cystoisospora)
Tuberculosis (rare)
Most com­mon eti­­ol­ogy world­wide;
Strongyloides infec­tion should always be
con­sid­ered due to world­wide dis­tri­bu­tion
and the poten­tial for fatal dis­sem­i­na­tion with
ste­roid ther­apy41
Autoimmune and immunodysregulatory
dis­or­ders
Inflammatory bowel dis­ease, sar­coid­o­sis, IgG4
dis­ease
Clinical sequelae of eosin­o­philia may or may
not be pres­ent and can be dif­fi­cult to ­
dis­tin­guish from the man­i­fes­ta­tions of the
under­ly­ing dis­or­der
Neoplasia
Leukemia, lym­phoma, solid tumors (espe­cially
ade­no­car­ci­noma)
Although any leu­ke­mia/lym­phoma can pres­ent
with HE/HES, pre–B-cell acute lym­pho­blas­tic
leu­ke­mia can be par­tic­u­larly dif­fi­cult to
diag­nose
Inborn errors of immu­nity
Omenn syn­drome, DOCK8 defi­ciency,
Loeys-Dietz syn­drome
Usually diag­nosed in child­hood, recur­rent or
unusual infec­tions com­mon42
Rare hypereosinophilic syn­dromes
Eosinophilic mye­loid neo­plasms, lym­pho­cytic
var­i­ant HE/HES, idi­o­pathic HE/HES, famil­ial
HE/HES, sin­gle-organ HE/HES, and other
over­lap dis­or­ders
See Figure 2 for addi­tional details
Other
Radiation, hypoadrenalism, cho­les­terol emboli,
admin­is­tra­tion of IL-2
COVID-19, coronavirus dis­ease 2019; HIV, human immu­no­de­fi­ciency virus.
fatal hyperinfection syn­
drome. Bone mar­
row biopsy and
chest/abdo­
men/pel­
vis imag­
ing should be strongly con­
sid­
ered in any patient with AEC ≥1.5 × 109/L and no clear sec­ond­ary
cause of the HE. Additional test­ing, includ­ing test­ing for par­a­
sitic infec­tions other than Strongyloides, should be guided by
the clin­i­cal his­tory and dis­ease man­i­fes­ta­tions.
CLINICAL CASE (Con­t in­u ed)
A bone mar­
row biopsy spec­
i­
men was hypercellular with
increased eosin­o­phils and plasma cells. Testing for FIP1L1::PDGFRA was neg­a­tive. She was treated with pred­ni­sone 60mg daily
for pre­sumed idi­o­pathic throm­bo­cy­to­pe­nic pur­pura with clin­i­cal
and hema­to­logic improve­ment. However, as the pred­ni­sone was
tapered, the eosin­o­philia, pru­ri­tus, and edema returned. She was
referred to Mayo Clinic for fur­ther eval­u­a­tion, which included an
inde­ter­mi­nate sero­logic test for Strongyloides. She was treated
with 2 doses of iver­mec­tin and a 2-week course of albendazole.
During this time, the pred­ni­sone was slowly tapered despite
a ris­ing AEC, peaking at 26.0×109/L on pred­ni­sone 5mg every
other day, and recur­rent symp­toms. The pred­ni­sone dose was
increased to 25mg daily, and hydroxy­urea ther­apy (500mg twice
daily) was ini­ti­ated. This was inef­fec­tive, and she was referred to
the National Institutes of Health for fur­ther eval­u­a­tion.
48 | Hematology 2022 | ASH Education Program
At the time of refer­ral, she complained of fatigue, swell­ing,
and extreme pru­ri­tus. Physical exam­i­na­tion revealed sym­met­
ric nonpitting edema of the thighs and exco­
ri­
a­
tions pre­
dom­
i­
nantly on the lower legs. Laboratory test­ing was nota­ble for AEC
3.01×109/L; plate­lets 114 000; mark­edly ele­vated IgG, IgM, and IgE;
and nor­mal serum B12 and tryptase. Computed tomog­ra­phy (CT)
scan was nota­ble for bor­der­line spleno­meg­aly and min­i­mal dif­
fuse lymph­ade­nop­a­thy. Repeat bone mar­row again showed only
increased eosin­o­phils. Mast cells were not increased, and test­ing
for D816V KIT was neg­a­tive. T-cell recep­tor test­ing by poly­mer­ase
chain reac­tion showed a clonal pat­tern, and flow cytom­e­try was
nota­ble for an aber­rant CD3–CD4+CD10+ T-cell pop­u­la­tion, con­
sis­tent with a diag­no­sis of lym­pho­cytic var­i­ant hypereosinophilic
syn­drome (HES). B-cell clonality stud­ies were neg­a­tive. She was
started on inter­feron α 1mU daily with res­o­lu­tion of eosin­o­philia,
plate­lets >50 k and symp­tom­atic ­improve­ment.
Definition and clas­si­fi­ca­tion of HES
The def­i­ni­tion of HES has evolved over time since Chusid’s land­
mark descrip­tion of 14 patients with idi­o­pathic HE and var­ied clin­
i­cal man­i­fes­ta­tions in 1975.4 Whereas the cur­rent World Health
Organization def­i­ni­tion uses the term HES to describe only idi­
o­pathic HE with clin­i­cal man­i­fes­ta­tions,5 a recently updated
con­sen­sus def­i­ni­tion pro­vi­des a broader approach that rec­og­
nizes the over­lap in clin­i­cal pre­sen­ta­tion between idi­o­pathic and
other types of HES, the imper­fect sen­si­tiv­ity and spec­i­fic­ity of
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/47/2021769/47klion.pdf by guest on 09 December 2022
Category
Atopic dis­or­ders
Table 2. Definitions of hypereosinophilic syn­drome
WHO def­i­ni­tion of HES
Consensus def­i­ni­tion
AEC >1.5×109/L for greater than 6 months
AEC >1.5×109/L on 2 exam­i­na­tions at least 1 month apart and/or
tis­sue HE
Evidence of end-organ man­i­fes­ta­tions
Organ dam­age and/or dys­func­tion attrib­ut­­able to tis­sue HE and ­
exclu­sion of other con­di­tions as a major rea­son for organ dam­age
“Idiopathic”—exclu­sion of reac­tive HE, lym­pho­cytic var­i­ant HES,
CEL-NOS, WHO-defined malig­nan­cies, eosin­o­philia-asso­ci­ated MPNs,
or AML/ALL with rearrangements of PDGFRA, PDGFRB, FGFR1, or
PCM1-JAK2
Includes mul­ti­ple clin­i­cal clas­si­fi­ca­tions, includ­ing idi­o­pathic, pri­mary
(clonal/neo­plas­tic), sec­ond­ary (pre­sumed cyto­kine driven—includ­ing
lym­pho­cytic var­i­ant), and hered­i­tary
ALL, acute lymphocytic leukemia; AML, acute myeloid leukemia; CEL-NOS, chronic eosinophilic leukemia-not otherwise specified; MPN,
myeloproliferative neoplasm; WHO, World Health Organization.
the periph­eral blood. Expanded sur­face phenotyping is often
nec­es­sary to dem­on­strate and/or con­firm the aber­rant pop­u­la­
tion.12,13 It is impor­tant to rec­og­nize that the sur­face phe­no­type
of the clonal pop­u­la­tion in LHES can be indis­tin­guish­able from
that seen in T-cell malig­nan­cies (espe­cially angioimmunoblastic
T-cell lym­phoma and cuta­ne­ous T-cell lym­phoma).14 Thus, LHES
is a diag­no­sis of exclu­sion. Moreover, pro­gres­sion of LHES to a
lym­phoid malig­nancy occurs in approx­i­ma­tely 10% of patients,
some­times after many years of sta­ble dis­ease.15-17 Consequently,
at a min­i­mum, patients with LHES should undergo assess­ment for
occult lym­phoma at diag­no­sis and in the set­ting of an increase in
size of the clonal T-cell pop­u­la­tion or devel­op­ment of resis­tance
to pre­vi­ously effec­tive ther­apy.
Approach to ther­apy
Despite the dif­fer­ences in def­i­ni­tions, the gen­eral approach to
HE is very sim­i­lar between World Health Organization and the
con­sen­sus group (Figure 1). Since sec­ond­ary causes of eosin­
o­philia, such as hel­minth infec­tion, typ­i­cally require a dif­fer­ent
ther­a­peu­tic approach, these should be con­sid­ered early in the
diag­nos­tic pro­cess. If an under­ly­ing eti­­ol­ogy is iden­ti­fied or
highly suspected, spe­cific treat­ment should be ini­ti­ated. Primary
(clonal/neo­plas­tic) eosin­o­philia is also impor­tant to iden­tify
early due to prog­nos­tic and ther­a­peu­tic impli­ca­tions.18 Finally,
the pres­ence and sever­ity of clin­i­cal man­i­fes­ta­tions should be
assessed as this will affect both the nature and urgency of ther­
a­peu­tic inter­ven­tion. For exam­ple, care­ful mon­i­tor­ing with­out
ther­apy may be appro­pri­ate for asymp­tom­atic HE with­out evi­
dence of end-organ involve­ment (hypereosinophilia of unde­ter­
mined sig­nif­i­cance),19 whereas urgent inter­ven­tion is needed in
the con­text of myo­car­di­tis or throm­bo­em­bo­lism.
Prednisone remains the main­stay of ther­apy in the acute
set­ting for severe and/or life-threat­en­ing man­i­fes­ta­tions
of HES. If the eosin­
o­
philia does not dra­
mat­
i­
cally decrease
within 24 to 48 hours, addi­tional ther­apy should be con­sid­
ered depending on the suspected clin­i­cal sub­type (ie, imati­
nib for patients with clin­i­cal find­ings sug­ges­tive of a mye­loid
neo­plasm, cyclo­phos­pha­mide for patients with man­i­fes­ta­tions
sug­ges­tive of eosin­o­philic granulomatosis with polyangiitis).
The use of eosin­
o­
phil-targeting bio­
log­
ics in the acute set­
ting remains con­tro­ver­sial but is supported by case reports
and small series.20 Once the patient is sta­ble, fur­ther eval­u­a­
tion should focus on the iden­ti­fi­ca­tion of the most likely clin­
i­
cal sub­
type (Table 3). With the excep­
tion of patients with
mye­loid HES,18 most patients with symp­tom­atic HES respond
Hypereosinophilia and hypereosinophilic syn­dromes | 49
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/47/2021769/47klion.pdf by guest on 09 December 2022
avail­­able diag­nos­tic test­ing, and the iden­ti­fi­ca­tion of new eti­­
ol­o­gies of HES over time6 (Table 2). In this con­sen­sus def­i­ni­tion,
a diag­no­sis of HE requires AEC >1.5×109/L on 2 exam­i­na­tions at
least 1 month apart (to exclude lab­o­ra­tory error but allow diag­
no­sis with­out a delay of 6 months) and/or tis­sue HE (to rec­og­
nize the arbi­trary nature of the AEC cut­off in the set­ting of clear
eosin­o­phil-medi­ated dis­ease). HES is defined as HE with evi­
dence of end-organ dys­func­tion attrib­ut­­able to the eosin­o­philia,
irrespective of the cause. To address the het­ero­ge­ne­ity of dis­or­
ders included in this umbrella def­i­ni­tion of HES and help guide
diag­nos­tic pro­ce­dures and ther­a­peu­tic choices, the fol­low­ing
clin­
i­
cal sub­
types have been pro­
posed (Table 3)7: (1) mye­
loid
HE/HES (suspected or proven eosin­o­philic mye­loid neo­plasm,
includ­ing those asso­ci­ated with rearrangements of PDGFRA and
other recur­rent molec­u­lar abnor­mal­i­ties), (2) lym­pho­cytic var­i­ant
HE/HES (pres­ence of a clonal or phe­no­typ­i­cally aber­rant T-cell
pop­u­la­tion that pro­duces cyto­kines that drive the eosin­o­philia),
(3) over­lap HES (sin­gle-organ-restricted eosin­o­philic dis­or­ders
and clin­
i­
cally defined eosin­
o­
philic syn­
dromes that over­
lap in
pre­sen­ta­tion with idi­o­pathic HES; ie, eosin­o­philic gas­tro­in­tes­ti­
nal dis­or­ders, eosin­o­philic granulomatosis with polyangiitis), (4)
asso­ci­ated HE/HES (in the con­text of a defined dis­or­der, such
as a hel­minth infec­tion, neo­plasm, immu­no­de­fi­ciency, or hyper­
sen­si­tiv­ity reac­tion), (5) famil­ial HE/HES (occur­rence in >1 fam­
ily mem­ber exclud­ing asso­ci­ated HE/HES), and (6) idi­o­pathic
HE/HES (unknown cause and exclu­sion of other sub­types).
Lymphocytic var­i­ant HES (LHES) was first described in 1994 in
a 30-year-old man with pru­ri­tus and cough and a large CD2+CD3–
CD4+ T-cell pop­u­la­tion that pro­duced inter­leu­kin (IL) 4 and IL-5
in response to stim­u­la­tion.8 Since that time, other sur­face phe­
no­
types have been described, and there have been sev­
eral
infor­ma­tive case series describ­ing the clin­i­cal and lab­o­ra­tory
find­ings of patients with LHES.9-11 Equally fre­quent in males and
females, LHES most often pres­ents with der­ma­to­logic man­i­fes­
ta­tions. That said, any organ sys­tem can be involved, and some
patients with asymp­tom­atic HE have clonal aber­rant T-cell pop­
u­la­tions indis­tin­guish­able from those with symp­tom­atic LHES.
Serum IgM, IgE, and serum and thy­mus and acti­va­tion-reg­u­lated
chemokine lev­els are ele­vated in most patients with LHES and
can pro­vide use­ful diag­nos­tic clues.12 Whereas the gold stan­dard
for diag­no­sis of LHES is iden­ti­fi­ca­tion of a clonal and/or aber­
rant pop­u­la­tion of T cells pro­duc­ing type 2 cyto­kines, intra­cel­
lu­lar flow cytom­e­try is not avail­­able at most cen­ters, and the
diag­no­sis most often relies on a com­pat­i­ble clin­i­cal pic­ture and
dem­on­stra­tion of a clonal and/or aber­rant T-cell pop­u­la­tion in
Table 3. Initial assess­ment of the patient with hypereosinophilia
Comments
Comprehensive his­tory and phys­i­cal exam­i­na­tion
Including prior eosin­o­phil counts, med­i­ca­tions, travel/expo­sure his­tory
Complete blood count with dif­fer­en­tial and smear*
Dysplastic eosin­o­phils, other lin­e­age involve­ment, and/or pres­ence of
mye­loid pre­cur­sors are sug­ges­tive of (but not diag­nos­tic for) MHES
Routine chem­is­tries, includ­ing liver func­tion tests*
To assess end organ involve­ment
Quantitative serum immu­no­glob­u­lin lev­els
IgE lev­els are typ­i­cally ele­vated in a vari­ety of con­di­tions (ie, LHES,
EGPA, par­a­sitic infec­tions, and some immunodeficiencies); IgM lev­els
are ele­vated in LHES and epi­sodic angioedema and eosin­o­philia
Serum tryptase and B12 lev­els
Elevated serum B12 lev­els can be seen in many mye­loid neo­plasms;
ele­vated serum tryptase is near uni­ver­sal in PDGFRA and KIT-asso­ci­ated
dis­ease
T- and B-cell recep­tor rearrangement stud­ies*; lym­pho­cyte phenotyping
by flow cytom­e­try* (see Carpentier et al11)†
Clonal and/or aber­rant T-cell pop­u­la­tions are char­ac­ter­is­tic of LHES
and some types of lym­phoma. Clonal B cells are sus­pi­cious for B-cell
neo­plasm, includ­ing pre–B-cell acute lym­pho­blas­tic leu­ke­mia in ­
chil­dren/ado­les­cents.
Serum tro­po­nin,* elec­tro­car­dio­gram, and echo­car­dio­gram
If abnor­mal, car­diac MRI should be con­sid­ered
Chest/abdo­men/pel­vis CT*
To assess for spleno­meg­aly, lymph­ade­nop­a­thy, asymp­tom­atic
pul­mo­nary involve­ment, and occult neo­plasms
Biopsy of affected tis­sues (if pos­si­ble)*
Cardiac tis­sue involve­ment can be patchy, lim­it­ing the util­ity of car­diac
biopsy
Selected patients with HE/HES
Pulmonary func­tion tests*
Any patient with suspected pul­mo­nary involve­ment or abnor­mal
find­ings on chest CT
Bone mar­row aspi­rate and biopsy*
All patients with AEC >5.0×109/L and/or fea­tures sug­ges­tive of LHES or
MHES; patients with clear diag­noses, such as EGPA or par­a­sitic infec­tion,
may not need bone mar­row test­ing despite AEC >5.0×109/L
Testing for BCR::ABL1, FIP1L1::PDGFRA, and trans­lo­ca­tions/muta­tions
involv­ing PDGFRB, JAK2, FGFR1, and KIT
Testing should be guided by results of ini­tial test­ing and bone mar­row
exam­i­na­tion; all­patients with ele­vated serum tryptase and/or B12 lev­els
should be tested for FIP1L1::PDGFRA
NGS mye­loid panel; targeted or whole-exome sequenc­ing; other
genetic test­ing
Depending on ini­tial eval­u­a­tion
PET scan,* EBV viral load
Particularly in patients with suspected LHES
Other test­ing for sec­ond­ary causes
As indi­cated by clin­i­cal his­tory and phys­i­cal exam­i­na­tion
*Can be dra­mat­i­cally affected by cor­ti­co­ste­roid ther­apy.
†Not all­patients with LHES will have clonal or aber­rant T-cell pop­u­la­tions detect­able by rou­tine test­ing.
CT, com­puted tomog­ra­phy; EBV, Epstein-Barr virus; EGPA, eosin­o­philic granulomatosis and polyangiitis; MHES, mye­loid var­i­ant hypereosinophilic
syn­drome; MRI, mag­netic res­o­nance imag­ing; NGS, next-gen­er­a­tion sequenc­ing; PET, pos­i­tron emis­sion tomog­ra­phy.
rap­idly to ­cor­ti­co­ste­roid ther­apy, although tox­ic­ity and resis­
tance limit the util­ity of this ther­apy in the long term.3 Conven­
tional sec­ond-line agents, includ­ing hydroxy­urea and inter­feron
α, are fraught with sim­i­lar issues but have advan­tages in select
pop­u­la­tions/clin­i­cal HES sub­types (Tables 3 and 4).
CLINICAL CASE (Con­t in­u ed)
Over the next 4 years, she remained rel­a­tively sta­ble on inter­
feron α ther­apy with par­tially con­trolled symp­toms, AEC <1.0 to
2.0 × 109/L, but was unable to taper pred­ni­sone below 12.5 mg
daily with­out sig­nif­i­cant wors­en­ing. Her clonal T-cell pop­u­
la­tion increased to approx­i­ma­tely 30% of total lym­pho­cytes,
prompting repeat CT scan and bone mar­
row, which were
unchanged, and pos­i­tron emis­sion tomog­ra­phy (PET)/CT scan,
50 | Hematology 2022 | ASH Education Program
which showed no evi­dence of lym­phoma. She was enrolled on
a phase 2 pla­cebo-con­trolled trial of benralizumab.
Eosinophil-targeted ther­a­pies and HES
The avail­
abil­
ity of bio­
log­
ics targeting IL-5 (mepolizumab and
reslizumab) and its recep­
tor (benralizumab), all­of which are
approved by the US Food and Drug Administration for the
treat­ment of asthma, has pro­foundly altered the approach to
the treat­ment of idi­o­pathic, lym­pho­cytic, and over­lap var­i­ants
of HES. Whereas cor­ti­co­ste­roids are still recommended as ini­
tial ther­apy in most cases, eosin­o­phil-targeting bio­log­ics have
shown excel­lent safety and effi­cacy pro­files in the treat­ment of
HES,21-23 lead­ing to the recent approval of mepolizumab for HES
and eosin­o­philic granulomatosis with polyangiitis and the ini­ti­
a­tion of phase 3 tri­als of benralizumab for the same indi­ca­tions.
Of note, mepolizumab and reslizumab cause mat­u­ra­tional arrest
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/47/2021769/47klion.pdf by guest on 09 December 2022
All patients with con­firmed HE
in the bone mar­row with dra­matic but incom­plete reduc­tion of
blood and tis­sue eosin­o­philia. In con­trast, benralizumab, an afu­
cosylated mono­clo­nal anti­body to IL-5 recep­tor α, tar­gets eosin­
o­phils, baso­phils, and their pre­cur­sors for anti­body-depen­dent
cell-medi­
ated cyto­
tox­
ic­
ity, resulting in com­
plete (or near-­
com­
plete) deple­
tion in all­tis­
sues stud­
ied to date. Although
defin­i­tive data are lacking, what lit­tle infor­ma­tion is avail­­able
concerning the effi­cacy of these agents in mye­loid forms of HES
is dis­cour­ag­ing.22,24,25
Other agents that directly or indi­rectly decrease blood and
tis­sue eosin­o­philia are in devel­op­ment for HES and/or approved
for select eosin­o­philic indi­ca­tions (Table 4). These include liren­
telimab (an afucosylated anti­
body to Siglec-8 that depletes
eosin­
o­
phils and baso­
phils and inhib­
its mast cell acti­
va­
tion),26
dupilumab (a mono­clo­nal anti­body to IL-4 recep­tor α approved
for asthma, atopic der­ma­ti­tis, chronic rhinosinusitis, and eosin­
o­
philic esophagitis that blocks IL-4 and IL-13 sig­
nal­
ing and
eotaxin-medi­ated tis­sue migra­tion of eosin­o­phils),27 and dex­
pramipexole (an orally avail­­able small mol­e­cule that causes mat­
u­ra­tional arrest and eosin­o­phil deple­tion through an unknown
mech­a­nism).28
CLINICAL CASE (Con­t in­u ed)
After an ini­tial response to benralizumab, inter­feron α was dis­
continued. Two weeks later, eosin­
o­
philia and severe symp­
toms returned. Benralizumab was discontinued, and she was
started on pred­ni­sone 60 mg in addi­tion to inter­feron α. She
sub­se­quently devel­oped acute sen­so­ri­neu­ral hear­ing loss
that resolved with ces­sa­tion of inter­feron α and a pred­ni­sone
burst. Cyclosporine was added. Due to per­sis­tent symp­toms
and inabil­ity to taper pred­ni­sone below 20 mg daily, she was
enrolled on a phase 3 pla­cebo-con­trolled trial of mepolizumab.
Her symp­toms improved, and she was ­able to taper ­pred­ni­sone
to 9 mg daily, at which point she devel­oped cough and short­
ness of breath requir­
ing hos­
pi­
tal admis­
sion. Evaluation was
nota­ble for ground-glass infil­trates, dif­fuse lymph­ade­nop­a­
thy, and spleno­
meg­
aly. Lymph node biopsy (approx­
i­
ma­
tely
10 years after her ini­tial diag­no­sis) revealed angioimmunoblastic
T-cell lym­phoma. She was treated with CHOP-R (cyclophospha­
mide, doxorubicin hydrochloride, vincristine sulfate [Oncovin],
prednisone, and rituximab) with tran­sient response. Romidep­
sin was added, but she devel­oped right-sided heart fail­ure and
died of respi­ra­tory fail­ure.
Predictors of response to targeted ther­apy
There is cur­rently lit­tle avail­­able infor­ma­tion to guide the ini­tial
choice of bio­logic for a patient with HES. Although AEC has been
shown to pre­dict response to IL-5/IL-5 recep­tor targeting agents
in patients with asthma, nei­ther serum IL-5 lev­els nor eosin­o­phil
count at ini­ti­a­tion of treat­ment were found to pre­dict response
to mepolizumab in the phase 3 trial in patients with PDGFRAneg­a­tive, cor­ti­co­ste­roid-respon­sive HES,29 and nei­ther his­toric
peak nor base­line AEC predicted response to benralizumab in
a phase 2 trial in patients with PDGFRA-neg­a­tive, treat­mentrefrac­tory HES.22 The only con­sis­tent find­ing across tri­als has
been dif­fer­ences across clin­i­cal HES sub­types, with decreased
response rates and/or increased relapse rates in patients with
lym­pho­cytic var­i­ant HES22,25,30-32 (Figure 2). That said, responses
to the dif­fer­ent bio­log­ics targeting the IL-5 axis are var­i­able, and
a lack of response to one agent does not pre­clude suc­cess with
another.32
Potential risks of eosin­o­phil deple­tion
Over the past 10 to 15 years, it has become increas­ingly appar­
ent that eosin­
o­
phils play an impor­
tant role in homeo­
static
pro­cesses, includ­ing tis­sue remodeling, tumor sur­veil­lance,
Hypereosinophilia and hypereosinophilic syn­dromes | 51
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/47/2021769/47klion.pdf by guest on 09 December 2022
Figure 1. Initial approach to the patient with hypereosinophilic syndrome.
Table 4. Selected ther­a­peu­tic agents for the treat­ment of hypereosinophilic syn­dromes
Selected agents
Target
Comments
Oral pred­ni­sone and equiv­a­lents
Topical CS (ie, swallowed
budesonide)
NA
First-line ther­apy in most types
of HES; although most patients
respond ini­tially, tox­ic­ity and
resis­tance are lim­it­ing long term
May be suf­fi­cient in sin­gle-organ
eosin­o­philic dis­or­ders, such as
eosin­o­philic esophagitis, and
eosin­o­philic der­ma­ti­tis
Cytotoxic
Hydroxyurea
Methotrexate
Cyclophosphamide
Dividing cells
Conventional sec­ond-line ther­apy
for idi­o­pathic HES* and some
MHES; can be used at high dose to
rap­idly lower counts; inex­pen­sive
but sig­nif­i­cant tox­ic­ity
Steroid-spar­ing agent most ­
com­monly used in EGPA and other
rheumatologic over­lap dis­or­ders*
Mostly used for ste­roid-refrac­tory
or life-threat­en­ing EGPA
Immunomodulatory
Interferon α
Cyclosporine
Biologics
Mepolizumab*
Reslizumab
Benralizumab
Dupilumab
Lirentelimab
IL-5
IL-5
IL-5 recep­tor
IL-4 recep­tor
Siglec-8
Approved for the treat­ment of HES
and EGPA at 300mg monthly19,21
Likely to be com­pa­ra­ble to
mepolizumab; approved for
asthma with weight-based dos­ing
In phase 3 tri­als for HES; approved
for asthma
Blocks tis­sue eosin­o­philia but
may cause periph­eral eosin­o­philia
and, rarely, eosin­o­philic com­pli­
ca­tions; approved for chronic
rhinosinusitis, atopic der­ma­ti­tis,
and, most recently, eosin­o­philic
esophagitis
Depletes eosin­o­phils and pre­vents
mast cell degran­u­la­tion; in clin­i­cal
devel­op­ment
Tyrosine kinase inhib­i­tors
Imatinib
Ruxolitinib
Multiple
JAK
Approved for the treat­ment of
HES; near 100% effi­cacy in PDGFRasso­ci­ated mye­loid neo­plasms;
some effi­cacy in other HES with
mye­loid fea­tures17
Approved for sev­eral mye­loid
dis­or­ders typ­i­cally asso­ci­ated
with JAK2 muta­tions; case reports
sug­gest that it may be use­ful in
JAK2-asso­ci­ated HES and LHES
Novel targeted
Dexpramipexole
Causes mat­u­ra­tional arrest at
eosin­o­philic promyelocyte stage
via unknown mech­a­nism
In clin­i­cal devel­op­ment for HES;
phase 2 study prom­is­ing
Conventional sec­ond-line ther­apy
for LHES and some idi­o­pathic HES;
sub­stan­tial tox­ic­ity*
Alternative sec­ond-line agent,
espe­cially in LHES; renal tox­ic­ity is
a sig­nif­i­cant prob­lem*
*Mepolizumab, which is approved for the treat­ment of HES, has bet­ter effi­cacy and lower tox­ic­ity but is expen­sive and not avail­­able in all­countries.
Primary out­come of the phase 3 trial of mepolizumab for PDGFRA-neg­a­tive, ste­roid-respon­sive HES in adults: reduc­tion of dis­ease flares over a
32-week period in patients receiv­ing mepolizumab and sta­ble back­ground ther­apy com­pared with those receiv­ing pla­cebo and sta­ble back­ground
ther­apy.21 Dual pri­mary out­comes of the phase 3 trial of mepolizumab for relaps­ing or treat­ment-refrac­tory EGPA in adults: total accrued weeks of
remis­sion, defined as a Birmingham vas­cu­li­tis score of 0 on less than 4mg pred­ni­sone daily for 52 weeks, and pro­por­tion of par­tic­i­pants in remis­sion
at weeks 36 and 48.23
CS, cor­ti­co­ste­roid; EGPA, eosin­o­philic granulomatosis and polyangiitis; MHES, mye­loid var­i­ant hypereosinophilic syn­drome.
†
52 | Hematology 2022 | ASH Education Program
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/47/2021769/47klion.pdf by guest on 09 December 2022
Class
CS
con­firm the safety of these agents over the long term and to
assess the impact of blocking mul­ti­ple lin­e­ages and/or path­ways
on homeo­static pro­cesses.
Acknowledgment
This work was funded by the Division of Intramural Research,
National Institute of Allergy and Infectious Diseases, National
Institutes of Health.
Conflict-of-inter­est dis­clo­sure
Amy D. Klion: no com­pet­ing finan­cial inter­ests to declare.
Off-label drug use
Figure 2. Clinical subtypes of hypereosinophilic syndrome:
­frequency distribution of 554 patients referred to the National
Institutes of Health for unexplained eosinophilia. EGID, eosino­
philic gastrointestinal disease; EGPA, eosinophilic granulomato­
sis and polyangiitis; EO FASCIITIS, eosinophilic fasciitis.
Correspondence
Amy D. Klion, Human Eosinophil Section, Deputy Chief, Labo­
ratory of Parasitic Diseases, National Institute of Allergy and In­
fectious Diseases, National Institutes of Health, Building 4, Room
B1-27, 4 Memorial Drive, Bethesda, MD 20892; e-mail: aklion@
nih​­.gov.
References
met­a­bolic func­tion, and immu­no­reg­u­la­tion.33 Although this has
led to the­o­ret­i­cal con­cerns about the effects of long-term ther­
apy with bio­log­ics and other agents that sig­nif­i­cantly reduce
eosin­o­phils in blood and tis­sue, data to date sug­gest that eosin­
o­phil deple­tion in humans is safe.25,34-37 Importantly, there have
been no reports of an increased inci­
dence of malig­
nancy or
auto­im­mune dis­ease related to eosin­o­phil deple­tion, and even
sub­tle immu­no­logic con­se­quences dem­on­strated in murine
mod­els, such as impaired vac­cine responses, have not been rep­
li­cated in human stud­ies.22,34,38,39 This lack of sig­nif­i­cant tox­ic­ity is
likely due, at least in part, to the redun­dancy and com­plex­ity of
the human immune sys­tem, which raises poten­tial con­cerns as
the num­ber of targeted ther­a­pies and bio­log­ics increases and
the use of com­bi­na­tion ther­a­pies becomes more com­mon. It is
also impor­tant to note that toxicities may be restricted to spe­
cific pop­u­la­tions or clin­i­cal set­tings (ie, patients exposed to hel­
minth infec­tion, infected with coronavirus dis­ease 2019, or at
high risk of auto­im­mune dis­ease), for which there are lit­tle to no
pro­spec­tive data to date.
Conclusions
Whereas eosin­o­philia is com­mon in the gen­eral pop­u­la­tion, HES
are a het­ero­ge­neous and com­plex group of rare dis­or­ders with
clin­i­cal man­i­fes­ta­tions that span the range of med­i­cal subspe­
cialties. Comprehensive clin­i­cal eval­u­a­tion is nec­es­sary both
to assess end-organ man­
i­
fes­
ta­
tions and deter­
mine the most
likely eti­­ol­ogy and/or clin­i­cal sub­type of HES, as this infor­ma­
tion has impor­tant ther­a­peu­tic and prog­nos­tic impli­ca­tions.
Although cor­
ti­
co­
ste­
roids con­
tinue to be first-line ther­
apy in
most sit­u­a­tions, novel targeted ther­a­pies are rap­idly replacing
con­ven­tional cyto­toxic and broad immu­no­sup­pres­sive agents
as sec­ond-line agents of choice for the treat­ment of eosin­o­philasso­ci­ated clin­i­cal man­i­fes­ta­tions. Despite the lack of safety sig­
nals to date, vig­i­lance and pro­spec­tive stud­ies are needed to
1.
Sunadome H, Sato S, Matsumoto H, et al. Similar dis­tri­bu­tion of periph­
eral blood eosin­o­phil counts in Euro­pean and East Asian pop­u­la­tions from
inves­ti­ga­tions of large-scale gen­eral pop­u­la­tion stud­ies: the Nagahama
Study. Eur Respir J. 2021;57(1):2004101.
2. Crane MM, Chang CM, Kobayashi MG, Weller PF. Incidence of mye­lo­pro­
lif­er­a­tive hypereosinophilic syn­drome in the United States and an esti­
mate of all­hypereosinophilic syn­drome inci­dence. J Allergy Clin Immunol.
2010;126(1):179-181.
3. Ogbogu PU, Bochner BS, Butterfield JH, et al. Hypereosinophilic syn­
drome: a mul­ti­cen­ter, ret­ro­spec­tive anal­y­sis of clin­i­cal char­ac­ter­is­tics and
response to ther­apy. J Allergy Clin Immunol. 2009;124(6):1319-1325.e3.
4. Chusid MJ, Dale DC, West BC, Wolff SM. The hypereosinophilic syndrome:
analysis of fourteen cases with review of the literature. Medicine (Baltimore). 1975;54(1):1-27.
5. Shomali W, Gotlib J. World Health Organization-defined eosin­o­philic dis­or­
ders: 2022 update on diag­no­sis, risk strat­i­fi­ca­tion, and man­age­ment. Am J
Hematol. 2022;97(1):129-148.
6. Valent P, Klion AD, Horny H-P, et al. Contemporary con­sen­sus pro­posal on
cri­te­ria and clas­si­fi­ca­tion of eosin­o­philic dis­or­ders and related syn­dromes.
J Allergy Clin Immunol. 2012;130(3):607-612.e9.
7. Klion AD. How I treat hypereosinophilic syn­dromes. Blood. 2015;126(9):10691077.
8. Cogan E, Schandené L, Crusiaux A, Cochaux P, Velu T, Goldman M. Brief
report: clonal pro­lif­er­a­tion of type 2 helper T cells in a man with the hyper­
eosinophilic syn­drome. N Engl J Med. 1994 Feb 24;330(8):535-538.
9. Simon HU, Plötz SG, Dummer R, Blaser K. Abnormal clones of T cells pro­duc­
ing inter­leu­kin-5 in idi­o­pathic eosin­o­philia. N Engl J Med. 1999;341(15):11121120.
10. Carpentier C, Verbanck S, Schandené L, et al. Eosinophilia asso­ci­ated with
CD3-CD4+ T cells: char­ac­ter­iza­tion and out­come of a sin­gle-cen­ter cohort
of 26 patients. Front Immunol. 2020;11:1765.
11. Lefèvre G, Copin M-C, Staumont-Sallé D, et al. The lym­phoid var­i­ant of
hypereosinophilic syn­drome: study of 21 patients with CD3-CD4+ aber­rant
T-cell phe­no­type. Medicine. 2014;93(17):255-266.
12. Carpentier C, Schandené L, Dewispelaere L, Heimann P, Cogan E, Roufosse
F. CD3-CD4+ lym­pho­cytic var­i­ant hypereosinophilic syn­drome: diag­nos­tic
tools revisited. J Allergy Clin Immunol Pract. 2021;9(6):2426-2439.e72439e7.
13. Ledoult E, Groh M, Kahn J-E, et al; CEREO—French National Reference Cen­
ter for Hypereosinophilic Syndromes. Assessment of T-cell polar­i­za­tion on
the basis of sur­face marker expres­sion: diag­no­sis and poten­tial ther­a­peu­tic
impli­ca­tions in lym­pho­cytic var­i­ant hypereosinophilic syn­drome. J Allergy
Clin Immunol Pract. 2020;8(3):1110-1114.e21114e2.
Hypereosinophilia and hypereosinophilic syn­dromes | 53
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/47/2021769/47klion.pdf by guest on 09 December 2022
Amy D. Klion: all of the drugs discussed, with the exception of
imatinib and mepolizumab, are considered off-label for the treat­
ment of hypereosinophilic syndromes.
54 | Hematology 2022 | ASH Education Program
29. Rothenberg ME, Roufosse F, Faguer S, et al. Mepolizumab reduces hyper­
eosinophilic syn­drome flares irrespective of blood eosin­o­phil count and
IL-5 [published online ahead of print 12 May 2022]. J Allergy Clin Immunol
Pract.
30. Williams AK, Dou C, Chen LYC. Treatment of lym­pho­cyte-var­i­ant hypereo­
sinophilic syn­drome (L-HES): what to con­sider after confirming the elu­sive
diag­no­sis. Br J Haematol. 2021;195(5):669-680.
31. Roufosse F, de Lavareille A, Schandené L, et al. Mepolizumab as a cor­ti­co­
ste­roid-spar­ing agent in lym­pho­cytic var­i­ant hypereosinophilic syn­drome.
J Allergy Clin Immunol. 2010;126(4):828-835.e3.
32. Chen MM, Roufosse F, Wang SA, et al. An inter­na­tional, ret­ro­spec­tive study
of off-label bio­logic use in the treat­ment of hypereosinophilic syn­dromes.
J Allergy Clin Immunol Pract. 2022;10(5):1217-1228.e31228e3.
33. Wen T, Rothenberg ME. The reg­u­la­tory func­tion of eosin­o­phils. Microbiol
Spectr. 2016;4(5):10.1128/microbiolspec.MCHD-0020-2015.
34. Kuang FL, Bochner BS. Lessons learned from targeting eosin­
o­
phils in
human dis­ease. Semin Immunopathol. 2021;43(3):459-475.
35. Khatri S, Moore W, Gib­son PG, et al. Assessment of the long-term safety
of mepolizumab and dura­bil­ity of clin­i­cal response in patients with severe
eosin­o­philic asthma. J Allergy Clin Immunol. 2019;143(5):1742-1751.e71751e7.
36. Korn S, Bourdin A, Chupp G, et al. Integrated safety and effi­cacy among
patients receiv­ing benralizumab for up to 5 years. J Allergy Clin Immunol
Pract. 2021;9(12):4381-4392.e44392e4.
37. Beck LA, Deleuran M, Bissonnette R, et al. Dupilumab pro­vi­des accept­
able safety and sustained effi­cacy for up to 4 years in an open-label study
of adults with mod­er­ate-to-severe atopic der­ma­ti­tis. Am J Clin Dermatol.
2022;23(3):393-408.
38. Zeitlin PL, Leong M, Cole J, et al; ALIZE study inves­ti­ga­tors. Benralizumab
does not impair anti­body response to sea­sonal influ­enza vac­ci­na­tion in
ado­les­cent and young adult patients with mod­er­ate to severe asthma:
results from the phase IIIb ALIZE trial. J Asthma Allergy. 2018;11:181-192.
39. Manetz S, Maric I, Brown T, et al. Successful preg­nancy in the set­ting of
eosin­o­phil deple­tion by benralizumab. J Allergy Clin Immunol Pract.
2021;9(3):1405-1407.e31407e3.
40. Hama N, Abe R, Gib­son A, Phil­lips EJ. Drug-induced hyper­sen­si­tiv­ity syn­
drome (DIHS)/drug reac­
tion with eosin­
o­
philia and sys­
temic symp­
toms
(DRESS): clin­i­cal fea­tures and path­o­gen­e­sis. J Allergy Clin Immunol Pract.
2022;10(5):1155-1167.e51167e5.
41. O’Connell EM, Nutman TB. Eosinophilia in infec­tious dis­eases. Immunol
Allergy Clin North Am. 2015;35(3):493-522.
42. Olbrich P, Ortiz Aljaro P, Freeman AF. Eosinophilia asso­ci­ated with immune
defi­ciency. J Allergy Clin Immunol. 2022;10(5):1140-1153.
DOI 10.1182/hema­tol­ogy.2022000367
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/47/2021769/47klion.pdf by guest on 09 December 2022
14. Moerman-Herzog A, Mehdi SJ, Wong HK. Gene expres­sion com­par­i­son
between Sézary syn­drome and lym­pho­cytic-var­i­ant hypereosinophilic
syn­drome refines bio­mark­ers for Sézary syn­drome. Cells. 2020;9(9):1992.
15. Lefèvre G, Copin M-C, Roumier C, et al; French Eosinophil Network.
CD3-CD4+ lym­
phoid var­
i­
ant of hypereosinophilic syn­
drome: nodal and
extranodal his­to­path­o­log­i­cal and immunophenotypic fea­tures of a periph­
eral indo­lent clonal T-cell lymphoproliferative dis­or­der. Haematologica.
2015;100(8):1086-1095.
16. Roufosse F, de Leval L, van Krieken H, van Deuren M. Lymphocytic var­i­
ant hypereosinophilic syn­drome progressing to angioimmunoblastic T-cell
lym­phoma. Leuk Lymphoma. 2015;56(6):1891-1894.
17. Shi Y, Wang C. What we have learned about lym­pho­cytic var­i­ant hyper­
eosinophilic syn­drome: a sys­tem­atic lit­er­a­ture review. Clin Immunol.
2022;237:108982.
18. Gotlib J. Available and emerging therapies for bona fide advanced sys­
temic mastocytosis and primary eosinophilic neoplasms. Hematology Am
Soc Hematol Educ Program. 2022;2022:34-46.
19. Chen Y-Y, Khoury P, Ware J-M, et al. Marked and per­
sis­
tent eosin­
o­
philia in the absence of clin­i­cal man­i­fes­ta­tions. J Allergy Clin Immunol.
2014;133(4):1195-1202.
20. Dellon ES, Simon D, Wechsler ME. Controversies in allergy: the poten­tial
role of bio­log­ics as first-line ther­apy in eosin­o­philic dis­or­ders. J Allergy Clin
Immunol Pract. 2022;10(5):1169-1176.
21. Roufosse F, Kahn J-E, Rothenberg M-E, et al; HES Mepolizumab Study
Group. Efficacy and safety of mepolizumab in hypereosinophilic syn­drome:
a phase III, ran­dom­ized, pla­cebo-con­trolled trial. J Allergy Clin Immunol.
2020;146(6):1397-1405.
22. Kuang FL, Legrand F, Makiya M, et al. Benralizumab for PDGFRA-neg­a­tive
hypereosinophilic syn­drome. N Engl J Med. 2019;380(14):1336-1346.
23. Wechsler ME, Akuthota P, Jayne D, et al; EGPA Mepolizumab Study Team.
Mepolizumab or pla­cebo for eosin­o­philic granulomatosis with polyangiitis.
N Engl J Med. 2017;376(20):1921-1932.
24. Klion AD, Law MA, Noel P, Kim Y-J, Haverty T-P, Nutman T-B. Safety and
effi­cacy of the mono­clo­nal anti-inter­leu­kin-5 anti­body SCH55700 in the
treat­
ment of patients with hypereosinophilic syn­
drome. Blood. 2004;
103(8):2939-2941.
25. Kuang FL, Fay MP, Ware J, et al. Long-term clin­i­cal out­comes of high-dose
mepolizumab treat­
ment for hypereosinophilic syn­
drome. J Allergy Clin
Immunol Pract. 2018;6(5):1518-1527.e51527e5.
26. Dellon ES, Peterson KA, Murray JA, et al. Anti-Siglec-8 anti­body for eosin­o­
philic gas­tri­tis and duodenitis. N Engl J Med. 2020;383(17):1624-1634.
27. Hirano I, Dellon ES, Hamilton JD, et al. Efficacy of dupilumab in a phase 2
ran­dom­ized trial of adults with active eosin­o­philic esophagitis. Gastroenterology. 2020;158(1):111-122.e10122e10.
28. Panch SR, Bozik ME, Brown T, et al. Dexpramipexole as an oral ste­roidspar­ing agent in hypereosinophilic syn­dromes. Blood. 2018;132(5):501-509.
ANXIETY PROVOKING CONSULTATIONS: MAST CELLS AND EOSINOPHILS
Cem Akin
Division of Allergy and Clinical Immunology, Department of Medicine, University of Michigan, Ann Arbor, MI
Mast cell disorders include mastocytosis and mast cell activation syndromes. Mastocytosis is a rare clonal disorder of
the mast cell, driven by KIT D816V mutation in most cases. Mastocytosis is diagnosed and classified according to World
Health Organization criteria. Mast cell activation syndromes encompass a diverse group of disorders and may have clonal
or nonclonal etiologies. Hematologists may be consulted to assist in the diagnostic workup and/or management of mast
cell disorders. A consult to the hematologist for mast cell disorders may provoke anxiety due to the rare nature of these
diseases and the management of nonhematologic mast cell activation symptoms. This article presents recommendations
on how to approach the diagnosis and management of patients referred for common clinical scenarios.
LEARNING OBJECTIVES
• Review the diagnostic criteria and classification of mastocytosis
• Identify which patients with mast cell activation symptoms and elevated tryptase levels need further workup for
a mast cell disorder
CLINICAL CASE
A 35­year­old man is referred for evaluation of a mast cell
disorder. He was admitted to the hospital for hypoten­
sive syncope after sustaining a wasp sting while biking.
He felt flushed and light­headed and collapsed within
10 minutes of the sting. When emergency medical ser­
vices arrived, his systolic blood pressure was noted to
be 50 mmHg, with a heart rate of 120/min. There were no
hives or angioedema, and the skin exam otherwise was
unremarkable. His past medical history is significant for
occasional flushing and lightheadedness with vigorous
exercise. A serum tryptase level obtained 1 hour after the
episode in the emergency department was 120 ng/mL.
Another tryptase level 1 week later when he is at his base­
line was 28 ng/mL.
Introduction
Consults with a hematologist for a “mast cell disorder” can
provoke anxiety for a number of different reasons. First,
these disorders are rare, and the provider may not have
enough practical experience to direct a diagnostic workup,
choose the right tests, and interpret them correctly. Sec­
ond, treatment guidelines for many disease categories are
evolving. Third, many patients with mast cell disorders
have no abnormalities in other hematologic lineages, and
therefore the hematologist may feel the treatment options
to be out of their range of expertise. Likewise, there may
be questions about when and how to initiate a workup for
a patient referred to rule out a mast cell disorder due to
symptoms of mast cell activation or elevated tryptase.
The hematologist may receive consultations or referrals
for mast cell disorders due to a number of different clinical
scenarios:
1. Patient has an established diagnosis of advanced systemic
mastocytosis (advSM) and is in need of cytoreductive ther­
apy of the mast cell disease and any associated non–mast
cell neoplastic component such as a myelodysplastic syn­
drome (MDS) or myeloproliferative neoplasm (MPN).
2. Patient has skin lesions of maculopapular mastocytosis,
and a bone marrow biopsy is needed to confirm or rule
out systemic mastocytosis and categorize the disease.
3. Patient has a diagnosis of cutaneous or nonadvanced
(indolent) systemic mastocytosis, and recommendations
for management and follow­up are needed.
Dr Prakash Singh Shekhawat
Mast cell disorders | 55
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/55/2021719/55akin.pdf by guest on 09 December 2022
How to evaluate the patient with a suspected
mast cell disorder and how/when to manage
symptoms
4. Patient has an ele­vated tryptase level.
5. Patient has symp­toms of mast cell acti­va­tion, and an opin­ion
is needed regard­ing whether the patient is in need of fur­ther
hema­to­logic workup.
Before discussing each of these sce­nar­ios, a primer on the
diag­no­sis and clas­si­fi­ca­tion of mastocytosis and mast cell acti­va­
tion dis­or­ders would be help­ful.
Mast cell dis­or­ders
Mast cell dis­or­ders can be broadly cat­e­go­rized into those involv­
ing mast cell acti­va­tion and those involv­ing pro­lif­er­a­tion, with a
sig­nif­i­cant over­lap between the 2 categories (Figure 1).1 One can
also encoun­ter the ter­mi­nol­ogy of “clonal” and “nonclonal” mast
cell dis­or­ders in the lit­er­a­ture, the for­mer refer­ring to mastocyto­
sis and the lat­ter refer­ring to patients presenting with ­recur­rent
idi­o­pathic ana­phy­laxis or other mast cell acti­va­tion dis­or­ders
with­out evi­dence of mastocytosis.2 The pro­to­typ­i­cal pro­lif­er­a­tive
mast cell dis­or­der is mastocytosis.3,4 Although patients with rare
myelomastocytic leu­ke­mia or reac­tive mast cell hyper­pla­sia pres­
ent with increased mast cells in bone mar­row biop­sies, these are
not “dis­or­ders of the mast cell” and are not discussed here.5
Diagnosis and clas­si­fi­ca­tion of mastocytosis
Mastocytosis is a clonal dis­
ease of the mast cell pro­
gen­
i­
tor,
most often driven by a somatic gain-of-func­tion muta­tion in KIT,
resulting in the path­o­logic accu­mu­la­tion and acti­va­tion of mast
cells in tis­sues. It can be diag­nosed in chil­dren and adults. The
driver muta­tion is KIT D816V in more than 90% of adults and in
about 30% of chil­dren.6
Pediatric-onset dis­ease
Pediatric-onset mastocytosis usu­ally pres­ents in the first year of
life with typ­i­cal maculopapular skin lesions, also known as urti­caria
pigmentosa. Pediatric mastocytosis gen­er­ally has a self-lim­ited
course, with spon­ta­ne­ous res­o­lu­tion or sig­nif­i­cant regres­sion by
ado­les­cence.7 Systemic involve­ment can be diag­nosed in about
10% of cases. These chil­dren pres­ent with pro­gres­sively increas­
ing tryptase lev­els, liver or spleen enlarge­ment, lymph­ade­nop­
a­thy, or unex­plained abnor­mal­i­ties in the com­plete blood count
with dif­fer­en­tial. The periph­eral blood KIT D816V muta­tion may
56 | Hematology 2022 | ASH Education Program
be pos­i­tive. Children with typ­i­cal self-lim­ited pedi­at­ric mastocy­
tosis have poly­mor­phic skin lesions, while chil­dren with sys­temic
dis­ease have mono­mor­phic, smaller lesions resem­bling adultonset skin lesions. Other rare types of skin involve­
ment in
­chil­dren include mastocytomas and dif­fuse cuta­ne­ous masto­
cytosis (CM).8 A bone mar­row biopsy is usu­ally not nec­es­sary
in chil­dren with typ­i­cal poly­mor­phic skin lesions and a tryptase
level within nor­mal range unless the child pres­ents with increas­
ing tryptase lev­els or one of the red flags men­tioned above, or
the skin lesions fail to improve by ado­les­cence. It is very unusual
for pedi­at­ric patients with mastocytosis to pres­ent with symp­
toms of mast cell acti­va­tion with­out skin lesions (as opposed to
some adult patients). Therefore, in a pedi­at­ric patient referred
for mast cell acti­va­tion symp­toms with­out skin lesions, a bone
mar­row biopsy is usu­ally not nec­es­sary unless there is another
indi­ca­tion. If such a patient has an ele­vated base­line tryptase
(nor­
mal range is con­
sid­
ered <11.5 ng/mL in most com­
mer­
cial
assays), hered­i­tary alpha tryptasemia (HaT) should be con­sid­
ered first (see below for more detailed dis­cus­sion). If HaT is not
found, a bone mar­row biopsy can be con­sid­ered.
Adult-onset dis­ease
Adult-onset mastocytosis can be diag­nosed in the third decade
of life or later.9 Patients may pres­ent with maculopapular CM
lesions, symp­toms of mast cell acti­va­tion such as recur­rent ana­
phy­laxis, signs and symp­toms of a hema­to­logic dis­or­der, or skel­
e­tal abnor­mal­i­ties on imag­ing rais­ing sus­pi­cion for a met­a­static
dis­ease or mye­loma. Adult-onset mastocytosis is almost always
sys­temic, and a bone mar­row biopsy is needed to estab­lish the
diag­no­sis. Systemic dis­ease refers to the pres­ence of neo­plas­tic
mast cells in extracutaneous tis­sue.
The World Health Organization (WHO) diag­nos­tic cri­te­ria for
sys­temic mastocytosis are shown in Table 1.10,11 A major cri­te­rion
plus 1 minor cri­te­ria or 3 minor cri­te­ria are required to estab­lish
the diag­no­sis.
Patients fulfilling only 1 or 2 minor clonality cri­
te­
ria (KIT
muta­tion and/or CD25 expres­sion) with symp­toms of mast cell
acti­va­tion are termed to have mono­clo­nal mast cell acti­va­tion
syn­drome (MMAS).11 MMAS rep­re­sents a low-bur­den clonal mast
cell dis­ease and is man­aged sim­i­larly to sys­temic mastocytosis.
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/55/2021719/55akin.pdf by guest on 09 December 2022
Figure 1. Mast cell disorders can be separated into those involving proliferation and activation. Mastocytosis is a clonal proliferative
disorder of the mast cell and its progenitor. Reactive mast cell hyperplasia can be seen in a number of inflammatory and neoplastic
conditions. Mast cell activation disorders (MCADs) include a broad range of conditions stemming from primary (mastocytosis/mono­
clonal MCAS), secondary (IgE and non–IgE-mediated mast cell activation due to allergic and nonallergic inflammatory and neoplastic
diseases), and idiopathic origins. MCAS is a subgroup of MCADs usually presenting with severe episodic symptoms, which may be
similar to anaphylaxis (see text for explanation). There may be overlap in patients with mastocytosis presenting with recurrent ana­
phylaxis symptoms or in patients with monoclonal MCAS.
Table 1. WHO diag­nos­tic cri­te­ria for sys­temic mastocytosis
Major cri­te­rion
Comments
Multifocal com­pact infil­trates of mast cells (>15 cells per infil­trate) in a
tis­sue biopsy other than skin (such as bone mar­row).
Mast cells are best visu­al­ized by tryptase or CD117 immu­no­his­to­chem­i­cal
stains in biopsy sec­tions. These infil­trates are gen­er­ally found in
perivascular and paratrabecular loca­tions.
Minor cri­te­ria
CD25 is the most spe­cific neo­plas­tic mast cell marker. CD2 may be
absent in some advanced SM cases. CD30 has been recently added as
a marker in the lat­est WHO doc­u­ment. These mark­ers can be assessed
by serial sec­tions by immu­no­his­to­chem­is­try or by flow cytom­e­try.
However, mast cell flow cytom­e­try should be treated as a rare event
anal­y­sis with acqui­si­tion of ide­ally 1 mil­lion or more events. Typical
leu­ke­mia/lym­phoma FC pan­els do not con­tain enough cells to gate on
mast cells.
Morphological abnor­mal­i­ties in mast cells, such as spin­dle-shaped,
elon­gated mast cells with hypogranulation, cyto­plas­mic pro­jec­tion,
and an off-cen­tric or multilobated nucleus.
More than 25% of mast cells in the infil­trate should be mor­pho­log­i­cally
aber­rant. A bone mar­row aspi­rate smear is the best sam­ple to eval­u­ate
for these aber­rant mast cell forms. Mast cells are usu­ally found
embed­ded in or in close prox­im­ity to spic­ules. There is insuf­fi­cient data
on other tis­sue biop­sies to assess mast cell mor­phol­ogy. A detailed
pho­to­graphic guide to these abnor­mal­i­ties is presented in Sperr et al.42
Detection of KIT D816V muta­tion or another gain of func­tion KIT
muta­tion in blood, bone mar­row, or another noncutaneous tis­sue.
Mutation detec­tion should be done by a high-sen­si­tiv­ity test such as
allele spe­cific PCR or drop­let dig­i­tal PCR with a sen­si­tiv­ity to detect
mutated allele fre­quency of <0.1%. NGS pan­els or sequenc­ing-based
assays lack this sen­si­tiv­ity and are often falsely neg­a­tive.
Baseline serum or plasma tryptase level of >20 ng/ml.
Tryptase is a highly spe­cific marker for mast cell bur­den and acti­va­tion.
It should be mea­sured when the patient is at base­line and not after
an ana­phy­lac­tic or mast cell acti­va­tion event, dur­ing which it may be
found ele­vated regard­less of mastocytosis. This cri­te­rion is not valid if
the patient has another mye­loid neo­plasm as tryptase can be found in
smaller quan­ti­ties in mye­loid pro­gen­i­tor cells.
An impor­tant point has to be made about KIT D816V muta­tion
detec­tion: next gen­er­a­tion sequenc­ing (NGS) or sequenc­ingbased muta­tion detec­tion meth­ods are often not ade­quate, and
KIT D816V allele-spe­cific poly­mer­ase chain reac­tion (PCR) or dig­
i­tal drop­let PCR are required, espe­cially if periph­eral blood is
used for screen­ing (see sce­nario 4 below). However, NGS would
be use­ful for eval­u­at­ing for addi­tional muta­tions that one would
typ­i­cally see in advSM, espe­cially hema­to­logic non–mast cell
neo­plasm (SM-AHN), as these muta­tions may have fur­ther prog­
nos­tic sig­nif­i­cance.
Bone mar­row biopsy and aspi­rate is the gold stan­dard to
estab­lish the diag­no­sis by checking for the molec­u­lar and his­
to­
path­
o­
logic mark­
ers below, as the bone mar­
row is almost
uni­formly involved in SM. However, work­ing with a gas­tro­en­ter­
ol­o­gist may often be required in the fur­ther workup of symp­toms
such as abdom­i­nal pain and diar­rhea. In this regard, immu­no­
his­
to­
chem­
i­
cal stains for tryptase, CD117, and CD25 should
be employed on endo­
scopic biop­
sies to deter­
mine whether
abdom­i­nal symp­toms may be due to mast cell infil­tra­tion or
medi­a­tor release or both. Some mast cells in gas­tro­in­tes­ti­nal
(GI) tract biop­sies may be neg­a­tive for tryptase and pos­i­tive for
CD117. It should be noted that an increased num­ber of mast cells
alone in the GI tract are not diag­nos­tic of mastocytosis and that
bands or clus­ters of aber­rant mast cells fulfilling WHO pathol­ogy
cri­te­ria are required to prove GI involve­ment.12
Well-dif­fer­en­ti­ated sys­temic mastocytosis (WDSM) is a rare
his­to­logic sub­type in which mast cells have a mature, round mor­
phol­ogy, do not express aber­rant CD25 or CD2, and u
­ su­ally lack
the typ­i­cal KIT D816V muta­tion, mak­ing it a chal­leng­ing diag­no­
sis.13,14 Patients with WDSM can have a his­tory of pedi­at­ric-onset
cuta­ne­ous dis­ease that may have resolved or been per­sis­tent.
Mast cells in WDSM express CD30, which should not be pres­ent
in nor­mal mast cells.
Classification of mastocytosis
The most recent WHO clas­si­fi­ca­tion of mastocytosis is shown in
Table 2.
CM is the most com­mon cat­e­gory in chil­dren with involve­
ment lim­ited to the skin. Patients with CM most often pres­ent
with poly­mor­phic skin lesions in the first year of life. Other less
com­mon skin man­i­fes­ta­tions of CM include mastocytomas and
dif­fuse CM. These chil­dren do not require a bone mar­row biopsy
unless one of the red flag signs discussed above is pres­ent. In
con­
trast, adult patients with skin lesions almost always have
bone mar­row involve­ment. If an adult patient with skin lesions
does not have a bone mar­row biopsy, that patient should be
clas­si­fied as “mastocytosis in the skin” rather than CM, as the
prob­a­bil­ity of sys­temic dis­ease is high.
Systemic mastocytosis means the pres­
ence of neo­
plas­
tic
mast cells meet­ing WHO diag­nos­tic cri­te­ria in noncutaneous
tis­sue, char­ac­ter­is­ti­cally bone mar­row. Systemic mastocytosis
can be fur­
ther subdivided into nonadvanced SM (bone mar­
row mastocytosis [BMM], indo­lent SM [ISM], and smol­der­ing SM
[SSM]) and advanced SM (SM-AHN, aggres­sive SM [ASM], and
mast cell leu­ke­mia [MCL]) based on his­to­path­o­logic cri­te­ria,
end organ dys­func­tion, and the pres­ence of another asso­ci­ated
Dr Prakash Singh Shekhawat
Mast cell dis­or­ders | 57
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/55/2021719/55akin.pdf by guest on 09 December 2022
Mast cells co-express CD25, CD2, and CD30.
Table 2. WHO clas­si­fi­ca­tion of mastocytosis
Category
Comment
Cutaneous mastocytosis
Most com­mon cat­e­gory in chil­dren. Subtypes include MPCM,
mastocytoma, and dif­fuse CM.
Systemic mastocytosis
Presence of neo­plas­tic mast cell infil­trates in extracutanous tis­sues. Most
com­mon in adults.
The most com­mon cat­e­gory of SM, representing >80% of patients with
SM. Patients in this cat­e­gory have mast cell col­lec­tions meet­ing WHO
cri­te­ria for sys­temic dis­ease in the bone mar­row but do not have an
asso­ci­ated hema­to­logic neo­plasm or advanced dis­ease find­ings or 2 or
more B find­ings of SSM. Patients with ISM have a life expec­tancy that is
com­pa­ra­ble to the gen­eral age-matched pop­u­la­tion with less than 5%
risk of pro­gres­sion to advanced dis­ease.
Bone mar­row mastocytosis (BMM)
This cat­e­gory was added in the most recent WHO doc­u­ment and was
for­merly included in ISM. BMM is diag­nosed in a patient with­out skin
lesions, tryptase level <125, no mast cell infil­trates in extramedullary
tis­sue, and oth­er­wise fits the cri­te­ria for ISM. This cat­e­gory is believed
to have a lower risk of pro­gres­sion than ISM and SSM.
Smoldering sys­temic mastocytosis (SSM)
This is a rare cat­e­gory of nonadvanced SM with higher mast cell bur­den
as evidenced by 2 or more B find­ings: i. tryptase level ≥200 ng/mL or
bone mar­row biopsy infil­tra­tion by mast cells of ≥30% or KIT D816V var­i­
ant allele frac­tion of ≥10% in bone mar­row or periph­eral blood;
ii. spleno­meg­aly with­out hypersplenism and/or hepa­to­meg­aly with­out
liver dys­func­tion and/or lymph­ade­nop­a­thy >2 cm; iii. myeloproliferation
or sub­tle mor­pho­logic abnor­mal­i­ties in mye­loid cells with­out meet­ing
the cri­te­ria for a WHO-clas­si­fied neo­plasm. Patients with SSM may have
a higher rate of pro­gres­sion to advSM, which still remains <10%.
Systemic mastocytosis with asso­ci­ated hema­to­logic non–mast cell
neo­plasm (SM-AHN)
These patients have an AHN (usu­ally an MPN or MDS) meet­ing the WHO
cri­te­ria in addi­tion to SM. Prognosis is deter­mined by the AHN but is
gen­er­ally poor, with a median sur­vival time of about 2 years after
diag­no­sis.
Aggressive sys­temic mastocytosis (ASM)
Patients with ASM have high-level mast cell bur­den and tis­sue ­­
dys­func­tion due to infil­trat­ing mast cells (also known as C find­ings).
One or more C find­ings attrib­ut­­able to MC infil­tra­tion are required for
diag­no­sis: cytopenias (hemo­glo­bin <10 g/dL, plate­lets <100 000, and
neutropenia <1000), liver dys­func­tion with por­tal hyper­ten­sion, ele­vated
liver func­tion tests, asci­tes, mal­ab­sorp­tion, and diar­rhea with exten­sive
GI infil­trates, spleno­meg­aly with hypersplenism, large (≥2 cm) lytic bone
lesions with path­o­log­i­cal bone frac­tures. It should be noted that
oste­o­po­ro­sis and smaller lytic and scle­rotic bone lesions are com­mon
in all­categories of SM and are not con­sid­ered a C find­ing. Patients with
ASM have a reduced life expec­tancy, with <3 years of aver­age sur­vival
after diag­no­sis.
Mast cell leu­ke­mia (MCL)
MCL is diag­nosed when ≥10% mast cells are found in periph­eral
cir­cu­la­tion or ≥20% in bone mar­row aspi­rate smears in a nonspicular
area. It should be noted that 20% infil­tra­tion grade refers to bone
mar­row aspi­rate and not to bone mar­row biopsy. Patients with­out
cir­cu­lat­ing mast cells are referred to as aleukemic MCL. Patients with
typ­i­cal MCL also have C find­ings sim­i­lar to ASM and carry a very poor
prog­no­sis. A chronic form of MCL with­out C-find­ings or cytopenias have
recently been rec­og­nized with more favor­able sur­vival rates.
Mast cell sar­coma
Rare inva­sive solid mast cell tumor with poor prog­no­sis.
non–mast cell clonal dis­ease).11 Patients with advanced SM usu­
ally have other mye­loid muta­tions detected in NGS, in addi­tion
to KIT D816V.15
As men­tioned in Table 2, oste­o­po­ro­sis or scle­rotic lesions
should not be misinterpreted as advanced dis­ease as they are
com­mon find­ings in ISM. To that end, a rou­tine dual-energy x-ray
absorptiometry bone den­sity scan is recommended for each
patient diag­nosed with SM. Furthermore, radio­graphic imag­ing
of areas with bone pain may be con­sid­ered in indi­vid­ual patients.
A bone scan may show increased uptake focally or appear as a
58 | Hematology 2022 | ASH Education Program
superscan, although it is not rou­tinely recommended. We would
rec­om­mend treat­ment of oste­o­po­ro­sis in col­lab­o­ra­tion with an
endo­cri­nol­o­gist. Bisphosphonates or denosumab has been suc­
cess­fully used in these patients.16
Mast cell sar­coma is an exceed­ingly rare sub­type that does not
meet the diag­nos­tic cri­te­ria for SM but is char­ac­ter­ized by a highgrade inva­sive solid MC tumor that car­ries a poor prog­no­sis.17
Categories of sys­temic mastocytosis include ISM (most com­
mon), BMM, SSM, SM-AHN, ASM, and MCL. See Table 2 for impor­
tant com­ments about these categories.
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/55/2021719/55akin.pdf by guest on 09 December 2022
Indolent sys­temic mastocytosis (ISM)
Mast cell acti­va­tion syn­dromes
in this sce­nario is fairly straight­for­ward. If the patient is an adult
with skin lesions, the like­li­hood of sys­temic dis­ease is high, and a
bone mar­row biopsy and aspi­ra­tion is indi­cated to estab­lish the
diag­no­sis and cat­e­go­rize the dis­ease. In pedi­at­ric patients pre­
senting with typ­i­cal CM, a bone mar­row biopsy is gen­er­ally not
indi­cated due to the low risk of sys­temic dis­ease unless the child
has liver or spleen enlarge­ment, has unex­plained per­sis­tent or
pro­gres­sive lymph­ade­nop­a­thy, has periph­eral blood abnor­mal­
i­ties not explain­able by another pro­cess, dem­on­strates con­sis­
tently increas­ing tryptase lev­els in repeated mea­sure­ments, has
a pos­i­tive periph­eral blood KIT D816V muta­tion, or has per­sis­tent
or pro­gres­sive skin dis­ease after ado­les­cence.
3. The patient has a diag­no­sis of cuta­ne­ous or nonadvanced
(indo­lent) sys­temic mastocytosis, and rec­om­men­da­tions for
man­age­ment and fol­low-up are needed. In this sce­nario, patients
do not have a hema­to­logic abnor­mal­ity but pres­ent for man­age­
ment of var­i­ous mast cell acti­va­tion symp­toms, such as flushing,
abdom­i­nal pain, diar­rhea, pep­tic symp­toms, tachy­car­dia, hypo­
ten­sive recur­rent ana­phy­laxis, neurocognitive prob­lems such
as brain fog, mus­cu­lo­skel­e­tal pain, and fatigue. Symptoms vary
greatly from patient to patient; while some have min­i­mal or no
symp­toms, oth­ers may have severe and dis­abling pre­sen­ta­tions.
Symptom bur­den does not cor­re­late with the extent of bone
mar­row infil­tra­tion by mast cells. Up to 50% of adult patients
may expe­ri­ence ana­phy­laxis dur­ing the course of their dis­ease.28
A pre­scrip­tion of mul­ti­ple doses of self-inject­able epi­neph­rine
is there­fore recommended for all­patients diag­nosed with SM.
Approximately 1 of 3 patients have clin­i­cally sig­nif­i­cant oste­o­
po­ro­sis,16 and about 10% may have an asso­ci­ated IgE-medi­ated
hyme­nop­tera (bee or ves­pid) venom allergy, which can be life
threat­en­ing.29 In this sce­nario, hema­tol­o­gists may feel they lack
the expe­ri­ence or exper­tise to man­age these symp­toms, and it
is rea­son­able to refer these patients to an aller­gist with spe­cial
exper­tise in treating mast cell acti­va­tion symp­toms.
Common man­age­ment strat­e­gies include the avoid­ance of
trig­gers known to cause symp­toms for each spe­cific patient and
the use of anti–mast cell medi­a­tor–targeting ther­a­pies such as
H1 and H2 anti­his­ta­mines, antileukotriene drugs, and mast cell
sta­bi­liz­ers such as cromolyn.30,31 The anti-IgE mono­clo­nal anti­
body omalizumab has been shown to reduce recur­rent ana­phy­
lac­tic symp­toms in patients who do not respond to first-line
antimediator ther­a­pies.32
Patients with SM may be more sus­cep­ti­ble to perioperative
MC acti­va­tion events and ana­phy­laxis, although most patients
undergo suc­
cess­
ful sur­
ger­
ies with premedication and the
selec­tion of agents less likely to cause mast cell acti­va­tion and
the avoid­ance of those known to cause symp­toms in indi­vid­ual
patients.33 Multidisciplinary man­age­ment involv­ing sur­geons,
aller­gists, and anes­the­si­­ol­o­gists is cru­cial to mit­i­gate MC acti­
va­tion in patients need­ing sur­gery. Pregnancy is gen­er­ally well
tol­er­ated in SM, and the man­age­ment of SM in preg­nancy is
discussed else­where.34
KIT D816V–targeting TKIs are cur­
rently approved for the
treat­ment of advanced SM and are in the clin­i­cal trial stage for
patients with nonadvanced SM who do not respond to opti­mized
antimediator man­age­ment.35 Hematological exper­tise may be
required for the treat­ment and mon­i­tor­ing of these patients with
TKIs in the future if they are approved for nonadvanced SM indi­
ca­tions. Avapritinib, cur­rently approved for advSM, was eval­u­
ated in a mul­ti­cen­ter pla­cebo-con­trolled clin­i­cal trial. According
Dr Prakash Singh Shekhawat
Mast cell dis­or­ders | 59
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/55/2021719/55akin.pdf by guest on 09 December 2022
Mast cell acti­va­tion syn­dromes (MCASs) rep­re­sent a het­ero­ge­
neous group of dis­or­ders char­ac­ter­ized by (1) the epi­sodic pres­
ence of mast cell acti­va­tion symp­toms in more than 2 organ
sys­tems, such as cuta­ne­ous, car­dio­vas­cu­lar, GI, pul­mo­nary, and
naso-ocu­lar; (2) response of symp­toms to mast cell medi­a­tor–
targeting drugs; and (3) the detec­tion of a val­i­dated marker of
mast cell acti­va­tion dur­ing the symp­tom­atic phase.2,18 The best
val­i­dated sur­ro­gate marker of mast cell acti­va­tion is tryptase.19
Tryptase should be checked at base­line and within 4 hours of a
suspected mast cell acti­va­tion event. A for­mula of 20% of base­line
plus 2ng/mL is used to cal­cu­late the min­i­mal increase required to
diag­nose mast cell acti­va­tion.20 Urinary metab­ol­ites of other mast
cell medi­a­tors such as his­ta­mine, pros­ta­glan­din D2, and leu­ko­tri­
ene C4 mea­sured in a 24-hour or spot col­lec­tion can also be used
to doc­u­ment mast cell acti­va­tion if tryptase lev­els are not avail­­
able; how­ever, the sen­si­tiv­ity and spec­i­fic­ity of these mark­ers as
well as the min­i­mal increases and cut­off lev­els diag­nos­tic for mast
cell acti­va­tion have not been established.21 A patient with recur­
rent ana­phy­laxis is a pro­to­typ­i­cal pre­sen­ta­tion of MCAS; how­ever,
less severe man­i­fes­ta­tions can meet the above cri­te­ria.
MCAS can be pri­mary or clonal when it is asso­ci­ated with
mastocytosis or MMAS, sec­ond­ary when asso­ci­ated with immu­
no­glob­u­lin (Ig) E or non–IgE-medi­ated mast cell acti­va­tion trig­
gers, or idi­o­pathic when no under­ly­ing cause has been found.
There is quite an exten­sive con­tro­versy surrounding the diag­
no­sis of MCAS, and alter­na­tive diag­nos­tic schemes with broader
inclu­sion cri­te­ria may result in diag­nos­ing nearly 17% of the gen­eral
pop­u­la­tion with MCAS.22 However, this can become prob­lem­atic
for patients whose symp­toms may be caused by another entity
that may sim­ply be asso­ci­ated with reac­tive mast cell acti­va­tion
because focus­ing on mast cell acti­va­tion may leave the under­ly­ing
entity undi­ag­nosed. Contributing to this con­tro­versy some­what
is the International Classification of Diseases ICD-10-CM clas­si­fi­
ca­tion of diag­nos­tic codes that cur­rently includes addi­tional cat­
egories of mast cell acti­va­tion, includ­ing “Mast cell acti­va­tion, not
oth­er­wise spec­i­fied,” which until recently did not have diag­nos­tic
guid­ance.23 A detailed dis­cus­sion of mast cell acti­va­tion dis­or­ders
is out­side the scope of this arti­cle, as hema­tol­o­gists are gen­er­ally
not expected to diag­nose and treat these dis­or­ders.
With this back­ground infor­ma­tion in mind, let us exam­ine
spe­cific con­sul­ta­tion sce­nar­ios men­tioned at the begin­ning of
the text.
1. Patients with advSM. This group of patients rep­re­sents a
clas­si­cal indi­ca­tion for refer­ral to a hema­tol­o­gist. Patients with
advanced SM are in need of cytoreductive ther­apy and treat­
ment of non–mast cell hema­to­logic neo­pla­sias such as MPNs and
MDSs.4,24 The recent approval of KIT D816V–targeting tyro­sine
kinase inhib­i­tors (TKIs) rev­o­lu­tion­ized the treat­ment of patients
with advSM, who typ­i­cally have a reduced life expec­tancy.25-27
Since most of these patients pres­ent with muta­tions in addi­tion
to KIT D816V, addi­tional ther­a­pies in those who do not respond
or lose their response to TKIs may be needed, and targeted or
pal­li­a­tive treat­ment options for the asso­ci­ated AHN regard­less of
mastocytosis should be con­sid­ered. These options are discussed
in a sep­a­rate arti­cle by Gotlib in this edu­ca­tional pro­gram.43
2. The patient has skin lesions of maculopapular mastocytosis,
and a bone mar­row biopsy is needed to con­firm or rule out sys­
temic mastocytosis and cat­e­go­rize the dis­ease. This is another
com­mon sce­nario for refer­ral to the hema­tol­o­gist. The approach
ti­tion, skin rashes, and venom aller­gies.37 Subsequent stud­ies,
how­ever, failed to con­sis­tently con­firm many of the non­al­ler­gic
phe­no­types.38 HaT alone is not a mast cell acti­va­tion dis­or­der;
how­
ever, some reports indi­
cate that HaT may be a dis­
easemod­i­fy­ing ­fac­tor in those with con­cur­rent hyme­nop­tera venom
allergy, idi­o­pathic ana­phy­laxis and mastocytosis account­ing
for more severe mast cell acti­va­tion symp­toms in HaT car­ri­ers,
although pro­spec­tive stud­ies with no refer­ral bias are needed
to con­firm these asso­ci­a­tions.39 Interestingly, HaT is 2 to 3 times
more prev­a­lent in patients with sys­temic mastocytosis than in
the gen­eral pop­u­la­tion.37 Patients with SM and con­cur­rent HaT
tend to pres­ent more often with mast cell acti­va­tion symp­toms
rather than skin lesions or hema­to­logic abnor­mal­i­ties.40 Whether
this asso­ci­a­tion is due to a mech­a­nis­tic link between HaT and
SM or selec­tion bias needs to be inves­ti­gated in future stud­ies.
There is no spe­cific treat­ment avail­­able (or needed) for HaT,
other than the treat­ment of under­ly­ing mastocytosis symp­toms
or man­age­ment of ana­phy­laxis.
Testing for HaT (TPSAB1 gene copy num­ber) is cur­rently com­
mer­cially avail­­able in the US. In this test, copy num­bers of alpha
and beta alleles are reported, and their sum should be 4 in most
indi­vid­u­als. Possible com­bi­na­tions include 4 beta and 0 alpha,
3 beta and 1 alpha, and 2 beta and 2 alpha. A quick inter­pre­ta­
tion would be to con­sider any num­ber above 4 to be an extra
alpha allele, pro­vided there are at least 2 alpha alleles. For exam­
ple, in a patient with 3 beta and 2 alpha alleles, there is 1 extra
alpha allele. However, as stated above, a pos­i­tive test for HaT
does not rule out con­cur­rent mastocytosis. Patients with HaT
and mastocytosis tend to have higher tryptase lev­els than what
might be expected for HaT alone. A cor­rec­tion of tryptase level
according to HaT geno­type has been pro­posed as tryptase level
divided by the extra alpha allele num­ber plus 1. For exam­ple, if
the serum tryptase is 15 ng/mL and the patient has an extra copy
Figure 2. Hereditary alpha tryptasemia.
60 | Hematology 2022 | ASH Education Program
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/55/2021719/55akin.pdf by guest on 09 December 2022
to part 1 data avail­­able from the PIONEER study (clinicaltrials​­.gov,
NCT03731260), a dose of 25 mg/d was selected to move for­ward
in an expanded part 2 cohort based on safety and effi­cacy data
show­ing an approx­i­ma­tely 30% reduc­tion in patient-reported
symp­
tom scores and no seri­
ous adverse events.36 Avapritinib
has also been shown to be highly effec­tive in improv­ing skin
lesions in SM, and these improve­ments were asso­ci­ated with
objec­tive reduc­tions in mast cell dis­ease bur­den. The PIONEER
trial is cur­rently closed to new patient enroll­ment. Other clin­i­cal
tri­als cur­rently open for enroll­ment to eval­u­ate D816V-­­selec­tive
KIT inhib­i­tors in patients with ISM involve BLU-263 (HARBOR,
NCT04910685) and bezuclastinib (SUMMIT, NCT05186753).
4. The patient has an ele­vated tryptase level. In this sce­nario
a patient with or with­out mast cell acti­va­tion symp­toms and no
skin lesions is referred due to an ele­vated tryptase level to rule
out mastocytosis. The most com­mon cause of an ele­vated trypt­
ase level in the gen­eral pop­u­la­tion is HaT, which is seen in up
to 7% of the pop­u­la­tion in the US (Figure 2).37 HaT is an auto­
so­mal dom­i­nantly trans­mit­ted genetic poly­mor­phism of uncer­
tain clin­i­cal sig­nif­i­cance due to copy num­ber var­i­a­tions of the
TPSAB1 gene encoding the alpha tryptase gene. Alpha tryptase,
a pro­en­zyme with no pro­teo­lytic activ­ity, accounts for the bulk
of mea­sur­able serum tryptase in base­line con­di­tions. A median
nor­mal tryptase level is around 4.5 to 5 ng/mL. Patients with
HaT are gen­er­ally mea­sured to have tryptase lev­els higher than
8 ng/mL. The ele­va­tion in tryptase level is pro­por­tion­ate to the
num­ber of alpha-encoding TPSAB1 genes. Due to the auto­so­maldom­i­nant mode of trans­mis­sion, patients with HaT are expected
to have at least 1 par­ent with an ele­vated tryptase level and
have a 50% chance of pass­ing it on to their chil­dren. The ini­tial
descrip­tion of HaT asso­ci­ated this genetic event with a num­ber
of seem­ingly unre­lated phe­no­typic fea­tures, such as irri­ta­ble
bowel ­syn­drome, skel­e­tal abnor­mal­i­ties, retained pri­mary den­
The clin­i­cal case presented at the begin­ning of this arti­cle
is an exam­ple of a patient with an ele­vated base­line tryptase
and severe hypo­ten­sive ana­phy­laxis. Even though he is fairly
asymp­tom­atic other than the ana­phy­laxis epi­sode, his ele­vated
base­line tryptase and high REMA score indi­cate the need for
a bone mar­row biopsy. While one can also check HaT sta­tus, a
pos­i­tive HaT test does not nec­es­sar­ily rule out mastocytosis in
this case. The patient should also be referred to an aller­gist for
venom allergy test­ing and con­sid­er­ation for venom immu­no­
ther­apy and should be pre­scribed self-inject­able epi­neph­rine
for as-needed use.
5. The patient has symp­
toms of mast cell acti­
va­
tion with
a nor­
mal base­
line tryptase level, and an opin­
ion is needed
regard­
ing whether the patient is in need of fur­
ther hema­
to­
logic workup. This sce­nario is prob­a­bly the most con­fus­ing for
the hema­tol­o­gist, as most of these patients pres­ent with­out an
obvi­ous hema­to­logic abnor­mal­ity, and hema­tol­o­gists may feel
they lack suf­fi­cient exper­tise to eval­u­ate and treat mast cell acti­
va­tion or ana­phy­laxis. While a nor­mal tryptase level does not
­nec­es­sar­ily rule out sys­temic mastocytosis, SM is exceed­ingly
rare in patients with a base­line tryptase lower than 4 ng/mL.
Further guid­ance can be obtained by assessing the patient for
REMA cri­te­ria (Figure 3).41 These cri­te­ria were devel­oped by the
Span­
ish Network of Mastocytosis and have a high pre­
dic­
tive
value for deter­min­ing under­ly­ing mastocytosis in patients pre­
senting with ana­phy­lac­tic symp­toms. Patients with hypo­ten­sive
Figure 3. REMA score to predict clonal mast cell disease (mastocytosis) in patients presenting with mast cell activation symptoms
and/or anaphylaxis. Reproduced with permission from Alvarez-Twose et al.41
Dr Prakash Singh Shekhawat
Mast cell dis­or­ders | 61
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/55/2021719/55akin.pdf by guest on 09 December 2022
of the alpha tryptase allele, the corrected tryptase level would
be 7.5 ng/mL.
Our approach to a patient with an ele­vated tryptase level
higher than 8 ng/mL with­out skin lesions of mastocytosis or
any other rea­son to explain ele­vated tryptase (such as mye­
loid neo­pla­sia, chronic renal fail­ure, etc) is as fol­lows: If the
patient has a his­tory of severe hypo­ten­sive ana­phy­laxis and a
score of 2 points or greater in REMA score (see below),41 or if
the corrected tryptase is greater than 8 ng/mL, we con­sider
a bone mar­row biopsy and aspi­ra­tion to rule out mastocyto­
sis. Some experts also pro­pose a high-sen­si­tiv­ity periph­eral
blood KIT D816V test in patients reluc­tant to have a bone mar­
row biopsy; how­ever, a neg­a­tive result in this test would not
nec­es­sar­ily rule out mastocytosis. If the patient, how­ever, is
asymp­tom­atic or has non­spe­cific symp­toms such as fatigue,
mus­cu­lo­skel­e­tal pain, abdom­i­nal pain, or mul­ti­ple food intol­
er­ances, HaT test­ing is done. If this test shows pos­i­tive results
and the corrected tryptase is lower than 8 ng/mL, a bone mar­
row biopsy is optional, and the patient can be followed yearly
with symp­tom­atic assess­ment and a com­plete blood count
with dif­fer­en­tial and tryptase lev­els. If the patient shows an
increas­ing trend in tryptase lev­els or devel­ops new symp­toms
such as hypo­ten­sive ana­phy­laxis, a bone mar­row biopsy is
con­sid­ered. If HaT test­ing is neg­a­tive in a patient with an ele­
vated tryptase level, a bone mar­row biopsy is recommended
regard­less of symp­tom­atic sta­tus.
Conclusions
Mast cell dis­or­ders encom­pass mastocytosis and MCASs. Due to
their rar­ity and var­i­ous forms of pre­sen­ta­tion with­out a hema­to­
logic dis­ease, hema­tol­o­gists may feel they lack suf­fi­cient exper­
tise to diag­nose and treat these dis­or­ders. However, using the
approach outlined above for dif­fer­ent sce­nar­ios, most of these
patients can be triaged to an appro­pri­ate diag­nos­tic workup. A
hematopathologist with expe­ri­ence in reviewing bone mar­row
and other tis­sue biopsy sam­ples as well as an aller­gist knowl­
edge­able about the symp­tom­atic man­age­ment of these patients
can be help­ful resources dur­ing this pro­cess. A refer­ral to a cen­
ter with more exper­tise can be con­sid­ered for man­age­ment rec­
om­men­da­tions and eval­u­a­tion for clin­i­cal tri­als. A list of these
aca­demic cen­ters is avail­­able through the Amer­i­can Initiative
on Mastocytosis (www​­.aimcd​­.net). Finally, the patient sup­port
group the Mast Cell Disease Society (www​­.tmsforacure​­.org) has
many use­ful infor­ma­tional mate­ri­als for patients diag­nosed with
a mast cell dis­or­der.
Conflict-of-inter­est dis­clo­sure
Cem Akin: con­sul­tancy: Blueprint Medicines, Cogent; research
funding: Blueprint Medicines, Cogent.
Off-label drug use
Cem Akin: omalizumab is discussed. In addition, avapritinib,
BLU-263, and bezuclastinib are in clinical trials for ISM.
Correspondence
Cem Akin, Division of Allergy and Clinical Immunology, Depart­
ment of Medicine, University of Michigan, 24 Frank Lloyd Wright Dr,
Lobby H-2100, Ann Arbor, MI 48106; e-mail: cemakin@umich​­.edu.
References
1.
Theoharides TC, Valent P, Akin C. Mast cells, mastocytosis, and related dis­
or­ders. N Engl J Med. 2015;373(2):163-172.
2. Valent P, Akin C, Arock M, et al. Definitions, cri­te­ria and global clas­si­fi­ca­
tion of mast cell dis­or­ders with spe­cial ref­er­ence to mast cell acti­va­tion
syn­dromes: a con­sen­sus pro­posal. Int Arch Allergy Immunol. 2012;157(3):
215-225.
3. Valent P, Akin C, Gleixner KV, et al. Multidisciplinary chal­lenges in mastocy­
tosis and how to address with per­son­al­ized med­i­cine approaches. Int J Mol
Sci. 2019;20(12):2976.
62 | Hematology 2022 | ASH Education Program
4. Pardanani A. Systemic mastocytosis in adults: 2021 update on diag­no­sis,
risk strat­i­fi­ca­tion and man­age­ment. Am J Hematol. 2021;96(4):508-525.
5. Horny HP, Sotlar K, Reiter A, Valent P. Myelomastocytic leu­ke­mia: his­to­
path­o­log­i­cal fea­tures, diag­nos­tic cri­te­ria and dif­fer­en­tial diag­no­sis. Expert
Rev Hematol. 2014;7(4):431-437.
6. Hoermann G, Sotlar K, Jawhar M, et al. Standards of genetic test­ing in the
diag­no­sis and prog­nos­ti­ca­tion of sys­temic mastocytosis in 2022: rec­om­
men­da­tions of the EU-US Cooperative Group. J Allergy Clin Immunol Pract.
2022;10(8):1953-1963.
7. Lange M, Hartmann K, Carter MC, et al. Molecular back­ground, clin­i­cal fea­
tures and man­age­ment of pedi­at­ric mastocytosis: sta­tus 2021. Int J Mol Sci.
2021;22(5):2586.
8. Hartmann K, Escribano L, Grattan C, et al. Cutaneous man­i­fes­ta­tions in
patients with mastocytosis: con­sen­sus report of the Euro­pean Compe­
tence Network on Mastocytosis; the Amer­i­can Academy of Allergy, Asthma
and Immunology; and the Euro­pean Academy of Allergology and Clinical
Immunology. J Allergy Clin Immunol. 2016;137(1):35-45.
9. Brockow K. Epidemiology, prog­
no­
sis, and risk fac­
tors in mastocytosis.
Immunol Allergy Clin N Am. 2014;34(2):283-295.
10. Valent P, Horny HP, Escribano L, et al. Diagnostic cri­te­ria and clas­si­fi­ca­tion
of mastocytosis: a con­sen­sus pro­posal. Leuk Res. 2001;25(7):603-625.
11. Valent P, Akin C, Hartmann K, et al. Updated diag­nos­tic cri­te­ria and clas­
si­fi­ca­tion of mast cell dis­or­ders: a con­sen­sus pro­posal. Hemasphere.
2021;5(11):e646.
12. Doyle LA, Sepehr GJ, Hamilton MJ, Akin C, Castells MC, Hornick JL. A clin­
i­co­path­o­logic study of 24 cases of sys­temic mastocytosis involv­ing the
gas­tro­in­tes­ti­nal tract and assess­ment of muco­sal mast cell den­sity in irri­
ta­
ble bowel syn­
drome and asymp­
tom­
atic patients. Am J Surg Pathol.
2014;38(6):832-843.
13. Akin C, Fumo G, Yavuz AS, Lipsky PE, Neckers L, Metcalfe DD. A novel form
of mastocytosis asso­
ci­
ated with a trans­
mem­
brane c-kit muta­
tion and
response to imatinib. Blood. 2004;103(8):3222-3225.
14. Álvarez-Twose I, Jara-Acevedo M, Morgado JM, et al. Clinical, immuno­
phenotypic, and molec­u­lar char­ac­ter­is­tics of well-dif­fer­en­ti­ated sys­temic
mastocytosis. J Allergy Clin Immunol. 2016;137(1):168-178.e1178e1.
15. Schwaab J, Schnittger S, Sotlar K, et al. Comprehensive muta­tional pro­fil­
ing in advanced sys­temic mastocytosis. Blood. 2013;122(14):2460-2466.
16. Greene LW, Asadipooya K, Corradi PF, Akin C. Endocrine man­i­fes­ta­tions of
sys­temic mastocytosis in bone. Rev Endocr Metab Disord. 2016;17(3):419431.
17. Ryan RJ, Akin C, Castells M, et al. Mast cell sar­coma: a rare and poten­tially
under-rec­og­nized diag­nos­tic entity with spe­cific ther­a­peu­tic impli­ca­tions.
Mod Pathol. 2013;26(4):533-543.
18. Akin C, Valent P, Metcalfe DD. Mast cell acti­va­tion syn­drome: pro­posed
diag­nos­tic cri­te­ria. J Allergy Clin Immunol. 2010;126(6):1099-1104.e4.
19. Schwartz LB. Diagnostic value of tryptase in ana­phy­laxis and mastocytosis.
Immunol Allergy Clin North Am. 2006;26(3):451-463.
20. Valent P, Bonadonna P, Hartmann K, et al. Why the 20% +2 tryptase for­mula
is a diag­nos­tic gold stan­dard for severe sys­temic mast cell acti­va­tion and
mast cell acti­va­tion syn­drome. Int Arch Allergy Immunol. 2019;180(1):44-51.
21. Butterfield JH. Nontryptase uri­nary and hema­to­logic bio­mark­ers of mast
cell expan­sion and mast cell acti­va­tion: sta­tus 2022. J Allergy Clin Immunol
Pract. 2022;10(8):1974-1984.
22. Afrin LB, Ackerley MB, Bluestein LS, et al. Diagnosis of mast cell acti­va­tion
syn­drome: a global “con­sen­sus-2.” Diagnosis (Berl). 2021;8(2):137-152.
23. Valent P, Hartmann K, Bonadonna P, et al. Global clas­si­fi­ca­tion of mast cell
acti­va­tion dis­or­ders: an ICD-10-CM-Adjusted Proposal of the ECNM-AIM
Consortium. J Allergy Clin Immunol Pract. 2022;10(8):1941-1950.
24. Reiter A, George TI, Gotlib J. New devel­
op­
ments in diag­
no­
sis, prog­
nos­
ti­
ca­
tion, and treat­
ment of advanced sys­
temic mastocytosis. Blood.
2020;135(16):1365-1376.
25. Gotlib J, Kluin-Nelemans HC, George TI, et al. Efficacy and safety of mido­
staurin in advanced sys­temic mastocytosis. N Engl J Med. 2016;374(26):25302541.
26. DeAngelo DJ, Radia DH, George TI, et al. Safety and effi­cacy of avapritinib
in advanced sys­temic mastocytosis: the phase 1 EXPLORER trial. Nat Med.
2021;27(12):2183-2191.
27. Gotlib J, Reiter A, Radia DH, et al. Efficacy and safety of avapritinib in
advanced sys­temic mastocytosis: interim anal­y­sis of the phase 2 PATH­
FINDER trial. Nat Med. 2021;27(12):2192-2199.
28. Brockow K, Jofer C, Behrendt H, Ring J. Anaphylaxis in patients with masto­
cytosis: a study on his­tory, clin­i­cal fea­tures and risk fac­tors in 120 patients.
Allergy. 2008;63(2):226-232.
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/55/2021719/55akin.pdf by guest on 09 December 2022
ana­phy­laxis epi­sodes (par­tic­u­larly after bee/wasp stings or idi­o­
pathic) with­out urti­caria or angioedema are at par­tic­u­lar risk. It
should be noted that while patients with sev­eral unex­plain­able
symp­toms such as chronic fatigue; fibromyalgia; head­aches; irri­
ta­ble bowel syn­drome-like symp­toms; dysautonomia, includ­ing
pos­tural ortho­static tachy­car­dia syn­drome; Ehlers-Danlos syn­
drome; and mul­ti­ple food, drug, and envi­ron­men­tal intol­er­ances
are con­sid­ered for a diag­no­sis of mast cell dis­or­der, there are
no mech­a­nis­tic stud­ies prov­ing a causal link between a pri­mary
mast cell dis­or­der, such as mastocytosis or MCAS, and these enti­
ties, and such patients, espe­cially those with nor­mal tryptase
lev­els, are not can­di­dates for a hema­to­logic workup. Some of
these patients may have other local­ized man­i­fes­ta­tions of mast
cell acti­va­tion com­mon to the gen­eral pop­u­la­tion, such aller­gic
rhi­ni­tis, urti­caria, and even flushing, and these patients could be
eval­u­ated for MCAS with the help of an aller­gist for these spe­cific
enti­ties. Patients with a his­tory of ana­phy­laxis should be eval­u­
ated and man­aged by an aller­gist.
38. Chollet MB, Akin C. Hereditary alpha tryptasemia is not asso­ci­ated with
spe­cific clin­i­cal phe­no­types. J Allergy Clin Immunol. 2022;149(2):728-735.e2.
39. Lyons JJ, Chovanec J, O’Connell MP, et al. Heritable risk for severe ana­phy­
laxis asso­ci­ated with increased α-tryptase-encoding germline copy num­
ber at TPSAB1. J Allergy Clin Immunol. 2021;147(2):622-632.
40. Greiner G, Sprinzl B, Górska A, et al. Hereditary α tryptasemia is a valid
genetic bio­marker for severe medi­a­tor-related symp­toms in mastocytosis.
Blood. 2021;137(2):238-247.
41. Alvarez-Twose I, González de Olano D, Sánchez-Muñoz L, et al. Clinical, bio­
log­i­cal, and molec­u­lar char­ac­ter­is­tics of clonal mast cell dis­or­ders present­
ing with sys­temic mast cell acti­va­tion symp­toms. J Allergy Clin Immunol.
2010;125(6):1269-1278.e2.
42. Sperr WR, Escribano L, Jordan JH, et al. Morphologic prop­er­ties of neo­
plas­tic mast cells: delin­ea­tion of stages of mat­u­ra­tion and impli­ca­tion for
cyto­log­i­cal grad­ing of mastocytosis. Leuk Res. 2001;25(7):529-536.
43. Gotlib J. Available and emerging therapies for bona fide advanced systemic
mastocytosis and primary eosinophilic neoplasms. Hematology Am Soc
Hematol Educ Program. 2022;2022:34-46.
© 2022 by The Amer­i­can Society of Hematology
DOI 10.1182/hema­tol­ogy.2022000366
Dr Prakash Singh Shekhawat
Mast cell dis­or­ders | 63
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/55/2021719/55akin.pdf by guest on 09 December 2022
29. Bonadonna P, Scaffidi L. Hymenoptera ana­phy­laxis as a clonal mast cell
dis­or­der. Immunol Allergy Clin North Am. 2018;38(3):455-468.
30. Castells M, Butterfield J. Mast cell acti­va­tion syn­drome and mastocytosis:
ini­tial treat­ment options and long-term man­age­ment. J Allergy Clin Immunol Pract. 2019;7(4):1097-1106.
31. Gotlib J, Gerds AT, Bose P, et al. Systemic Mastocytosis, Version 2.2019,
NCCN Clinical Practice Guidelines in Oncology. J Natl Compr Canc Netw.
2018;16(12):1500-1537.
32. Jendoubi F, Gaudenzio N, Gallini A, Negretto M, Paul C, Bulai Livideanu C.
Omalizumab in the treat­ment of adult patients with mastocytosis: a sys­
tem­atic review. Clin Exp Allergy. 2020;50(6):654-661.
33. Dewachter P, Castells MC, Hepner DL, Mouton-Faivre C. Perioperative man­
age­ment of patients with mastocytosis. Anesthesiology. 2014;120(3):753759.
34. Lei D, Akin C, Kovalszki A. Management of mastocytosis in preg­nancy: a
review. J Allergy Clin Immunol Pract. 2017;5(5):1217-1223.
35. Akin C, Arock M, Valent P. Tyrosine kinase inhib­i­tors for the treat­ment of
indo­lent sys­temic mastocytosis: are we there yet? J Allergy Clin Immunol.
2022;149(6):1912-1918.
36. Akin C, Sabato V, Gotlib J, et al. PIONEER: a ran­dom­ized, dou­ble-blind,
pla­cebo-con­trolled, phase 2 study of avapritinib in patients with indo­lent
or smol­der­ing sys­temic mastocytosis (SM) with symp­toms inad­e­quately
con­trolled by stan­dard ther­apy. J Allergy Clin Immunol. 2020;145(2):AB336.
37. Lyons JJ, Yu X, Hughes JD, et al. Elevated basal serum tryptase identifies a
mul­ti­sys­tem dis­or­der asso­ci­ated with increased TPSAB1 copy num­ber. Nat
Genet. 2016;48(12):1564-1569.
ARE ALTERNATIVE DONORS NOW MAINSTREAM IN ALLOGENEIC TRANSPLANT ?
In 2022, which is preferred: haploidentical
or cord transplant?
Division of Hematology, Bone Marrow Transplantation, and Hemato-Oncology Center, Chaim Sheba Medical Center, Tel HaShomer, Ramat Gan,
Israel; 2Sackler Faculty of Medicine, Tel Aviv University, Tel Aviv, Israel; 3Acute Leukemia Working Party, European Society for Blood and Marrow
Transplantation, Paris, France; 4Department of Clinical Hematology and Cellular Therapy, Saint-Antoine Hospital, Assistance Publique-Hȏpitaux de
Paris, Sorbonne University, Paris, France; and 5Sorbonne University, INSERM, Saint-Antoine Research Center, Paris, France
1
Allogeneic hematopoietic stem cell transplantation is the treatment of choice for high-risk hematological malignancies
such as acute myeloid and lymphocytic leukemia, myelodysplastic syndrome, and myeloproliferative disorders. Alternative donor transplantation from either haploidentical (haplo-SCT) or cord blood donor (CBT) is an established therapeutic
alternative for patients who need transplants but lack a human leukocyte antigen–matched donor. Although haplo-SCT
(mainly non–T-cell-depleted haplo-SCT with posttransplant cyclophosphamide) is increasing while CBT is decreasing
worldwide (Figure 1), recent developments in CBT, especially cord blood expansion and other strategies to improve
engraftment and immune reconstitution post-CBT, make CBT still a valuable option. This article discusses the 2 options
based on the currently available data, focusing on adults, and tries to give some clues to help the transplant physician
choose a haploidentical vs a cord blood donor. Given the limited numbers of published or ongoing well-designed randomized controlled trials comparing haplo-SCT to CBT and the overall similar clinical results in the available, mostly registrybased, and single-center studies, with substantial heterogeneity and variability, the decision to perform haplo-SCT or CBT
in a given patient depends not only on the patient, disease, and donor characteristics and donor availability (although
most if not all patients should have in principle an alternative donor) but also on the transplant physician’s discretion and,
most importantly, the center’s experience and preference and ongoing protocols and strategies.
LEARNING OBJECTIVES
• Review the available published literature on haploidentical vs cord blood transplantation
• Discuss the available literature that should guide decisions regarding haploidentical vs cord blood transplantation
for an adult patient with a high-risk hematological malignancy in need of allogeneic transplantation
CLINICAL CASE 1
A 50-year-old man hospitalized because of acute myeloid
leukemia (AML) presented with subfebrile fever, flu-like
symptoms, bleeding tendency and leukocytosis, neutropenia, anemia, and thrombocytopenia. Bone marrow
(BM) aspiration revealed 70% myeloblasts. Cytogenetic
and mutational analysis disclosed monosomy 7 and fmslike tyrosine kinase 3 internal tandem duplication positivity while nucleophosmin 1 was negative. Conventional
induction chemotherapy with cytarabine and daunorubicin (7 + 3) in combination with midostaurin yielded the first
complete remission (CR1; by morphology and cytogenetics), but molecular remission was achieved only post highdose cytarabine-midostaurin consolidation. The patient
had no human leukocyte antigen (HLA)-matched brothers or sisters, and a search for an unrelated donor that
64 | Hematology 2022 | ASH Education Program
began at the time of diagnosis failed to allocate a 10 of
10 HLA-matched donor. The patient had 4 brothers and
sisters (2 of them haploidentical [haplo] and 1 with an
HLA DRB1 mismatch), 3 children, and parents with no
comorbidities. In parallel, the search for unrelated donors
yielded few potential cord blood units (CBUs) with a 6 out
of 8 HLA match (considering HLA loci A, B, C, and DRB1).
Two of the potential CBUs contained a class II mismatch
on HLA DRB1. The total nucleated cell dose of the units
ranged between 1 to 2 × 107/kg with 1 to 1.2 × 105 CD34+/kg.
Screening by Luminex assay for anti-HLA donor-specific
antibodies (DSAs) was positive for 2 of the potential CBUs,
with a mean fluorescence intensity level greater than
1000. The patient had no DSAs to the potential haplo
donors. The patient was referred by his primary hematologist to a transplant expert for a consultation about
allogeneic transplantation, which is strongly indicated for
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/64/2022007/64nagler.pdf by guest on 09 December 2022
Arnon Nagler1–3 and Mohamad Mohty3–5
Haplo and cord blood transplantaon in Europe (EBMT)
A.
Haplo&CBT 2000-2020
B.
Haplo&CBT 2000-2020
Haplo vs CB 2011-2020
Haplo vs CB 2000-2010
CB allo
2006-2010
46%
N - 4641
CB allo
2011-2020
19%
N - 4779
Haplo
2000-2010
54%
Haplo
2011-2020
81%
N - 20207
N - 5535
EBMT Survey 2020
Haplo CBT
Haplo&CBT 2000-2020
in acute leukemia
AML
CBT
8%
N - 311
Haplo vs CB
Haplo vs CB in adults
CBT
4%
N - 117
ALL
Haplo vs CB in peds
CBT
19%
N - 194
HAPLO
81%
N - 825
HAPLO
96%
N - 2537
HAPLO
92%
N - 3362
Haplo vs CB in AML
CBT
7%
Haplo vs CB in ALL
CBT
9%
N - 70
N - 98
Haplo CBT Haplo CBT
Haplo CBT Haplo CBT
HAPLO
91%
N - 676
HAPLO
93%
N - 1346
Haplo vs CB in NHL
Haplo vs CB in MDS/MPN
CBT
8%
N - 31
HAPLO
92%
N - 354
CBT
3%
N-8
HAPLO
97%
N - 248
Figure 1. Haploidentical and cord blood transplantation activities in Europe, the US, and China. (A) Haplo-SCT and CBT for AML and
ALL in Europe (Acute Leukemia Working Party of the EBMT). Data include non–T-cell- depleted and T-cell depleted haplo-SCT, single
CTB, and double CBT in 2000-2020. (B) Haplo-SCT and CBT in Europe in adult and pediatric populations in 2000-2020. Data include
haplo-SCT and CBT for various hematological malignancies (AML, ALL, MDS, MPN [myeloproliferative neoplasms], NHL) in 2020. (C)
Haplo-SCT and CBT in the US (CIBMTR) in 2000-2020 for various hematological malignancies (AML, ALL, MDS/MPN, NHL, HL) and
nonmalignant diseases, excluding SAA, modified from CIBMTR 2021 summary slides. Data include adults (>18 years) and children and
adolescents (<18 years) and BM and PB (peripheral blood) grafts, with MAC or RIC/NMA (nonmyeloablative conditioning). (D) HaploSCT (HID) and CBT in China (Chinese Blood and Marrow Transplantation Registry Group) in 2008-2021. Data are for haplo-SCT and
CBT in various hematological malignancies (AML, ALL, MDS, NHL) in 2021. peds, pediatric patients; SCBT; single cord blood transplantation; URD, unrelated donor transplant.
high-risk AML. As no HLA-matched sib­ling or unre­lated donor
was avail­­able, the poten­tial rel­e­vant donors included a haplo
donor,1,2 CB with 2 CBUs or ex vivo expanded CB in a clin­i­
cal trial, or a com­bi­na­tion of a graft from a haplo donor and
a cord (haplo-cord).3-5 The the­o­ret­i­cal con­sid­er­ations, advan­
tages, and dis­ad­van­tages of haplo trans­plants (haplo-SCT) vs
CB trans­plants (CBT) were discussed (Table 1).
The remain­der of this arti­cle focuses mainly on a com­par­i­son of
non–T-cell-depleted haplo-SCT with posttransplant cyclo­phos­
pha­mide (PTCy) and CBT.
Theoretical con­sid­er­ations for haplo vs CB grafts
As depicted in Table 1, CB and haplo donors as alter­na­tive hema­
to­poi­etic stem cell sources have poten­tial strengths as well as lim­
i­ta­tions.1-4 Some of these his­tor­i­cal advan­tages and dis­ad­van­tages
have sub­se­quently been over­come with recent devel­op­ments in
the field, includ­ing dou­ble-unit CBT (DCBT) and the ex vivo expan­
sion of CBUs, over­com­ing the cell-dose lim­i­ta­tion that resulted in
slow and low engraft­ment as well as slow immune recon­sti­tu­tion
and lead to the high infec­tion rates and early trans­plant-related
mor­tal­ity rate asso­ci­ated with CBT.3,4 In par­
al­
lel, the non–Tcell-depleted haplo plat­
form, espe­
cially with PTCy, improves
engraft­
ment rates and immune recon­
sti­
tu­
tion posttransplant,
sig­nif­i­cantly reduc­ing the his­tor­i­cal lim­i­ta­tions of T-cell-depleted
haplo-SCT, which are mainly nonengraftment, high infec­tion rates,
high trans­plant-related com­pli­ca­tions and mor­tal­ity, and high
relapse inci­dence (RI).1,2 Still, haplo grafts pos­sess some advan­
tages in com­par­i­son to CBU grafts, includ­ing a shorter time for
donor iden­ti­fi­ca­tion, the pos­si­bil­ity for a repeat stem cell dona­
tion, donor lym­pho­cyte infu­sion (DLI), and lower cost (Table 1).
Dr Prakash Singh Shekhawat
Haploidentical vs cord trans­plan­ta­tion | 65
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/64/2022007/64nagler.pdf by guest on 09 December 2022
Haplo&CBT by me period
Haplo CBT
2500
2000
Haplo
CB
Haplo and cord blood transplantaon
in the US (CIBMTR)
1500
1000
500
Number of Transplants
300
Haplo > 18 Years
BM
PB
BM
PB
1200
900
600
300
0
250
Haplo < 18 Years
200
150
100
50
0
700
600
500
AML
ALL
MDS / MPN
NHL/HL
Non-malignant disease*
Number of Transplants
Number of Transplants
800
400
300
200
100
0
500
450
400
350
300
250
200
150
100
50
0
CB > 18 Years
400
CB < 18 Years
350
300
250
200
150
100
50
0
2000
1800
1600
1400
1200
1000
800
600
400
200
0
MAC
RIC/NMA
Figure 1. Continued
Haplo-STC vs CBT: reg­is­try-based stud­ies
Several reg­is­try-based stud­ies have com­pared alter­na­tive donor
trans­plants, includ­ing haplo-SCT, to CBT (Supplemental Table 1).
Ruggeri et al,6 on behalf of the Euro­pean Society for Blood and
Marrow Transplantation (EBMT), com­pared non–T-cell-depleted
haplo-SCT to CBT in 1446 patients with AML (n=918) and lym­pho­
cytic leu­ke­mia (ALL; n=528). CBT was asso­ci­ated with delayed
engraft­ment and higher graft fail­ures in both AML and ALL. CBT
resulted in a lower inci­dence of chronic graft-ver­sus-host dis­
ease (cGVHD) in both AML and ALL. The main take-home mes­
sage of this large-cohort reg­is­try-based anal­y­sis was that there
were no sta­tis­ti­cal dif­fer­ences in main out­comes after haplo-SCT
and CBT. Specifically, RI did not dif­fer between the two graft
sources. Nonrelapse mor­tal­ity (NRM) and leu­ke­mia-free sur­vival
66 | Hematology 2022 | ASH Education Program
(LFS) were com­pa­ra­ble as well (Supplemental Table 1).6 A sub­
se­quent anal­y­sis, by Versluis et al,7 assessed the out­comes of
poor-risk AML in CR1 from var­i­ous donor categories. The study
included 193 patients who under­went haplo-SCT and 333 under­
go­ing CBT. The 2-year OS was sig­nif­i­cantly lower for umbil­i­cal
cord blood (UCB) grafts com­pared to haplo-SCT, with no dif­fer­
ences in 2-year NRM. The results were con­firmed by mul­ti­var­i­
able anal­y­sis (MVA). The author con­cludes that espe­cially haplo
but also CB donors are alter­na­tive graft sources for high-risk
AML patients who need hema­to­poi­etic stem cell trans­plan­ta­tion
(HSCT) and do not have an HLA-matched donor. However, suf­fi­
cient num­bers and fol­low-up to define a hier­ar­chy are lacking.7 In
a more recent EBMT anal­y­sis that included 106 188 patients with
dif­fer­ent hema­to­log­i­cal malig­nan­cies and assessed lon­gi­tu­di­nal
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/64/2022007/64nagler.pdf by guest on 09 December 2022
1500
Number of Transplants
1800
Number of Transplants
0
Number of Transplants
Number of Transplants
C.
out­comes of HSCT from var­i­ous graft sources com­par­ing the
results of trans­
plants performed in 2006-2010 to those performed in 2011-2015, the authors dem­on­strated improved 3-year
OS for both haplo-SCT and CBT. Interestingly, the improve­ment
in OS was mostly driven by a reduc­tion in NRM, includ­ing for
haplo-SCT, with the excep­
tion of CBT recip­
i­
ents, who had a
lower RI.8 Both haplo-SCT and CBT dem­on­strated a greater haz­
ard for NRM, with overlapping risk between the 2 donor types.
The lead­ing cause of NRM in both haplo-SCT and CBT was infec­
tions.8 A com­bined EBMT-Eurocord study focused on patients
with AML receiv­
ing thio­
tepa, busul­
fan, and fludarabine (TBF)
con­
di­
tion­
ing before both non–T-cell-depleted haplo-SCT and
sin­gle CBT. It dem­on­strated increased NRM, delayed engraft­
ment, and higher graft fail­ure in the CBT vs haplo-SCT, which
trans­lated into reduced OS and GVHD-free, relapse-free sur­vival
(GRFS) in the CBT group com­pared to the haplo group.9 The
Adult AML and the Donor/Source Working Group of the Japan
Society for Hematopoietic Cell Transplantation also performed
a reg­is­try-based study ret­ro­spec­tively com­par­ing the results of
haplo-SCT to CBT in 1313 adult patients with inter­me­di­ate- or
poor-risk AML in CR (Supplemental Table 1).10 Between 2007 and
2018 in Japan, 211 patients received haplo-SCT and 1102, sin­
gle CBT. The cumu­la­tive inci­dences of neu­tro­phil and plate­let
recov­ery were sig­nif­i­cantly lower in CBT when com­pared with
haplo-SCT. Grade 2 to 4 acute GVHD (aGVHD) was sig­nif­i­cantly
higher in CBT, while exten­sive cGVHD and cyto­meg­a­lo­vi­rus antigenemia were higher in haplo-SCT. In the MVA and pro­pen­sity
scores matching grade 3 to 4, aGVHD as well as RI, NRM, LFS,
OS, and GRFS did not dif­fer between the two graft sources.10
The Kyoto Stem Cell Transplantation Group ret­ro­spec­tively com­
Dr Prakash Singh Shekhawat
Haploidentical vs cord trans­plan­ta­tion | 67
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/64/2022007/64nagler.pdf by guest on 09 December 2022
Figure 1. Continued
Table 1. Haplo-SCT vs CBT: advan­tages and dis­ad­van­tages
Donor
CBT
HAPLO-transplant
SCBT
DCBT
TCD
TCR
Indications
Nonmalignant and
malig­nant
Nonmalignant and
malig­nant
Mainly malig­nant
Mainly malig­nant
Age
Mainly pedi­at­ric
Adult and pedi­at­ric
Adult and pedi­at­ric
Adult and pedi­at­ric
Donor avail­abil­ity
High
High
High
High
1-2 mis­matches
1-2 mis­matches
2-3 mis­matches
2-3 mis­matches
Important
Important
Important
Important
Time for donor
iden­ti­fi­ca­tion
Slower
<1 month
Slower com­pared
to SCBT
Faster (imme­di­ate)
Faster (imme­di­ate)
Time of graft acqui­si­tion
15-30 days
15-30 days
15-30 days
15-30 days
Type of graft
Cryopreserved CBU
2 cryopreserved CBUs
Mobilized PB
Mobilized PB and/or BM
Limitation of graft
acqui­si­tion
Cell dose
Improved com­pared
to SCBT
Poor mobi­li­za­tion
Improved com­pared to TCD haplo
Graft manip­u­la­tion
Ex vivo expan­sion
None
In vivo TCD
None
Donor safety (risk to the donor)
No risk
No risk
Very low
Very low
Easy to repeat stem cell dona­tion
Not pos­si­ble
Not pos­si­ble
Possible
Possible
Conditioning
MAC and RIC
MAC and RIC
MAC
MAC and RIC
Engraftment
Delayed
Improvement com­
pared to SCBT
Comparable to
MRD or MUD
Comparable to MRD or MUD
Use of ATG
Determinantal omit­ted
Determinantal
omit­ted
Indicated (can be
replaced by PTCy)
Indicated (can be replaced
by PTCy)
GVHD
Lower inci­dence with
less strin­gent HLA
matching needed
Incidence increased
com­pared to SCBT
Low inci­dence
Low inci­dence with BM and PTCy;
higher inci­dence with PB
NRM
High due to graft fail­ure
and infec­tions
Improved com­pared
to SCBT
High
Improved com­pared to TCD-haplo
RI
Comparable to MSD or
MUD
Low com­pared to
SCBT
High
Comparable to MSD or MUD (higher
in ini­tial report) with BM and RIC
DLI
Not pos­si­ble
Not pos­si­ble
Possible
Possible
Immune recon­sti­tu­tion
Slow
Improved com­pared
to SCBT
Slow
Improved com­pared to TCD-haplo
Infections inci­dence
High
Improved com­pared
to SCBT
High
Improved com­pared to TCD-Haplo
OS
Lower than other stem
cell sources
Improved com­pared
to SCBT
Lower than other
stem cell sources
Improved com­pared to TCD-Haplo
Cost asso­ci­ated with donor
search and graft pro­cure­ment
High
Very high
Lower
Lower
SCBT, sin­gle cord blood trans­plan­ta­tion; TCD, T-cell deple­tion; TCR, T-cell reple­tion.
pared 57 non–T-cell-depleted haplo-SCT patients to 460 patients
receiv­ing sin­gle CBT between 2013 and 2019 and dem­on­strated
sim­i­lar results: sig­nif­i­cantly delayed neu­tro­phil and plate­let
engraft­ment for CBT recip­i­ents but sim­i­lar OS, NRM, GVHD, and
RI. Interestingly, in the AML sub­group RI was lower with CBT
in com­
par­
i­
son to the haplo-SCT group (Supplemental Table
1).11 The Center for International Blood and Marrow Transplant
Research (CIBMTR) has also recently com­pared 375 non–T-celldepleted haplo-SCT with PTCy to 333 CBT performed between
2012 and 2017 with myeloablative con­di­tion­ing (MAC) in patients
(adults and pedi­at­ric) with dif­fer­ent hema­to­log­i­cal malig­nan­
68 | Hematology 2022 | ASH Education Program
cies. Three-year OS and LFS were sim­i­lar. RI was lower and NRM
was higher in fewer than or equal to 5 out of 8 HLA-matched
CBT patients vs 6 to 8 out of 8 HLA-matched CBT and haplo-SCT
patients. The engraft­
ment of both neu­
tro­
phils and plate­
lets
was delayed with CBT vs haplo-SCT. aGVHD was higher with
CBT, while cGVHD did not dif­fer between the 2 graft sources.
Of note, the authors did not find a cen­ter effect, and no dif­fer­
ence was observed between cen­
ters that con­
trib­
uted more
than 10 vs 6 to 10 cases to the study (Supplemental Table 1).12
In a recent anal­y­sis by the CIBMTR focus­ing on ALL and com­
par­ing 393 non–T-cell-deleted haplo-SCT to other graft sources,
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/64/2022007/64nagler.pdf by guest on 09 December 2022
Degree of HLA mis­match
DSA
Haplo-SCT vs CBT: ran­dom­ized con­trolled tri­als
The 2 ran­dom­ized con­trolled tri­als (RCTs) com­par­ing haplo-SCT
to CBT are rather lim­ited. The CTN1101 trial (NCT01597778) was
performed by the Blood and Marrow Transplant Clinical Trials
Network (BMT CTN) and published in Blood in 2021 (Supplemental Table 1 and Table 2).15 This was a phase 3 trial that ran
between 2012 to 2018, with a rather long recruiting period, and
included 368 patients aged 18 to 70 years with var­i­ous hema­to­
log­i­cal malig­nan­cies in remis­sion (CR; it also included patients
beyond the first CR). Of these, 182 patients under­went haplo
with BM graft, while 186 under­went CBT with 2 CBUs. Both
groups received low-dose TBI/fludarabine/Cy as a con­di­tion­
ing reg­i­men, and haplo recip­i­ents received PTCy as GVHD pro­
phy­laxis. The trial dem­on­strated sim­i­lar 2-year pro­gres­sion-free
sur­vival (PFS) (the study pri­mary end point) between haplo-SCT
and CBT. However, the 2-year NRM was sig­nif­i­cantly higher in
the CBT group com­pared to the haplo group (Supplemental
Table 1). This resulted in sig­nif­i­cantly bet­ter 2-year over­all sur­
vival (OS) for haplo-SCT vs CBT (Supplemental Table 1). Both
were sec­ond­ary end points in the study, with the lim­i­ta­tion
that, usu­ally, stud­ies are not powered on the basis of sec­ond­
ary end points, and rarely are sta­
tis­
ti­
cal adjust­
ments made
for the issue of mul­ti­ple com­par­i­sons. In agree­ment with the
pre­vi­ous lit­er­a­ture, neu­tro­phil recov­ery was lower in the CBT
group com­pared to the haplo group. Platelet recov­ery was no
dif­fer­ent between CBT and haplo-SCT. Of note, the inci­dence
of both aGVHD and cGVHD, includ­ing the severe and exten­sive
forms, was sim­i­lar between the 2 study arms. As for the graftver­sus-leu­ke­mia effect (GVL) ini­tially believed to be some­what
sep­a­rated from GVHD in CBT, inter­est­ingly, the 2-year RI was
almost iden­ti­cal post haplo and CBT (Supplemental Table 1).
Also, the main cause of death in both study arms was the recur­
rence of the malig­nancy. The trial did not assess immu­no­log­i­
cal recon­sti­tu­tion. Notably, mor­tal­ity from GVHD and infec­tions
was com­pa­ra­ble in haplo-SCT and CBT. As the most impor­
tant RTC in the field, even this study faced severe lim­i­ta­tions,
includ­ing a rather long run­ning time and, most impor­tantly,
early ter­mi­na­tion as the trial failed to com­plete accrual, high­
light­ing the main hur­dles in run­ning this type of ran­dom­ized
study. A sec­ond anal­y­sis of the trial dem­on­strated that trans­
plant cen­ter expe­ri­ence mat­ters and that cen­ters with pre­vi­ous
expe­ri­ence in haplo BM trans­plants (≥10 trans­plants) had bet­ter
results, while those with no such expe­ri­ence (≤10 trans­plants)
showed sim­i­lar results for haplo-SCT and CBT. Therefore, cen­
ters with no expe­ri­ence in performing DCBT may do bet­ter performing BM haplo with PTCy.16 Moreover, the CIBMTR recently
reported on an extended fol­low-up of trial par­tic­i­pants beyond
2 years. A com­par­i­son of trial dou­ble-unit UCB (dUCB) and nontrial haplo-BM trans­plan­ta­tion recip­i­ents dem­on­strated higher
PFS and OS at 5 years after haplo-BM trans­plan­ta­tion. A com­
par­i­son to real-world trans­plants disclosed sim­i­lar 5-year OS in
reg­is­try-based and nontrial DCBT and between trial and nontrial
haplo-BM trans­plants, respec­tively.17 Notably, this addi­
tional
reg­is­try-based study com­par­ing real-world trans­plan­ta­tion
out­comes in sub­stan­tially larger num­bers of nontrial haplo-BM
trans­plants to trial results with lon­ger fol­low-up revealed that
5-year sur­
vival was sig­
nif­i­
cantly higher after nontrial haplo
BM trans­
plants com­
pared with trial dUCB, while the lon­
ger
­fol­low-up of the orig­i­nal trial cohorts failed to dem­on­strate a
dif­fer­ence in 5-year PFS and OS between the treat­ment groups.
Further, the type of graft mat­
ters, as nontrial patients who
under­went haplo trans­plants using mobi­lized PBSC trans­plants
dem­on­strated sig­nif­i­cantly higher 5-year sur­vival com­pared to
trial haplo BM and nontrial haplo BM. Similarly, sur­vival was bet­
ter after haplo-PBSC com­pared to trial UCB and nontrial UCB.17
The sec­ond RCT to ran­dom­ize haplo-SCT vs CBT was performed by Sanz et al (NCT02386332) (Supplemental Table 1 and
Table 2).18 The study pro­spec­tively ran­dom­ized CBT with sin­gle
CBUs and non–T-cell-depleted haplo with PBSCs, both fol­low­
ing MAC consisting of TBF. Of note, sin­gle UCB is not a stan­
dard of care in North America. Anti-GVHD pro­phy­laxis included
antithymocyte glob­u­lin (ATG) in the CBT group and PTCy in the
haplo-SCT group. The study report includes 22 patients who
under­went haplo-SCT and 23 who under­went CBT.18 Interestingly, 4 patients allo­cated to the haplo arm crossed over to
the CBT arm because no haplo donor could be allo­cated (in
the CTN1101 trial, CB was used for 11 of 182 patients ran­dom­
ized to the haplo arm due to donors being inel­i­gi­ble or declin­
ing to donate, whereas only 1 of 186 patients ran­dom­ized to
receive CB crossed over to the haplo study arm, due to patient
­pref­er­ence). In this rather small study, neu­tro­phil and plate­
let recov­ery was slower and lower with CBT com­pared to the
haplo-SCT. Acute and chronic GVHD, as well as RI, was sim­i­lar
between the haplo-SCT and the CBT groups. NRM was higher
in CBT in com­par­i­son to haplo-SCT. Two-year dis­ease-free sur­
vival (DFS), OS, and GRFS were sig­nif­i­cantly bet­ter in haplo-SCT
com­pared to CBT. This study empha­sizes once again the tech­ni­
cal dif­fi­cul­ties of run­ning donor-based RTC stud­ies and also the
het­ero­ge­ne­ity, even in RTC stud­ies, in com­par­ing haplo-SCT to
CBT not just in patients and dis­ease char­ac­ter­is­tics but also in
the trans­plant plat­form, includ­ing sin­gle vs dou­ble CB grafts,
BM vs PBSC graft, MAC vs reduced-inten­sity con­di­tion­ing (RIC),
and GVHD pro­phy­laxis.
Haplo-SCT vs CBT: meta-anal­y­sis
Several meta-ana­
ly­
ses have com­
pared haplo to other graft
sources, includ­ing CBT, as stem cell sources for allo­ge­neic trans­
plan­ta­tion.19 In 1 such meta-anal­y­sis recently reported by Owattanapanich et al com­par­ing haplo-SCT and CBT in patients with
ALL in mostly reg­is­try-based stud­ies, the pooled odds ratios of
OS and LFS in the haplo-SCT and CBT groups were not sta­tis­
ti­cally dif­fer­ent. Similarly, the NRM, RI, and aGVHD grade 2 to
4 out­comes of the 2 groups did not sta­tis­ti­cally dif­fer. Another
meta-anal­
y­
sis included 7 stud­
ies with 1311 adults under­
go­
ing
Dr Prakash Singh Shekhawat
Haploidentical vs cord trans­plan­ta­tion | 69
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/64/2022007/64nagler.pdf by guest on 09 December 2022
includ­ing 305 CBTs in the years 2013 to 2017, CBT had infe­rior OS
and LFS before 18 months and sim­i­lar OS and LFS after 18 months.
The rates of RI and cGVHD after 18 months were no dif­fer­ent
between CBT and haplo-SCT, but CBT had increased inci­dences
of NRM and grade 3 to 4 aGVHD (Supplemental Table 1).13 Finally,
the health care bur­den, espe­cially the long-term health care bur­
den, is becom­ing a very rel­e­vant issue with mod­ern, sophis­ti­
cated, and com­plex ther­a­pies. Recently, the Minnesota group
ana­
lyzed the 5-year health care bur­
den in 1077 con­
sec­
u­
tive
adults under­go­ing HSCT from var­i­ous graft sources, includ­ing
BM, periph­eral blood stem cells (PBSCs), and CBUs. Notably,
adults under­go­ing CBT had a lower long-term health care bur­
den com­pared with the BM and PBSC graft sources, prob­a­bly
due to a lower inci­dence of cGVHD, as well as other poten­tial
unknown fac­tors, which trans­lated into a bet­ter qual­ity of life.14
Dr Prakash Singh Shekhawat
Ohio State University
Comprehensive Cancer
Center,
Colum­bus, OH, US
Algemeen Ziekenhuis
Sint-Jan,
Brugge, Belgium
Institut Jules Bordet,
Brussels, Belgium
Universitair Ziekenhuis
Gasthuisberg,
Leuven, Belgium
Hamilton Niag­ara Regional
Haemophilia Centre,
Hamilton, Ontario, Canada
(and 5 more)
NCT02188290
Data used with per­mis­sion from ClinicalTrials​­.gov.
ATIR, haploidentical, naive cell-enriched T-cell prod­uct, depleted of recip­i­ent-alloreactive T cells.40
178
adult,
older
adult
(18y to
65y)
368
adult,
older
adult
(18y to
70y)
CIBMTR, Medical College
of Wisconsin, US
University of Alabama
at Birmingham,
Birmingham, AL, US
Arizona Cancer Center,
Phoenix, AZ, US
City of Hope National
Medical Center,
Duarte, CA, US
(and 35 more)
NCT01597778
A Multi-Center, Phase III,
Randomized Trial of RIC and
Transplantation of (dUCB) Versus
HLA-Haplo Related Bone Marrow
for Patients With Hematologic
Malignancies
An Observational Cohort Study
on Transplant-Related Mortality
in Patients Receiving Either
a Hematopoietic Stem Cell
Transplantation Without ATIR
or an Umbilical Cord Blood
Transplantation
206
adults,
≥18 and
≤55y
Instituto de Investigacion
Sanitaria La Fe, Valencia,
Spain
NCT02386332
A Randomized Multicenter Study
Comparing Unrelated Umbilicalcord Blood Transplant Versus
Human Leukocyte Antigen (HLA)Haploidentical Related Hematopoietic
Stem Cell Transplant for Adult Patients
With Hematologic Malignancies
Retrospective
Randomized
Horowitz M
Completed
AML
ALL
Bur­kitt
lym­phoma
Follicular
lym­phoma
Hodgkin
lym­phoma
Mantle cell
lym­phoma
Non-Hodgkin
lym­phoma
AML
ALL
MDS
Percentage of
par­tic­i­pants
with PFS (time
frame: 2y)
Transplantrelated mor­tal­ity
(time frame: up
to 12 mo after
trans­plan­ta­tion)
Sep­tem­ber
2014-6
Octo­ber
2015
12 May
2012-11
Sep­tem­ber
2020
Sanz MA
Unknown
Hematologic
malig­nan­cies
DFS
(time frame: 2y)
Completed
Rovers J
Jia Chen M.
11 March
2015-March
2020
Recruiting
Hematologic
dis­eases
Estimated OS at
3y post SCT
Primary
inves­ti­ga­tor
1 June
2017- 31
Decem­ber
2022
Randomized
100
child,
adult
First Affiliated Hospital
of Soochow University
Suzhou, Jiangsu, China
NCT03719534
Haplo SCT vs Haplo-cord SCT
for Patients With Hematological
Disorders
Status
Indication
Primary out­come
mea­sures
Study
period
Allocation
Patients, N
Institution/
coun­try
Clinical trial
iden­ti­fier
Study title
Table 2. Randomized con­trolled tri­als of haplo-SCT vs CBT
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/64/2022007/64nagler.pdf by guest on 09 December 2022
70 | Hematology 2022 | ASH Education Program
Fuchs
et al15
Brunstein
et al16
Sanz et al18
References
CBT and 915 adults under­go­ing haplo-SCT and showed no major
dif­fer­ence in aGVHD and cGVHD and no dif­fer­ence in 2-year RI,
NRM, and DFS (mea­sures of sur­vival could not be pooled because
of the dif­fer­ent def­i­ni­tions used in the dif­fer­ent stud­ies included).
The authors con­cluded that CBTs and haplo-SCT can be con­sid­
ered equally effec­tive options for adult patients with­out HLAmatched donors.20 A third meta-anal­y­sis included 12 stud­ies that
com­pared haplo-SCT to CBT in patients with var­i­ous hema­to­log­i­
cal malig­nan­cies that dif­fered from the above 2 meta-ana­ly­ses. It
reported supe­rior OS and lower NRM and aGVHD with haplo-SCT
com­pared to the CBT group, while RI was sim­i­lar and cGVHD
higher, respec­tively.21
CBT.29 More recently, Childs et al conducted a pro­spec­tive phase
2 trial of haplo-cord (co-infus­ing a sin­gle UCB unit with CD34+selected cells from a haplo rel­a­tive) in 29 patients, includ­ing 10
with hypo­plas­tic MDS post severe aplastic ane­mia (SAA) and
10 SAA patients. Neutrophil recov­ery occurred with a median of
10 days and plate­let recov­ery, with a median of 32 days. Grade 2
to 4 aGVHD was 21% while cGVHD was 41%. The 7.5-year OS was
83% and GRFS was 69%. The authors con­cluded that haplo-cord
trans­plants pro­vide an alter­na­tive option for patients with hypo­
plas­tic MDS and refrac­tory SAA who lack HLA-matched donors.30
CLINICAL CASE 1 (Con­t in­u ed)
The patient was informed, after going over the avail­­able lit­er­a­
ture, includ­ing RCTs, reg­is­try-based stud­ies, and meta-­­anal­y­sis,
that both haplo trans­
plan­
ta­
tion and CBT were legit­
i­
mate
options. However, the poten­tial CBUs allo­cated to him were
below the recommended cell dose. Additionally, given the
patient’s DSAs and the referred trans­
plant cen­
ter’s lack of
expe­ri­ence in CBT and the lack of an ongo­ing clin­i­cal trial with
ex vivo expanded or haplo-cord, the cen­ter recommended a
non–T-cell-depleted haplo trans­plant with TBF con­di­tion­ing
and PTCy as anti-GVHD pro­
phy­
laxis.3,22 The pre­
ferred haplo
donor was his brother with the HLA-DRB1 mis­match in view of
the emerg­ing data dem­on­strat­ing that HLA-DRB1 mis­matches in
haplo-SCT may be asso­ci­ated with a lower risk of dis­ease recur­
rence and that HLA-DRB1 mismatching with HLA-DQB1 matching and HLA-B leader matching in spe­cial sce­nar­ios cor­re­lates
with improved DFS and OS.23 Furthermore, the patient was told
that due to har­bor­ing mono­somy 7 and fms-like tyro­sine kinase
3 inter­nal tan­dem dupli­ca­tion muta­tion defin­ing high-risk leu­
ke­mia, which has been shown to be a poor prog­nos­tic fac­tor
in alter­na­tive donor trans­plants,24,25 the risk of posttransplant
recur­rence was sub­stan­tial. As per recent pub­li­ca­tions, HLA-B
leader might also have impli­ca­tions for NK immu­no­ther­apy,26
which may be impor­tant for the GVL effect. For the same rea­
son, posttransplant sorafenib was recommended.27
Haplo-SCT vs CBT: myelodysplastic syn­drome
and sec­ond­ary AML
Few stud­ies have com­pared haplo-SCT and CBT in myelodysplastic syn­
drome (MDS) and sec­
ond­
ary AML. The EBMT performed a reg­is­try-based study com­par­ing non–T-cell-depleted
haplo-SCT with PTCy to CBT and mismatched unre­lated trans­
plants (MMUDs) in patients with MDS (Supplemental Table 1).28
The haplo-SCT group consisted of 222 patients while the CBT
group com­prised 168 patients. Engraftment was lower with CBT
com­pared to haplo-SCT, while GVHD was lower in haplo-SCT,
and 3-year cGVHD and RI did not dif­fer. OS and NRM were bet­
ter in haplo-SCT com­pared to CBT (Supplemental Table 1). Additional anal­y­sis was performed for 409 sec­ond­ary AML patients
receiv­
ing either CBT (n=163) or haplo-SCT (n=246). CBT was
asso­ci­ated with a higher risk of grade 2 to 4 aGVHD and lower
GRFS com­pared to haplo-SCT. Of note, cGVHD, RI, NRM, LFS,
and OS were not sta­tis­ti­cally dif­fer­ent between haplo-SCT and
A 32-year-old man presented with clin­i­cally mean­ing­ful lymph­
ade­nop­a­thies. A mor­pho­log­i­cal, immu­no­his­to­chem­i­cal, and
molec­u­lar eval­u­a­tion of a lymph node biopsy disclosed highrisk lym­phoma with MYC trans­lo­ca­tion and TP53 muta­tion. The
dis­ease was refrac­tory to mul­ti­ple lines of immune che­mo­ther­
apy as well as HSCT from an HLA-matched sib­ling donor (MSD)
fol­low­ing MAC. Within 3 weeks of trans­plant, the dis­ease progressed to a leu­ke­mic phase refrac­tory to high-dose cytarabine and mitoxantrone and DLI. The patient was admit­ted and
received chi­me­ric anti­gen recep­tor (CAR) CD19 T-cell ther­apy.
One month after cell infu­sion, pos­i­tron emis­sion tomog­ra­phy/
com­puted tomog­ra­phy showed sig­nif­i­cant improve­ment, with
the res­o­lu­tion of pre­vi­ous F-fluorodeoxyglucose-avid lesions
con­sis­tent with a com­plete met­a­bolic response. In view of the
ultra-high-risk dis­ease and the short per­sis­tence of the CAR T
cells and the pre­vi­ous relapse 3 weeks post HLA-iden­ti­cal sib­
ling trans­plant, a sec­ond trans­plant from an alter­na­tive donor
was con­sid­ered. As no HLA-com­pat­i­ble unre­lated donor was
avail­­able, the alter­na­tives were haplo-SCT or CBT.
Haplo-SCT vs CBT: non-Hodgkin, Hodgkin lym­phoma,
and other lym­phatic malig­nan­cies
Although the num­
ber of HSCTs in malig­
nant lym­
pho­
mas is
decreas­
ing due to the emer­
gence of novel drug-con­
ju­
gated
and bispecific novel mono­
clo­
nal antibodies and CAR T cells,
HSCT still has a poten­tial role in a selected group of young lym­
phoma patients. Few sin­gle-cen­ter and reg­is­try-based stud­ies
have com­pared haplo-SCT to CBT for non-Hodgkin lym­phoma
(NHL) and Hodgkin lym­phoma (HL). A joint study by the EBMT,
Eurocord, and the CIBMTR ana­lyzed 740 patients with NHL and
HL who received either haplo-SCT or CBT between 2009 and
2016. Haplo trans­plants with either BM or PBSCs as a graft source
resulted in sig­nif­i­cantly lower NRM and bet­ter PFS and OS (Supplemental Table 1).31 Cornillon et al com­pared haplo-SCT to CBT
in 95 patients with T-cell lym­phoma, dem­on­strat­ing sim­i­lar PFS
and OS and show­ing that the donor source, whether haplo or
CBU, had no major impact on the trans­plant out­come (Supplemental Table 1).32 Messer et al sys­tem­at­i­cally reviewed some of
the early stud­ies on alter­na­tive and unre­lated donor trans­plan­
ta­tion in HL. In the 2 or 3 stud­ies that included haplo-SCT and
CBT, out­comes did not dif­fer at large; how­ever, the group was
very het­
ero­
ge­
neous, and some of the out­
come param­
e­
ters
were miss­ing.33 Additional stud­ies com­bined the 2 graft sources
in high-risk lym­phoma and chronic lym­pho­cytic leu­ke­mia with
inter­est­ing results.34,35 Another addresses haplo-SCT and CBT as
Dr Prakash Singh Shekhawat
Haploidentical vs cord trans­plan­ta­tion | 71
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/64/2022007/64nagler.pdf by guest on 09 December 2022
CLINICAL CASE 2
well as haplo-cord trans­plants in pedi­at­ric ALL, which is beyond
the scope of this review.36-38
Conflict-of-inter­est dis­clo­sure
CLINICAL CASE 2 (Con­t in­u ed)
Future direc­tions
The future opti­mi­za­tion of CBT should include a bet­ter under­
stand­ing of the clin­i­cal sig­nif­i­cance of the HLA bar­rier, hope­
fully lead­ing to improve­ments in the cur­rent selec­tion of the
opti­
mal CBU, increas­
ing CBT cell dose, improv­
ing engraft­
ment by ex vivo expan­sion, and omit­ting in vivo T-cell deple­
tion (ATG) from the pre­
par­
a­
tive pro­
to­
cols. Some of these
strat­e­gies have already been implemented, lead­ing to sig­nif­
i­cant improve­ments in hema­to­poi­etic engraft­ment post CBT
and a sub­stan­tially reduced nonengraftment rate, short­en­ing
the length of posttransplant pan­cy­to­pe­nia, reduc­ing infec­tion
rates, improv­
ing immune recon­
sti­
tu­
tion post CBT, impres­
sively reduc­ing day-100 mor­tal­ity post CBT, and improv­ing
out­comes.39 As for haplo trans­plants, future direc­tions must
focus on opti­miz­ing con­di­tion­ing pro­to­cols, defin­ing the
opti­mal dose and sched­ule of PTCy in non–T-cell-depleted
trans­plants, and facil­i­tat­ing posttransplant immune recon­sti­
tu­tion.40 Optimizing con­di­tion­ing pro­to­cols with com­pounds
such as treosulfan and espe­cially using phar­ma­co­ki­net­icsbased con­di­tion­ing improves haplo-SCT suc­cess rates in the
advanced-age patient pop­u­la­tion.2 The shift from BM grafts,
the orig­
i­
nal stem cell source in non–T-cell-­
depleted haplo-SCT with PTCy, to PBSC grafts may speed engraft­ment and
immune recon­sti­tu­tion and reduce the relapse rate, with the
cost of higher inci­dences of both GVHD and cyto­kine release
syn­drome (CRS).2,41 Notably, patients who devel­oped CRS fol­
low­ing a non–T-cell-depleted haplo-SCT with PBSC grafts and
PTCy dem­on­strated a higher inci­dence of grade 2 to 4 acute
GVHD com­pared to those who did not.41 Finally, the opti­mal
dose and sched­ule of PTCy are yet to be defined, and it is
con­ceiv­able that lower doses reduce some of the reported
trans­plant-asso­ci­ated tox­ic­ity, includ­ing cardiotoxicity and
CRS.41,42 Based on the prom­is­ing results of the ongo­ing stud­
ies aiming to improve engraft­ment in the set­ting of CBT and to
fas­ten immune recon­sti­tu­tion post haplo-SCT,39,40 future out­
comes with both trans­plant options look encour­ag­ing.
Acknowledgments
The authors thank Annalisa Ruggeri for help­ful dis­cus­sion and crit­
i­cal read­ing of the man­u­script; Jacob Passweg, Helen B
­ aldomero,
the Acute Leukemia Working Party (ALWP) office, Xiao Jun Huang,
and Marcelo Pasquini for pro­vid­ing trans­plant activ­i­ties in the
EBMT, Chi­nese Blood and Marrow Transplantation Registry, and
Arnon Nagler: no com­pet­ing finan­cial inter­ests to declare.
Mohamad Mohty: no com­pet­ing finan­cial inter­ests to declare.
Off-label drug use
Arnon Nagler: nothing to disclose.
Mohamad Mohty: nothing to disclose.
Correspondence
Arnon Nagler, Division of Hematology, Bone Marrow Transplantation, and Hemato-Oncology Center, Chaim Sheba Medical Center, Tel HaShomer, 52621 Ramat Gan, Israel; e-mail: a​
­.nagler@sheba​­.health​­.gov​­.il.
References
1.
McCurdy SR, Luznik L. How we per­form haploidentical stem cell trans­plan­
ta­tion with posttransplant cyclo­phos­pha­mide. Blood. 2019;134(21):18021810.
2. Nagler A, Ruggeri A. Haploidentical stem cell trans­plan­ta­tion (HaploSCT)
for patients with acute leu­ke­mia—an update on behalf of the ALWP of the
EBMT. Bone Marrow Transplant. 2019;54(suppl 2):713-718.
3. Politikos I, Davis E, Nhaissi M, et al; Amer­i­can Society for Transplantation
and Cellular Therapy Cord Blood Special Interest Group. Guidelines for
cord blood unit selec­tion. Biol Blood Marrow Transplant. 2020;26(12):21902196.
4. Baron F, Ruggeri A, Nagler A. Methods of ex vivo expan­sion of human cord
blood cells: chal­lenges, successes and clin­i­cal impli­ca­tions. Expert Rev
Hematol. 2016;9(3):297-314.
5. van Besien K, Artz A, Champlin RE, et al. Haploidentical vs haplo-cord
trans­plant in adults under 60 years receiv­ing fludarabine and mel­pha­lan
con­di­tion­ing. Blood Adv. 2019;3(12):1858-1867.
6. Ruggeri A, Labopin M, Sanz G, et al; Eurocord, Cord Blood Committee of
Cellular Therapy and Immunobiology Working Party—Euro­pean Society
for Blood and Marrow Transplantation; ALWP-EBMT Study Investigators.
Comparison of out­comes after unre­lated cord blood and unma­nip­u­lated
haploidentical stem cell trans­plan­ta­tion in adults with acute leu­ke­mia. Leukemia. 2015;29(9):1891-1900.
7. Versluis J, Labopin M, Ruggeri A, et al. Alternative donors for allo­ge­neic
hema­to­poi­etic stem cell trans­plan­ta­tion in poor-risk AML in CR1. Blood
Adv. 2017;1(7):477-485.
8. Shouval R, Fein JA, Labopin M, et al. Outcomes of allo­ge­neic haematopoietic stem cell trans­plan­ta­tion from HLA-matched and alter­na­tive donors:
a Euro­pean Society for Blood and Marrow Transplantation reg­is­try ret­ro­
spec­tive anal­y­sis. Lancet Haematol. 2019;6(11):e573-e584.
9. Giannotti F, Labopin M, Shouval R, et al. Haploidentical trans­plan­ta­tion
is asso­ci­ated with bet­ter over­all sur­vival when com­pared to sin­gle cord
blood trans­plan­ta­tion: an EBMT-Eurocord study of acute leu­ke­mia patients
con­di­tioned with thio­tepa, busul­fan, and fludarabine. J Hematol Oncol.
2018;11(1):110.
10. Konuma T, Kanda J, Yamasaki S, et al. Single cord blood trans­
plan­
ta­
tion ver­sus unma­nip­u­lated haploidentical trans­plan­ta­tion for adults with
acute mye­loid leu­ke­mia in com­plete remis­sion. Transplant Cell Ther.
2021;27(4):334.1-334334.e11.
11. Wada F, Kanda J, Yoshioka S, et al; Kyoto Stem Cell Transplantation Group.
Single cord blood trans­plan­ta­tion ver­sus HLA-haploidentical-related donor
trans­plan­ta­tion using posttransplant cyclo­phos­pha­mide in patients with
hema­to­log­i­cal malig­nan­cies. Transplantation. 2022;106(6):1279-1287.
12. Wagner JE, Ballen KK, Zhang MJ, et al. Comparison of haploidentical and
umbil­i­cal cord blood trans­plan­ta­tion after myeloablative con­di­tion­ing.
Blood Adv. 2021;5(20):4064-4072.
13. Wieduwilt MJ, Metheny L, Zhang MJ, et al. Haploidentical vs sib­ling, unre­
lated, or cord blood hema­to­poi­etic cell trans­plan­ta­tion for acute lym­pho­
blas­tic leu­ke­mia. Blood Adv. 2022;6(1):339-357.
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/64/2022007/64nagler.pdf by guest on 09 December 2022
Following broad con­
sul­
ta­
tion with the patient and his fam­
ily eval­u­at­ing the lim­ited ther­a­peu­tic alter­na­tives in a highly
aggres­sive TP53-pos­
i­
tive lym­
phoma and the course of the
patient’s dis­ease, haplo-SCT with TBF con­di­tion­ing and PTCy
was performed. The donor was a haplo brother because it was
felt that a haplo donor, rather than a CB donor, offered the fast
engraft­ment needed. Unfortunately, the patient died early after
the sec­ond trans­plant from trans­plant-related tox­ic­ity.
72 | Hematology 2022 | ASH Education Program
Center for International Blood and Marrow Transplant Research,
respec­tively; Ivetta Danylesko and Katia Beider for help­ing with
the man­u­script tables and illus­tra­tions, and Myriam Labopin and
other col­leagues from the ALWP of the EBMT.
with sec­ond­ary acute mye­loid leu­ke­mia, a com­par­a­tive study from Eurocord
and the ALWP EBMT. Bone Marrow Transplant. 2019;54(12):1987-1994.
30. Childs RW, Tian X, Vo P, et al. Combined haploidentical and cord blood
trans­plan­ta­tion for refrac­tory severe aplastic anae­mia and hypo­plas­tic
myelodysplastic syn­drome. Br J Haematol. 2021;193(5):951-960.
31. Fatobene G, Rocha V, St Martin A, et al. Nonmyeloablative alter­
na­
tive
donor trans­plan­ta­tion for Hodgkin and non-Hodgkin lym­phoma: from the
LWP-EBMT, Eurocord, and CIBMTR. J Clin Oncol. 2020;38(14):1518-1526.
32. Cornillon J, Daguenet E, Tournilhac O, et al. Allogeneic hema­
to­
poi­
etic
stem cell trans­plan­ta­tion from unma­nip­u­lated haploidentical donor and
unre­lated cord blood for T-cell lym­phoma: a ret­ro­spec­tive study from the
Société Francophone de Greffe de Moelle et de Therapie Cellulaire. Bone
Marrow Transplant. 2021;56(11):2849-2856.
33. Messer M, Steinzen A, Vervölgyi E, et al. Unrelated and alter­na­tive donor
allo­
ge­
neic stem cell trans­
plant in patients with relapsed or refrac­
tory
­Hodgkin lym­phoma: a sys­tem­atic review. Leuk Lymphoma. 2014;55(2):296306.
34. Zhou B, Xu M, Lu S, et al. Clinical out­comes of B cell acute lym­pho­blas­tic
leu­ke­mia patients treated with haploidentical stem cells com­bined with
umbil­i­cal cord blood trans­plan­ta­tion. Transplant Cell Ther. 2022;28(3):173.1173173.e6.
35. Hsu J, Artz A, Mayer SA, et al. Combined haploidentical and umbil­
i­
cal
cord blood allo­ge­neic stem cell trans­plan­ta­tion for high-risk lym­phoma
and chronic lym­pho­blas­tic leu­ke­mia. Biol Blood Marrow Transplant.
2018;24(2):359-365.
36. Ponvilawan B, Owattanapanich W, Charoenngam N, et al. Effects of dif­fer­
ent types of allo­ge­neic hema­to­poi­etic stem cell trans­plan­ta­tion donors on
Philadelphia chro­mo­some-pos­i­tive acute lym­pho­blas­tic leu­ke­mia dur­ing
the tyro­sine kinase inhib­i­tor era: a sys­tem­atic review and meta-anal­y­sis
[published online ahead of print 2 June 2022]. Hematol Oncol Stem Cell
Ther.doi:10.1016/j.hemonc.2021.09.007.
37. Gómez-Santos C, González-Vicent M, Molina B, et al. Comparison of clin­i­cal
out­comes between unre­lated sin­gle umbil­i­cal cord blood and “ex-vivo”
T-cell depleted haploidentical trans­plan­ta­tion in chil­dren with hema­to­log­
i­cal malig­nan­cies. World J Pediatr. 2021;17(6):609-618.
38. Mo XD, Tang BL, Zhang XH, et al. Comparison of out­comes after umbil­i­
cal cord blood and unma­nip­u­lated haploidentical hema­to­poi­etic stem cell
trans­plan­ta­tion in chil­dren with high-risk acute lym­pho­blas­tic leu­ke­mia. Int
J Cancer. 2016;139(9):2106-2115.
39. Horwitz ME, Stiff PJ, Cutler C, et al. Omidubicel vs stan­dard myeloablative umbil­i­cal cord blood trans­plan­ta­tion: results of a phase 3 ran­dom­ized
study. Blood. 2021;138(16):1429-1440.
40. Roy DC, Walker I, Maertens J, et al. ATIR101 admin­is­tered after T-cell-depleted haploidentical HSCT reduces NRM and improves over­all sur­vival in
acute leu­ke­mia. Leukemia. 2020;34(7):1907-1923.
41. Solán L, Landete E, Bailén R, et al. Cytokine release syn­drome after allo­
ge­neic stem cell trans­plan­ta­tion with posttransplant cyclo­phos­pha­mide.
Hematol Oncol. 2020;38(4):597-603.
42. Duléry R, Mohty R, Labopin M, et al. Early car­diac tox­ic­ity asso­ci­ated with
post-trans­plant cyclo­phos­pha­mide in allo­ge­neic stem cell trans­plan­ta­tion.
JACC CardioOncol. 2021;3(2):250-259.
© 2022 by The Amer­i­can Society of Hematology
DOI 10.1182/hema­tol­ogy.2022000327
Dr Prakash Singh Shekhawat
Haploidentical vs cord trans­plan­ta­tion | 73
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/64/2022007/64nagler.pdf by guest on 09 December 2022
14. Garcia JG, Grillo S, Cao Q , et al. Low 5-year health care bur­den after umbil­
i­cal cord blood trans­plan­ta­tion. Blood Adv. 2021;5(3):853-860.
15. Fuchs EJ, O’Donnell PV, Eapen M, et al. Double unre­lated umbil­i­cal cord
blood vs HLA-haploidentical bone mar­row trans­plan­ta­tion: the BMT CTN
1101 trial. Blood. 2021;137(3):420-428.
16. Brunstein CG, O’Donnell PV, Logan B, et al; Blood and Marrow Transplant
Clinical Trials Network. Impact of cen­ter expe­ri­ence with donor type on
out­comes: a sec­ond­ary anal­y­sis, Blood and Marrow Transplant Clinical Trials Network 1101. Transplant Cell Ther. 2022;28(7):406.e1-406.e6.
17. O’Donnell PV, Brunstein CG, Fuchs EJ, et al. Umbilical cord blood or
HLA-haploidentical trans­plan­ta­tion: real-world out­comes ver­sus ran­dom­
ized trial out­comes. Transplant Cell Ther. 2022;28(2):109.1-109109.e8.
18. Sanz J, Montoro J, Solano C, et al. Prospective ran­dom­ized study com­par­
ing myeloablative unre­lated umbil­i­cal cord blood trans­plan­ta­tion ver­sus
HLA-haploidentical related stem cell trans­plan­ta­tion for adults with hema­
to­logic malig­nan­cies. Biol Blood Marrow Transplant. 2020;26(2):358-366.
19. Owattanapanich W, Leelakanok N, Sanpakit K, Buaboonnam J. A com­par­
i­son of the clin­i­cal out­comes of haploidentical trans­plan­ta­tion and other
graft sources in acute lym­pho­blas­tic leu­ke­mia: a sys­tem­atic review and
meta-anal­y­sis. Clin Lymphoma Myeloma Leuk. 2022;22(3):174-191.
20. Li D, Li X, Liao L, Li N. Unrelated cord blood trans­plan­ta­tion ver­sus haploidentical trans­plan­ta­tion in adult and pedi­at­ric patients with hema­to­log­
i­cal malig­nan­cies—a meta-anal­y­sis and sys­tem­atic review. Am J Blood Res.
2020;10(1):1-10. https:​­/​­/pubmed​­.ncbi​­.nlm​­.nih​­.gov​­/32206440​­/. Accessed 1
February 2020.
21. Wu R, Ma L. Haploidentical hema­to­poi­etic stem cell trans­plan­ta­tion ver­
sus umbil­i­cal cord blood trans­plan­ta­tion in hema­to­logic malig­nan­cies:
a sys­tem­atic review and meta-anal­y­sis. Cell Transplant. 2020;29(JanDec):963689720964771.
22. Ciurea SO, Cao K, Fernandez-Vina M, et al. The Euro­pean Society for Blood
and Marrow Transplantation (EBMT) con­sen­sus guide­lines for the detec­
tion and treat­ment of donor-spe­cific anti-HLA antibodies (DSA) in haploidentical hema­to­poi­etic cell trans­plan­ta­tion. Bone Marrow Transplant.
2018;53(5):521-534.
23. Fuchs EJ, McCurdy SR, Sol­o­mon SR, et al. HLA informs risk pre­dic­tions after
haploidentical stem cell trans­plan­ta­tion with posttransplantation cyclo­
phos­pha­mide. Blood. 2022;139(10):1452-1468.
24. Canaani J, Labopin M, Huang XJ, et al. T-cell replete haploidentical stem
cell trans­plan­ta­tion atten­u­ates the prog­nos­tic impact of FLT3-ITD in acute
mye­loid leu­ke­mia: a report from the Acute Leukemia Working Party of the
Euro­pean Society for Blood and Marrow Transplantation. Am J Hematol.
2018;93(6):736-744.
25. Ustun C, Giannotti F, Zhang MJ, et al. Outcomes of UCB trans­plan­ta­tion are
com­pa­ra­ble in FLT3+ AML: results of CIBMTR, EUROCORD and EBMT col­lab­
o­ra­tive anal­y­sis. Leukemia. 2017;31(6):1408-1414.
26. Petersdorf EW, Gooley T, Volt F, et al. Use of the HLA-B leader to opti­mize
cord blood trans­plan­ta­tion. Haematologica. 2021;106(12):3107-3114.
27. Bazarbachi A, Bug G, Baron F, et al. Clinical prac­tice rec­om­men­da­tion
on hema­to­poi­etic stem cell trans­plan­ta­tion for acute mye­loid leu­ke­mia
patients with FLT3-inter­nal tan­dem dupli­ca­tion: a posi­tion state­ment from
the Acute Leukemia Working Party of the Euro­pean Society for Blood and
Marrow Transplantation. Haematologica. 2020;105(6):1507-1516.
28. Robin M, Porcher R, Ruggeri A, et al. HLA-mismatched donors in patients
with myelodysplastic syn­drome: an EBMT reg­is­try anal­y­sis. Biol Blood Marrow Transplant. 2019;25(1):114-120.
29. Ruggeri A, Labopin M, Savani B, et al. Hematopoietic stem cell trans­plan­ta­
tion with unre­lated cord blood or haploidentical donor grafts in adult patients
ARE ALTERNATIVE DONORS NOW MAINSTREAM IN ALLOGENEIC TRANSPLANT ?
New strategies for mismatched unrelated donor
(MMUD) hematopoietic cell transplant (HCT)
City of Hope National Medical Center, Duarte, CA
With increasing numbers of patients with hematologic malignancies requiring allogeneic hematopoietic cell transplant
(HCT), including minority racial and ethnic groups, the limited availability of matched related donors and matched unrelated donors remains a significant obstacle. Hence, the use of alternative donors such as haploidentical and mismatched
unrelated donors (MMUDs) is on the rise. Herein, we present case studies to outline a rational and stepwise approach
with a focus on the use of MMUD for HCT in patients with hematologic malignancies. We also review novel approaches
used to reduce the incidence of severe graft-versus-host disease and improve HCT outcomes in patients undergoing
MMUD HCT.
LEARNING OBJECTIVES
• Describe the need for mismatched unrelated donors (MMUDs) for hematopoietic cell transplant (HCT) when
matched donors are not available for patients with hematologic malignancies
• Illustrate the impact of different human leukocyte antigen locus mismatch on HCT outcomes
• Defne novel strategies used to improve HCT outcomes in the MMUD setting
Introduction
MMUDs
Allogeneic hematopoietic cell transplantation (HCT) offers
a potentially curative treatment for patients with most
hematologic malignancies.1 Historically, HCT from a human
leukocyte antigen (HLA) matched sibling donor (MSD) is
considered the gold standard with the lowest nonrelapse
mortality (NRM) and graft-vs-host disease (GVHD)-related
mortality and morbidity.2 Comparable transplant outcomes have been reported with matched unrelated donors
(MUDs), which became a widely acceptable alternative for
patients with no MSD.3,4 The likelihood of identifying an 8/8
MUD varies with ethnicity and race; for example, this could
be as high as 70% for Caucasians but falls to below 20%
for African American and other ethnic minorities and even
more challenging for mixed-race individuals.5 Other factors
affecting donor availability include typing issues, medical
deferral, and donor or donor center availability.5 With the
recent advancement and improvement in outcomes, alternative donors have been increasingly used in close to 36%
of cases where match donors are not available (Figure 1).
These include mismatched unrelated donors (MMUDs),
haploidentical donors (Haplo), and umbilical cord blood.
Matching at HLA loci plays an integral role in selecting
donors for HCT. When a matched donor (MSD or MUD) is
not available, alternative donors are sought such as haploidentical donors, MMUDs, and cord blood units.
74 | Hematology 2022 | ASH Education Program
Outcomes using standard (conventional) GVHD
prophylaxis for MMUD transplants
Large single-center and registry retrospective comparative
studies have shown worse outcome with higher degree of
mismatch in the setting of conventional GVHD prophylaxis, regardless of disease type or risk, HCT conditioning
intensity, or stem cell source (Table 1). Conventional GVHD
prophylaxis usually includes a calcineurin inhibitor (ie, tacrolimus or cyclosporine) combined with mycophenolate
mofetil, methotrexate, or sirolimus. In 1 study, mostly in
patients undergoing myeloablative conditioning (MAC) for
bone marrow (BM) HCT,6 difference in survival is less when
disease risk is higher, but this was not confrmed in other
studies when expanded to other conditioning intensities
and stem cell source. For example, patients with poor-risk
acute myeloid leukemia (AML) in complete remission 1 who
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/74/2021682/74arslan.pdf by guest on 09 December 2022
Shukaib Arslan and Monzr M. Al Malki
Table 1. Studies using con­ven­tional GVHD pro­phy­laxis for MMUD HCTs
Study
Population
Design
Comparison
Key find­ings
Gupta et al (2010)7
Poor-risk AML in CR1 under­went
HCT from MSD (n=226), 8/8 MUD
(n=254), and MMUD (n=104)
CIBMTR—ret­ro­spec­tive
NA
1. Similar LFS and OS for MSD and
8/8 MUD trans­plants but LFS and
OS were worse for MMUD HCT
Al Malki et al (2020)9
482 patients with
hema­to­logic malig­nan­cies
under­went MUD/MMUD HCT
with tacrolimus and sirolimus
com­bi­na­tion for GVHD ­
pro­phy­laxis
Single-insti­tu­tion
ret­ro­spec­tive study
NA
1. 5-year OS: 39% for MMUD vs
50.7% for MUD, P = .022
2. NRM: 33.5% for MMUD vs 21.7%
for MUD, P = .038
3. Severe GVHD: 20.8% for MMUD
vs 12.8% for MUD, P = .22
4. Infection 33.0% for MMUD vs
18.1% for MUD, P < .01
5. No improve­ment in rates of
relapse with MMUD HCT
Kornblit et al (2020)10
77 patients who under­went
RIC PBSC HCT from MMUD for
hema­to­logic malig­nan­cies (76
evaluable patients)
Multicenter phase 2 trial
Historical
1. Day 100 CI of grade II to IV
aGVHD was 36%
2. OS at 4 years=62%
3. NRM=18%
4. Relapse/pro­gres­sion=30%
CR1, complete remission 1; NA, not available.
under­went HCT from MSD (n=226), 8/8 MUD (n=254), and MMUD
(n=104) had sim­i­lar leu­ke­mia-free sur­vival (LFS) and over­all sur­
vival (OS) to MSD and 8/8 MUD trans­plants, but LFS and OS were
worse for MMUD HCT.7 Also, in patients with other dis­eases, such
as myelodysplastic syn­drome, sim­i­lar results were reported with
worse out­comes in MMUD HCT.8
A large sin­gle-cen­ter ret­ro­spec­tive study com­pared out­
comes of MUD and MMUD with tacrolimus and sirolimus as
GVHD pro­phy­laxis in 482 periph­eral blood stem cell (PBSC)
HCTs. With a long-term fol­low-up (median = 6.2 years), 5-year
OS was sig­
nif­i­
cantly worse in patients under­
go­
ing MMUD
HCT (39% vs 50.7%, P = .022), mostly due to higher risk of
NRM (33.5% vs 21.7%, P = .038), with more severe GVHD
(20.8% vs 12.8%, P = .22) and infec­tion (33.0% vs 18.1%, P < .01).
There was no improve­ment in rates of relapse with MMUD
HCT.9 A mul­ti­cen­ter phase 2 trial used a trip­let of cyclo­spor­
ine, mycophenolate mofetil, and sirolimus as GVHD pro­
phy­laxis for reduced inten­sity con­di­tion­ing (RIC) trans­plant
using MMUD for hema­to­logic malig­nan­cies.10 A com­bi­na­
tion of fludarabine and total body irradiation (2-3 Gy) was
used as con­di­tion­ing ther­apy. The study included 76 evaluable patients with a median age of 63 years. With a median
fol­low-up of 47 months (4-94 months), the cumu­la­tive inci­
dence of day 100 grade II to IV acute GVHD (aGVHD) was 36%.
The study reported a cumu­la­tive inci­dence of NRM (18%),
relapse/pro­gres­sion (30%), and OS (62%) after 4 years.
Mismatch type, num­ber, and loca­tion
Multiple stud­ies have attempted to cor­re­late HCT out­comes with
mis­match at cer­tain HLA loci (Table 2). In 1 study, out­comes of
1874 donor-recip­
i­
ent pairs under­
go­
ing myeloablative BM HCT
were ana­lyzed with respect to high-res­o­lu­tion typ­ing at HLA-A, -B,
-C, -DRB1, -DQ , and -DP.11 Survival was worse with higher num­ber
of mismatched loci regard­less of the mismatched locus loca­tion
Dr Prakash Singh Shekhawat
MMUD trans­plant PTCy | 75
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/74/2021682/74arslan.pdf by guest on 09 December 2022
Figure 1. CIBMTR-reported relative proportion of allogeneic HCTs in the United States by donor type.
Table 2. Landmark stud­ies for MMUD trans­plant (HCT)
Population
Study design
Comparison
Key find­ings
1874 patients in NMDP
who under­went HCT
for HMs and non­ma­lig­
nant dis­or­ders
Retrospective
1. Mismatches at loci HLA-A
(n=374), -B (n=477), -C
(n=749), or -DRB1 (n=311),
-DQ (n=415), and -DP
(n=1648)
2. Study included 108
patients with 8/8 match
1. Reduced OS after MMUD HCT with mis­
matches at loci HLA-A, -B, -C, or -DRB1
2. Survival was worse with increas­ing num­ber of
mis­matches.
3. Mismatches at loci HLA-A, -B, -C, or -DRB1 all­
showed more fre­quent aGVHD
4. The inci­dence of grade III/IV aGVHD was
much higher with mis­match at HLA-A as com­
pared to mis­match at HLA-B, -C, or -DRB1.
5. Mismatches at HLA-DQ and -DP had lit­tle
effect on HCT out­comes.
6. Class I HLA mis­matches with or with­out HLADQ mis­match had sim­i­lar out­comes.
Lee et al (2007)6
3857 patients with
HMs (early stage,
inter­me­di­ate stage,
and advanced/higher
stage) who under­
went MAC HCT
• 94% received BMSCs
• 78% received T-cell
replete grafts
• GVHD pro­phy­laxis
included calcineurin
inhib­i­tor
Retrospective
(CIBMTR)
1. 8/8 MUD (n=1840)
2. Single (n=985) or
mul­ti­ple (n=1032)
MMUD
1. Higher-risk dis­ease at the time of HCT had
a higher asso­ci­a­tion with poor sur­vival as
com­pared to increased num­ber of HLA mis­
matches.
2. No sta­tis­ti­cally sig­nif­i­cant dif­fer­ence in sur­
vival out­comes between sin­gle-anti­gen mis­
match vs allele mis­match on the same locus.
No sta­tis­ti­cally sig­nif­i­cant dif­fer­ence between
8/8 and 7/8 with respect to engraft­ment,
relapse, and chronic GVHD.
3. A sin­gle mis­match at HLA-A, -C, or -DRB1 was
asso­ci­ated with sig­nif­i­cantly reduced OS
com­pared to 8/8 matches.
4. Multiple mis­matches were asso­ci­ated with
poorer sur­vival out­comes.
5. Single mis­match at HLA-DQ was not asso­ci­
ated with any effect on sur­vival.
6. HLA-DP mis­match did not affect sur­vival but
was asso­ci­ated with increased risk of aGVHD.
Saber et al (2012)4
2223 adult patients
with AML who
under­went alloHCT
between 2002 and
2006
Retrospective
(CIBMTR)
1. MSD (n=624)
2. 8/8 MUD (n=1193)
3. 7/8 MUD (MMUD)
(n=406)
1. Significantly lower 100-day cumu­la­tive inci­
dence of grade II to IV aGVHD in MSD HCT
recip­i­ents than in 8/8 MUD and 7/8 MUD HCT
recip­i­ents (33%, 51%, and 53%, respec­tively;
P<.001).
2. 8/8 MUD HCT recip­i­ents had a sim­i­lar sur­vival
rate com­pared with MSD HCT recip­i­ents (RR,
1.03; P=.62).
3. 7/8 MUD HCT recip­i­ents had higher early
mor­tal­ity than MSD HCT recip­i­ents (RR, 1.40;
P<.001), but beyond 6 months after HCT, their
sur­vival rates were sim­i­lar (RR, 0.88; P=.30).
Pidala et al (2014)12
8003 patients with
HMs (early stage,
inter­me­di­ate stage,
and advanced/higher
stage) who under­
went MUD HCT with
MAC between 1999
and 2011
Retrospective
(CIBMTR)
1. 8/8 MUD (n=5449)
2. 7/8 MMUD (n=2071)
3. 6/8 MMUD (n=483)
1. 6/8-7/8 led to sig­nif­i­cantly increased risk
of grade II to IV and III to IV aGVHD, chronic
GVHD, TRM, and reduced OS as com­pared to
8/8 matched pairs.
2. For 8/8 HCTs, mis­match at both HLA-DQB1
and HLA-DPB1 loci led to an increased inci­
dence of aGVHD and mis­match at HLA-DPB1
resulted in reduced relapse. Both PR and
no-PR HLA-DPB1 allele mis­matches led to
sig­nif­i­cantly increased inci­dence of grade II
to IV and grade III to IV aGVHD and reduced
risk of relapse.
3. Significantly increased risk of TRM in no-PR
pairs as com­pared to matched pairs or pairs
with PR DPB1 allele mis­matches.
4. Adjusted OS for 8/8 matched HCT was best
among the HLA-DPB1 allele mismatched pairs
followed by fully matched HLA-DPB1, with
worst adjusted OS for non­per­mis­sive mis­
matches (P=.015).
5. No dif­fer­ence in out­comes for 1 or dou­ble
allele mis­matches at HLA-DPB1 for per­mis­sive
or non­per­mis­sive mis­matches.
76 | Hematology 2022 | ASH Education Program
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/74/2021682/74arslan.pdf by guest on 09 December 2022
Study
Flomenberg et al
(2004)11
Table 2. Landmark stud­ies for MMUD trans­plant (HCT) (Continued)
Population
Study design
Comparison
Key find­ings
310 patients
RIC HCT
Retrospective
1. 12/12 MUD
2. 11/12 MUD (DPB1mismatched)
3. 10/12 MUD (DPB1mismatched)
1. Better OS and relapse in 11/12 when com­
pared to 12/12 and bet­ter OS as com­pared
to 10/12.
2. Among the 11/12 pairs, when com­pared with
per­mis­sive DP mis­match (PR), non­per­mis­
sive DP mis­match (no-PR) is asso­ci­ated with
increased risk of grade II to IV aGVHD and
NRM.
3. When DP expres­sion was con­sid­ered,
using high DP expressors as donors for
low-expressor recip­i­ents was asso­ci­ated
with increased risk of mor­tal­ity. DP PR 11/12
mis­match was asso­ci­ated with higher OS
when both donor and patient were DP low
expressors when com­pared with other
com­bi­na­tions.
BMSC, bone mar­row stem cell; HM, hematological malignancies; NMDP, National Marrow Donor Program; PR, per­mis­sive match; no-PR, non­per­mis­
sive match; TRM, trans­plant-related mor­tal­ity.
(A, B, C, or DR) or level of test­ing (anti­gen vs allele).11 Although
mis­matches at loci HLA-A, -B, -C, or -DRB1 all­showed more fre­
quent aGVHD, the inci­
dence of grade III/IV GVHD was much
higher with mis­match at HLA-A (in this par­tic­u­lar study) com­pared
with mis­match at HLA-B, -C, or -DRB1. Mismatches at HLA-DQ and
-DP showed lit­tle effect on HCT out­comes. Another study showed
that for the selec­tion of an appro­pri­ate MMUD, pref­er­ence should
always be to match at high res­o­lu­tion with HLA-A, -B, -C, and -DR,
allowing for HLA-DQ and/or -DP mis­match.11
HLA-DPB1 mis­match.13 Adjusted OS for 8/8 matched HCT was best
among patients with per­mis­sive HLA-DPB1 allele mismatched pairs
followed by fully matched HLA-DPB1 with worst adjusted OS for
non­per­mis­sive mis­matches (P = .015). Same results were obtained
in a smaller cohort with RIC under­go­ing PBSC MUD HCT.14
In addi­tion, since risk of GVHD asso­ci­ated with HLA-DPB1 mismatching is influ­
enced by the HLA-DPB1 rs9277534 expres­
sion
marker, patients with higher HLA DPB1 expres­sion allele have a
high risk for GVHD when receiv­ing trans­plants from donors with
the low-expres­sion allele, regard­less of the con­di­tion­ing ­inten­sity.15
HLA-B leader match in HLA-B mis­match
CLINICAL CASE 1
A 65-year-old woman in complete remission 1 after induc­tion
ther­apy for FLT3+ acute mye­loid leu­ke­mia, with no suit­able sib­
ling donors, presented in con­sul­ta­tion. A MUD search revealed
four 8/8 or 10/10 HLA matched donors. Patient’s high-res­o­lu­tion
HLA typ­ing showed that HLA-DPB1 was homo­zy­gous at 04:01,
which are low expresser HLA-DPB1 alleles. Review of the HLA
typ­ing of the 3 poten­tial donors revealed one 30-year-old donor
with 11/12 match with per­mis­sive HLA-DPB1 at 02:01, which is a
low expressor allele that was acti­vated. Patient under­went HCT
with con­
di­
tion­
ing ther­
apy consisting of fludarabine and mel­
pha­lan, as well as GVHD pro­phy­laxis of tacrolimus and sirolimus.
Patient is at 2.5 months posttransplant, in con­tin­ued complete
remission (CR) and no evi­dence of GVHD.
Role of HLA-DPB1 mis­match
HLA-DPB1 mis­match was stud­ied in mul­ti­ple dif­fer­ent large stud­ies.
In a study by Pidala et al,12 8003 donor-recip­i­ent pairs under­went
MUD HCT with MAC between 1999 and 2011. Mismatch at A, B, C, or
DRB1 led to sig­nif­i­cantly increased risk of grade II to IV and III to IV
aGVHD, chronic GVHD, trans­plant-related mor­tal­ity, and reduced
OS com­pared with 8/8 matched pairs, confirming prior results.
GVHD was higher with no impact on OS/NRM in patients with HLADQB1 and/or HLA-DPB1 mis­match. However, mis­match at HLA-DPB1
resulted in reduced rate of relapse. Based on T-cell epi­tope matching algo­rithm, risk of NRM was higher in pairs with non­per­mis­sive
HLA-B has high poly­
mor­
phism at the pep­
tide bind­
ing site.
Mismatch at HLA-B leads to increased risk of GVHD and lower
dis­ease-free sur­vival.16 Leader pep­
tides have methi­
o­
nine (M)
or thre­
o­
nine (T) at posi­
tion 2 containing sequence dimor­
phisms in exon 1 of HLA-B.16 These leader pep­tides are bound
by HLA-E, which serves as a ligand for the nat­u­ral killer recep­
tors. The 2 HLA-B allo­types deter­mine the leader geno­type in
a per­son. T- and M-lead­ers can affect the T-cell and nat­u­ral killer–cell responses by chang­ing the expres­sion of HLA-E.16-19 These
changes could affect alloimmune responses. A large ret­ro­spec­
tive study reported out­comes of 33 982 patients who under­went
alloHCT from a MUD/MMUD between 1988 and 2016. The study
included 1457 trans­plants with mis­match at HLA-B and eval­u­
ated out­comes based on B-leader match/mis­match. Presence
of M-leader in recip­i­ents or donors increases the risk of grade III
to IV aGVHD in patients who under­went HCT from an HLA-B mismatched donor. GVHD risk is also increased with the pres­ence of
dif­fer­ent lead­ers at the mismatched HLA-B a
­ llo­type.16
In this set­ting, if avail­­able donors are mismatched at HLA-B,
the selec­tion of the most suit­able donor is based on the recip­i­
ent’s leader geno­type. For sin­gle-locus HLA-B mismatched hematopoietic cell transplants, 94.6% of the recip­i­ents had the MT or
TT leader geno­type. Patients with the MT leader geno­type had
2 leader-matched donor pos­si­bil­i­ties: a donor matched at the
M-leader allo­type and mismatched for the T-leader allo­type (TTM)
or a donor mismatched at the M-leader allo­type and matched
for the T-leader allo­type (MMT). Authors reported that TTM trans­
plants had a higher risk of severe aGVHD com­pared with MMT
Dr Prakash Singh Shekhawat
MMUD trans­plant PTCy | 77
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/74/2021682/74arslan.pdf by guest on 09 December 2022
Study
Malki et al (2020)14
(odds ratio, 1.99; 95% con­fi­dence inter­val [CI], 1.10-3.59; P=.022),
pre­fer­ring MMT over TTM to improve HCT out­comes.16 The donor
choices for recip­i­ents with the TT leader geno­type include leader
matched (TTT) and leader mismatched (TMT), and TTT pairing is
pre­ferred over TMT for improved GVHD out­comes.16
HLA-C mis­match and role of per­mis­sive C mis­match
CLINICAL CASE 2
A 69-year-old Med­i­ter­ra­nean woman in complete remission 2
after induc­tion ther­apy for NPM1-pos­i­tive AML with hypomethylating agents with venetoclax was seen in con­sul­ta­tion. HLA
typ­ing for patient and 2 sib­lings did not reveal MSD. No MUD
was found in the reg­is­try. She had mul­ti­ple high titers (Mean
Fluorescent Intensity >5000-10 000) and donor-spe­cific antibodies (DSAs) against HLA-A, -B, and -DRB1 loci in her chil­
dren. Unrelated donor search revealed a young 7/8 donor with
HLA-C mis­match that was per­mis­sive, for which the patient
had no DSAs. Patient under­went PBSC MMUD HCT with HLA-C
mismatched donor using GVHD pro­phy­laxis of tacrolimus and
sirolimus. The patient devel­oped grade II skin GVHD, which
responded to ste­roid. Day +100 BM biopsy spec­i­men showed
minimal residual disease CR and full donor chi­me­rism. Patient is
at >6 months post-HCT in con­tin­ued remis­sion and no evi­dence
of GVHD.
Donor selec­tion for patients with high-risk HLA antibodies
Antibodies to HLA can be detected at ele­
vated fre­
quen­
cies
in patients with prior sen­si­tiz­ing events such as preg­nancy or
trans­fu­sion,23 but HLA antibodies are also some­times pres­ent in
oth­er­wise healthy indi­vid­u­als in the absence of known sen­si­tiz­
ing events. Although the impact of humoral sen­si­ti­za­tion in HCT
out­
comes remains poorly under­
stood com­
pared with the sig­
nif­i­
cance of DSAs devel­
op­
ing in solid organ trans­
plan­
ta­
tion,24
recent stud­ies of HLA antibodies pre- and early post-HCT have
been shown to exert a neg­a­tive impact on engraft­ment and graft
func­tion.25,26 To avoid graft fail­ure or graft dys­func­tion, screen­ing
patients for HLA antibodies when under­go­ing workup prior to HLA
78 | Hematology 2022 | ASH Education Program
Novel approaches to improve out­comes in MMUD HCT
The role of posttransplant cyclo­
phos­
pha­
mide (PTCy) in haploidentical trans­plant is well established.30 Alloreactive T cells
are sen­si­tive to the kill­ing effect of cyclo­phos­pha­mide, while
mem­ory and reg­u­la­tory T cells, with expres­sion of alde­hyde
dehy­dro­ge­nase, are resis­tant, mak­ing PTCy a highly effec­tive
ther­apy to pre­vent GVHD while pre­serv­ing T-cell reg­u­la­tory and
mem­ory func­tion.31-34 A phase 2 pro­spec­tive mul­ti­cen­ter trial
was conducted to study out­comes of MMUD HCT using PTCy.35
All patients (n=80) received fresh BM grafts on day 0, PTCy on
days 3 and 4 after HCT, and sirolimus with mycophenolate starting on day 5. The study included 48% of the patients who were
racial or eth­nic minor­i­ties, and 39% of the patients received a
BM graft with more than 1 HLA allele mis­match (4-6/8). One-year
OS for the whole group was 76%, as well as 72% and 79% for
MAC and RIC, respec­tively. Degree of HLA mis­match did not
affect the OS (75% for 7/8 and 77% for 4-6/8) in this small cohort,
which was later con­firmed in a lon­ger fol­low-up pre­sen­ta­tion.
The inci­dence of day 100 grade II to IV aGVHD, III to IV aGVHD,
and 1-year chronic GVHD was 43%, 18%, and 36% for patients
receiv­ing MAC and 33%, 0%, and 18% for patients who received
RIC, respec­tively.35 These results were com­pa­ra­ble to the out­
come of a his­tor­i­cal con­trol under­go­ing Haplo HCT and reported
to the Center for International Blood, and Marrow Transplant
Research (CIBMTR) around the same time.35 Similar ret­ro­spec­
tive but timely com­par­i­son was also published by the Euro­pean
Society for Blood and Marrow Transplantation36 in patients with
AML in CR with patients under­go­ing MMUD HCT (n=155) com­
pared with Haplo BM (n=647) and Haplo PB (n=949) in the set­ting
of PTCy-based GVHD pro­phy­laxis. Haplo BM and Haplo PB had
a higher NRM com­pared with MMUD (haz­ard ratio, 2.28; 95% CI,
1.23-4.24; P<.01 and haz­ard ratio, 2.65; 95% CI, 1.46-4.81; P<.01,
respec­tively), with lower LFS and OS. All raised the need for a
ran­dom­ized study com­par­ing the 2 graft sources in a pro­spec­
tive and sys­temic man­ner.
A con­cur­rently conducted, sin­gle-cen­ter, pro­spec­tive study
of 38 patients who under­went PBSC MMUD HCT (Figure 2) with
a sim­i­lar design showed sim­i­lar out­comes.29 Median num­ber of
HLA mis­matches was 2 (range, 1-4) of 12. Racial and eth­nic minor­
i­ties made up 61% of the study pop­u­la­tion. The 1-year OS and
GVHD-free/relapse-free sur­vival was 87% and 68%, respec­tively.
The inci­dences of grade II to IV and III to IV aGVHD at day +100
and chronic GVHD at 1 year were 50%, 18%, and 48%, respec­
tively.29 Combined together, these stud­ies showed that MMUD
HCT with PTCy is fea­si­ble and well tol­er­ated with prom­is­ing
results using either BM or PBSCs. A large National Marrow Donor
Program-spon­sored study (ACCESS) is ongo­ing to con­firm these
results using PBSC in adults and BM in chil­dren (NCT04904588).
The Euro­pean Society for Blood and Marrow Transplantation
conducted a ret­ro­spec­tive anal­y­sis on 272 patients with AML who
under­went 9/10 HLA matched HCT with GVHD pro­phy­laxis consisting of PTCy-based (n = 93) or antithymocyte glob­u­lin–based
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/74/2021682/74arslan.pdf by guest on 09 December 2022
For HCTs involv­ing mis­match at HLA-C, the immu­no­ge­nic­ity is
affected by the level of expres­sion of mismatched HLA-C in the
patient.20 A large ret­ro­spec­tive study com­pared out­comes of 1933
unre­
lated donor-recip­
i­
ent pairs who under­
went PBSC alloHCT
for hema­to­logic malig­nan­cies. This study included 189 pairs mismatched at the HLA-C anti­gen. HLA-C mis­match at high res­o­lu­tion
was asso­ci­ated with sig­nif­i­cantly lower leu­ke­mia-free sur­vival (rel­
a­tive risk [RR], 1.36; 95% CI, 1.13-1.64; P = .0010) and sig­nif­i­cantly
increased risk for mor­tal­ity (RR, 1.41; 95% CI, 1.16-1.70; P = .0005),
trans­plant-related mor­tal­ity (RR, 1.61; 95% CI, 1.25-2.08; P = .0002),
and grade III to IV aGVHD (RR, 1.98; 95% CI, 1.50-2.62; P < .0001).21
Single-allele mis­match at HLA-C, C*03:03/C*03:04, was very fre­
quent in a reg­
is­
try study of patients with Cau­
ca­
sian ances­
try
(68.7%).22 When com­pared with 8/8 matched HCTs, 7/8 allele
level match with C*03:03/C*03:04 mis­match was asso­ci­ated with
com­pa­ra­ble out­comes such as OS, NRM, dis­ease-free sur­vival, and
risk of aGVHD.22 Due to this find­ing, C*03:03/C*03:04 mis­match is
referred to as a per­mis­sive mis­match and is pre­ferred at least in
the set­ting of myeloablative BM HCT.
mismatched (related or unre­lated) HCT has become a stan­dard
prac­tice.23 Patients with clin­i­cally sig­nif­i­cant DSAs may undergo
DSA desen­si­ti­za­tion to mit­i­gate risk and facil­i­tate engraft­ment in
most cases but not all­.27 With recent advances in GVHD pro­phy­
laxis and improve­ment in out­come of MMUD,27-29 avoiding the mismatched loci with DSAs and mismatching to a non-DSA locus has
become one of the advan­tages of this donor type.28
(n=179) reg­i­mens. HLA mis­match involved class I in 74% and class
II in 26%; half of the patients received MAC and other half received
RIC. Use of PTCy was asso­ci­ated with a lower inci­dence of severe
aGVHD and higher LFS and GVHD-free/relapse-free sur­vival.37
Abatacept (ABA) is a CTLA-4-IgG1 fusion pro­tein that binds
to CD80/CD86 on the sur­face of anti­gen presenting cells and
blocks the CD28-medi­ated costimulation.38 A sin­gle-arm fea­
si­bil­ity study of ABA as pro­phy­laxis of GVHD was conducted
enroll­ing 10 patients with hema­to­logic malig­nan­cies using
unre­lated donors.39 Four patients under­went HCT from 8/8 HLA
matched donors, and 6 patients received 7/8 HLA matched
donor grafts. Two patients received a BM graft, and 8 patients
received PBSCs. Preparative reg­i­men consisted of MAC in 7 and
RIC in 3 patients. GVHD pro­phy­laxis was abatacept, cyclo­spor­
ine, and meth­o­trex­ate.39 This helped estab­lish proof of con­cept
and pre­pared for a larger phase 2 study eval­u­at­ing the use of
ABA in com­bi­na­tion with calcineurin inhib­i­tor (CNI) and meth­
o­
trex­
ate in HLA matched unre­
lated donors for hema­
to­
logic
malig­nan­cies.40 Thirty-eight patients in this study reported to
have received MMUD, mostly MAC, with 48% receiv­ing PBSCs.
This was reported in com­par­i­son to a his­tor­i­cal cohort from
CIBMTR with or with­out antithymocyte glob­u­lin. Abatacept in
com­bi­na­tion with CNI and meth­o­trex­ate (MTX) and in com­par­i­
son to a his­tor­i­cal cohort was shown to be effec­tive in decreas­
ing day 100 grade III to IV and day 180 severe aGVHD-free
sur­vival with a min­i­mal effect on the inci­dence and sever­ity of
chronic GVHD.40 A large mul­ti­cen­ter pro­spec­tive study is ongo­
ing to con­firm these results and to test if extended expo­sure
to ABA would affect inci­dence and sever­ity of chronic GVHD
(NCT04380740).
A recent abstract, presented at an Amer­
i­
can Society of
Hematology meet­
ing in 2021, com­
pared the real-world out­
come of 7/8 HLA MMUD HCTs for hema­
to­
logic malig­
nan­
cies
reported to CIBMTR between 2011 and 2018 using either CNI +
MTX with (n=54) or with­out ABA (n=162).41 The OS at day +180
for the cohort receiv­ing ABA was 98% com­pared with 75% for
CNI + MTX alone (P=.0028). In an explor­atory anal­y­sis focused
on short-term end point (OS at 180 days), out­come of patients
under­go­ing MMUD HCT with ABA with CNI + MTX was com­pa­ra­
ble to patients under­go­ing PTCy.41 Further long-term anal­y­sis will
be needed to define the effect of this novel GVHD pro­phy­laxis
reg­i­men on chronic GVHD and ulti­mately on OS and the effi­cacy
of this reg­i­men in a higher degree of mis­match (<7/8).
Other novel GVHD pro­phy­laxis approaches in the MMUD HCT
set­ting, used recently, with a short-term fol­low-up include graft
manip­u­la­tion, such as CD34+ cells with add back of mem­
ory
CD45RA+ T cells, α/β T-cell recep­tor deple­tion, and CD34+ cell
deple­tion with add back of a conventional T cells and regulatory
T cells approach. Those prom­is­ing approaches are still under
inves­ti­ga­tion (Table 3).
Dr Prakash Singh Shekhawat
MMUD trans­plant PTCy | 79
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/74/2021682/74arslan.pdf by guest on 09 December 2022
Figure 2. Study schema. (A) RIC regimen. Fludarabine was administered at the daily dose of 25 mg/m2 from days −7 to −3 before HCT.
Melphalan was given on day −2 at 140 mg/m2 or 100 mg/m2 for patients ≥60 years old. (B) MAC regimen consisted of daily fludarabine
at 30 mg/m2, from days −7 to −5 before HCT. Total body irradiation was administered in 8 fractions of 150 cGy, 2 times a day, from
days −4 to −1, for a total of 1200 cGy. Graft source was PBSCs for both strata. ANC, absolute neutrophils count; G-CSF, granulocyte
colony stimulating factor; TBI, total body irradiation.
Table 3. Studies using PTCy or abatacept for GVHD pro­phy­laxis for MMUD HCT
Population
Design
Comparison
Key find­ings
Shaw et al (2021)35
80 patients who
under­went RIC/MAC
BMSC MMUD HCT with
PTCY/MMF/sirolimus-based
GVHD pro­phy­laxis
Multicenter, pro­spec­tive,
phase 2 trial
NA
1. The 1-year OS was 76% (72%
for MAC and 79% for RIC)
2. Degree of HLA mis­match did
not affect the OS (75% for
7/8 and 77% for 4-6/8).
3. The day +100 inci­dence of
grade II to IV and grade III to
IV aGVHD was 43% and 18%
for MAC and 33% and 0%
for RIC.
4. The 1-year inci­dence of
chronic GVHD MAC and RIC
was 36% and 18%,
respec­tively.
Al Malki et al (2021)29
38 patients who
under­went PBSC MMUD
HCT
Single-cen­ter, pro­spec­tive
study
Battipaglia et al (2022)36
Patients with AML in CR
under­go­ing MMUD HCT
(n=155) com­pared to Haplo
BM (n=647) and Haplo PB
(n=949) with PTCy-based
GVHD pro­phy­laxis
Retrospective study
Battipaglia et al (2019)37
272 patients with AML
under­went 9/10 HLA
matched HCT with GVHD
pro­phy­laxis consisting of
PTCy-based (n=93) or
ATG-based (n=179)
reg­i­mens.
HLA mis­match involved
class I in 74% and class II
in 26%.
Half of the patients
received MAC, and the
other half received RIC.
Retrospective study
Watkins et al (2021)40
38 received MMUD, mostly
MAC, with 48% receiv­ing
PBSCs and 52% received
BMSCs. GVHD pro­phy­laxis
with abatacept in com­bi­
na­tion with CNI and MTX
Phase 2
Kean et al (2021)41
7/8 HLA MMUD HCTs for
hema­to­logic malig­nan­cies
between 2011 and 2018
using either CNI + MTX
with (n=54) or with­out ABA
(n=162)
CIBMTR ret­ro­spec­tive
1. The 1-year OS was 87%.
2. The 1-year GRFS 68%.
3. The day +100 inci­dence of
grade II to IV and III to IV
aGVHD was 50% and 18%,
respec­tively.
4. The 1-year inci­dence of
chronic GVHD was 48%.
Haplo HCT
1. Haplo BMSC and Haplo PBSC
had a higher NRM com­
pared to MMUD (HR, 2.28;
95% CI, 1.23-4.24; P<.01 and
HR, 2.65; 95% CI, 1.46-4.81;
P<.01, respec­tively) with
lower LFS and OS.
1. Use of PTCy was ­
asso­ci­ated with lower
inci­dence of severe aGVHD
and higher LFS and GRFS.
Historical cohort
from CIBMTR with or
with­out ATG
1. The inci­dence of grade
II to IV aGVHD was 2.3%
(CNI/MTX plus abatacept,
inten­tion-to-treat pop­u­la­
tion), which com­pared favor­
ably with a nonrandomized
matched cohort of CNI/MTX
(30.2%, P<.001), and the
SGFS was bet­ter (97.7% vs
58.7%, P<.001).
1. The OS at day +180 for the
cohort receiv­ing ABA was
98% com­pared to 75% for
CNI + MTX alone (P=.0028).
2. In an explor­atory anal­y­
sis focused on short-term
end point (OS at 180 days),
out­come of patients under­
go­ing MMUD HCT with ABA
with CNI + MTX was com­pa­
ra­ble to patients under­go­ing
PTCy.
ATG, antithymocyte glob­u­lin; GRFS, GVHD free relapse-free sur­vival; HR, haz­ard ratio; MMF, mycophenolate mofetil; SGFS, severe aGVHD-free
­sur­vival.
80 | Hematology 2022 | ASH Education Program
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/74/2021682/74arslan.pdf by guest on 09 December 2022
Study
Conclusion
Recent advances in GVHD pro­phy­laxis and donor selec­tion have
reignited inter­est in MMUD as a fea­si­ble and suit­able option for
patients with­out a matched donor for life-sav­ing HCT. Careful
selec­tion of opti­mal MMUDs involv­ing high-res­o­lu­tion HLA ­typ­ing
and HLA anti­body mea­sure­ment to detect DSAs helps guide in
the selec­tion of HLA mis­matches to improve out­comes. Novel
GVHD pro­
phy­
laxis approaches such as PTCy, abatacept, and
graft manip­u­la­tion seem to improve GVHD out­comes, although
pro­spec­tive stud­ies are under way. There is con­stant need for
novel ther­a­pies to improve the out­comes using MMUDs.
Acknowledgments
Conflict-of-inter­est dis­clo­sure
Shukaib Arslan: no com­pet­ing finan­cial inter­ests to declare.
Monzr M. Al Malki: no com­pet­ing finan­cial inter­ests to declare.
Off-label drug use
Shukaib Arslan: cyclophosphamide is discussed.
Monzr M. Al Malki: cyclophosphamide is discussed.
Correspondence
Monzr M. Al Malki, Department of Hematology and HCT, City of
Hope National Medical Center, 1500 E. Duarte Rd, Duarte, CA
91010, USA; e-mail: malmalki@coh​­.org.
References
1.
Copelan EA. Hematopoietic stem-cell trans­
plan­
ta­
tion. N Engl J Med.
2006;354(17):1813-1826.
2. Beatty PG, Clift RA, Mickelson EM, et al. Marrow trans­plan­ta­tion from related
donors other than HLA-iden­ti­cal sib­lings. N Engl J Med. 1985;313(13):765771.
3. Hows JM, Passweg JR, Tichelli A, et al; IMUST Study Participating Centers;
Late Effects Working Party of the Euro­pean Group for Blood and Marrow
Transplantation (EBMT). Comparison of long-term out­comes after allo­ge­
neic hema­to­poi­etic stem cell trans­plan­ta­tion from matched sib­ling and
unre­lated donors. Bone Marrow Transplant. 2006;38(12):799-805.
4. Saber W, Opie S, Rizzo JD, Zhang M-J, Horowitz M-M, Schriber J. Outcomes
after matched unre­lated donor ver­sus iden­ti­cal sib­ling hema­to­poi­etic
cell trans­plan­ta­tion in adults with acute mye­log­e­nous leu­ke­mia. Blood.
2012;119(17):3908-3916.
5. Gragert L, Eapen M, Williams E, et al. HLA match like­li­hoods for hema­to­poi­
etic stem-cell grafts in the U.S. reg­is­try. N Engl J Med. 2014;371(4):339-348.
6. Lee SJ, Klein J, Haagenson M, et al. High-res­o­lu­tion donor-recip­i­ent HLA
matching con­trib­utes to the suc­cess of unre­lated donor mar­row trans­plan­
ta­tion. Blood. 2007;110(13):4576-4583.
7. Gupta V, Tallman MS, He W, et al. Comparable sur­vival after HLA-wellmatched unre­lated or matched sib­ling donor trans­plan­ta­tion for acute
mye­loid leu­ke­mia in first remis­sion with unfa­vor­able cyto­ge­net­ics at diag­
no­sis. Blood. 2010;116(11):1839-1848.
8. Saber W, Cutler CS, Nakamura R, et al. Impact of donor source on hema­
to­poi­etic cell trans­plan­ta­tion out­comes for patients with myelodysplastic
syn­dromes (MDS). Blood. 2013;122(11):1974-1982.
9. Al Malki MM, Gendzekhadze K, Yang D, et al. Long-term out­come of allo­ge­
neic hema­to­poi­etic stem cell trans­plan­ta­tion from unre­lated donor using
tacrolimus/sirolimus-based GvHD pro­phy­laxis: impact of HLA mis­match.
Transplantation. 2020;104(5):1070-1080.
10. Kornblit B, Storer BE, Andersen NS, et al. Sirolimus with CSP and MMF
as GVHD pro­phy­laxis for allo­ge­neic trans­plan­ta­tion with HLA anti­genmismatched donors. Blood. 2020;136(13):1499-1506.
11. Flomenberg N, Baxter-Lowe LA, Confer D, et al. Impact of HLA class I and
class II high-res­o­lu­tion matching on out­comes of unre­lated donor bone
Dr Prakash Singh Shekhawat
MMUD trans­plant PTCy | 81
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/74/2021682/74arslan.pdf by guest on 09 December 2022
The authors thank Steven Devine, MD, for dis­cus­sion, feed­back,
and crit­i­cal read­ing of the man­u­script and Sally Mokhtari, PhD,
for reviewing the man­u­script.
mar­row trans­plan­ta­tion: HLA-C mismatching is asso­ci­ated with a strong
adverse effect on trans­plan­ta­tion out­come. Blood. 2004;104(7):1923-1930.
12. Pidala J, Lee SJ, Ahn KW, et al. Nonpermissive HLA-DPB1 mis­
match
increases mor­tal­ity after myeloablative unre­lated allo­ge­neic hema­to­poi­
etic cell trans­plan­ta­tion. Blood. 2014;124(16):2596-2606.
13. Fleischhauer K, Shaw BE, Gooley T, et al; International Histocompatibility Working Group in Hematopoietic Cell Transplantation. Effect of
T-cell-epi­
tope matching at HLA-DPB1 in recip­
i­
ents of unre­
lated-donor
haemopoietic-cell trans­plan­ta­tion: a ret­ro­spec­tive study. Lancet Oncol.
2012;13(4):366-374.
14. Malki MMA, Gendzekhadze K, Stiller T, et al. Protective effect of HLADPB1 mis­
match remains valid in reduced-inten­
sity con­
di­
tion­
ing unre­
lated donor hema­to­poi­etic cell trans­plan­ta­tion. Bone Marrow Transplant.
2020;55(2):409-418.
15. Petersdorf EW, Malkki M, O’hUigin C, et al. High HLA-DP expres­sion and
graft-ver­sus-host dis­ease. N Engl J Med. 2015;373(7):599-609.
16. Petersdorf EW, Carrington M, O’hUigin C, et al; International Histocompatibility Working Group in Hematopoietic Cell Transplantation. Role of HLA-B
exon 1 in graft-ver­sus-host dis­ease after unre­lated haemopoietic cell trans­
plan­ta­tion: a ret­ro­spec­tive cohort study. Lancet Haematol. 2020;7(1):e50e60.
17. Horowitz A, Djaoud Z, Nemat-Gorgani N, et al. Class I HLA hap­lo­types
form two schools that edu­cate NK cells in dif­fer­ent ways. Sci Immunol.
2016;1(3):eaag1672.
18. Braud V, Jones EY, McMichael A. The human major his­to­com­pat­i­bil­ity com­
plex class Ib mol­e­cule HLA-E binds sig­nal sequence-derived pep­tides with
pri­mary anchor res­i­dues at posi­tions 2 and 9. Eur J Immunol. 1997;27(5):11641169.
19. Romagnani C, Pietra G, Falco M, et al. Identification of HLA-E-spe­cific alloreactive T lym­pho­cytes: a cell sub­set that undergoes pref­er­en­tial expan­
sion in mixed lym­pho­cyte cul­ture and dis­plays a broad cyto­lytic activ­ity
against allo­ge­neic cells. Proc Natl Acad Sci U S A. 2002;99(17):11328-11333.
20. Petersdorf EW, Gooley TA, Malkki M, et al; International Histocompatibility
Working Group in Hematopoietic Cell Transplantation. HLA-C expres­sion
lev­els define per­mis­si­ble mis­matches in hema­to­poi­etic cell trans­plan­ta­
tion. Blood. 2014;124(26):3996-4003.
21. Woolfrey A, Klein JP, Haagenson M, et al. HLA-C anti­gen mis­match is asso­
ci­ated with worse out­come in unre­lated donor periph­eral blood stem cell
trans­plan­ta­tion. Biol Blood Marrow Transplant. 2011;17(6):885-892.
22. Fernandez-Viña MA, Wang T, Lee SJ, et al. Identification of a per­mis­si­
ble HLA mis­match in hema­to­poi­etic stem cell trans­plan­ta­tion. Blood.
2014;123(8):1270-1278.
23. Powers A, Stowell CP, Dzik WH, Saidman SL, Lee H, Makar RS. Testing only
donors with a prior his­tory of preg­nancy or trans­fu­sion is a log­i­cal and
cost-effec­tive trans­fu­sion-related acute lung injury pre­ven­tion strat­egy.
Transfusion. 2008;48(12):2549-2558.
24. Terasaki P, Lachmann N, Cai J. Summary of the effect of de novo HLA antibodies on chronic kid­ney graft fail­ure. Clin Transpl. 2006:455-462.
25. Focosi D, Zucca A, Scatena F. The role of anti-HLA antibodies in hema­
to­poi­etic stem cell trans­plan­ta­tion. Biol Blood Marrow Transplant.
2011;17(11):1585-1588.
26. Ciurea SO, Thall PF, Wang X, et al. Donor-spe­cific anti-HLA Abs and graft
fail­ure in matched unre­lated donor hema­to­poi­etic stem cell trans­plan­ta­
tion. Blood. 2011;118(22):5957-5964.
27. Ciurea SO, Al Malki MM, Kongtim P, et al. Treatment of allosensitized patients
receiv­ing allo­ge­neic trans­plan­ta­tion. Blood Adv. 2021;5(20):4031-4043.
28. Ciurea SO, Cao K, Fernandez-Vina M, et al. The Euro­pean Society for Blood
and Marrow Transplantation (EBMT) con­sen­sus guide­lines for the detec­
tion and treat­ment of donor-spe­cific anti-HLA antibodies (DSA) in haploidentical hema­to­poi­etic cell trans­plan­ta­tion. Bone Marrow Transplant.
2018;53(5):521-534.
29. Al Malki MM, Tsai N-C, Palmer J, et al. Posttransplant cyclo­phos­pha­mide as
GVHD pro­phy­laxis for periph­eral blood stem cell HLA-mismatched unre­
lated donor trans­plant. Blood Adv. 2021;5(12):2650-2659.
30. Luznik L, O’Donnell PV, Symons HJ, et al. HLA-haploidentical bone mar­
row trans­plan­ta­tion for hema­to­logic malig­nan­cies using nonmyeloablative
con­di­tion­ing and high-dose, posttransplantation cyclo­phos­pha­mide. Biol
Blood Marrow Transplant. 2008;14(6):641-650.
31. Mayumi H, Tokunaga K. Cyclophosphamide-induced chi­mera-type tol­er­
ance to allo­grafts: an over­view of drug-induced immu­no­log­i­cal tol­er­ance.
Fukuoka Igaku Zasshi. 1990;81(1):20-39.
32. Mayumi H, Umesue M, Nomoto K. Cyclophosphamide-induced immu­no­
log­i­cal tol­er­ance: an over­view. Immunobiology. 1996;195(2):129-139.
82 | Hematology 2022 | ASH Education Program
38. Malmström V, Catrina AI, Klareskog L. The immunopathogenesis of sero­
pos­i­tive rheu­ma­toid arthri­tis: from trig­ger­ing to targeting. Nat Rev Immunol. 2017;17(1):60-75.
39. Koura DT, Horan JT, Langston AA, et al. In vivo T cell costimulation block­ade
with abatacept for acute graft-ver­sus-host dis­ease pre­ven­tion: a first-indis­ease trial. Biol Blood Marrow Transplant. 2013;19(11):1638-1649.
40. Watkins B, Qayed M, McCracken C, et al. Phase II trial of costimulation
block­
ade with abatacept for pre­
ven­
tion of acute GVHD. J Clin Oncol.
2021;39(17):1865-1877.
41. Kean LS, Burns LJ, Kou TD, et al. Improved over­all sur­vival of patients
treated with abatacept in com­bi­na­tion with a calcineurin inhib­i­tor and
meth­o­trex­ate fol­low­ing 7/8 HLA-matched unre­lated allo­ge­neic hema­to­
poi­etic stem cell trans­plan­ta­tion: anal­y­sis of the Center for International
Blood and Marrow Transplant Research Database. Blood. 2021;138(suppl 1):
3912.
© 2022 by The Amer­i­can Society of Hematology
DOI 10.1182/hema­tol­ogy.2022000398
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/74/2021682/74arslan.pdf by guest on 09 December 2022
33. Kanakry CG, Ganguly S, Zahurak M, et al. Aldehyde dehy­dro­ge­nase expres­
sion drives human reg­u­la­tory T cell resis­tance to posttransplantation cyclo­
phos­pha­mide. Sci Transl Med. 2013;5(211):211ra157.
34. Wachsmuth LP, Patterson MT, Eckhaus MA, Venzon DJ, Gress RE, Kanakry
CG. Post-trans­plan­ta­tion cyclo­phos­pha­mide pre­vents graft-ver­sus-host
dis­ease by induc­ing alloreactive T cell dys­func­tion and sup­pres­sion. J Clin
Invest. 2019;129(6):2357-2373.
35. Shaw BE, Jimenez-Jimenez AM, Burns LJ, et al. National Marrow Donor
Program-spon­sored mul­ti­cen­ter, phase II trial of HLA-mismatched unre­
lated donor bone mar­
row trans­
plan­
ta­
tion using post-trans­
plant cyclo­
phos­pha­mide. J Clin Oncol. 2021;39(18):1971-1982.
36. Battipaglia G, Galimard J-E, Labopin M, et al. Post-trans­plant cyclo­phos­
pha­mide in one-anti­gen mismatched unre­lated donor trans­plan­ta­tion ver­
sus haploidentical trans­plan­ta­tion in acute mye­loid leu­ke­mia: a study from
the Acute Leukemia Working Party of the EBMT. Bone Marrow Transplant.
2022;57(4):562-571.
37. Battipaglia G, Labopin M, Kröger N, et al. Posttransplant cyclo­phos­pha­
mide vs antithymocyte glob­u­lin in HLA-mismatched unre­lated donor trans­
plan­ta­tion. Blood. 2019;134(11):892-899.
ARE ALTERNATIVE DONORS NOW MAINSTREAM IN ALLOGENEIC TRANSPLANT ?
Hematology 2022—what is complete HLA
match in 2022?
Center for International Blood and Marrow Transplant Research, National Marrow Donor Program/Be The Match, Minneapolis, MN
Allogeneic hematopoietic cell transplantation (alloHCT) often represents the only curative treatment for various malignant and nonmalignant disorders. Initially, the only suitable donors were considered human leukocyte antigen (HLA)–
matched or partially matched relatives. The founding of international unrelated donor and umbilical cord blood registries
expanded unrelated donor options and access for patients. In the absence of a matched sibling donor (MSD) with 13% to
51% availability, the current consensus recommends use of a matched unrelated donor (MUD) at HLA-A, B, C, and DRB1
with consideration of matching at HLA-DPB1 and -DQB1. MUD donor availability (donor willing and available to donate)
ranges from 29% to 78% with African American patients on the lower end and white non-Hispanic patients with the
highest likelihood of a match. Recent studies comparing donor to no-donor treatment options in malignant disease consistently point to substantially better outcomes following alloHCT. In the absence of an MSD or MUD, alternative donor
choices turn to haploidentical related (Haplo), mismatched unrelated donor (MMUD), and umbilical cord blood (UCB).
Novel strategies for alloHCT, including the use of posttransplant cyclophosphamide-based graft vs host disease prophylaxis, have expanded the safety and effectiveness of transplant procedures across HLA barriers using Haplo and MMUD.
The less restrictive matching requirements for UCB transplant are well documented and allow for transplant across multiply mismatched HLA alleles. When all donor options are considered, nearly all patients have an available donor. Here
we discuss the likelihood of donor availability, complete HLA match by available donor type, and current controversies
warranting future research.
LEARNING OBJECTIVES
• Describe the degree of HLA matching and association of matching/mismatching on outcomes for available
allogeneic donor options
• Understand the likelihood of availability and hierarchy of donor type selection for allogeneic hematopoietic cell
transplantation
Introduction
CLINICAL CASE
A 60-year-old Hispanic man presents with de novo acute
myeloid leukemia (M2;inv 3/t(3;3)). Induction and consolidation therapy achieved complete remission with no
measurable residual disease based on next-generation
sequencing. Family studies identifed 1 sibling as a full
human leukocyte antigen (HLA) match. However, they
have a history of malignant melanoma and were deemed
ineligible to donate. One additional (65-year-old) sibling
is haploidentical to the patient and is willing and eligible to donate. An unrelated donor search identifes multiple young (<30 years old) well-matched (8/8-considering
HLA-A, B, C, and DRB1) and single mismatched (7/8) unrelated donors and multiple cord blood units available.
Allogeneic hematopoietic cell transplantation (alloHCT)
often represents the only curative treatment for various
malignant and nonmalignant disorders. Over the approximately 50 years since the frst alloHCTs were attempted and
demonstrated engraftment and curative potential, the defnition of optimal matching has continued to evolve. Initially,
the only suitable donors were considered HLA matched or
partially matched relatives. The founding of the National Marrow Donor Program (NMDP), Anthony Nolan Registry, and
other international donor registries expanded the potential
to identify suitable donors outside of the patient’s immediate family. Unrelated donor registries have grown to include
over 40 million volunteer donors and cord blood registries
to over 800 000 banked umbilical cord blood units worldwide (World Marrow Donor Association stats). HLA testing
Dr Prakash Singh Shekhawat
HLA and donor selection | 83
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/83/2021755/83spellman.pdf by guest on 09 December 2022
Stephen R. Spellman
Table 1. HLA typ­ing tech­niques and resulting res­o­lu­tion
Approach
Interpretation
Resolution
Application
Results
Serology
Cellular assay based on
com­ple­ment fix­a­tion by
HLA-spe­cific antibodies
Cell death—yes/no
Low
Family screen­ing
Null allele
con­fir­ma­tion
A2, A24
Sequence-spe­cific
prim­ers (PCR-SSP)
HLA sequence-spe­cific
PCR prim­ers
Amplification—yes/no
Low to high,
depen­dent on
DNA sequence
cov­er­age
Family screen­ing
Verification typ­ing
Low—A*02:XX, A*24:XX or
A*02AB, A*24:BC
High—A*0201g, A*24:02g
Sequence-spe­cific
oli­go­nu­cle­o­tide
probes (PCR-SSOP)
HLA sequence-spe­cific
oli­go­nu­cle­o­tide probes
that bind to poly­mor­phic
sequences of ampli­fied
DNA
Probe bind­ing—yes/no
Low to high
depending on
DNA sequence
cov­er­age
Family screen­ing
Verification typ­ing
Low—A*02:XX, A*24:XX or
A*02AB, A*24:BC
High—A*02:01G, A*24:02G
Sanger sequencebased typ­ing (SBT)
HLA amplicon sequenc­ing
using base ter­mi­na­tion
Base pair reads and
con­sen­sus align­ment
High to allele level
depending on
cov­er­age
All
High—A*02:01G, A*24:02G
Allele—A*02:01:01:03,
A*24:02:01:01
Next-gen­er­a­tion
sequenc­ing (NGS)
Multiple plat­forms, based
on mas­sive par­al­lel
sequenc­ing reac­tions
Base pair call­ing and
con­sen­sus align­ment
High to allele-level
depending on
cov­er­age
All
High—A*02:01G, A*24:02G
Allele—A*02:01:01:03,
A*24:02:01:01
tech­nol­o­gies and matching strat­e­gies have evolved sig­nif­i­cantly
as well, mak­ing the char­ac­ter­iza­tion of match between unre­lated
indi­vid­u­als more rou­tine.1 The advent of DNA-based typ­ing tech­
nol­o­gies and enhanced data­bases of well-char­ac­ter­ized HLA allele
sequences have increased the pre­ci­sion and accu­racy of typ­ing,
lead­ing to improved matching (Table 1). Yet, the ques­tion remains,
what is the opti­mal match? The answer is not straight­for­ward,
as the opti­mal HLA match for a given patient will vary based on
the donor options avail­­able at the time of need and clin­i­cal fac­
tors (eg, tim­ing of trans­plant, patient size/weight, and prior HLA
sen­si­ti­za­tion).
HLA typ­ing for search and match deter­mi­na­tion
To facil­
i­
tate HLA match assess­
ment, patients and poten­
tial
donors (related and unre­lated) should be high-res­o­lu­tion HLA
typed using DNA-based meth­ods for HLA-A, B, C, DRB1, and
DPB1.2 Extended typ­ing can include HLA-DRB3/4/5, DQB1, DQA1,
and DPA1, espe­cially in the case of highly sen­si­tized patients to
avoid mis­matches that may be a tar­get of anti-HLA antibodies
(eg, donor-spe­cific antibodies) to min­i­mize the risk of graft fail­
ure.3 Traditional DNA-based tech­niques for HLA typ­ing focused
on the anti­
gen rec­
og­
ni­
tion domain (ARD).2 Technological
advances now make the rou­tine sequenc­ing of the entire HLA
gene fea­si­ble. A recent study found that mismatching out­side
of the ARD was asso­ci­ated with increased risk of acute graftvs-host dis­ease (GVHD) grades II to IV but with no increased
risk of trans­plant-related mor­tal­ity or decreased over­all sur­vival
(OS),4 suggesting that matching for alle­lic var­i­a­tion out­side the
ARD has lim­ited impact in the matched unre­lated donor (MUD)
set­ting.
HLA typ­ing should be ver­i­fied via con­fir­ma­tory test­ing prior
to final­iz­ing donor selec­tion for alloHCT through an Amer­i­can
Society of Histocompatibility and Immunogenetics or Euro­pean
Federation for Immunogenetics accredited lab­o­ra­tory, pref­er­
a­bly with guid­ance from an accredited his­to­com­pat­i­bil­ity and
immu­no­ge­net­ics lab­o­ra­tory direc­tor to assist the clin­i­cal team
with inter­pre­ta­tion of the typ­ing results and match assess­ment
between the pro­spec­tive donor(s) and patient.
84 | Hematology 2022 | ASH Education Program
What are the avail­­able donor options and like­li­hood
of avail­abil­ity?
The con­ven­tional opti­mal donor choice is an HLA-matched fam­
ily mem­ber, most often a matched sib­ling. However, the rate of
matched sib­ling donor (MSD) avail­abil­ity can vary sub­stan­tially
based on eth­nic­ity and age of the patient with rates rang­ing
from 13% to 51%.5 For those patients with­out a matched sib­ling,
the first-choice alter­na­tive donor option is an MUD. This rec­om­
men­da­tion is supported by the find­ings of the annual Center for
International Blood and Marrow Transplant Research (CIBMTR)
Center Specific Analysis. The 2021 report included >25 000 unre­
lated and related alloHCTs performed between 2017 and 2019
in the United States with fol­low-up through 1 year after alloHCT.
A com­par­i­son of alter­na­tive donor types can be assessed using
the odds ratio of OS at 1 year com­pared with MSD as the base­line
with an odds ratio (OR) for MUD of 0.87 (95% con­fi­dence inter­
val [CI], 0.78-0.97, P=.010), haploidentical related (Haplo) OR of
0.76 (0.68-0.84, P<.001), mismatched unre­lated donor (MMUD)
(7/8) OR of 0.69 (0.59-0.81, P<.001), and umbil­i­cal cord blood
(UCB) (mul­ti­ple UCB ≥4/6) OR of 0.50 (0.39-0.64, P<.001) (see
the visual abstract). This very large, con­tem­po­rary, mul­ti­cen­ter
anal­y­sis of real-world data clearly dem­on­strates the hier­ar­chy for
donor selec­tion pri­or­i­tiz­ing matched sib­lings, MUDs, and then
mismatched graft sources.
Despite the large inter­na­tional pool of >40 mil­lion vol­un­teer
unre­lated donors, MUD avail­abil­ity varies and is most sub­stan­
tially affected by the eth­nic back­ground of the patient. MUD
donor avail­abil­ity (donor will­ing and avail­­able to donate) ranges
from 29% to 78%, with Afri­can Amer­i­can patients on the lower
end and white non-His­panic patients with the highest like­li­hood
of a match (NMDP inter­nal anal­y­sis). In the absence of an MUD,
alter­na­tive donor choices turn to Haplo, MMUD, and UCB. Novel
strat­e­gies for alloHCT, includ­ing the use of posttransplant cyclo­
phos­pha­mide-based graft-vs-host dis­ease pro­phy­laxis, have
expanded the safety and effec­tive­ness of trans­plant pro­ce­dures
across HLA bar­ri­ers using Haplo6 and MMUD.7,8 The less restric­tive
matching require­ments for UCB trans­plant are well documented
and allow for trans­plant across mul­ti­ply mismatched HLA alleles.9
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/83/2021755/83spellman.pdf by guest on 09 December 2022
Methodology
Matched sib­ling donors
As noted above, an MSD is the bench­mark against which all­alter­
na­tive donor options are eval­u­ated. Donor choice is gen­er­ally lim­
ited in the MSD set­ting and with donor age tightly cor­re­lated to
patient age. Given that mul­ti­ple stud­ies sug­gest that lower donor
age is asso­ci­ated with bet­ter OS in both the matched unre­lated
and haploidentical alloHCT set­ting,12-14 ques­tions have been raised
about whether an older sib­ling donor is still the best option when
mul­ti­ple donor choices are avail­­able. A CIBMTR study of older
MSD (≥50 years old) vs youn­ger MUD (<50 years old) alloHCT for
lym­phoma and leu­ke­mia reported supe­rior out­comes in MSD.15
A recent anal­y­sis com­par­ing older MSD (≥50 years old) to youn­
ger MUD (≤35 years old) alloHCT for myelodysplastic syn­dromes
found that MUD was asso­ci­ated with lower relapse and improved
dis­
ease-free sur­
vival (DFS), suggesting that in cer­
tain cir­
cum­
stances (eg, high risk of relapse), MUD may be advan­ta­geous.16
In addi­tion, avail­abil­ity of youn­ger Haplo or MMUD may war­rant
inves­ti­ga­tion as an alter­na­tive to an older matched sib­ling as out­
comes for all­donor types con­tinue to improve.
One poten­tial con­cern in older donors is the higher rates of
clonal hema­to­poi­e­sis of indeterminant poten­tial (CHIP) asso­ci­
ated with increas­ing age.17 Prior stud­ies dif­fer on the find­ings
regard­ing risk/ben­e­fit of CHIP in the alloHCT set­ting. Frick et al18
reported increased risk of chronic GVHD fol­low­ing alloHCT with
a related donor har­bor­ing CHIP, while Gib­son et al19 reported
poten­tial ben­e­fits of DNMT3A var­i­ants in faster engraft­ment and
bet­ter dis­ease con­trol. While these data raise poten­tial con­cerns
about the suit­abil­ity of older donors for opti­mal trans­plant out­
comes, there are scarce data to sup­port pri­or­i­tiz­ing alter­na­tive
donors over an avail­­able MSD at pres­ent.
HLA matching in MUD
These data were reviewed and sum­
ma­
rized in the NMDP/
CIBMTR donor selec­tion guide­lines published in Blood in 2019
and pro­vided rec­om­men­da­tions for opti­mal MUD matching in
the con­text of con­ven­tional calcineurin inhib­i­tor-based GVHD
pro­phy­laxis.20 The opti­mal level of match in the MUD set­ting
was established in the sem­i­nal pub­li­ca­tion by Lee et al.21 This
piv­
otal study dem­
on­
strated that matching for HLA-A, B, C,
and DRB1 was asso­ci­ated with supe­rior OS and lower rates of
acute GVHD21 (Figure 1). This level of match is often referred to
as 8/8 matching. HLA-DQB1 match was not asso­ci­ated with any
out­comes as an iso­lated mis­match and did not con­fer any addi­
tional risk when paired with HLA-A, B, C, or DRB1. Inclusion of
DQB1 in matching is often referred to as 10/10 matching. Most
(>95%) cases matched at 8/8 will also be matched at 10/10, so
gen­er­ally bring along a match at HLA-DQB1. These find­ings were
rep­li­cated in sub­se­quent ana­ly­ses in a more con­tem­po­rary pop­
u­la­tion,22 as well as by graft type,23 non­ma­lig­nant dis­ease,24 and
inter­na­tional cohorts.25
HLA-DPB1 mismatching in the con­text of an 8/8 or a 10/10
match is asso­ci­ated with increased risk of nonrelapse mor­tal­
ity driven by higher GVHD and lower risk of relapse with no
asso­ci­a­tion with OS.21,22 A model pioneered by Fleischhauer
et al26 based on the immu­no­ge­nic­ity of HLA-DPB1 in the con­text
of 3 T-cell epi­tope (TCE) groups pro­vi­des a meth­od­ol­ogy for
miti­gat­ing the risks asso­ci­ated with HLA-DPB1 mismatching. In
mul­ti­ple ana­ly­ses, the selec­tion of HLA-DPB1 TCE per­mis­sive
mis­matches was asso­ci­ated with improved OS and reduced
risk of GVHD, fur­ther adding to the strat­egy for selecting an
opti­mal 8/8 MUD.22,26,27 In addi­tion, a recent anal­y­sis found that
Figure 1. Survival of patients with early-stage disease depending on degree of HLA matching (8/8, 7/8, and 6/8) for HLA-A, B, C,
and DRB1. (Figure courtesy of Lee et al.21)
Dr Prakash Singh Shekhawat
HLA and donor selec­tion | 85
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/83/2021755/83spellman.pdf by guest on 09 December 2022
When all­donor options are con­sid­ered, there is a high like­li­hood
that all­patients will have an avail­­able donor.10
How should alter­na­tive donor options be pri­or­i­tized? Varying
reports sug­gest sit­u­a­tions where 1 donor type may be pre­ferred
over oth­ers (eg, Haplo vs dou­ble UCB in adult patients receiv­
ing alloHCT for lym­phoma or acute leu­ke­mia based on BMT CTN
1101,11 youn­ger alter­na­tive donors in the con­text of older MSD,
MMUD vs Haplo), but much con­tro­versy remains. The BMT CTN
1702 pro­to­col, “Clinical Transplant-Related Long-Term Outcomes
of Alternative Donor Allogeneic Transplantation (CTRL-ALT-D)”
(NCT03904134), is inves­
ti­
gat­
ing the out­
comes of alloHCT fol­
low­ing bio­log­i­cal assign­ment to an MUD or alter­na­tive donor
(Haplo, UCB, or MMUD) based on like­li­hood of MUD avail­abil­ity.
This trial reflects cur­rent con­sen­sus that an MUD, if avail­­able in
a timely man­ner, is the pre­ferred alter­na­tive to an MSD and will
shed fur­ther light on alter­na­tive donor choice and out­comes.
This includes an eval­u­a­tion of the urgency of need and poten­tial
to obtain var­i­ous donor types.
Table 2. HLA matching in matched unre­lated donor allo­ge­neic hema­to­poi­etic cell trans­plan­ta­tion
Reference
Consensus guide­lines
Dehn et al20 (2019)
1. 8/8 match at ARD level for HLA-A, B, C, DRB1
2. Younger donor age
3. Match or per­mis­sive HLA-DPB1 TCE
4. Minimize mis­matches at HLA-DRB3/4/5 and DQB1
5. Avoid DSA tar­gets includ­ing DQA1 and DPA1
New research published, since release of guide­lines
Reference
Study type
Comparison
Key find­ings
Mayor et al (2021),
ultra-high res­o­lu­tion
N=5140 10/10 MUD, first
alloHCT for ALL, AML, or
MDS. Myeloablative or RIC
Observational—mul­ti­cen­ter
12/12 Ultra-high res­o­lu­tion
vs ≤11/12
UHR matching asso­ci­ated
with lower risk of acute
GVHD II-IV, no asso­ci­a­tions
with sur­vival out­comes
Arrieta-Bolaños et al28
(2022), TCE core
N=5140 10/10 MUD, first
alloHCT for ALL, AML, or
MDS. Myeloablative or RIC
Observational—mul­ti­cen­ter
10/10 MUD HLA-DPB1
per­mis­sive core alleles vs
noncore alleles vs
non­per­mis­sive
10/10 MUD HLA-DPB1 TCE3
per­mis­sive core asso­ci­
ated with reduced risk of
acute GVHD II-IV and TRM
com­pared to non­per­mis­sive
mis­matches
ALL, acute lym­pho­blas­tic leu­ke­mia; AML, acute mye­loid leu­ke­mia; DSA, donor-spe­cific anti-HLA antibodies; MDS, myelodysplastic syn­dromes;
RIC, reduced inten­sity con­di­tion­ing; TRM, transplant-related mortality; UHR, ultrahigh resolution; 8/8, high-res­o­lu­tion match at HLA-A, B, C,
and DRB1; 10/10, high-res­o­lu­tion match at HLA-A, B, C, DRB1, and DQB1; 12/12, HLA match at HLA-A, B, C, DRB1, DQB1, and DPB1.
the TCE3 group could be fur­ther cat­e­go­rized into core and
noncore alleles to reduce the risk of GVHD and transplantrelated mortality com­pared with non­per­mis­sive mis­matches.28
The poten­tial to find an MUD with an HLA-DPB1 allele match
may be lim­ited, but the poten­tial to iden­tify a per­mis­sive mis­
match is highly likely and imput­able to sup­port iden­ti­fi­ca­tion
of donors with miss­ing DPB1 typ­ing29,30 (Table 2).
The 1 non-HLA donor char­ac­ter­is­tic that con­sis­tently asso­ci­
ates with improved sur­vival after alloHCT is youn­ger donor age
and is pri­or­i­tized above extended matching at HLA-DPB1 and
other loci in the cur­rent donor selec­tion guide­lines.12,13,20 International donor reg­is­tries now pri­or­i­tize recruit­ment of youn­ger
donors (≤40 years old), and NMDP has established donor age
as a met­ric of prod­uct qual­ity empha­siz­ing the use of donors
≤35 years old.
Further stud­
ies are required to val­
i­
date the effects of
extended HLA matching and youn­ger donor age in MUD in the
con­text of novel GVHD pro­phy­laxis strat­e­gies, such as posttransplant cyclo­phos­pha­mide (ptCy), abatacept, and advanced graft
engi­neer­ing.
HLA matching in mismatched graft sources
Related haploidentical donors
The selec­tion and pri­or­i­ti­za­tion of Haplo donors is lim­ited by
fam­
ily size. The Euro­
pean Blood and Marrow Transplantation
group recently published con­sen­sus rec­om­men­da­tions for
donor selec­tion in Haplo alloHCT focused mainly on non-HLA
donor char­ac­ter­is­tics.31 Novel HLA match and mis­match asso­
ci­
a­
tions with Haplo alloHCT out­
comes were published after
these rec­om­men­da­tions. Sol­o­mon et al. reported an asso­ci­a­tion
between mismatching at HLA class II loci and decreased relapse
and improved OS in a sin­gle-cen­ter study of T-replete Haplo
alloHCT using ptCy.32 A recent large mul­ti­cen­ter study reported
HLA locus-spe­cific asso­ci­a­tions with var­i­ous out­comes and
pro­posed a model for Haplo donor selec­tion to opti­mize DFS.
86 | Hematology 2022 | ASH Education Program
The study eval­u­ated over­all degree of high-res­o­lu­tion matching,
impact of indi­vid­ual loci, HLA-DPB1 TCE matching, and matching
for the con­served exon 1 leader sequence of HLA-B (B leader)
pre­vi­ously found to asso­ci­ate with acute GVHD in the MMUD set­
ting.33 The model rec­om­mends pri­or­i­tiz­ing HLA-B leader match,
HLA-DPB1 TCE non­per­mis­sive mis­match (oppo­site of MUD), DRB1
mis­match, and DQB1 match for opti­mal DFS and was cod­i­fied in
an online tool (Table 3).34
Mismatched unre­lated donors
Recommendations for the pri­
or­
i­
ti­
za­
tion of HLA match and
mis­match in MMUD are pre­dom­i­nately based on past expe­ri­
ence in the set­ting of calcineurin inhib­i­tor–based GVHD pro­
phy­laxis strat­e­gies.20 While mul­ti­ple stud­ies inves­ti­gated the
poten­tial to apply algo­rithms that pri­or­i­tize mis­matches based
on struc­tural sim­i­lar­ity or pep­tide bind­ing affin­ity, most have
failed val­i­da­tion in large mul­ti­cen­ter stud­ies.35 FernandezViña et al36 described a per­mis­sive mis­match at HLA-C where
the alleles (C*03:03 and C*03:04) only dif­fer out­side the anti­
gen rec­og­ni­tion domain, but the exten­sion of these find­ings to
other mis­matches is lim­ited. Hurley et al37 reported that mis­
matches at HLA-A, B, C, and DRB1 lim­ited to the host-vs-graft
direc­tion were well tol­er­ated and sim­i­lar to a full match but
only appli­ca­ble in the set­ting of mismatching at a homo­zy­gous
locus in the patient. HLA-B leader matches asso­ci­ated with
lower risk of acute GVHD.33 Prioritizing HLA-DPB1 TCE per­mis­
sive mis­matches and min­i­miz­ing over­all level of mis­match at
extended HLA class II loci can also con­trib­ute to bet­ter out­
comes, includ­ing OS in the MMUD set­ting26,38 (Table 4).
The recently described asso­
ci­
a­
tions between HLA match
and mis­match in the mul­ti­ply mismatched Haplo ptCy set­ting
war­rant inves­ti­ga­tion in MMUD.34 The expanded donor pool in
the MMUD set­ting can sup­port addi­tional selec­tion cri­te­ria (eg,
donor-spe­cific anti-HLA antibodies in highly sen­si­tized patients
and poten­tial to pri­or­i­tize favor­able match/mis­matches).
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/83/2021755/83spellman.pdf by guest on 09 December 2022
Population
4
Table 3. HLA matching in haploidentical related donor allo­ge­neic hema­to­poi­etic cell trans­plan­ta­tion
Reference
Consensus guide­lines
Ciurea et al31 (2020)
T-replete Haplo using ptCy-based GVHD pro­phy­laxis:
1. Avoid DSA (MFI <1000)
2. Younger donor over older
3. Male donor for male recip­i­ent
4. Sibling or off­spring over par­ent
5. Father pre­ferred over mother
6. ABO match over minor over major mis­match
New research published, since release of guide­lines
Population
Trial/study design
Comparison(s)
Key find­ings
N=208 con­sec­u­tive first
alloHCT for hema­to­logic
malig­nancy receiv­ing T-cell
replete Haplo with ptCy
GVHD pro­phy­laxis
Observational—sin­gle
cen­ter
1. Total num­ber of
mis­matches
2. Individual locus
match/mis­match
effects on out­comes
1. HLA-DRB1 and HLA-DPB1
non­per­mis­sive mis­match
inde­pen­dently asso­ci­ated with
improved OS
2. HLA-A mis­match increased
chronic GVHD
Fuchs et al34 (2022)
N=1484, first alloHCT for
ALL, AML, or MDS, adult
patients receiv­ing T-cell
replete Haplo with ptCybased GVHD pro­phy­laxis
Observational—mul­ti­cen­ter
1. Total num­ber of
mis­matches
2. Individual locus
match/mis­match
effects on out­comes
1. No asso­ci­a­tion with num­ber of
mis­matches
2. HLA-B leader match improves
OS and DFS; HLA-DRB1 GVH
vec­tor mis­match decreases
relapse and improves DFS;
HLA-DPB1 non­per­mis­sive
mis­match improves OS and
DFS; HLA-C match decreases
chronic GVHD
MFI, mean fluorescence intensity.
Table 4. HLA matching in mismatched unre­lated donor allo­ge­neic hema­to­poi­etic cell trans­plan­ta­tion
Reference
Consensus guide­lines
Dehn et al (2019)
1. 7/8 match at ARD level for HLA-A, B, C, DRB1
2. Younger donor age
3. Match or per­mis­sive HLA-DPB1 TCE
4. Minimize mis­matches at HLA-DRB3/4/5 and DQB1
5. Avoid DSA tar­gets includ­ing DQA1 and DPA1
20
New research published, since release of guide­lines
Study
Population
Trial/study design
Comparison
Results
Petersdorf et al33 (2020)
N=33 982 total cohort with
N=1457 HLA-B mismatched,
first alloHCT for mul­ti­ple
dis­eases, CNI-based GVHD
pro­phy­laxis
Observational study
Exon 1 B leader match vs
mis­match in 9/10 MMUD
B leader match asso­ci­ated with
decreased acute GVHD, no
asso­ci­a­tion between match
sta­tus and sur­vival
Shaw et al7 (2021)
N=80 7/8-4/8 MMUD
alloHCT for hema­to­logic
malig­nan­cies, ptCy-based
GVHD pro­phy­laxis with
bone mar­row grafts
Prospective trial
7/8 matched vs <7/8
matched
No dif­fer­ences in any out­comes
CNI, calcineurin inhib­i­tor; 7/8, sin­gle high-res­o­lu­tion mis­match at HLA-A, B, C, and DRB1; 9/10, sin­gle high-res­o­lu­tion mis­match at HLA-A, B, C, DRB1,
and DQB1.
Umbilical cord blood
Traditionally, UCB match is con­sid­ered at HLA-A, B at the anti­gen
level and HLA-DRB1 at high res­o­lu­tion to achieve a min­i­mum of
a 4/6 match to the patient. In the sin­gle UCB alloHCT set­ting,
matching at HLA-C and at high res­o­lu­tion for HLA-A, B, C, and
DRB1 is asso­ci­ated with improved out­comes.39 This is not always
prac­ti­cal in the set­ting of UCB HCT where there is a stron­ger
empha­sis placed on achiev­ing a min­i­mum total nucle­ated and/or
CD34 cell dose, lim­it­ing UCB choice. Expert rec­om­men­da­tions
from the NMDP/CIBMTR20 and Amer­i­can Society of Transplant
and Cellular Therapy9 sug­gest selecting the best match con­sid­
er­ing HLA-A, B, C, and DRB1 and avoiding UCB <4/8 matched
at high res­o­lu­tion or <4/6 using the tra­di­tional match stan­dards
(Table 5).
Dr Prakash Singh Shekhawat
HLA and donor selec­tion | 87
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/83/2021755/83spellman.pdf by guest on 09 December 2022
Study
Sol­o­mon et al32 (2020)
Table 5. HLA matching in umbil­i­cal cord blood allo­ge­neic hema­to­poi­etic cell trans­plan­ta­tion
Reference
Consensus guide­lines
Dehn et al20 (2019)
≥4/6 HLA-A and -B anti­gen match, HLA-DRB1 high-res­o­lu­tion match, ≥4/8 high-res­o­lu­tion match meet­ing min­i­mum
cryopreserved cell dose for sin­gle unit alloHCT of TNC ≥2.5×107/kg and CD34+ cells ≥1.5×105/kg and dou­ble-unit
alloHCT TNC ≥1.5×107/kg and CD34+ cells ≥1.0×105/kg per unit
Politikos et al9 (2020)
≥4/6 HLA-A and -B anti­gen match, HLA-DRB1 high-res­o­lu­tion match, ≥4/8 high-res­o­lu­tion match meet­ing min­i­mum
cryopreserved cell dose for sin­gle-unit alloHCT of TNC ≥3.0×107/kg and CD34+ cells ≥2.0×105/kg and dou­ble-unit
alloHCT CD34+ cells ≥1.5×105/kg per unit
TNC, total nucle­ated cell count.
As trans­plant con­tin­ues to evolve and growth inev­i­ta­bly occurs
in the HLA mismatched set­ting thanks to safer, more effec­tive
approaches to min­i­mize acute and chronic GVHD with­out risking increased relapse and infec­tions, we will gain more insights
into the matching/mismatching to opti­mize out­comes. Prioritization of alter­na­tive donor types (Haplo, MMUD, and UCB) in the
absence of an MSD or MUD remains con­tro­ver­sial, with vary­ing
pro­grams pre­fer­ring one approach over oth­ers. Until these data
mature, the opti­mal match will vary based on the donor type
avail­­able to a given patient and the choice of donors within that
selec­tion pool.
CLINICAL CASE (Con­t in­u ed)
The patient was enrolled on BMT CTN 1702 and bio­
log­
i­
cally
assigned to the MUD donor arm due to their good search prog­no­
sis score.40 The MUD donor options included mul­ti­ple 8/8 donors
either matched or per­mis­sively mismatched for HLA-DPB1 TCE
and matched for HLA-DQB1. The trans­plant team selected the
youn­
gest 10/10 HLA-DPB1 TCE per­
mis­
sive mismatched donor
avail­­able and suc­cess­fully proceeded to trans­plant.
Acknowledgments
The author thanks Bronwen Shaw, Jeffery Auletta, and Steven Devine
for helpful discussions and critical review of the manuscript.
Conflict-of-inter­est dis­clo­sure
Stephen R. Spellman: no com­pet­ing finan­cial inter­ests to declare.
Off-label drug use
Stephen R. Spellman: nothing to disclose.
Correspondence
Stephen R. Spellman, Center for International Blood and Marrow Transplant Research, National Marrow Donor Program/Be
The Match, 500 5th N Street, Minneapolis, MN 55401-1206;
e-mail:sspellma@nmdp​­.org.
References
1.
Spellman S, Setterholm M, Maiers M, et al. Advances in the selec­tion of
HLA-com­pat­i­ble donors: refine­ments in HLA typ­ing and matching over the
first 20 years of the National Marrow Donor Program Registry. Biol Blood
Marrow Transplant. 2008;14(9 suppl):37-44.
88 | Hematology 2022 | ASH Education Program
2. Nunes E, Heslop H, Fernandez-Vina M, et al. Definitions of his­to­com­pat­i­bil­
ity typ­ing terms. Blood. 2011;118(23):e180-e183.
3. Ciurea SO, Cao K, Fernandez-Vina M, et al. The Euro­
pean Society for
Blood and Marrow Transplantation (EBMT) con­sen­sus guide­lines for the
detec­tion and treat­ment of donor-spe­cific anti-HLA antibodies (DSA) in
haploidentical hema­to­poi­etic cell trans­plan­ta­tion. Bone Marrow Transplant. 2018;53(5):521-534.
4. Mayor NP, Wang T, Lee SJ, et al. Impact of pre­vi­ously unrec­og­nized HLA
mis­matches using ultra­high res­o­lu­tion typ­ing in unre­lated donor hema­to­
poi­etic cell trans­plan­ta­tion. J Clin Oncol. 2021;39(21):2397-2409.
5. Besse K, Maiers M, Confer D, Albrecht M. On mod­el­ing human leu­ko­cyte
anti­gen-iden­ti­cal sib­ling match prob­a­bil­ity for allo­ge­neic hema­to­poi­etic
cell trans­plan­ta­tion: esti­mat­ing the need for an unre­lated donor source.
Biol Blood Marrow Transplant. 2016;22(3):410-417.
6. Luznik L, O’Donnell PV, Symons HJ, et al. HLA-haploidentical bone mar­
row trans­plan­ta­tion for hema­to­logic malig­nan­cies using nonmyeloablative
con­di­tion­ing and high-dose, posttransplantation cyclo­phos­pha­mide. Biol
Blood Marrow Transplant. 2008;14(6):641-650.
7. Shaw BE, Jimenez-Jimenez AM, Burns LJ, et al. National Marrow Donor
Program-spon­sored mul­ti­cen­ter, phase II trial of HLA-mismatched unre­
lated donor bone mar­
row trans­
plan­
ta­
tion using post-trans­
plant cyclo­
phos­pha­mide. J Clin Oncol. 2021;39(18):1971-1982.
8. Al Malki MM, Tsai N-C, Palmer J, et al. Posttransplant cyclo­phos­pha­mide as
GVHD pro­phy­laxis for periph­eral blood stem cell HLA-mismatched unre­
lated donor trans­plant. Blood Adv. 2021;5(12):2650-2659.
9. Politikos I, Davis E, Nhaissi M, et al; Amer­i­can Society for Transplantation
and Cellular Therapy Cord Blood Special Interest Group. Guidelines for cord
blood unit selec­tion. Biol Blood Marrow Transplant. 2020;26(12):2190-2196.
10. Gragert L, Eapen M, Williams E, et al. HLA match like­li­hoods for hema­to­poi­
etic stem-cell grafts in the U.S. reg­is­try. N Engl J Med. 2014;371(4):339-348.
11. Fuchs EJ, O’Donnell PV, Eapen M, et al. Double unre­lated umbil­i­cal cord
blood vs HLA-haploidentical bone mar­row trans­plan­ta­tion: the BMT CTN
1101 trial. Blood. 2021;137(3):420-428.
12. Kollman C, Spellman SR, Zhang M-J, et al. The effect of donor char­ac­ter­
is­tics on sur­vival after unre­lated donor trans­plan­ta­tion for hema­to­logic
malig­nancy. Blood. 2016;127(2):260-267.
13. Shaw BE, Logan BR, Spellman SR, et al. Development of an unre­lated donor
selec­tion score pre­dic­tive of sur­vival after HCT: donor age mat­ters most.
Biol Blood Marrow Transplant. 2018;24(5):1049-1056.
14. DeZern AE, Franklin C, Tsai H-L, et al. Relationship of donor age and rela­
tion­ship to out­comes of haploidentical trans­plan­ta­tion with posttransplant
cyclo­phos­pha­mide. Blood Adv. 2021;5(5):1360-1368.
15. Alousi AM, Le-Rademacher J, Saliba RM, et al. Who is the bet­ter donor
for older hema­to­poi­etic trans­plant recip­i­ents: an older-aged sib­ling or a
young, matched unre­lated vol­un­teer? Blood. 2013;121(13):2567-2573.
16. Guru Murthy GS, Kim S, Hu Z-H, et al. Relapse and dis­ease-free sur­vival in
patients with myelodysplastic syn­drome under­go­ing allo­ge­neic hema­to­
poi­etic cell trans­plan­ta­tion using older matched sib­ling donors vs youn­ger
matched unre­lated donors. JAMA Oncol. 2022;8(3):404-411.
17. Jaiswal S, Natarajan P, Silver AJ, et al. Clonal hema­to­poi­e­sis and risk of ath­
ero­scle­rotic car­dio­vas­cu­lar dis­ease. N Engl J Med. 2017;377(2):111-121.
18. Frick M, Chan W, Arends CM, et al. Role of donor clonal hema­to­poi­e­
sis in allo­ge­neic hema­to­poi­etic stem-cell trans­plan­ta­tion. J Clin Oncol.
2019;37(5):375-385.
19. Gib­son CJ, Kim HT, Zhao L, et al. Donor clonal hema­to­poi­e­sis and recip­i­ent
out­comes after trans­plan­ta­tion. J Clin Oncol. 2022;40(2):189-201.
20. Dehn J, Spellman S, Hurley CK, et al. Selection of unre­lated donors and
cord blood units for hema­to­poi­etic cell trans­plan­ta­tion: guide­lines from
the NMDP/CIBMTR. Blood. 2019;134(12):924-934.
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/83/2021755/83spellman.pdf by guest on 09 December 2022
Changing land­scape in HCT and impli­ca­tions for opti­mal
alter­na­tive donor selec­tion
32. Sol­o­mon SR, Aubrey MT, Zhang X, et al. Class II HLA mis­match improves
out­comes fol­low­ing haploidentical trans­plan­ta­tion with posttransplant
cyclo­phos­pha­mide. Blood Adv. 2020;4(20):5311-5321.
33. Petersdorf EW, Carrington M, O’hUigin C, et al; International Histocompatibility Working Group in Hematopoietic Cell Transplantation. Role of
HLA-B exon 1 in graft-ver­sus-host dis­ease after unre­lated haemopoietic cell
trans­plan­ta­tion: a ret­ro­spec­tive cohort study. Lancet Haematol. 2020;7(1):
e50-e60.
34. Fuchs EJ, McCurdy SR, Sol­o­mon SR, et al. HLA informs risk pre­dic­tions after
haploidentical stem cell trans­plan­ta­tion with posttransplantation cyclo­
phos­pha­mide. Blood. 2022;139(10):1452-1468.
35. Spellman S, Klein J, Haagenson M, et al. Scoring HLA class I mis­matches
by HistoCheck does not pre­dict clin­i­cal out­come in unre­lated hema­to­poi­
etic stem cell trans­plan­ta­tion. Biol Blood Marrow Transplant. 2012;18(5):
739-746.
36. Fernandez-Viña MA, Wang T, Lee SJ, et al. Identification of a per­mis­si­
ble HLA mis­match in hema­to­poi­etic stem cell trans­plan­ta­tion. Blood.
2014;123(8):1270-1278.
37. Hurley CK, Woolfrey A, Wang T, et al. The impact of HLA uni­di­rec­tional mis­
matches on the out­come of myeloablative hema­to­poi­etic stem cell trans­
plan­ta­tion with unre­lated donors. Blood. 2013;121(23):4800-4806.
38. Fernández-Viña MA, Klein JP, Haagenson M, et al. Multiple mis­
matches
at the low expres­sion HLA loci DP, DQ , and DRB3/4/5 asso­ci­ate with
adverse out­comes in hema­to­poi­etic stem cell trans­plan­ta­tion. Blood.
2013;121(22):4603-4610.
39. Eapen M, Klein JP, Ruggeri A, et al; Center for International Blood and Marrow Transplant Research, Netcord, Eurocord, and the Euro­pean Group for
Blood and Marrow Transplantation. Impact of allele-level HLA matching on
out­comes after myeloablative sin­gle unit umbil­i­cal cord blood trans­plan­ta­
tion for hema­to­logic malig­nancy. Blood. 2014;123(1):133-140.
40. Wadsworth K, Albrecht M, Fonstad R, Spellman S, Maiers M, Dehn J. Unrelated donor search prog­nos­tic score to sup­port early HLA con­sul­ta­tion and
clin­i­cal deci­sions. Bone Marrow Transplant. 2016;51(11):1476-1481.
© 2022 by The Amer­i­can Society of Hematology
DOI 10.1182/hema­tol­ogy.2022000326
Dr Prakash Singh Shekhawat
HLA and donor selec­tion | 89
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/83/2021755/83spellman.pdf by guest on 09 December 2022
21. Lee SJ, Klein J, Haagenson M, et al. High-res­o­lu­tion donor-recip­i­ent HLA
matching con­trib­utes to the suc­cess of unre­lated donor mar­row trans­plan­
ta­tion. Blood. 2007;110(13):4576-4583.
22. Pidala J, Lee SJ, Ahn KW, et al. Nonpermissive HLA-DPB1 mis­
match
increases mor­tal­ity after myeloablative unre­lated allo­ge­neic hema­to­poi­
etic cell trans­plan­ta­tion. Blood. 2014;124(16):2596-2606.
23. Woolfrey A, Klein JP, Haagenson M, et al. HLA-C anti­gen mis­match is asso­
ci­ated with worse out­come in unre­lated donor periph­eral blood stem cell
trans­plan­ta­tion. Biol Blood Marrow Transplant. 2011;17(6):885-892.
24. Horan J, Wang T, Haagenson M, et al. Evaluation of HLA matching in unre­
lated hema­to­poi­etic stem cell trans­plan­ta­tion for non­ma­lig­nant dis­or­ders.
Blood. 2012;120(14):2918-2924.
25. Fürst D, Müller C, Vucinic V, et al. High-res­o­lu­tion HLA matching in hema­
to­poi­etic stem cell trans­plan­ta­tion: a ret­ro­spec­tive col­lab­o­ra­tive anal­y­sis.
Blood. 2013;122(18):3220-3229.
26. Fleischhauer K, Shaw BE, Gooley T, et al; International Histocompatibility Working Group in Hematopoietic Cell Transplantation. Effect of
T-cell-epi­
tope matching at HLA-DPB1 in recip­
i­
ents of unre­
lated-donor
haemopoietic-cell trans­plan­ta­tion: a ret­ro­spec­tive study. Lancet Oncol.
2012;13(4):366-374.
27. Malki MMA, Gendzekhadze K, Stiller T, et al. Protective effect of HLADPB1 mis­
match remains valid in reduced-inten­
sity con­
di­
tion­
ing unre­
lated donor hema­to­poi­etic cell trans­plan­ta­tion. Bone Marrow Transplant.
2020;55(2):409-418.
28. Arrieta-Bolaños E, Crivello P, He M, et al. A core group of struc­tur­ally sim­i­lar
HLA-DPB1 alleles drives per­mis­sive­ness after hema­to­poi­etic cell trans­plan­
ta­tion. Blood. 2022;140(6):659-663.
29. Tram K, Stritesky G, Wadsworth K, Ng J, Anasetti C, Dehn J. Identification
of DPB1 per­mis­sive unre­lated donors is highly likely. Biol Blood Marrow
Transplant. 2017;23(1):81-86.
30. Sajulga R, Madbouly A, Fingerson S, et al. Predicting HLA-DPB1 per­mis­sive
prob­a­bil­i­ties through a DPB1 pre­dic­tion ser­vice towards the opti­mi­za­tion
of HCT donor selec­tion. Hum Immunol. 2021;82(12):903-911.
31. Ciurea SO, Al Malki MM, Kongtim P, et al. The Euro­pean Society for Blood
and Marrow Transplantation (EBMT) con­sen­sus rec­om­men­da­tions for
donor selec­tion in haploidentical hema­to­poi­etic cell trans­plan­ta­tion. Bone
Marrow Transplant. 2020;55(1):12-24.
AUTOIMMUNE HEMOLYTIC ANEMIAS
Cold AIHA and the best treatment strategies
Department of Pediatrics, Division of Hematology/Oncology, Baylor College of Medicine, Houston, TX
Cold-reactive autoimmune hemolytic anemia (AIHA) is rare among the hemolytic anemias. It results when 1 of a variety of processes causes the generation of immunoglobulin M (IgM) autoantibodies against endogenous erythrocytes,
resulting in complement activation and predominantly intravascular hemolysis. Cold AIHA is typically a primary lymphoproliferative disorder with marrow B-cell clones producing pathogenic IgM. More rarely, secondary cold AIHA (cAIHA)
can develop from malignancy, infection, or other autoimmune disorders. However, in children cAIHA is typically post
infection, mild, and self-limited. Symptoms include a sequelae of anemia, fatigue, and acrocyanosis. The severity of disease is variable and highly dependent on the thermal binding range of the autoantibody. In adults, treatment has most
commonly focused on reducing antibody production with rituximab-based regimens. The addition of cytotoxic agents
to rituximab improves response rates, but at the expense of tolerability. Recent insights into the cause of cold agglutinin
disease as a clonal disorder driven by complement form the basis of newer therapeutic options. While rituximab-based
regimens are still the mainstay of therapy, options have now expanded to include complement-directed treatments and
other B-cell-directed or plasma-cell-directed therapies.
LEARNING OBJECTIVES
• Understand the role of novel therapeutics in the treatment of primary cold agglutinin disease
• Compare the management of adult and pediatric cold autoimmune hemolytic anemia
CLINICAL CASE 1
A 65­year­old woman is admitted for anemia identified dur­
ing a workup for fatigue. Her hemoglobin level is 6.5 g/dL,
and she is sent to the hospital for admission. She reports
purple discoloration of her fingers during the winter. Her
exam demonstrates pallor, scleral icterus, and jaundice.
She has a positive polyspecific direct antiglobulin test
(DAT), negative immunoglobulin G (IgG) and positive C3d,
undetectable haptoglobin, lactate dehydrogenase (LDH)
of 620 U/L, unconjugated hyperbilirubinemia at 4.8 mol/L,
and an elevated absolute reticulocyte count (ARC) at
240 × 109/L. Her C3 and C4 are undetectable, and cold
agglutinin titer is 1024 at 4 °C. Bone marrow (BM) biopsy
does not identify malignancy but is consistent with cold
agglutinin­associated lymphoproliferative BM disease,
with nodular B­cell aggregates, the absence of paratra­
becular growth, and the fibrosis and lymphoplasmacyt­
oid cells seen in lymphoplasmacytic lymphoma. Serum
protein electrophoresis demonstrates monoclonal IgMκ.
The sample is negative for MYD88 L265P mutation. Flow
cytometry on her BM sample demonstrates a ratio of κ/λ­
positive B cells of 6. She is diagnosed with primary cold
90 | Hematology 2022 | ASH Education Program
agglutinin disease (CAD). She receives a warmed erythro­
cyte transfusion and is started on rituximab and benda­
mustine. As a bridge pending B­cell depletion, she starts
weekly sutimlimab infusions, which are spaced to every
2 weeks after the first 2 doses. Her hemolysis resolves and
she remains in remission at 3 months. Her sutimlimab is
discontinued without any further hemolytic anemia.
CLINICAL CASE 2
An 11­year­old boy is admitted for severe anemia and
hyperbilirubinemia. Ten days prior, he had nasal conges­
tion, rhinorrhea, cough, and low­grade fever. His symp­
toms progressed to include headache, dizziness, pallor,
worsening fatigue, dark urine, and yellowing of his eyes.
His pediatrician referred him to the emergency room. On
exam, he is tachycardic with a systolic ejection murmur.
He has diffuse pallor, jaundice, and scleral icterus, as well
as mild hepatosplenomegaly. His labs show a white blood
cell count of 344 × 109/L, a hemoglobin level of 5.4 g/dL, a
mean corpuscular volume of 104.3 fL, and a platelet count
of 344 × 109/L. He has an inappropriately low reticulocyte
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/90/2021751/90despotovic.pdf by guest on 09 December 2022
Jenny McDade Despotovic and Taylor Olmsted Kim
Table 1. Suggested diagnostic evaluation for CAIHA
Suggested adult CAIHA workup
Hematologic test­ing
Required for pedi­at­rics
CBC
Yes
Peripheral smeara
Yes
DAT
Yes
Hemolysis screena
Yes
a
a
LDH, hap­to­glo­bin, bil­i­ru­bin, retic­u­lo­cyte
count, uri­nal­y­sis
Cold agglu­ti­nin titera
Thermal ampli­tude
Introduction
C3 and C4 lev­els
Based on clin­i­cal sce­nario
Immunoglobulin quanitification
Based on clin­i­cal sce­nario
Serum pro­tein elec­tro­pho­re­sis +
immunofixationa
Cold auto­im­mune hemo­lytic ane­mia (cAIHA) is caused by IgM
autoantibodies whose κ light chains bind eryth­ro­cyte I (or i) anti­
gens at tem­per­a­tures below 37 °C.1 IgM-bound red blood cells
(RBCs) agglu­ti­nate and acti­vate com­ple­ment. C3b on the RBC
sur­face trig­gers phago­cy­to­sis via the hepatic retic­u­lo­en­do­the­
lial sys­tem. Terminal com­ple­ment is acti­vated in the vas­cu­la­ture,
lead­ing to intra­vas­cu­lar hemo­ly­sis.2
Cold AIHA is com­monly pri­mary but can be sec­ond­ary to
another dis­or­der. In adults, CAD is a clonal dis­or­der driven by
low-grade B-cell pro­lif­er­a­tion in the absence of an overt malig­
nancy. Proliferating B cells pro­duce IgM, which drives hemo­ly­sis
through com­ple­ment acti­va­tion (Visual Abstract).2 CAD is more
com­mon than cold agglu­ti­nin syn­drome (CAS), which is defined
as cold hemo­lytic ane­mia aris­ing sec­ond­ary to another dis­or­der
such as auto­im­mune dis­ease, infec­tion, or malig­nancy.3
Bone mar­row stud­iesa
Histologic exam­i­na­tion
Flow cytometry
Infectious workup
Mycoplasma
Recommended
Viral test­ing: HIV, HBV. HCV, EBV
Recommended
Autoimmune screen­ing
Antinuclear antibodiesa
Based on clin­i­cal sce­nario
dsDNA
Based on clin­i­cal sce­nario
Malignancy screen­inga
CT chest abdo­men and pel­vis
Minimum required for diag­no­sis of CAD vs CAS; other stud­ies based on
clin­i­cal sce­nario.
EBV, Epstein-Barr virus; HBV, hep­a­ti­tis B virus; HCV, hep­a­ti­tis C virus.
a
Diagnostic workup
The recommended workup for cAIHA is outlined in Table 1. The
patient his­
tory and phys­
i­
cal exam require atten­
tion to signs
of malig­nancy or infec­tious eti­­ol­ogy, includ­ing eval­u­a­tion for
lymph­ade­nop­a­thy and hepatosplenomegaly. Evidence of acro­
cyanosis is impor­tant in both diag­no­sis and for guid­ing treat­
ment choices. Patients may report worse symp­toms in colder
tem­per­a­tures or exac­er­ba­tions dur­ing illnesses.4
Direct anti­body test­ing
The ini­tial lab­o­ra­tory workup dem­on­strates hemo­ly­sis with an
ele­vated retic­u­lo­cyte count and pos­i­tive hemo­lytic mark­ers. A
pos­i­tive DAT con­firms immune-medi­ated hemo­ly­sis. The DAT is
typ­i­cally strongly C3d pos­i­tive but may also be weakly IgG pos­
i­tive for rea­sons not fully under­stood. IgM read­ily acti­vates the
clas­si­cal com­ple­ment path­way on the eryth­ro­cyte sur­face. Upon
warming, the anti­body detaches from the RBC sur­face before
it can be detected in the DAT assay, but bound C3b remains.
C3b-bound eryth­ro­cytes are phago­cy­tosed by the hepatic retic­
u­lo­en­do­the­lial sys­tem (extra­vas­cu­lar hemo­ly­sis). For cells that
sur­vive phago­cy­to­sis, sur­face C3b is degraded into less active
com­ple­ment com­po­nents, includ­ing C3d, which is reported in
clin­i­cal DAT results. Intravascular hemo­ly­sis occurs via ter­mi­nal
com­ple­ment acti­va­tion with lysis from the action of the C5b-C9
mem­brane attack com­plex on the RBC sur­face (Visual Abstract).5
The clas­
sic teach­
ing is that cAIHA is pri­
mar­
ily intra­
vas­
cu­
lar;
how­ever, extra­vas­cu­lar hemo­ly­sis in the liver pre­dom­i­na­tes,
espe­cially in sta­ble dis­ease. Baseline activ­ity of the reg­u­la­tory
pro­teins CD55 and CD59 impedes com­ple­ment acti­va­tion on the
RBC sur­face, and intra­vas­cu­lar hemo­ly­sis plays a larger role in
dis­ease exac­er­ba­tion or severe dis­ease.5
Performing a DAT and anti­body iden­ti­fi­ca­tion can be tech­ni­
cally chal­leng­ing. With a full under­stand­ing of the clin­i­cal sce­
nario, trans­
fu­
sion med­
i­
cine may be ­
able to opti­
mize test­
ing
and also pro­vide hema­tol­ogy insights into DAT strength or the
behav­ior of antibodies in the sam­ple. Collaboration with trans­
fu­sion med­i­cine in the man­age­ment of cAIHA is also crit­i­cal for
treat­ments such as trans­fu­sion or plas­ma­phe­re­sis.
Other lab­o­ra­tory fea­tures
The mean cor­pus­cu­lar vol­ume is ele­vated, reflecting reticulocy­
tosis, or falsely ele­vated due to RBC agglu­ti­na­tion.6 RBC agglu­
ti­na­tion is appar­ent on periph­eral smear (Figure 1). As with the
patient in Clinical Case 1, com­ple­ment lev­els are typ­i­cally low
due to con­sump­tion.7 The cold agglu­ti­nin titer is at least 64 in
CAD and often sub­
stan­
tially higher. Cold agglu­
ti­
nin ther­
mal
prop­
er­
ties have impli­
ca­
tions for the clin­
i­
cal course: a higher
ther­mal ampli­tude, which is the highest tem­per­a­ture at which
Dr Prakash Singh Shekhawat
Cold AIHA treat­ment | 91
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/90/2021751/90despotovic.pdf by guest on 09 December 2022
count (2.8%; ARC, 38 000). The periph­eral smear shows normo­
cytic ane­mia with spherocytes, clumped cells, and an absence
of schistocytes. DAT IgG is neg­a­tive and DAT C3 2 is pos­i­tive,
with cold auto­an­ti­body iden­ti­fied. He has a mild transamini­
tis, uncon­ju­gated hyperbilirubinemia, unde­tect­able hap­to­glo­
bin, and ele­vated LDH. He is trans­fused at pre­sen­ta­tion using
a blood warmer and is started empir­i­cally on a 5-day course of
azithromycin. He is found to be myco­plasma IgM and IgG pos­
i­tive. He requires a sec­ond trans­fu­sion 2 days later for a hemo­
glo­bin level of 6.4 g/dL. He remains hos­pi­tal­ized for 5 days with
his room kept warm and is advised against cold show­ers and
drink­ing cold bev­er­ages. At the time of dis­charge, his hemo­glo­
bin level is sta­ble at 8.1 g/dL, his ARC is 152 000 (5.7%), and his
hemo­glo­bin­uria and jaun­dice have resolved. His hemo­glo­bin
nor­mal­izes within 2 weeks dur­ing out­pa­tient fol­low-up.
and deep venous throm­bo­sis, which should be con­sid­ered in
mak­ing ther­apy deci­sions.1
Supportive man­age­ment
the cold agglu­ti­nin binds, equates to antibodies bind­ing at a
wider range of tem­per­a­tures and more severe dis­ease. This test­
ing is tech­ni­cally dif­fi­cult, requir­ing the sam­ple to be maintained
at 37 °C to 40 °C until the serum has been removed.4
Patients should be eval­u­ated for infec­tion, the pres­ence
of an overt malig­nancy, or sys­temic auto­im­mu­nity. CAS refers
to cold anti­body–medi­ated hemo­lytic ane­mia that has devel­
oped sec­ond­ary to another con­di­tion. CAD refers to pri­mary
cAIHA driven by the low-grade lymphoproliferation of clonal
B cells pro­duc­ing IgM. Serum pro­tein elec­tro­pho­re­sis con­
firms a CAD diag­no­sis and typ­i­cally shows mono­clo­nal IgMκ.
Required for adult patients in order to dis­tin­guish CAS and
CAD, BM assess­
ment dem­
on­
strates clonal lymphoprolifera­
tive infil­tra­tion, and immunophenotyping shows an abnor­mal
B-cell clone with a B-lym­pho­cyte κ/λ ratio greater than 3.5.8
A com­puted tomo­graphic scan of the chest, abdo­men, and
pel­vis is nec­es­sary to rule out under­ly­ing malig­nancy, par­tic­
u­larly lym­phoma.
Considerations for treat­ment
CAD man­age­ment is based on the degree of ane­mia and symp­
tom­atol­ogy. Those with mild or com­pen­sated ane­mia may not
need treat­ment. Disease exac­er­ba­tions occur in infec­tion, sur­
gery, or cold expo­sure and may neces­si­tate inter­ven­tion.1,9 In a
study of 232 CAD patients, 36% had mild (hemo­glo­bin level >10)
or fully com­pen­sated ane­mia.1 Another 37% had mod­er­ate ane­
mia (hemo­glo­bin 8-10 g/dL), and 27% had severe ane­mia with
hemo­glo­bin lev­els lower than 8.1 Some patients are trans­fu­sion
depen­
dent, with an esti­
mated 40% of patients hav­
ing been
trans­fused.1,7,10 Those liv­ing in colder cli­ma­tes expe­ri­ence coldinduced ische­mic symp­toms, includ­ing cya­no­sis of areas far­ther
from the body core such as the tips of the ears, the nose, and the
dig­its. While treat­ment for mod­er­ate to severe ane­mia or both­
er­some acral symp­toms is accepted, fatigue as an indi­ca­tion for
ther­apy is less clear.11 Fatigue is a prominent com­plaint for many
patients and thought to cor­re­late with com­ple­ment activ­ity.5
Independent of ane­mia sever­ity, CAD car­ries an increased risk for
throm­bo­em­bolic events, includ­ing stroke, myo­car­dial infarc­tion,
92 | Hematology 2022 | ASH Education Program
Therapy to avoid
In warm AIHA, opsonized RBCs are destroyed by splenic mac­ro­
phages. In cAIHA, C3b-coated RBCs are pri­mar­ily phago­cy­tosed
by the hepatic retic­u­lo­en­do­the­lial cells; there­fore, sple­nec­tomy
is not indi­cated in most CAD cases.14
Corticosteroids, the main­stay of treat­ment in warm AIHA, are
inef­fec­tive in cAIHA and should be avoided.17 Despite a lack of
supporting data and an unfa­vor­able side effect pro­file, cor­ti­co­
ste­roids are widely employed.1,7,18 Intravenous Ig is inef­fec­tive in
AIHA, in con­trast to its use in other immune cytopenias.19 Other
immune sup­pres­sive agents are of lim­ited util­ity.17
Pharmacologic man­age­ment
B-cell directed ther­apy
Therapies for CAD are summarized in Table 2, and a proposed
treatment algorithm is presented in Figure 2. Rituximab is the
main­stay of treat­ment for CAD. It is tol­er­ated well and induces
responses in 45% to 60% of patients. Complete responses are
rare.20,21 The median time to response to rituximab is approx­i­ma­
tely 1.5 to 3 months, with a median remis­sion dura­tion of approx­
i­ma­tely 6 months.20,21 Most patients expe­
ri­
ence relapse upon
B-cell repopulation, typ­i­cally within 1 year.21
Given its mod­er­ate effi­cacy and high relapse rate with mono­
therapy, rituximab is used in com­
bi­
na­
tion with the cyto­
toxic
agents fludarabine and bendamustine. Rituximab with fludara­
bine has supe­rior effi­cacy over rituximab alone and may induce
responses in those refrac­tory to rituximab monotherapy.22 How­
ever, fludarabine has higher tox­ic­ity; in a pro­spec­tive uncon­
trolled trial, 57% of patients devel­oped an infec­tion, and 41% had
grade 3 to 4 hema­to­logic tox­ic­ity (14% with grade 4 neutrope­
nia).22 Rituximab with bendamustine has an approx­i­ma­tely 70%
over­all response rate with a more favor­able side effect pro­file.
For those with comorbidities, rituximab monotherapy is the
recommended first-line ther­apy, while for fit indi­vid­u­als likely to
tol­er­ate cyto­toxic ther­apy, rituximab with bendamustine is the
treat­ment of choice.17 For those who fail ini­tial treat­ment or have
relaps­ing symp­toms, sec­ond-line ther­a­pies include com­ple­ment
inhib­i­tors, bortezomib, or, for those well enough to tol­er­ate it,
ibrutinib or rituximab with fludarabine.17
Bortezomib tar­
gets long-lived plasma cell pop­
u­
la­
tions.
In a trial of 21 refrac­
tory patients, 32% gained trans­
fu­
sion
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/90/2021751/90despotovic.pdf by guest on 09 December 2022
Figure 1. Peripheral blood smear from patient with cold agglutinin disease showing red blood cell clumping (100×). Image
courtesy of Dr Tarek M. Elghetany.
Patients should avoid cold tem­per­a­tures, with atten­tion to keep­
ing acral areas warm.10,12 For symp­tom­atic ane­mia, trans­fu­sion
can be given using an in-line blood warmer. Without warming,
hemo­ly­sis is exac­er­bated, lead­ing to line occlu­sion dur­ing the
trans­fu­sion.12
Plasmapheresis may be indi­cated for patients with severe,
symp­tom­atic ane­mia requir­ing rapid inter­ven­tion. A 1 to 1.5 times
plasma vol­ume exchange decreases IgM and improves ane­mia.13
For chil­dren, plas­ma­phe­re­sis is used as a tem­po­riz­ing mea­
sure pend­ing spon­ta­ne­ous dis­ease res­o­lu­tion. While in adults
with CAD, plas­ma­phe­re­sis should be followed by more dura­ble
­ther­apy.14
Other sup­port­ive mea­sures used for hemo­lytic ane­mia include
folate sup­ple­men­ta­tion and, poten­tially, eryth­ro­poi­e­tin.15,16
Table 2. Pharmacologic management of CAD
Dose
Rituximab
Rituximab-fludarabine
375 mg/m2
weekly×4
45%-60%
76%
Median dura­tion
of response
Clinical
con­sid­er­ations
CR: nor­mal­ized hemo­glo­bin level
and absence of hemo­ly­sis or CAD
symp­toms; PR: hemo­glo­bin
increase ≥1-2 g/dL (varies by study),
trans­fu­sion inde­pen­dence and
improved CAD symp­toms, 50%
reduc­tion in IgM con­cen­tra­tion
6.5-11 months
(observed)
Risk of rituximab
infu­sion reac­tions
CR: nor­mal­ized hemo­glo­bin
level, absence of hemo­ly­sis, CAD
symp­toms, unde­tect­able IgM, no
evi­dence of clonal pro­lif­er­a­tion; PR:
hemo­glo­bin increase >/=2 g/dL,
trans­fu­sion inde­pen­dence and
improved CAD symp­toms, 50%
reduc­tion in IgM con­cen­tra­tion
>66 months
(observed)
Definition of response
Confirm vac­ci­na­tion
sta­tus prior to dos­ing
Months for effect
Risk of rituximab
infu­sion reac­tions
Not appro­pri­ate for
frail patients with
comorbidities due
to risk for infec­tion,
neutropenia
Months for effect
Fludarabine
40 mg/m2 on days
1-5, 29-34, 57-61,
and 85-89
Rituximab-bendamustine
Rituximab
375 mg/m2 q28
days×4
71%
CR: nor­mal­ized hemo­glo­bin
level, absence of hemo­ly­sis, CAD
symp­toms, unde­tect­able IgM, no
evi­dence of clonal pro­lif­er­a­tion; PR:
hemo­glo­bin increase >/=2 g/dL,
trans­fu­sion inde­pen­dence and
improved CAD symp­toms, 50%
reduc­tion in IgM con­cen­tra­tion
>32 months
(observed)
Risk of rituximab
infu­sion reac­tions
Not appro­pri­ate for
frail patients with
comorbidities due
to risk for infec­tion,
neutropenia
Months for effect
Bendamustine
90 mg/m2 on days
1, 2 q28 days×4
Eculizumab
600 mg weekly×4,
then 900 mg
every other week
through week 26
54%
Decrease in LDH level pre and post
ther­apy >/=250 U/L
Must be used
con­tin­u­ously to
main­tain effect
Rapid onset; may be
used as a tem­po­riz­ing
mea­sure
Vaccinate against
encap­su­lated bac­te­ria;
menin­go­coc­cal
pro­phy­laxis until
vac­ci­nated
Will not improve
acrocyanosis
Sutimlimab
6500 mg
weekly×2, then
every other week
though 26 weeks
54%
Hemoglobin increase >/=1.5 g/dL,
trans­fu­sion inde­pen­dence, no use of
CAD treat­ments
Must be used
con­tin­u­ously to
main­tain effect
Rapid onset; may be
used as a tem­po­riz­ing
mea­sure
Vaccinate against
encap­su­lated bac­te­ria;
menin­go­coc­cal
pro­phy­laxis until
vac­ci­nated
Will not improve
acrocyanosis
Ibrutinib
Bortezomib
420 mg/d
1.3 mg/m2 on days
1, 4, 8, 11
100%
32%
CR: hemo­glo­bin >12 g/dL;
PR 10-12 g/dL or ≥2 g/dL
Rise in hemo­glo­bin >/=2 g/dL,
trans­fu­sion inde­pen­dence
Studies reported
on ~ median of
12 months daily
use
Requires acy­clo­vir
pro­phy­laxis
16 months
(observed)
Requires acy­clo­vir
pro­phy­laxis
Useful for acrocyanosis
7500 mg for patients weighing >/=75 kg.
CR, com­plete response; PR, par­tial response.
a
Dr Prakash Singh Shekhawat
Cold AIHA treat­ment | 93
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/90/2021751/90despotovic.pdf by guest on 09 December 2022
Rituximab
375 mg/m2 on
days 1, 29, 57,
and 85
Overall
response rate
Other prom­is­ing/exper­i­men­tal ther­a­pies
Figure 2. Treatment algorithm for cold agglutinin disease.
Adapted with permission from S. Berentsen.17
inde­pen­dence, and a sub­set of indi­vid­u­als achieved dura­ble
remis­sion.23,24
Pediatric cAIHA
Complement-directed ther­apy
The intra­vas­cu­lar com­po­nent of CAD is driven by ter­mi­nal com­
ple­ment acti­va­tion on the RBC sur­face (Visual Abstract). Eculi­
zumab, an anti-CD5 mono­
clo­
nal anti­
body inhibiting ter­
mi­
nal
com­ple­ment, is used in other com­ple­ment-driven con­di­tions,
includ­ing par­ox­ys­mal noc­tur­nal hemo­glo­bin­uria and atyp­i­
cal hemo­lytic ure­mic syn­drome.25,26 The DECADE trial was an
open-label, pro­spec­tive nonrandomized phase 2 trial of eculi­
zumab in CAD patients.27 Eculizumab showed an over­all response
rate of 54%, with response defined as a decrease in LDH greater
than or equal to 250 U/L. The median increase in hemo­glo­bin
level was mod­est, at less than 1 g/dL over the 26-week study
period.28 Patients with a narrower ther­mal ampli­tude (indi­cat­ing
milder dis­ease) had a bet­ter response to eculizumab.27
Sutimlimab, a mono­clo­nal anti­body–inhibiting ser­ine pro­te­ase
C1s in the clas­si­cal com­ple­ment path­way, recently com­pleted
test­ing in a sin­gle-arm phase 3 clin­i­cal trial of nontransfused CAD
patients (CARDINAL).11,28,29 Patients tol­er­ated the drug well and
saw an increase in hemo­glo­bin level of approx­i­ma­tely 2.6 g/dL
from base­line by week 3, with a hemo­glo­bin level greater than
or equal to 11 g/dL maintained from week 3 to the end of the
26-week study period. Notably, the hemo­glo­bin level response
in this study was ear­lier and more robust than stud­ies using ecu­
lizumab. Sutimlimab also improved fatigue.11 These find­ings were
val­
i­
dated in a ran­
dom­
ized pla­
cebo-con­
trolled phase 3 study
(CADENZA), with 73% of patients meet­ing the pri­mary end point
of hemo­glo­bin lev­els ris­ing more than or equal to 1.5 g/dL with­out
the need for trans­fu­sion of addi­tional CAD-directed ther­a­pies.30
94 | Hematology 2022 | ASH Education Program
CAD is driven by clonal B-cell expan­
sion that pro­
duces IgM.
Bruton’s tyro­sine kinase inhib­i­tors, phosphatidylinositol 3-kinase
δ inhib­i­tors, or B-cell lym­phoma 2 (BCL2) inhib­i­tors are effi­ca­cious
in other clonal B-cell dis­or­ders and have poten­tial in CAD. Bruton’s
tyro­sine kinase inhib­i­tors such as ibrutinib are effec­tive in both
CAD and CAS.35 Daratumumab, an anti­body against plasma cell
CD38, used in mul­ti­ple mye­loma, has suc­cess­fully resolved hemo­
ly­sis, reduced agglu­ti­nin titers. and resolved non–com­ple­mentmedi­ated symp­toms in refrac­tory CAD.36,37
cAIHA is rare in chil­dren com­pared to adults. The major­ity of
cases develop sec­ond­ary to infec­tion, with Mycoplasma pneumoniae being the most com­
mon cause. Epstein-Barr virus,
influ­enza, cyto­meg­a­lo­vi­rus, and var­i­cella are also impli­cated.38
Infectious test­ing is recommended for all­chil­dren, and auto­im­
mune screen­ing is suggested (eg, anti­nu­clear antibodies titer)
in ado­les­cents and teens. An exten­sive workup to dis­tin­guish
CAD from CAS such as that described in Clinical Case 1 is not
nec­es­sary; cAIHA caused by low-grade clonal B-cell lymphop­
roliferation is not reported in chil­dren. Treatment for pedi­at­ric
cAIHA is focused on sup­port­ive care dur­ing the acute pre­sen­ta­
tion while awaiting spon­ta­ne­ous dis­ease res­o­lu­tion. Measures
include avoiding cold expo­sure and warming blood prod­ucts
or other intra­ve­nous infu­sions. Though in this case the child’s
Mycoplasma pneumoniae was treated, most infec­tions driv­ing
pedi­at­ric cAIHA are self-lim­ited. In an ­emer­gency, plas­ma­phe­
re­sis can be uti­lized. Rituximab and com­ple­ment-directed ther­
a­pies have not been stud­ied in pedi­at­ric cAIHA.39
Paroxysmal cold hemo­glo­bin­uria
Paroxysmal cold hemo­
glo­
bin­
uria (PCH) is uncom­
mon in chil­
dren and exceed­
ingly rare in adults.5 PCH is caused by the
Donath-Landsteiner anti­body directed against the eryth­ro­cyte
P anti­gen.6 IgG fixes com­ple­ment at cold tem­per­a­tures. After
cooling and rewarming the sam­ple, com­ple­ment is ampli­fied,
and RBCs undergo intra­vas­cu­lar hemo­ly­sis.5,39 PCH is trig­gered
by viral infec­tions and treat­ment is sup­port­ive. Because it is IgGmedi­ated, PCH patients may be ste­roid respon­sive.39
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/90/2021751/90despotovic.pdf by guest on 09 December 2022
Sutimlimab is being stud­ied in an open-label mul­ti­cen­ter trial for
trans­fused CAD patients (NCT03347396). Sutimlimab is the only
US Food and Drug Administration-approved drug for adults with
CAD to decrease the need for trans­fu­sion.
BIVV020, an anti-C1s anti­body, is in a phase 1 trial assessing
safety and tol­er­a­bil­ity in CAD patients (clinicaltrials​­.gov iden­ti­fier:
NCT4269551).29 Pegcetacoplan, a C3 inhib­i­tor, suc­cess­fully causes
the ces­sa­tion of hemo­ly­sis in in vitro mod­els.31 In par­ox­ys­mal noc­
tur­nal hemo­glo­bin­uria clin­i­cal tri­als, it is well tol­er­ated and has
supe­rior clin­i­cal out­comes com­pared to eculizumab.32,33 A phase
3, ran­dom­ized, dou­ble-blind, pla­cebo-con­trolled mul­ti­cen­ter trial
eval­u­at­ing pegcetacoplan in CAD is recruiting (NCT05096403).
Unlike rituximab-based pro­to­cols, com­ple­ment-directed treat­
ments require ongo­ing dos­ing. Because com­ple­ment inhib­i­tors
do not stop IgM bind­ing and RBC agglu­ti­na­tion, ische­mic symp­
toms are not alle­vi­ated with this class of drugs, though h
­ emo­ly­sis
may improve.34 The most appro­
pri­
ate use of com­
ple­
mentdirected drugs may be alle­vi­at­ing acute, severe ane­mia or bridg­
ing to dura­ble rituximab-based reg­i­mens.
Conclusion
CAD is a rare cause of AIHA. In adults it is typ­i­cally pri­mary,
driven by clonal lymphoproliferation leading to IgM production,
thereby triggering complement. The under­
stand­
ing of CAD
path­o­phys­i­­o­l­ogy is improv­ing, and treat­ment options have
expanded. Novel ther­a­peu­tics targeting clonal B cells or com­
ple­ment have shown suc­cess, and there are more new ther­a­pies
on the hori­zon. Increasingly, patients are ­able to receive ther­
apy to improve ane­mia, acral symp­toms, and fatigue. With more
well-tol­er­ated treat­ment options avail­­able, the risk/ben­e­fit
ratio of treating cAIHA is shifting toward improved symp­tom
con­trol for more patients.
Jenny McDade Despotovic: con­sul­tancy: Novartis; research fund­
ing: Novartis; hon­o­raria: Novartis, Dova, Amgen; roy­al­ties: Upto­
date.
Taylor Olmsted Kim: con­sul­tancy: Novartis.
Off-label drug use
All drugs are off-label for pediatric patients. For adults, sutim­
limab is the only FDA-approved therapy for cold agglutinin dis­
ease. The rest listed in Table 2 are off-label.
Correspondence
Jenny McDade Despotovic, Baylor College of Medicine, 6701 Fan­
nin St, Ste 1580, Houston, TX 77030; e-mail: jmdespot@txch​­.org
References
1.
Berentsen S, Barcellini W, D’Sa S, et al. Cold agglu­ti­nin dis­ease revisited: a
mul­ti­na­tional, obser­va­tional study of 232 patients. Blood. 2020;136(4):480488.
2. Randen U, Trøen G, Tierens A, et al. Primary cold agglu­ti­nin-asso­ci­ated
lymphoproliferative dis­ease: a B-cell lym­phoma of the bone mar­row dis­
tinct from lymphoplasmacytic lym­phoma. Haematologica. 2014;99(3):497504.
3. Jäger U, Barcellini W, Broome CM, et al. Diagnosis and treat­ment of auto­
im­mune hemo­lytic ane­mia in adults: rec­om­men­da­tions from the First Inter­
national Consensus Meeting. Blood Rev. 2020;41(May):100648.
4. Berentsen S. How I man­age patients with cold agglu­ti­nin dis­ease. Br J Haematol. 2018;181(3):320-330.
5. Berentsen S. New insights in the path­o­gen­e­sis and ther­apy of cold agglu­
ti­nin-medi­ated auto­im­mune hemo­lytic ane­mia. Front Immunol. 2020;11(7
April):590.
6. Kim TO, Despotovic JM. Primary and sec­ond­ary immune cytopenias: eval­
u­a­tion and treat­ment approach in chil­dren. Hematol Oncol Clin North Am.
2019;33(3):489-506.
7. Berentsen S, Ulvestad E, Langholm R, et al. Primary chronic cold agglu­ti­nin
dis­ease: a pop­u­la­tion based clin­i­cal study of 86 patients. Haematologica.
2006;91(4):460-466.
8. Berentsen S, Tjønnfjord GE. Diagnosis and treat­ment of cold agglu­ti­nin
medi­ated auto­im­mune hemo­lytic ane­mia. Blood Rev. 2012;26(3):107-115.
9. Ulvestad E, Berentsen S, Mollnes TE. Acute phase haemolysis in chronic
cold agglu­ti­nin dis­ease. Scand J Immunol. 2001;54(1-2):239-242.
10. Swiecicki PL, Hegerova LT, Gertz MA. Cold agglu­
ti­
nin dis­
ease. Blood.
2013;122(7):1114-1121.
11. Röth A, Barcellini W, D’Sa S, et al. Sutimlimab in cold agglu­ti­nin dis­ease.
N Engl J Med. 2021;384(14):1323-1334.
12. Gabbard AP, Booth GS. Cold agglu­
ti­
nin dis­
ease. Clin Hematol Int.
2020;2(3):95-100.
13. Szczepiorkowski ZM, Winters JL, Bandarenko N, et al; Apheresis Applica­
tions Committee of the Amer­i­can Society for Apheresis. Guidelines on
the use of ther­a­peu­tic apher­e­sis in clin­i­cal prac­tice—evi­dence-based
approach from the Apheresis Applications Committee of the Amer­i­can
Society for Apheresis. J Clin Apher. 2010;25(3):83-177.
14. Berentsen S. Complement acti­
va­
tion and inhi­
bi­
tion in auto­
im­
mune
hemo­
lytic ane­
mia: focus on cold agglu­
ti­
nin dis­
ease. Semin Hematol.
2018;55(3):141-149.
© 2022 by The Amer­i­can Society of Hematology
DOI 10.1182/hema­tol­ogy.2022000369
Dr Prakash Singh Shekhawat
Cold AIHA treat­ment | 95
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/90/2021751/90despotovic.pdf by guest on 09 December 2022
Conflict-of-inter­est dis­clo­sure
15. Fattizzo B, Michel M, Zaninoni A, et al. Efficacy of recom­bi­nant eryth­ro­poi­
e­tin in auto­im­mune hemo­lytic ane­mia: a mul­ti­cen­ter inter­na­tional study.
Haematologica. 2021;106(2):622-625.
16. Go RS, Winters JL, Kay NE. How I treat auto­im­mune hemo­lytic ane­mia.
Blood. 2017;129(22):2971-2979.
17. Berentsen S. How I treat cold agglu­ti­nin dis­ease. Blood. 2021;137(10):12951303.
18. Mullins M, Jiang X, Bylsma LC, et al. Cold agglu­ti­nin dis­ease bur­den: a lon­
gi­tu­di­nal anal­y­sis of ane­mia, med­i­ca­tions, trans­fu­sions, and health care uti­
li­za­tion. Blood Adv. 2017;1(13):839-848.
19. Flores G, Cunningham-Rundles C, Newland AC, Bussel JB. Efficacy of intra­
ve­nous immu­no­glob­u­lin in the treat­ment of auto­im­mune hemo­lytic ane­
mia: results in 73 patients. Am J Hematol. 1993;44(4):237-242.
20. Schöllkopf C, Kjeldsen L, Bjerrum OW, et al. Rituximab in chronic cold
agglu­ti­nin dis­ease: a pro­spec­tive study of 20 patients. Leuk Lymphoma.
2006;47(2):253-260.
21. Berentsen S, Ulvestad E, Gjertsen BT, et al. Rituximab for pri­mary chronic
cold agglu­ti­nin dis­ease: a pro­spec­tive study of 37 courses of ther­apy in 27
patients. Blood. 2004;103(8):2925-2928.
22. Berentsen S, Randen U, Vågan AM, et al. High response rate and dura­ble
remis­sions fol­low­ing fludarabine and rituximab com­bi­na­tion ther­apy for
chronic cold agglu­ti­nin dis­ease. Blood. 2010;116(17):3180-3184.
23. Rossi G, Gramegna D, Paoloni F, et al. Short course of bortezomib in ane­
mic patients with relapsed cold agglu­ti­nin dis­ease: a phase 2 pro­spec­tive
GIMEMA study. Blood. 2018;132(5):547-550.
24. Carson KR, Beckwith LG, Mehta J. Successful treat­ment of IgM-medi­ated
auto­im­mune hemo­lytic ane­mia with bortezomib. Blood. 2010;115(4):915.
25. Hillmen P, Young NS, Schubert J, et al. The com­ple­ment inhib­i­tor eculizumab
in par­ox­ys­mal noc­tur­nal hemo­glo­bin­uria. N Engl J Med. 2006;355(12):12331243.
26. Legendre CM, Licht C, Muus P, et al. Terminal com­
ple­
ment inhib­
i­
tor eculizumab in atyp­i­cal hemo­lytic-ure­mic syn­drome. N Engl J Med.
2013;368(23):2169-2181.
27. Röth A, Bommer M, Hüttmann A, et al. Eculizumab in cold agglu­ti­nin dis­
ease (DECADE): an open-label, pro­
spec­
tive, bicentric, nonrandomized
phase 2 trial. Blood Adv. 2018;2(19):2543-2549.
28. Nikitin PA, Rose EL, Byun TS, Parry GC, Panicker S. C1s inhi­
bi­
tion by
BIVV009 (sutimlimab) pre­vents com­ple­ment-enhanced acti­va­tion of auto­
im­mune human B cells in vitro. J Immunol. 2019;202(4):1200-1209.
29. Jäger U, D’Sa S, Schörgenhofer C, et al. Inhibition of com­ple­ment C1s
improves severe hemo­lytic ane­mia in cold agglu­ti­nin dis­ease: a first-inhuman trial. Blood. 2019;133(9):893-901.
30. Röth A, Berentsen S, Barcellini W, et al. Sutimlimab in patients with cold
agglu­ti­nin dis­ease: results of the ran­dom­ized pla­cebo-con­trolled phase 3
CADENZA trial. Blood. 2022;140(9):980-991.
31. Risitano AM, Ricklin D, Huang Y, et al. Peptide inhib­i­tors of C3 acti­va­tion as
a novel strat­egy of com­ple­ment inhi­bi­tion for the treat­ment of par­ox­ys­mal
noc­tur­nal hemo­glo­bin­uria. Blood. 2014;123(13):2094-2101.
32. de Castro C, Grossi F, Weitz IC, et al. C3 inhi­bi­tion with pegcetacoplan
in sub­jects with par­ox­ys­mal noc­tur­nal hemo­glo­bin­uria treated with eculi­
zumab. Am J Hematol. 2020;95(11):1334-1343.
33. Hillmen P, Szer J, Weitz I, et al. Pegcetacoplan ver­sus eculizumab in par­ox­
ys­mal noc­tur­nal hemo­glo­bin­uria. N Engl J Med. 2021;384(11):1028-1037.
34. Berentsen S. Cold agglu­ti­nins: fending off the attack. Blood. 2019;133(9):
885-886.
35. Jalink M, Berentsen S, Castillo JJ, et al. Effect of ibrutinib treat­ment on
hemo­lytic ane­mia and acrocyanosis in cold agglu­ti­nin dis­ease/cold agglu­
ti­nin syn­drome. Blood. 2021;138(20):2002-2005.
36. Tomkins O, Berentsen S, Arulogun S, et al. Daratumumab for dis­abling cold
agglu­ti­nin dis­ease refrac­tory to B-cell directed ther­apy [published online
11 July 2022]. Amer­i­can J Hematol.
37. Zaninoni A, Giannotta JA, Gallì A, et al. The immu­no­mod­u­la­tory effect and
clin­i­cal effi­cacy of daratumumab in a patient with cold agglu­ti­nin dis­ease.
Front Immunol. 2021;12(March):649441.
38. Kim TO, Bussel JB. Autoimmune Hemolytic Anemia in the Adolescent
Females. Springer Nature Switzerland AG; 2020.
39. Teachey DT, Lambert MP. Diagnosis and man­age­ment of auto­im­mune cyto­
penias in child­hood. Pediatr Clin North Am. 2013;60(6):1489-1511.
AUTOIMMUNE HEMOLYTIC ANEMIAS
Susan T. Johnson and Kathleen E. Puca
Clinical Education and Diagnostic Laboratories Immunohematology Reference Laboratory, Versiti Blood Center of Wisconsin, Milwaukee, WI
The serologic evaluation of autoimmune hemolytic anemia (AIHA) confirms the clinical diagnosis, helps distinguish the
type of AIHA, and identifies whether any underlying alloantibodies are present that might complicate the selection of
the safest blood for any needed transfusion. The spectrum of testing is generally dependent on the amount and class
(immunoglobulin G or M) of autoantibody as well as the resources and methodologies where testing is performed. The
approach may range from routine pretransfusion testing, including the direct antiglobulin test, to advanced techniques
such as adsorptions, elution, and red cell genotyping. When transfusion is needed, the selection of the optimal unit of
red blood cells is based on urgency and whether time allows for the completion of sophisticated serologic and molecular
testing methods. From the start of when AIHA is suspected until the completion of testing, communication among the
clinical team and medical laboratory scientists in the transfusion service and immunohematology reference laboratory
is critical as testing can take several hours and the need for transfusion may be urgent. The frequent exchange of information including the patient’s transfusion history and clinical status, the progress of testing, and any available results
is invaluable for timely diagnosis, ongoing management of the patient, and the safety of transfusion if required before
testing is complete.
LEARNING OBJECTIVES
• Describe the difficulties in determining ABO group/RhD type and in identifying alloantibodies in patients with
autoantibodies
• Discuss strategies for the selection of blood for safe transfusion, keeping in mind patient history and completed
laboratory testing
• Explain the importance of communication among clinicians and laboratory professionals in the approach to
transfusion
Introduction
Autoimmune hemolytic anemia (AIHA) is serologically
defined as warm or cold based on the class of immunoglobulin (Ig), either IgG or IgM, causing hemolysis. Warm
autoantibodies are IgG and primarily detected in vitro
at 37 °C and in the indirect antiglobulin test (IAT), while
cold autoantibodies (CAAs) are IgM and bind preferentially
to red blood cells (RBCs) at colder temperatures in clinical testing (Table 1). Despite these simplified definitions,
the serologic evaluation of patients presenting with these
strong autoantibodies is a challenge for any laboratory. The
96 | Hematology 2022 | ASH Education Program
extent of testing capabilities varies among hospital transfusion services and immunohematology reference laboratories (IRLs). An overview of the various tests employed and
a reference for the clinical cases are provided in Table 2.
CLINICAL CASE 1
A 35-year-old woman with a vague history of idiopathic thrombocytopenia presented to the emergency
department with a complaint of fatigue, malaise, head-
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/96/2021987/96johnson.pdf by guest on 09 December 2022
Evaluating patients with autoimmune hemolytic
anemia in the transfusion service and
immunohematology reference laboratory:
pretransfusion testing challenges and best
transfusion-management strategies
Table 1. Summary of clas­sic sero­logic find­ings in AIHA
Type of AIHA
Antibody class
Antibody spec­i­fic­ity
Optimum test phase
of anti­body reac­tiv­ity
Warm
IgG
Broadly reac­tive
Cold
IgM
Mixed Type
IgG and IgM
DAT result
Eluate result
IAT
2-4+
IgG and C3 or IgG
only
Positive—reac­tive
with all­/major­ity of
panel RBCs
I/i
RT, 37 °Ca
2-3+
C3 only
Not usu­ally performed
Broadly reac­tive
RT, 37 °C, IAT
3-4+
IgG and C3
Positive—reac­tive
with all­/major­ity of
panel RBCs
ache, and cough. Vital signs are tem­per­a­ture, 37.8 °C; blood
pres­sure, 114/55 mmHg; heart rate, 98/min; respi­ra­tory rate,
20; and oxy­gen sat­u­ra­tion, 100% on room air. Initial lab­o­ra­tory
stud­ies show a pos­i­tive COVID test, a white blood cell count of
10 100/µL (4200-11 000), a hemo­glo­bin (Hgb) level of 4.4 g/dL
(12.0-15.5), and a plate­let count of 6000/µL (140 000-450 000).
Type and screen is ordered to pre­pare 2 units of RBCs and 1
unit of apher­e­sis plate­lets for trans­fu­sion. The patient’s blood
type is O-pos­i­tive, but rou­tine anti­body screen­ing by a col­umn
agglu­ti­na­tion test is strongly pos­i­tive (4+) with all­cells, includ­
ing the patient’s own cells. The direct anti­glob­u­lin test (DAT) is
strongly pos­i­tive (4+) with polyspecific reagent. Monospecific
test­ing with anti-IgG and anti-C3d is not avail­­able at this lab­o­
ra­tory. With fur­ther test­ing, the patient sam­ple is strongly pos­
i­tive (3+) with all­panel cells at saline IAT.
The clin­i­cal team is noti­fied of the results and the need to
send out patient sam­ples to the IRL for fur­ther test­ing, with an
antic­i­pated delay of at least 6 to 8 hours. Subsequent labs show
retic­u­lo­cytes, 24.4%; abso­lute retic­u­lo­cyte count, 278 000/µL
(10 000-120 000); lac­tate dehy­dro­ge­nase, 692 U/L (82-240); total
bil­i­ru­bin, 1.5 mg/dL (0.2-1.0); hap­to­glo­bin less than 8 mg/dL (30200); and microspherocytes on periph­eral smear, sug­ges­tive of a
diag­no­sis of warm AIHA (WAIHA). Communication between the
lab­o­ra­tory staff and clin­i­cal team reveals the patient is clin­i­cally
sta­ble, has a his­tory of one missed abor­tion, and no prior trans­
fu­sions. The safety of releas­ing RBC units before com­ple­tion of
test­ing and the low risk of any reac­tions were discussed with
pro­vid­ers.
Pretransfusion test­ing in patients with WAIHA
For any RBC trans­
fu­
sion, ini­
tial rou­
tine tests include an ABO
group, an RhD type, and an anti­body detec­tion test (also known
as an anti­body screen). The pri­mary pur­pose of anti­body detec­
tion test­ing is to detect and iden­tify an anti­body that could
poten­
tially react with trans­
fused RBCs and cause hemo­
ly­
sis.
Antibody detec­tion involves the IAT to iden­tify whether clin­i­cally
sig­nif­i­cant alloantibodies or autoantibodies are pres­ent. For 90%
or more of pretransfusion sam­ples, the results are neg­a­tive and
inter­pre­ta­tion, straight­for­ward.1
The pres­ence of an anti­body that is broadly reac­tive, dem­
on­strat­ing sim­i­lar agglu­ti­na­tion with all­ reagent RBCs includ­ing
the patient’s own RBCs, is clas­si­fied as an auto­an­ti­body. A salient
point in the inter­pre­ta­tion of autoantibodies is their cor­re­la­tion
with clin­i­cal and other lab­o­ra­tory find­ings. An inde­pen­dent
assess­ment of the pres­ence or absence of hemo­lytic ane­mia is
essen­tial. “Benign” autoantibodies are fairly com­mon in sero­
logic work­ups but gen­er­ally not asso­ci­ated with strong pos­i­tive
DAT or hemo­ly­sis.2
The case dem­on­strates one of the more perplexing prob­lems
for med­i­cal lab­o­ra­tory sci­en­tists (MLS) work­ing in the hos­pi­tal
lab­o­ra­tory—distinguishing whether an under­ly­ing allo­an­ti­body
is pres­ent in a sam­ple containing a strong warm auto­an­ti­body
(WAA). As seen in Table 3, all­RBCs strongly agglu­ti­nate in the
pres­ence of a WAA at the IAT, eas­ily masking the detec­tion of an
under­ly­ing allo­an­ti­body. The prev­a­lence of RBC alloantibodies
is higher in patients with WAIHA, rang­ing from 12% to 54% with
a mean of 38% (Table 4). For this rea­son, fur­ther stud­ies to eval­
u­ate for the pres­ence of a clin­i­cally sig­nif­i­cant allo­an­ti­body are
nec­es­sary for the selec­tion of blood. Initially, a method that does
not enhance anti­body bind­ing (known as the no-enhance­ment
method or saline IAT) is under­
taken. This step may omit or
weaken the reac­
tiv­
ity of a low-titer auto­
an­
ti­
body and allow
detec­tion of any under­ly­ing alloantibodies.15
Direct anti­glob­u­lin test
The pres­ence of an auto­an­ti­body in the plasma is the result of
all­anti­gen sites bound on the patient’s RBCs and should trig­ger
the MLS to per­form a DAT. The DAT is a crit­i­cal diag­nos­tic tool for
AIHA and is ini­tially performed by test­ing the patient’s washed
RBCs with polyspecific anti­hu­man glob­u­lin (AHG), which con­
tains both anti-IgG and anti-C3d.
The use of a con­ven­tional test tube con­tin­ues to be the pri­
mary method for the DAT, requir­ing man­ual obser­va­tion for agglu­
ti­na­tion of the patient RBCs after the addi­tion of AHG. It is reli­able
and often used as con­fir­ma­tory test­ing for other DAT tech­niques.
In addi­tion, to detect C3d, a test tube method using anti-C3d is
opti­mal. With the expanding role of auto­ma­tion in hos­pi­tal lab­
o­ra­to­ries, a DAT can be performed using col­umn agglu­ti­na­tion
test­ing (known as gel or beads) or solid-phase meth­ods. While
these alter­na­tive meth­ods for the DAT may be eas­ier to per­form
and pro­
vide less sub­
jec­
tiv­
ity when interpreting the reac­
tion
results (mainly due to end point read­ing by auto­ma­tion), in the
set­ting of AIHA the authors have not observed the other meth­ods
to be more sen­si­tive than a well-performed DAT by tube tech­
nique. Correlation with patient his­tory is par­a­mount to assess the
sig­nif­i­cance of any pos­i­tive DAT result regard­less of method.
When the DAT is pos­
i­
tive with polyspecific AHG, fur­
ther
test­ing with mono­spe­cific reagents is required to dif­fer­en­ti­ate
whether IgG, C3d, or both are coat­ing the patient RBCs and to
Dr Prakash Singh Shekhawat
Serologic eval­u­a­tion of AIHA patients | 97
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/96/2021987/96johnson.pdf by guest on 09 December 2022
a
results after 15-to-30 min­ute incu­ba­tion at 37 °C.
2-4+, agglu­ti­na­tion strength; RT, direct agglutination at room temperature.
Table 2. Immunohematology meth­ods used in eval­u­at­ing patients with warm and cold autoantibodies
Method
Description/pur­pose
Comments/pit­falls
ABO group/RhD type
Determine patient’s ABO group and RhD type
• Presence of high titer CAA may cause
autoagglutination at room tem­per­a­ture. Steps to
resolve this may include:
o Warm/wash patient RBCs at 37 °C
o Perform test­ing at 37 °C
o Treat RBCs with thiol reagent (DTT/2-ME)
to remove IgM coat­ing patient RBCs
DAT
• Polyspecific AHG
Detects IgG and/or C3 (indi­rectly detects IgM) on
patient’s RBCs
• Positive in >90% of cases of AIHA
• Screening assay; when pos­i­tive must repeat test­ing
with anti-IgG and anti-C3d to define pro­tein coat­
ing patient RBCs
DAT
• Monospecific
o Anti-IgG
o Anti-C3d
Individual reagents to detect whether IgG, C3d, or
both on patient’s RBCs
• Both reagents plus con­trol must be performed
• False pos­i­tives can occur due to autoagglutination
(most com­mon with IgM)
Antibody detec­tion (screen)
Use of 2-3 reagent RBCs of known phe­no­type to
screen for auto and/or allo­an­ti­body in patient’s serum
• Detects IgM CAA when performed at IS/RT (direct
agglu­ti­na­tion) using a test tube method
• Detects IgG WAA at IAT with any method.
• Detection depends on test method used
(test tubes, CAT, SP)
• May be neg­a­tive if all­auto­an­ti­body bound to
patient’s RBCs
Antibody iden­ti­fi­ca­tion
Use of panel reagent RBCs (10-16) of known phe­no­type
to deter­mine spec­i­fic­ity of anti­body.
• Reactivity at IS/RT (direct agglu­ti­na­tion) con­firms
CAA. Some show autoanti-I or -i spec­i­fic­ity.
• Confirms broad reac­tiv­ity or spec­i­fic­ity of WAA at IAT.
• Subsequent test­ing when anti­body is detected to
iden­tify antibodies of clin­i­cal sig­nif­i­cance
• Identifies spec­i­fic­ity of any under­ly­ing
allo­an­ti­body once auto­an­ti­body removed from
patient serum
Crossmatch
Confirms ABO com­pat­i­bil­ity by either sero­logic (IS) or
elec­tronic/com­puter.
• If known clin­i­cally sig­nif­i­cant allo­an­ti­body, must
per­form sero­logic crossmatch at IAT (also known as
AHG or full crossmatch)
Phenotype
Serologically type patient RBCs using known antisera.
• Performed when patient has not been
trans­fused in last 3 months
• Strong pos­i­tive DAT may cause false pos­i­tive; not
all­antisera ­able to give accu­rate results when DAT
is pos­i­tive
• Partial anti­gens or var­i­ant anti­gens may not be
detected
Elution
• Elution removes IgG bound to the patient’s RBCs
using a dilute acid solu­tion. Resulting elu­ate is then
tested against a panel of reagent RBCs to check for
spec­i­fic­ity of the RBC bound IgG.
o Positive with all­cells con­firms WAA.
o If neg­a­tive, sus­pect drug-depen­dent
anti­body.
• Once diag­no­sis of AIHA, no need to per­form with
each sub­se­quent pretransfusion test
• Also used in suspected delayed trans­fu­sion reac­tion
to detect bind­ing of allo­an­ti­body to cir­cu­lat­ing
trans­fused RBCs
Adsorption—autol­o­gous
• Removes IgG auto­an­ti­body (when incu­bated at
37 °C) from patient plasma using patient RBCs; then
adsorbed plasma tested to iden­tify under­ly­ing allo­
an­ti­body.
• Removes IgM auto­an­ti­body (when incu­bated at
4C) from patient plasma using patient RBCs; then
adsorbed plasma tested to iden­tify under­ly­ing
allo­an­ti­body.
• Performed when patient has not been trans­fused in
last 3 months
• Patient RBCs are first treated to remove bound
auto­an­ti­body to allow sites for auto­an­ti­body in
plasma to bind
• Requires suf­fi­cient quan­tity of patient RBCs for
removal of auto­an­ti­body
Adsorption—allo­ge­neic
• Removes IgG auto­an­ti­body (when incu­bated at
37 °C) from patient plasma using phenotyped donor
RBCs; then adsorbed plasma tested to iden­tify under­
ly­ing allo­an­ti­body.
• Removes IgM auto­an­ti­body (when incu­bated at 4C)
from patient plasma using donor RBCs; then adsorbed
plasma tested to iden­tify under­ly­ing
allo­an­ti­body.
• Typically, 3 donor RBCs of known phe­no­type
(eg, R1R1, R2R2, rr) are selected
• May miss anti­body to high prev­a­lence anti­gen
Routine:
98 | Hematology 2022 | ASH Education Program
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/96/2021987/96johnson.pdf by guest on 09 December 2022
Advanced:
Table 2. Immunohematology meth­ods used in eval­u­at­ing patients with warm and cold autoantibodies (Continued)
Method
Description/pur­pose
Comments/pit­falls
Genotype
• Molecular meth­ods used to deter­mine patient’s prob­
a­ble phe­no­type.
• More accu­rate than sero­logic phenotyping; gen­er­ally,
pro­vi­des more com­pre­hen­sive anti­gen pro­file
• Recommended when patient has been trans­fused in
last 3 months
• Unusual, rare genetic var­i­ants or par­tial anti­gens
may not be detected
Prewarm tech­nique
• Wash patient RBCs with 37 °C saline prior to test­ing.
• Warm patient plasma and reagent RBCs at 37 °C
sep­a­rately prior to mixing and performing test­ing to
avoid CAA reac­tiv­ity.
• Avoids CAA from bind­ing to patient or donor RBCs
• Clinically sig­nif­i­cant, weakly reac­tive alloantibodies
may not be detected in a prewarm IAT
Thermal ampli­tude study
• Incubate patient serum with reagent RBCs for
2 hours at 30 °C and 37 °C.
• Determines if CAA has abil­ity to bind to reagent RBCs
at warm body tem­per­a­ture.
• Should be performed with adult RBCs (pos­i­tive for
I) and cord blood RBCs (pos­i­tive for i)
• Typically, only need to per­form once to aid in
diag­no­sis
Advanced:
Table 3. Typical anti­body detec­tion (screen­ing) results for autoantibodies and alloantibodies
Cold auto­an­ti­body
Warm auto­an­ti­body
No anti­body detected
Alloantibody detected
Reagent red
cells
RT
37C
IAT
RT
37C
IAT
RT
37C
IAT
RT
37C
IAT
1
4+
3+
1+
0
0
3+
0
0
0
0
0
0
2
4+
3+
1+
0
0
3+
0
0
0
0
0
3+
3
4+
3+
1+
0
0
3+
0
0
0
0
0
0
Autocontrol*
4+
3+
1+
0
0
4+
0
0
0
0
0
0
Results after mixing patient plasma with reagent RBCs. Agglutination is graded 1-4+; no agglutination is indicated as 0.
*Autocontrol contains patient plasma plus patient RBCs.
Italics indicate typical optimum reactivity of cold autoantibody. Bold indicates typical optimum reactivity of warm autoantibody. Italics and bold
indicate typical reactivity of alloantibody.
RT, room temperature (direct agglutination/immediate spin at room temperature); 37C, results after incubation at 37 °C for 10-30 minutes; IAT, results
after addition of AHG.
Table 4. Patients with WAIHA and under­ly­ing alloantibodies
Number of alloantibodies
detected
Reference
% sam­ples with under­ly­ing
alloantibodies
Number of sam­ples
Morel et al3 (1978)
8
20
40
Branch and Petz4 (1982)
5
14
36
Wallhermfechtel et al5 (1984)
19
125
15
Laine and Beattie6 (1985)
41
109
38
James et al (1988)
13
41
38
Issitt et al8 (alloadsorptions) (1996)
13
34
38
Issitt et al8 (autoadsorptions) (1996)
5
41
12
Leger and Garratty9 (1999)
105
263
40
7
Maley et al (2005)
39
126
31
Das and Chaudhary11 (2009)
7
23
30
10
Yürek et al (2015)
7
36
19
Park et al13 (2015)
87
161a
54
Shirey et al (2002)
8
20
40
TOTALS
389
1013
38%
12
14
124 were warm autoantibodies, 21 cold autoantibodies with wide ther­mal range, and 16 were cold-reac­tive only.
a
Dr Prakash Singh Shekhawat
Serologic eval­u­a­tion of AIHA patients | 99
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/96/2021987/96johnson.pdf by guest on 09 December 2022
CAT, col­umn agglu­ti­na­tion tech­nique; IS, imme­di­ate spin (direct agglu­ti­na­tion at room tem­per­a­ture); 2-ME, 2-mercaptoethanol; RT, room tem­per­a­
ture, direct test­ing; SP, solid phase tech­nique.
CLINICAL CASE 1 (Con­t in­u ed)
Testing at the IRL con­firms a strongly pos­i­tive DAT, includ­ing
4+ reac­tiv­ity with both anti-IgG and anti-C3d. Strongly reac­
tive auto­an­ti­body is pres­ent in the patient serum and elu­ate.
The auto­an­ti­body remains strongly pos­i­tive with all­panel cells
despite no enhance­ment at the IAT, neces­si­tat­ing the per­for­
mance of spe­cial­ized adsorp­tion tech­niques. RH and K phenotyping show that the patient’s RBCs are E− and K− and C+, c+,
and e+. RBC genotyping is ordered to deter­mine the extended
phe­no­type.
Adsorption—allo­ge­neic and autol­o­gous
Adsorption stud­ies are the opti­mal method to remove auto­an­ti­
body from the patient’s plasma and leave allo­an­ti­body behind,
if pres­ent. When a sam­ple is referred, the first ques­tion asked
by the IRL MLS is whether the patient has been trans­fused and
when. If trans­fused within the last 3 months, allo­ge­neic (using
donor cells of known phe­no­types) adsorp­tion is performed over
autol­o­gous (using the patient’s own RBCs) adsorp­tion. Even a
small amount of trans­fused RBCs in the patient’s sam­ple can
adsorb out alloantibodies.18 Another fac­tor for the use of alloadsorption is the patient’s degree of ane­mia. For patients with an
Hgb level less than 5g/dL or small pedi­at­ric patients, it is unlikely
that an ade­quate quan­tity of patient RBCs can be col­lected to
suc­cess­fully per­form repeated autoadsorptions.
Autologous adsorp­
tions are logis­
ti­
cally eas­
ier to per­
form and pre­ferred over alloadsorptions for 2 rea­sons. One,
remov­ing anti­body using the patient’s own RBCs proves the
auto­im­mune nature of the auto­an­ti­body. Second, autoadsorptions will not remove an alloantibody to a high prevalence
antigen. Patient RBCs must first be ­
chem­
i­
cally treated to
remove auto­an­ti­body. Chemicals such as ZZAP (dithiothreitol
(DTT)/2-mercaptoethanol + ficin/papain) or enzyme treating
(ficin/papain) alone can remove auto­an­ti­body to free up anti­
100 | Hematology 2022 | ASH Education Program
gen sites for more auto­an­ti­body bind­ing. Repeat adsorp­tions
(1 to 4) based on the strength of the auto­an­ti­body pres­ent
(1+-4+) in the patient’s serum are required, gen­er­ally tak­ing
3 to 4 hours.
Allogeneic adsorp­tions are argu­­ably one of the most com­pli­
cated immu­no­he­ma­tol­ogy tests to per­form and are ­con­fus­ing to
inter­pret. They are labor-inten­sive, gen­er­ally requir­ing 4 to 8 hours
to com­plete. An ade­quate inven­tory of phenotyped blood donors,
typ­i­cally only avail­­able in IRLs, is required for the proper anti­gen
makeup and vol­ume of donor RBCs to per­form the adsorp­tions.
Generally, 3 dif­fer­ent donor RBCs are selected, and the anti­gen
makeup of each must be care­fully con­sid­ered to reveal any clin­
i­cally sig­nif­i­cant allo­an­ti­body since all­donor RBCs adsorb the
auto­an­ti­body. If an allo­an­ti­body is pres­ent, as depicted in Figure 1,
it will bind to donor RBCs that express the cor­re­spond­ing anti­gen
but remain behind if donor RBCs lack the anti­gen. Typically, 2 or 3
rounds of adsorp­tions deplete the auto­an­ti­body.
Regardless of which adsorp­tion method is used, the adsorbed
plasma must be retested to detect and iden­tify any alloantibodies. If test­ing is neg­a­tive, one can be rel­a­tively assured that no
alloantibodies are pres­ent. In our case, 4 autoadsorptions and 4
alloadsorptions both failed to com­pletely remove the WAA. Any
fur­ther adsorp­tions would be imprac­ti­cal and likely increase the
risk of miss­ing weakly reac­tive alloantibodies. These rare sit­u­a­
tions illus­trate the impor­tance of RBC phenotyping/genotyping
in patients with WAIHA.
Patient RBC phenotyping and genotyping
Knowing the patient’s RBC phe­
no­
type can be a pow­
er­
ful
resource to pre­dict which alloantibodies the patient can make
and to guide the selec­tion of blood. At a min­i­mum, the patient
should be typed for RH (C, E, c, e) and K. Extended typ­ing for
addi­tional blood group anti­gens (Fya/Fyb, Jka/Jkb, M/N, and S/s)
is pos­si­ble today with the use of mono­clo­nal reagents. However,
when the patient has a strong pos­i­tive DAT, deter­min­ing the
extended phe­no­type can be prob­lem­atic due to autoagglutination. The patient’s RBCs can be chem­i­cally treated with EDTAgly­cine hydrochloric acid or chlo­ro­quine to remove IgG prior to
test­ing. These pre­treat­ment meth­ods are often tedious, and the
treated RBCs can­not be phenotyped for all­blood group anti­
gens as EDTA-gly­cine hydrochloric acid dena­tures KEL blood
group sys­
tem anti­
gens, and chlo­
ro­
quine treat­
ment reduces
RH anti­gen expres­sion.16 More impor­tantly, phe­no­type by sero­
logic method should not be performed if the patient has been
recently trans­fused.
RBC genotyping (RCG) is becom­ing an increas­ingly valu­able,
cost-effec­tive tool in trans­fu­sion prac­tice. As a send-out test for
most lab­o­ra­to­ries, results are unlikely to be avail­­able for 48 to
72 hours. RCG allows test­ing of more blood group sys­tems and
avoids the tech­ni­cal dif­fi­cul­ties that may occur with sero­logic
test­ing when the patient’s RBCs are coated with anti­body. In
addi­tion, recent trans­fu­sion does not pose a prob­lem, as DNA
for molec­
u­
lar anal­
y­
sis is iso­
lated from the patient’s mono­
nu­
clear cells. Using the patient’s predicted phe­no­type from RCG,
a strat­egy of pro­phy­lac­tic anti­gen-matched RBC units (RH [D,
C, E, c, e], K, and where fea­si­ble JK, FY) may pre­vent alloimmunization, sim­plify sub­se­quent pretransfusion test­ing, and pos­si­
bly reduce the fre­quency of adsorp­tion stud­ies.14,19 In a sur­vey of
aca­demic hos­pi­tal trans­fu­sion ser­vices and IRLs (n=54) on the
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/96/2021987/96johnson.pdf by guest on 09 December 2022
help clas­sify the type of hemo­lytic ane­mia (Table 1). In patients
with WAIHA, more com­
monly both IgG and C3d coat the
patient’s cells.16 While the DAT alone is not diag­nos­tic of hemo­
lytic ane­mia, stron­ger reac­tions (2-4+) denote higher anti­body
den­
sity on the RBC sur­
face and are more likely asso­
ci­
ated
with hemo­ly­sis. Similarly, the pres­ence and amount of C3 (C3b
and/or C3d) on the RBC mem­brane is an impor­tant pre­dic­tor
of immune hemo­ly­sis.2,17 Conversely, a neg­a­tive DAT does not
rule out AIHA and may be seen in 5% to 10% of patients with
hemo­lytic ane­mia.16,17
To con­firm the pos­i­tive DAT is due to an auto­an­ti­body, an
elu­
tion pro­
ce­
dure should be performed as part of the ini­
tial
inves­ti­ga­tion. The pre­dom­i­nant method uti­lized for elu­ate prep­
a­ra­tion removes cell-bound IgG auto­an­ti­body by the addi­tion of
a dilute acid solu­tion to the patient’s RBCs. The acid/RBC mix is
centri­fuged, and the super­na­tant (elu­ate) is buff­ered back to a
nor­mal pH. Antibody iden­ti­fi­ca­tion is performed on this elu­ate.
The results are key in dif­fer­en­ti­at­ing WAIHA from drug-induced
immune hemo­lytic ane­mia. In WAIHA the elu­ate con­tains the IgG
auto­an­ti­body, whereas with a neg­a­tive elu­ate, a drug-­depen­dent
anti­body would be suspected.
t­est­ing ­prac­tices for patients with WAA, 68% performed genotyping and 75% pro­vided pro­phy­lac­tic anti­gen-matched RBC
units, although pol­i­cies var­ied on indi­ca­tions for use.20
Transfusion sup­port in patients with AIHA
Generally, the deci­
sion to trans­
fuse AIHA patients should be
based on sim­
i­
lar indi­
ca­
tions as for ane­
mic patients with­
out
AIHA. In asymp­tom­atic patients, a restric­tive trans­fu­sion strat­
egy (ie, Hgb <7 g/dL) should be prac­ticed.21 The use of bed rest
and sup­ple­men­tal oxy­gen in patients with severe ane­mia (ie,
Hgb <6 g/dL) could help defer trans­fu­sion until test­ing is com­
plete. The patient’s clin­
i­
cal sta­
tus, comorbidities, and symp­
toms should guide the deci­sion to trans­fuse regard­less of Hgb
level and com­pat­i­bil­ity test results. When symp­toms of sig­nif­i­
cant hyp­oxia, con­fu­sion, angina, or hemo­dy­namic insta­bil­ity are
pres­ent, RBC trans­fu­sion should not be with­held, even when
test­ing is incom­plete. The pres­ence of reticulocytopenia in an
AIHA patient with severe ane­mia should be regarded as a med­i­
cal emer­gency. Reticulocytopenia is a poor prog­nos­tic indi­ca­tor
representing an inad­e­quate eryth­ro­poi­etic response and higher
trans­fu­sion require­ments.16,22
When trans­fu­sion is med­i­cally nec­es­sary, the patient’s phy­
si­cian should be assured that trans­fused RBCS are unlikely to
cause an acute hemo­lytic trans­fu­sion reac­tion. The risk for
a trans­fu­sion reac­tion in patients with AIHA appears to be
no higher than in other patients requir­ing trans­fu­sion. Chen
et al.23 found that reac­tions occurred in 1.6% of 450 trans­fused
AIHA patients, with febrile and aller­gic reac­tions the most
fre­quent and no reports of hemo­lytic adverse events. Most
patients with AIHA tol­er­ate sero­log­i­cally incom­pat­i­ble RBC
trans­fu­sions quite well, with­out a sig­nif­i­cant increase in their
under­ly­ing hemo­ly­sis.12,13,23
Dr Prakash Singh Shekhawat
Serologic eval­u­a­tion of AIHA patients | 101
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/96/2021987/96johnson.pdf by guest on 09 December 2022
Figure 1. Adsorption procedure to remove WAA using allogeneic RBCs. Example of the multistep process for allogeneic adsorptions to
remove WAA to identify whether an underlying alloantibody is present (eg, anti-Jka). Three donor RBCs of known phenotypes (ie, R1R1,
R2R2, and rr) are selected and treated with enzyme or ZZAP. The patient’s serum (or plasma) is added to an aliquot of each of the 3 donor
RBCs and incubated for 10 to 60 minutes at 37 °C. Three separate adsorptions are performed, each undergoing 1 to 4 rounds of adsorption in which the patient’s serum is repeatedly incubated with a new set of donor RBCs. Following an adequate number of adsorptions,
adsorbed serum samples are tested against reagent RBCs. If 1 or more aliquots are positive, an alloantibody can be identified based on
the pattern of reactivity and knowledge of the antigens expressed on each aliquot of donor RBCs used for adsorption.
Selection of blood for trans­fu­sion
Selection of the optimal RBC unit for transfusion depends on the
urgency, capabilities and resources of the hospital transfusion
service, and the status of the pretransfusion testing. When there
is life-threat­en­ing ane­mia and no time for com­ple­tion of test­ing,
emer­gency-release, group O RBC units should be given (Figure 2).
The avoid­ance of hemo­dy­namic col­lapse and clin­i­cal dete­ri­o­ra­
tion from pro­found ane­mia far out­weighs the risk of hemo­lytic
reac­tion from an allo­an­ti­body. If the patient has no his­tory of
trans­fu­sions or preg­nancy, the prob­a­bil­ity that under­ly­ing alloantibodies are pres­ent is quite low, and one can be reassured of
the safety of the trans­fused RBCs.
Extended phenotyping of the patient’s RBCs beyond ABO/Rh
D can be done rel­a­tively quickly and guides the selec­tion of pro­
phy­lac­tic anti­gen-matched blood for urgent trans­fu­sions (Figure
2). Even par­tial phenotyping, at a min­i­mum RH (C, c, E, e) and
K, and selecting the “best-matched” RBC unit from the avail­­able
inven­tory may pro­vide some mea­sure of safety and reduce the
risk of devel­op­ment of clin­i­cally sig­nif­i­cant alloantibodies.20,24 In
patients with WAA and con­com­i­tant alloantibodies, at least 50%
are directed against the RH and K anti­gens.25,26
102 | Hematology 2022 | ASH Education Program
Providing pro­phy­lac­tic phe­no­type-matched RBCs beyond RH
and K may be par­tic­u­larly advan­ta­geous for patients with ongo­
ing trans­fu­sion needs or pre­vi­ously iden­ti­fied alloantibodies.20,25
Delaney et al found that patients with WAA and preexisting
alloantibodies had a higher inci­dence of new allo­an­ti­body for­
ma­tion (51.8% vs 37.0%), suggesting that the use of pro­phy­lac­
tic anti­gen-matched RBC units in these patients could pre­vent
addi­tional alloimmunization.25 The degree of anti­gen-matching
depends on the extent of phenotyping performed, when RCG
results are avail­­able, the sup­ply of appro­pri­ate phenotyped/genotyped donor RBCs, and the urgency of trans­fu­sion.
The tech­
nique used for crossmatching blood for patients
with WAA varies based on local pol­i­cies.20 One crit­i­cal point
that can­not be overemphasized is that all­blood is sero­log­i­cally
incom­
pat­
i­
ble. Crossmatching RBC units to search for “least
incom­
pat­
i­
ble” should be dis­
cour­
aged as there is no clin­
i­
cal
ben­e­fit for the patient. Likewise, the use of this term does not
prop­erly con­vey the most impor­tant aspect of pretransfusion
test­
ing—the exclu­
sion of alloantibodies and/or the selec­
tion
of RBCs based on extended phe­no­type matching. Discussion
with the clin­i­cal team should include the extent of com­pat­i­bil­ity
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/96/2021987/96johnson.pdf by guest on 09 December 2022
Figure 2. Initial selection of blood for patients with WAIHA. The algorithm is a suggested approach for the selection of RBC units in
patients with WAIHA who are initially evaluated by the transfusion service. Serologic phenotyping of the patient’s RBCs and provision
of prophylactic antigen-matched RBC units are dependent on the available antisera at the institution and the patient’s recent transfusion history. If it is discovered that the patient has a known history of alloantibodies, provide corresponding antigen-negative RBC
units unless it would delay transfusion in urgent situations.
t­est­ing performed, assur­ance of the safety of the trans­fu­sion,
and the fact that the sur­vival of trans­fused RBCs are com­pa­ra­ble
to the patient’s own RBCs.27
CLINICAL CASE 1 (Con­t in­u ed)
Throughout a case, such as the one described, the need for
fre­quent, effec­tive dia­logue between the clin­i­cal team and trans­
fu­
sion ser­
vice for shared deci­
sion-mak­
ing on any trans­
fu­
sion
needs can­not be overemphasized. Information from the clin­i­cal
team should focus on the patient’s clin­i­cal sta­tus, urgency for
trans­fu­sion, and any his­tory of pre­vi­ous trans­fu­sions, preg­nan­
cies, or prior hos­pi­tal­i­za­tions to assess the like­li­hood of under­ly­
ing alloantibodies. Clear com­mu­ni­ca­tion to the pro­vider about
the extent of required test­ing, time for com­ple­tion, and options
for blood selec­tion is imper­a­tive to enhance under­stand­ing of
the risks and ben­e­fits when trans­fu­sion is nec­es­sary.28,29
CLINICAL CASE 2
A 73-year-old woman with a his­
tory of hyper­
ten­
sion and
cere­bro­vas­cu­lar dis­ease presented for fur­ther eval­u­a­tion of
chronic ane­mia, with a cur­rent Hgb level of 8.0 g/dL and nor­
mal iron stud­ies. She denied any sig­nif­i­cant chest pain, short­
ness of breath, or head­ache and had no epi­sodes of bleed­ing
or dark stools. Vital signs were blood pres­sure, 130/70 mmHg;
heart rate, 71/min; tem­per­a­ture, 36.1 °C; oxy­gen sat­u­ra­tion,
100% on room air. Further lab­o­ra­tory stud­ies showed the fol­
low­ing: white blood cell count, 6300/µL (4100-10 000); plate­
let count, 347 000/µL (150 000-450 000); retic­u­lo­cytes, 6.6%;
abso­
lute retic­
u­
lo­
cyte count, 156 000/µL (22 000-84 000);
hap­to­glo­bin lower than 3 mg/dL (50-220); lac­tate dehy­dro­
ge­nase, 270 U/L (88-207); total bil­i­ru­bin, 5.4 mg/dL (0-1.0);
and indi­
rect bil­
i­
ru­
bin, 4.6 mg/dL. Routine anti­
body detec­
tion by solid phase was neg­a­tive. Blood group­ing showed an
ABO dis­crep­ancy with extra pos­i­tive reac­tions in the patient’s
reverse group­ing. The DAT showed strong autoagglutination
in all­tubes includ­ing the saline con­trol. After mul­ti­ple washes
of the patient’s RBCs with 37 °C saline, the DAT with anti-IgG
was neg­
a­
tive and anti-C3d was 3+. CAA was iden­
ti­
fied at
room tem­per­a­ture (RT) and below.
Dr Prakash Singh Shekhawat
Serologic eval­u­a­tion of AIHA patients | 103
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/96/2021987/96johnson.pdf by guest on 09 December 2022
As the patient was sta­ble and did not have rap­idly drop­ping
Hgb, her trans­fer to a ter­tiary hos­pi­tal was arranged. She was
started on intra­ve­nous Ig and meth­yl­pred­nis­o­lone. Repeat labs
20 hours after her ini­tial pre­sen­ta­tion showed an Hgb level
of 3.6 g/dL. Assessment by the team noted that the patient
started her men­
ses and was complaining of mild head­
ache
and dyspnea. The adsorp­tion stud­ies by the IRL were not yet
com­plete. In col­lab­o­ra­tion with the clin­i­cal team, a deci­sion
was made by the trans­fu­sion team phy­si­cian to release par­
tially anti­gen-matched RBC units (E− and K−) for trans­fu­sion.
The patient received 4 RBC units over 3 days. RCG results were
avail­­able at the time of the fourth trans­fu­sion and indi­cated no
addi­tional anti­gen matching was required.
Cold AIHA (CAIHA), although sero­log­i­cally presenting with
unex­pected test results, is not as dif­fi­cult or time-con­sum­ing to
work up in the trans­fu­sion ser­vice or IRL. Automated meth­ods
for pretransfusion test­ing are designed to detect IgG antibodies,
and with their grow­ing use in hos­pi­tal trans­fu­sion ser­vices, CAA
may not be ini­tially detected in the anti­body screen.
As noted in the case, potent CAAs often cause ABO and Rh
typ­ing discrepancies. Patient RBCs that are heavily coated with
IgM CAA autoagglutinate at RT, caus­ing unex­pected pos­i­tive
results in for­ward ABO/RhD typ­ing. Washing patient RBCs with
warm 37 °C saline prior to test­ing usu­ally removes the autoagglutination. Rarely, if warming is unsuc­cess­ful, RBCs must be
chem­i­cally treated with DTT, a thiol reagent, to remove IgM
auto­
an­
ti­
body. DTT treat­
ment breaks disulfide bonds of IgM
antibodies and there­fore dis­si­pates autoagglutination in the
ABO for­ward group. The big­ger chal­lenge with CAA is unex­
pected pos­i­tive results in the reverse ABO group when the
patient’s plasma is mixed with A1 and B reagent RBCs. To elim­i­
nate this inter­fer­ence, the patient’s plasma is warmed to 37 °C
before adding to prewarmed A1 and B cells. In rare cases, autol­
o­gous or allo­ge­neic adsorp­tions are required to remove CAA to
obtain a valid reverse group.
Classically, in CAIHA the DAT is pos­i­tive with only C3d, which
infers IgM anti­body bind­ing. Similar to WAA, the eval­u­a­tion of
other lab­o­ra­tory stud­ies for hemo­ly­sis is vital to deter­mine the
sig­
nif­i­
cance of the CAA. Autoagglutination can occur with a
strong CAA and lead to false-pos­i­tive results when performing
a DAT, empha­siz­ing the impor­tance of simul­ta­neous test­ing of
con­trols. Resolution and removal of IgM coat­ing the patient’s
cells fol­low sim­i­lar steps as for ABO group dis­crep­ancy, with first
warm wash­ing of the cells and, if needed, DTT treat­ment.
CAAs rarely cause a chal­lenge in rul­ing out under­ly­ing alloantibodies because IgM auto­an­ti­body binds to RBCs pref­er­en­tially
at colder tem­per­a­tures but can carry over to 37 °C and the IAT,
while clin­i­cally impor­tant IgG alloantibodies are reac­tive only at
the IAT (Table 3). Eliminating the use of poten­ti­a­tors in tra­di­tional
test tube meth­ods and performing test­ing only at 37 °C (to avoid
auto­an­ti­body bind­ing at RT) is often suf­fi­cient to reduce reac­tiv­
ity by the CAA at the IAT. Negative reac­tions at the IAT indi­cate
that no IgG alloantibodies are pres­ent. Rarely, 1 or 2 autol­o­gous
or allo­ge­neic adsorp­tions with treated RBCs incu­bated at 4 °C
remove enough CAA to avoid inter­fer­ence in the IAT and allow
alloantibodies to be ruled out.
Once the under­
ly­
ing alloantibodies have been excluded,
the selec­tion of RBCs for trans­fu­sion typ­i­cally fol­lows stan­dard
prac­tice. The pro­vi­sion of pro­phy­lac­tic RH and K matching is
gen­er­ally recommended if the patient has an ongo­ing need for
trans­fu­sion. If ABO group­ing results can­not be resolved due to a
potent CAA, group O RBCs should be selected. In patients with
CAA with severe brisk hemo­lytic ane­mia, the use of an in-line
blood warmer as well as warming blan­kets may mit­i­gate any fur­
ther hemo­ly­sis.16
As part of the diag­nos­tic eval­u­a­tion, a cold agglu­ti­nin titer
may be ordered; how­ever, the titer result is not reflec­tive of the
abil­ity of the CAA to cause immune hemo­lytic ane­mia as the test­
ing is performed at 4°C. In the authors’ opin­ion, it is an overused test that pro­vi­des lit­tle clin­i­cal rel­e­vance. A bet­ter test to
deter­mine the path­o­log­i­cal nature of a CAA is a ther­mal ampli­
tude study. This test­ing assesses the abil­ity of the CAA to bind
to RBCs at 30 °C or warmer, reflecting body tem­per­a­ture. Some
may argue that nei­ther is nec­es­sary if stan­dard sero­logic test­ing
and clin­i­cal signs and symp­toms sup­port a path­o­log­i­cal CAA.
CLINICAL CASE 2 (Con­t in­u ed)
Our patient was asymp­tom­atic and did not require trans­fu­sion
at this time. Neither a cold agglu­ti­nin titer nor a ther­mal ampli­
tude study was performed as the find­ings from her sero­logic
test­
ing, includ­
ing a DAT, and other lab­
o­
ra­
tory stud­
ies were
con­sis­tent with CAIHA.
Both WAIHA and CAIHA pres­ent unique chal­lenges for pretransfusion test­ing. The pro­vi­sion of RBCs for trans­fu­sion is based
on the clin­i­cal sta­tus of the patient and the extent of test­ing
affected by the class (IgG or IgM) and con­cen­tra­tion of auto­an­ti­
body. Testing for patients with acute WAIHA is the most dif­fi­cult,
requir­ing com­pli­cated, lengthy pro­ce­dures such as adsorp­tion
stud­ies and genotyping that may not be com­pleted prior to
need­ing trans­fu­sion. Thus, early and fre­quent com­mu­ni­ca­tion is
essen­tial for the exchange of infor­ma­tion to effec­tively guide the
best approach for imme­di­ate and future trans­fu­sions. This extra
effort pro­vi­des assur­ance for the clin­i­cal and lab­o­ra­tory teams
that units released for trans­fu­sion, although incom­pat­i­ble, have
a high like­li­hood of being safe and effi­ca­cious for the patient.
Conflict-of-inter­est dis­clo­sure
Susan T. Johnson: book roy­al­ties: AABB Press.
Kathleen E. Puca: no com­pet­ing finan­cial inter­ests to declare.
Off-label drug use
Susan T. Johnson: nothing to disclose.
Kathleen E. Puca: nothing to disclose.
Correspondence
Susan T. Johnson, Clinical Education and Diagnostic Laboratories
Immunohematology Reference Laboratory, Versiti Blood Center of
Wisconsin, 638N 18th St, Milwaukee WI 53233; e-mail: stjohnson@
versiti​­.org.
References
1.
2.
3.
4.
5.
6.
7.
Hendrickson JE, Tormey CA. Understanding red blood cell alloimmunization
trig­gers. Hematology Am Soc Hematol Educ Program. 2016;2016(1):446451.
Wheeler CA, Calhoun L, Blackall DP. Warm reac­tive autoantibodies: clin­i­cal
and sero­logic cor­re­la­tions. Am J Clin Pathol. 2004;122(5):680-685.
Morel PA, Bergren MO, Frank BA. A sim­ple method for the detec­tion of allo­
an­ti­body in the pres­ence of warm auto­an­ti­body [ab­stract]. Transfusion.
1978;18(3):388.
Branch DR, Petz LD. A new reagent (ZZAP) hav­ing mul­ti­ple appli­ca­tions in
immu­no­he­ma­tol­ogy. Am J Clin Pathol. 1982;78(2):161-167.
Wallhermfechtel MA, Pohl BA, Chaplin H. Alloimmunization in patients with
warm autoantibodies. A ret­ro­spec­tive study employing three donor alloabsorptions to aid in anti­body detec­tion. Transfusion. 1984;24(6):482-485.
Laine ML, Beattie KM. Frequency of alloantibodies accom­pa­ny­ing autoantibodies. Transfusion. 1985;25(6):545-546.
James P, Rowe GP, Tozzo GG. Elucidation of alloantibodies in auto­im­mune
haemolytic anae­mia. Vox Sang. 1988;54(3):167-171.
104 | Hematology 2022 | ASH Education Program
© 2022 by The Amer­i­can Society of Hematology
DOI 10.1182/hema­tol­ogy.2022000406
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/96/2021987/96johnson.pdf by guest on 09 December 2022
Conclusion
8. Issitt PD, Combs MR, Bumgarner DJ, Allen J, Kirkland A, Melroy-Carawan H.
Studies of antibodies in the sera of patients who have made red cell autoantibodies. Transfusion. 1996;36(6):481-486.
9. Leger RM, Garratty G. Evaluation of meth­ods for detecting alloantibodies
under­ly­ing warm autoantibodies. Transfusion. 1999;39(1):11-16.
10. Maley M, Bruce DG, Babb RG, Wells AW, Williams M. The inci­dence of red
cell alloantibodies under­ly­ing panreactive warm autoantibodies. Immunohematology. 2005;21(3):122-125.
11. Das SS, Chaudhary R. Utility of adsorp­
tion tech­
niques in sero­
log­
i­
cal
eval­u­a­tion of warm auto­im­mune haemolytic anae­mia. Blood Transfus.
2009;7(4):300-304.
12. Yürek S, Mayer B, Almahallawi M, Pruss A, Salama A. Precautions surrounding blood trans­fu­sion in auto­im­mune haemolytic anae­mias are overestimated. Blood Transfus. 2015;13(4):616-621.
13. Park SH, Choe WH, Kwon SW. Red blood cell trans­fu­sion in patients with
autoantibodies: is it effec­tive and safe with­out increas­ing hemo­ly­sis risk?
Ann Lab Med. 2015;35(4):436-444.
14. Shirey RS, Boyd JS, Parwani AV, Tanz WS, Ness PM, King KE. Prophylactic anti­gen-matched donor blood for patients with warm autoantibodies:
an algo­rithm for trans­fu­sion man­age­ment. Transfusion. 2002;42(11):14351441.
15. Cohn CS, Delaney M, Johnson ST, Katz LM. AABB Technical Manual. 20th
ed. Association for the Advancement of Blood and Biotherapies; 2020:408409, 434-444.
16. Petz LD, Garratty G. Immune Hemolytic Anemias. 2nd ed. Churchill Livingstone; 2004:69-77,171,215-216,319,395.
17. Meulenbroek EM, Wouters D, Zeerleder SS. Lyse or not to lyse: clin­i­cal sig­
nif­i­cance of red blood cell autoantibodies. Blood Rev. 2015;29(6):369-376.
18. Laine EP, Leger RM, Arndt PA, Calhoun L, Garratty G, Petz LD. In vitro stud­
ies of the impact of trans­fu­sion on the detec­tion of alloantibodies after
autoadsorption. Transfusion. 2000;40(11):1384-1387.
19. El Kenz H, Efira A, Le PQ , et al. Transfusion sup­port of auto­im­mune hemo­
lytic ane­mia: how could the blood group genotyping help? Transl Res.
2014;163(1):36-42.
20. Ziman A, Cohn C, Carey PM, et al; Biomedical Excellence for Safer Transfusion Collaborative. Warm-reac­tive (immu­no­glob­u­lin G) autoantibodies and
lab­o­ra­tory test­ing best prac­tices: review of the lit­er­a­ture and sur­vey of
cur­rent prac­tice. Transfusion. 2017;57(2):463-477.
21. Carson JL, Guyatt G, Heddle NM, et al. Clinical prac­tice guide­lines from
the AABB: red blood cell trans­
fu­
sion thresh­
olds and stor­
age. JAMA.
2016;316(19):2025-2035.
22. Barcellini W, Zaninoni A, Giannotta JA, Fattio B. New insights in auto­
im­mune hemo­lytic ane­mia: from path­o­gen­e­sis to ther­apy. J Clin Med.
2020;9(12):3859.
23. Chen C, Wang L, Han B, Qin L, Ying B. Autoimmune hemo­lytic ane­mia in
hos­pi­tal­ized patients: 450 patients and their red blood cell trans­fu­sions.
Medicine. 2020;99(2):e18739.
24. Hill QA, Stamps R, Massey E, et al; Brit­ish Society for Haematology. The
diag­no­sis and man­age­ment of pri­mary auto­im­mune haemolytic anae­mia.
Br J Haematol. 2017;176(3):395-411.
25. Delaney M, Apelseth TO, Bonet Bub C, et al; Biomedical Excellence for
Safer Transfusion Collaborative. Red-blood-cell alloimmunization and pro­
phy­lac­tic anti­gen matching for trans­fu­sion in patients with warm autoantibodies. Vox Sang. 2020;115(6):515-524.
26. Garratty G, Petz LD. Approaches to selecting blood for trans­
fu­
sion to
patients with auto­im­mune hemo­lytic ane­mia. Transfusion. 2002;42(11):13901392.
27. Petz LD. “Least incom­pat­i­ble” units for trans­fu­sion in auto­im­mune hemo­
lytic ane­
mia: should we elim­
i­
nate this mean­
ing­
less term? A com­
men­
tary for cli­ni­cians and trans­fu­sion med­i­cine pro­fes­sion­als. Transfusion.
2003;43(11):1503-1507.
28. Petz LD. A phy­si­cian’s guide to trans­fu­sion in auto­im­mune haemolytic anae­
mia. Br J Haematol. 2004;124(6):712-716.
29. Ness PM. How do I encour­age cli­ni­cians to trans­fuse mismatched blood to
patients with auto­im­mune hemo­lytic ane­mia in urgent sit­u­a­tions? Transfusion. 2006;46(11):1859-1862.
AUTOIMMUNE HEMOLYTIC ANEMIAS
Warm autoimmune hemolytic anemia
and the best treatment strategies
Hematology Division, Massachusetts General Hospital, Harvard Medical School, Boston, MA
Warm autoimmune hemolytic anemia (wAIHA) is characterized by evidence of red blood cell (RBC) hemolysis and a
direct antiglobulin test positive for IgG and sometimes complement. While varying with the extent of the compensatory increase in RBC production, symptoms of anemia predominate, as does jaundice, the latter often exacerbated
by concurrent Gilbert’s syndrome. Initial treatment with corticosteroids is highly effective, with over 85% of patients
responding but with less than one-third maintaining that response upon weaning. Subsequent rituximab administration
in those failing corticosteroids provides complete remission in over 75% of patients and may be long-lasting. Over 50%
of patients failing rituximab respond to erythropoiesis-stimulating agents or immunosuppressive agents. Splenectomy is
best deferred if possible but does offer long-term remission in over two-thirds of patients. A number of new treatments
for wAIHA (fostamatinib, rilzabrutinib, and FcRn inhibitors) show promise. A treatment algorithm for wAIHA is proposed
to avoid the excessive use of corticosteroids.
LEARNING OBJECTIVES
• Recognize the clinical and laboratory features that determine the diagnosis and etiology of wAIHA
• Create a pathophysiology­based framework for the treatment of wAIHA with current and future novel therapies
CLINICAL CASE
A 69­year­old woman who was a Jehovah’s Witness pre­
sented with profound fatigue. Her previous complete
blood cell count (CBC) had been normal. Laboratory tests
showed the following values: hemoglobin (Hgb), 6 g/dL;
reticulocytes, 15%; lactate dehydrogenase (LDH), 560 IU/L;
total bilirubin, 2.3 mg/dL; and haptoglobin, less than
10 mg/dL. The smear showed spherocytes, and the direct
antiglobulin test (DAT) was strongly positive for immuno­
globulin G (IgG) and weak for C3. She was treated with
folate, 1 mg/d, and prednisone, 60 mg/d, for 3 weeks.
Her LDH and bilirubin decreased nearly to normal, and
her Hgb level rose to 13 g/dL, so prednisone was weaned
over the next 8 weeks to 5 mg/d. Her Hgb level remained
stable at 10 g/dL for 2 months on 5 mg of prednisone
every other day but then dropped to 7 g/dL. Prednisone
was increased to 20 mg/d, and 375 mg/m2 rituximab was
administered weekly for 4 weeks, with a return of the
Hgb to 12 g/dL. Steroids were again weaned to 5 mg/d,
but 3 months later her Hgb level again dropped below
9 g/dL. Mycophenolate (MMF) at 1000 mg twice a day was
started and the prednisone was weaned off, with Hgb lev­
els remaining stable at 10 g/dL for 12 months. The patient
developed increasing nausea, and the MMF was discon­
tinued, with the Hgb level falling slowly to 7.7 g/dL 10
weeks later. The patient refused splenectomy, resumed
prednisone at 5 mg/d, and was started on 200 mcg of
darbepoetin alpha every 3 weeks, with Hgb rising to 10.5
(the prednisone having been tapered off). After 9 months,
her insurance disallowed the darbepoetin alpha, and the
patient again resumed prednisone at 5 mg/d. The Hgb
level rose to 9.4 g/dL. She entered a clinical trial with fos­
tamatinib and has been maintained on 150 mg twice a day
for 2 years with an Hgb level of 12 g/dL off prednisone.
Introduction
The normal red blood cell (RBC) survives in the circulation
for approximately 120 days (Figure 1A). Hemolytic anemia
is considered when RBC survival drops below 100 days.1 In
evaluating a patient for hemolytic anemia, it is important
to first provide evidence for hemolysis. In patients with
hemolysis, a number of laboratory tests become abnormal
Warm autoimmune hemolytic anemia: diagnosis and treatment | 105
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/105/2021916/105kuter.pdf by guest on 09 December 2022
David J. Kuter
(Figure 1B). The retic­u­lo­cyte count and retic­u­lo­cyte index usu­ally
increase as a reflec­tion of the increased egress of young RBCs
from the bone mar­row as they “shift” into the cir­cu­la­tion. The
destruc­tion of RBCs results in the release of their LDH, and some­
times Hgb, into the cir­cu­la­tion. Free Hgb may be detect­able in
the urine and serum and can bind to and reduce lev­els of hap­
to­glo­bin; once fil­tered by the kid­ney, it may be catab­o­lized by
renal tubu­lar cells, which retain iron and pro­duce a pos­i­tive urine
hemo­sid­erin test. The catab­o­lism of RBCs increases bil­i­ru­bin; the
for­ma­tion of bil­i­ru­bin and the con­cur­rent pro­duc­tion of car­bon
mon­ox­ide in this pro­cess are per­haps the most accu­rate indi­ca­
tors of RBC turn­over.2,3
The above attri­butes of hemo­ly­sis have many excep­tions and
nuances. As discussed below, the retic­u­lo­cyte count and index
may be low in many patients with auto­im­mune hemo­lytic ane­
mia (AIHA) due to antibodies bind­ing the retic­u­lo­cyte or nutri­
tional or bone mar­row deficiencies. Given that Gilbert’s dis­or­der
affects 5% to 10% of the Cau­ca­sian pop­u­la­tion,2 even mild hemo­
ly­sis may result in strik­ingly ele­vated bil­i­ru­bin and symp­tom­atic
jaun­dice. In gen­eral, the pres­ence of urine hemo­sid­erin, urine,
or serum free Hgb reflects extra­
vas­
cu­
lar hemo­
ly­
sis, whereas
increased LDH, reduced hap­to­glo­bin, and ele­vated bil­i­ru­bin may
be seen in both intra­vas­cu­lar and extra­vas­cu­lar hemo­ly­sis.
106 | Hematology 2022 | ASH Education Program
The doc­u­men­ta­tion of RBC hemo­ly­sis is crit­i­cal for the diag­no­
sis of hemo­lytic ane­mia. In a small study of 100 ane­mic patients,
a hap­to­glo­bin lower than 25mg/dL was 83% spe­cific and 96%
sen­si­tive and had a pre­dic­tive value of 88% for hemo­ly­sis.4 When
the hap­to­glo­bin level is less than 25mg/dL and the LDH level is
greater than the upper limit of nor­mal, sen­si­tiv­ity drops to 50%,
but spec­i­fic­ity and pre­dic­tive value are 100%.5 Hemolytic ane­mia
is impre­cisely defined as an ane­mia with short­ened RBC sur­vival,
an LDH level greater than the upper limit of nor­mal, and a hap­to­
glo­bin level less than the lower limit of nor­mal.1
Many reg­u­la­tory mech­a­nisms may affect the phys­i­o­log­i­cal
com­
pen­
sa­
tion for increased RBC destruc­
tion. These include
eryth­
ro­
poi­
e­
tin and nutri­
ents such as folic acid, B12, and iron,
as well as the gen­eral health of the bone mar­row. Normal RBC
pro­
duc­
tion can increase 5-fold to 8-fold and pro­
vide sig­
nif­i­
cant com­pen­sa­tion for reduced RBC sur­vival. Indeed, for some
patients with any type of mild hemo­lytic ane­mia, the pro­vi­sion
of ade­quate nutri­tional resources and eryth­ro­poi­e­tin may allow
ade­quate com­pen­sa­tion with no need for other treat­ment.
There are many causes of hemo­lytic ane­mia. As described
in Table 1, they include intrin­sic defects of the RBC mem­brane,
enzymes, or Hgb. Many dis­or­ders extrin­sic to the RBC may also
cause short­ened RBC sur­vival. Nonimmune mech­a­nisms such as
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/105/2021916/105kuter.pdf by guest on 09 December 2022
Figure 1. RBC production and catabolism in states of normal (A) or accelerated (B) RBC turnover. The RBC half-life of 5 to 10 days is
simply illustrative. The changes denoted in RBC production and markers of catabolism vary with the extent of RBC turnover.
against mul­
ti­
ple RBC types. The con­
clu­
sions from the direct
­antiglobulin test (DAT) and elu­ate (using the DAT) are as fol­lows:
Table 1. Types of hemo­lytic ane­mia
Intrinsic
Membrane/cyto­skel­e­tal abnor­mal­i­ties—
hered­i­tary spherocytosis
Enzyme defects—G6PD, pyru­vate kinase
defi­ciency
Hgb defects—thal­as­se­mia, sickle cell dis­ease
Conclusions
C3 only
Immune com­plex, cold agglu­
ti­nin dis­ease, cold agglu­ti­nin
syn­drome
Increased shear—arti­fi­cial valves
DIC, TTP, HUS
Infection—malaria, bab­e­si­o­sis
Oxidative stress—dap­sone, rasburicase
Liver fail­ure—spur cell ane­mia
Hypersplenism
IgG, neg­a­tive elu­ate
Drug, absorbed IgG
IgG, elu­ate spe­cific
Alloimmune
IgG, elu­ate binds all­RBCs
wAIHA, drug
IgG, +/− C3, elu­ate binds all­RBCs
wAIHA
AIHA
Cold anti­body—cold AIHA: cold agglu­ti­nin
dis­ease, cold agglu­ti­nin syn­dromea
Warm anti­body—wAIHAb
Mixed
In sum­mary, AIHA is char­ac­ter­ized by evi­dence of hemo­ly­sis
and a pos­i­tive Coombs test. Eighty per­cent of patients have an
IgG (warm anti­body) that binds to pro­tein anti­gens on RBC sur­
faces at the core body tem­per­a­ture. Twenty per­cent of patients
have an IgM (cold anti­body) that binds to poly­sac­cha­ride anti­gens
on RBC sur­faces at tem­per­a­tures below the core body tem­per­a­
ture. A small pro­por­tion of patients may have both (mixed AIHA).
The rest of this review focuses on warm AIHA (wAIHA).
Extrinsic
Nonimmune
Immune
Intravascular hemo­ly­sis.
Extravascular hemo­ly­sis.
DIC, dis­sem­i­nated intra­vas­cu­lar coagulation; G6PD, glu­cose-6-phos­
phate dehy­dro­ge­nase defi­ciency; HUS, hemo­lytic ure­mic syn­drome;
TTP, throm­botic throm­bo­cy­to­pe­nic pur­pura.
a
b
Pathophysiology of wAIHA
dis­sem­i­nated intra­vas­cu­lar coagulation, infec­tion, and oxi­da­tive
stress are com­mon. The immune mech­a­nisms caus­ing AIHA are
much less com­mon but are char­ac­ter­ized by the dem­on­stra­tion
of spe­cific antibodies and/or com­ple­ment bind­ing to RBCs.
Critical to the diag­no­sis of AIHA is the iden­ti­fi­ca­tion of anti­
bodies against RBC anti­gens. This is rou­tinely done using many
var­i­a­tions of the test orig­i­nally designed by Robin Coombs.6,7
RBCs are ini­tially assessed for the pres­ence of both immu­no­
glob­u­lin and C3 using polyspecific antisera.8 If pos­i­tive, fur­ther
assess­ments with antibodies spe­cific for IgG and C3 are per­
formed. A com­
plete assess­
ment of RBCs test­
ing pos­
i­
tive for
IgG includes deter­min­ing whether elu­ate autoantibodies react
More than 95% of patients with wAIHA have an IgG anti­body that
binds to RBC anti­gens inde­pen­dently of tem­per­a­ture. Rarely,
antibodies may be IgA and even less com­
monly, IgM; these
may account for the approx­i­ma­tely 5% of patients who have
“Coombs-neg­a­tive” wAIHA. These antibodies are poly­clonal
even in clonal dis­or­ders such as chronic lym­pho­cytic leu­ke­mia
(CLL).9 About one-third fix com­ple­ment; IgG1 and IgG3 fix com­
ple­ment more effi­ciently than the oth­ers.
The antibodies bind two major RBC anti­
gens (Figure 2).
About 40 000 Rh pro­teins are fixed to the cyto­skel­e­ton in each
RBC and do not have a high enough den­sity to acti­vate com­ple­
ment. In con­trast, there are about 800 000 glycophorin pro­teins
Figure 2. RBC surface antigen targets in wAIHA.
Warm auto­im­mune hemo­lytic ane­mia: diag­no­sis and treat­ment | 107
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/105/2021916/105kuter.pdf by guest on 09 December 2022
DAT and elu­ate results
with mild hemo­ly­sis, in patients who have con­cur­rent Gilbert’s
syn­drome.2
Some wAIHA patients sim­
ply have a mild ane­
mia and
increased mean cor­pus­cu­lar vol­ume and are referred to hema­tol­
ogy for assess­ment for pos­si­ble myelodysplastic syn­drome. The
increased mean cor­pus­cu­lar vol­ume is related to the increased
num­bers of retic­u­lo­cytes (which are gen­er­ally larger than more
mature RBCs) and is not due to a pri­mary bone mar­row dis­or­der.
Incidence of wAIHA
Figure 3. Peripheral blood smear in patient with wAIHA failing
splenectomy. Increased numbers of spherocytes (red arrows)
and polychromatophilic reticulocytes (black arrows), as well as
one RBC with a Howell-Jolly body (green arrow), are indicated.
A, B, C, and D fixed to the mem­brane of each RBC, a high enough
den­sity to acti­vate com­ple­ment.
Once bound to an anti­body, the RBC can undergo mul­ti­ple
fates. Most will be bound to splenic mac­ro­phages via the FcγRIII
recep­tor, resulting in either phago­cy­to­sis of the entire RBC or,
more likely, removal of a sig­nif­i­cant por­tion of the RBC mem­
brane-pro­
duc­
ing spherocytes (also called microspherocytes)
(Figure 3). Unlike the nor­mal RBC, spherocytes are not deform­
able and, upon enter­ing the splenic Cords of Billroth, undergo
destruc­tion. This may account for some of the increased LDH and
depressed hap­to­glo­bin seen in a dis­or­der com­monly described
as “extra­vas­cu­lar.” About one-third of patients with wAIHA addi­
tion­ally have com­ple­ment bound to RBCs, and these C3b-coated
RBCs are then cleared by the C3b recep­tors on hepatic Kupffer
cells. In some of these RBCs, C3b is inactivated to C3d, thereby
pre­vent­ing their sub­se­quent destruc­tion and allowing a pop­u­la­
tion of anti­body/com­ple­ment-coated RBCs to per­sist.
In sum­mary, wAIHA is defined by the fol­low­ing: (1) evi­dence
of hemo­ly­sis; (2) an IgG anti­body bind­ing to pro­tein anti­gens on
the RBC sur­face at the core tem­per­a­ture (warm anti­body), dem­
on­strated by a pos­i­tive Coombs test for IgG in 95% of patients;
(3) spherocytes on the periph­eral blood smear. RBCs from 30%
of patients bind both IgG and C3, and less than 5% are Coombs
neg­a­tive.1,10
Presentation of wAIHA
The presenting char­ac­ter­is­tics are var­i­able. The typ­i­cal symp­
toms of ane­
mia are often pres­
ent, includ­
ing short­
ness of
breath with exer­cise, gen­er­al­ized fatigue, diz­zi­ness, syn­cope,
malaise, chest pres­sure/pain, and cog­ni­tive dys­func­tion. Most
patients have a reduced health-related qual­ity of life.11 Some
patients with mild ane­mia and sig­nif­i­cant com­pen­sa­tion have
no symp­toms.
Symptoms spe­
cific to hemo­
ly­
sis include jaun­
dice, spleno­
meg­aly, and gall­stones. Significant icterus may be pres­ent, even
108 | Hematology 2022 | ASH Education Program
Causes of wAIHA
Table 2 lists the many causes for wAIHA.14,16 Infections and a wide
vari­ety of drugs/tox­ins are com­monly asso­ci­ated with wAIHA,
but these hemo­lytic events are mostly tran­sient, usu­ally iden­
ti­
fied upon ini­
tial eval­
u­
a­
tion, and resolve with treat­
ment of
the under­ly­ing infec­tion or dis­con­tin­u­a­tion of the asso­ci­ated
drug/toxin. What is of con­cern then in this dis­cus­sion are those
patients in whom wAIHA is chronic and not obvi­ously asso­ci­
ated with tran­sient infec­tion or drug/toxin expo­sure. As listed
in Table 2, some patients none­the­less have some infec­tion or
some drug/toxin expo­
sure not ini­
tially iden­
ti­
fied. More com­
monly, about half of patients have no appar­ent under­ly­ing cause.
About a quar­ter have an iden­ti­fi­able lymphoproliferative dis­ease
(LPD); as high as 11% of patients with CLL develop wAIHA, with
this num­ber increas­ing after expo­sure to fludarabine.17 Another
quar­ter of all­ patients have an iden­ti­fi­able immune/auto­im­mune
dis­or­der, such as com­mon var­i­able immune defi­ciency, sys­temic
lupus erythematosus, Sjogren’s syn­drome, or auto­im­mune hep­
a­ti­tis. Although 29% of immune throm­bo­cy­to­pe­nia (ITP) patients
Table 2. Causes of chronic wAIHA
Idiopathic: ~45%
Malignancies: LPDs: ~24%
11% of CLL patients
Even higher after fludarabine treat­ment
Immune dis­or­ders: ~25%
Immune defi­ciency dis­or­ders (com­mon var­i­able immune defi­ciency)
Autoimmune dis­or­ders (sys­temic lupus, ulcer­a­tive coli­tis)
~10% of patients with sys­temic lupus
ITP patients (Evans syn­drome)
Others: ~5%
Viral infec­tion (usu­ally kids)
Drugs (huge num­ber)
Data reproduced with per­mis­sion from Hansen et al12 and Tranekær,
Hansen, and Frederiksen.14
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/105/2021916/105kuter.pdf by guest on 09 December 2022
As with many rare dis­eases, esti­ma­tes of the epi­de­mi­­ol­
ogy of wAIHA in adults are inex­act. The inci­dence in adults
ranges from 1.77 per 100 000 in Denmark to 2.4 per 100 000
in France, with a prev­a­lence in Denmark of 17 per 100 000.12,13
The median age of pre­sen­ta­tion in adults is 53 years and 60%
are female.14 The dis­ease inci­dence increases with age, with
0.35 per 100 000 patient-years in those more than 20 years
old; 0.51 per 100 000 patient-years in those 20 to 50 years
old, and 3.34 per 100 000 patient-years in patients over 50.15
have a pos­
i­
tive DAT, few have active hemo­
ly­
sis.18 Evans syn­
drome has an inci­dence of 1.8 per 1 000 000 patient-years, with
the wAIHA and ITP often not con­cur­rent.19
Evaluating the patient for wAIHA
History
Treatment of wAIHA
Who should be treated?
Signs of sig­
nif­i­
cant ane­
mia such as pal­
lor, tachy­
car­
dia, heart
fail­ure, pul­mo­nary edema, and periph­eral edema may be pres­
ent. Asymmetric lower-extrem­ity edema should raise con­cern
for throm­bo­sis. Icterus, par­tic­u­larly of the sclera, is com­monly
found. While mild spleno­meg­aly is com­mon, the spleen is usu­ally
not enlarged enough to be appre­ci­ated by pal­pa­tion. The pres­
ence of sig­nif­i­cant spleno­meg­aly on pal­pa­tion asso­ci­ated with
adenopathy should raise the con­cern of an LPD.
Some patients with min­i­mal ane­mia are suf­fi­ciently com­pen­
sated for their hemo­ly­sis and may not require any ther­apy. Such
asymp­tom­atic patients usu­ally have Hgb level higher than or
equal to 10g/dL and might expe­ri­ence a reduced qual­ity of life
with ther­apy. These patients are followed closely, supplemented
with folate, and advised to con­tact their hema­tol­o­gist should
they expe­ri­ence signs of wors­en­ing hemo­ly­sis (eg, darker urine,
more icterus, wors­en­ing fatigue), espe­cially when stressed by
infec­tion or other med­i­cal prob­lems. Generally, patients who
need treat­ment are symp­tom­atic, have Hgb lev­els lower than
10g/dL, or require RBC trans­fu­sion.
In patients with wAIHA sec­ond­ary to symp­tom­atic LPDs or
immune/auto­im­mune dis­or­ders, ther­apy directed at the under­
ly­ing dis­or­der may mit­i­gate the hemo­ly­sis.
Laboratory exam­i­na­tion
What are the goals of ther­apy for wAIHA?
Physical exam­i­na­tion
Table 3 lists the lab­o­ra­tory assess­ments of patients with wAIHA. A
CBC and a retic­u­lo­cyte count and tests to mea­sure LDH, bil­i­ru­bin,
and hap­to­glo­bin iden­tify those with hemo­lytic ane­mia. The DAT
and elu­ate assess­ment con­firm whether this is a wAIHA. A review
of the periph­eral blood smear is man­da­tory to affirm the pres­
ence of spherocytes and assess for LPDs. Subsequent tests for
under­ly­ing causes include anti­nu­clear anti­body and rheu­ma­toid
fac­tor to assess for auto­im­mune cau­sa­tion and a serum pro­tein
elec­tro­pho­re­sis to assess for chronic var­i­able immune defi­
ciency, as well as flow cytom­e­try and often a chest/­abdo­men/
pel­
vic com­
puted tomo­
graphic scan looking for LPD. Only in
the pres­ence of fea­tures (lymph­ade­nop­a­thy, other cytopenias,
other abnor­mal­i­ties on the periph­eral blood smear) sug­ges­tive
of other hema­to­logic dis­or­ders or in cases of Coombs-neg­a­tive
While desir­able, nor­mal­i­za­tion of Hgb is rarely required. Suc­
cessful ther­apy strives for a sta­ble Hgb level over 10g/dL, res­
o­lu­tion of symp­toms, and trans­fu­sion inde­pen­dence. Treatment
should min­i­mize the adverse effects of immu­no­sup­pres­sion. In
par­tic­u­lar, it is impor­tant to avoid the side effects of cor­ti­co­ste­
roids, with chronic pred­ni­sone dos­age never exceed­ing 5mg/d
and avoided entirely if pos­si­ble. Although sple­nec­tomy may be
a defin­i­tive treat­ment for many, most patients pre­fer to avoid it;
joint deci­sion-mak­ing in wAIHA is man­da­tory.
For clin­
i­
cal tri­
als, the First International Consensus Meet­
ing defined a “response” as an increase in Hgb by more than
2g/dL or nor­mal­i­za­tion of Hgb with­out a bio­chem­i­cal res­o­lu­tion
of hemo­ly­sis, along with an absence of trans­fu­sion for the last
7 days.1 A com­plete response (CR) was defined as nor­mal­i­za­tion
Table 3. Recommended lab­o­ra­tory tests for chronic wAIHA
CBC with dif­fer­en­tial
Folate level
Reticulocyte count
Erythropoietin levela
Review blood smear
Serum pro­tein elec­tro­pho­re­sis
Haptoglobin
ANA, rheu­ma­toid fac­tor
Bilirubin (direct/total)
Flow cytometry for clonal B cellsa
LDH level
Chest/abdo­men/pel­vic CT scan
Coombs test and elu­ate assess­ment
(Bone mar­row exam­i­na­tion)b
Fe/total iron-bind­ing capac­ity/fer­ri­tin
Prothrombin frag­ment 1.2a
B12 level
D-dimer
If test­ing avail­­able.
b
In cer­tain sit­u­a­tions (see text).
ANA, anti­nu­clear anti­body; CT, com­puted tomo­graphic.
a
Warm auto­im­mune hemo­lytic ane­mia: diag­no­sis and treat­ment | 109
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/105/2021916/105kuter.pdf by guest on 09 December 2022
The ini­tial eval­u­a­tion of the patient should include a review of the
signs and symp­toms of ane­mia, the prior bouts of such, and the
pres­ence of jaun­dice or dark-col­ored urine. A per­sonal or fam­ily
his­tory of other auto­im­mune con­di­tions is not uncom­mon. Symp­
toms of LPD such as adenopathy, night sweats, and spleno­meg­aly
should be sought. At this point it is also of par­a­mount impor­tance
to assess for recent infec­tions (eg, bab­e­si­o­sis) and expo­sure to
med­i­ca­tions or tox­ins. Evidence of blood loss ought to be con­sid­
ered. Assess for symp­toms of venous throm­bo­em­bo­lism.
wAIHA is bone mar­
row exam­
i­
na­
tion indi­
cated. In those with
Coombs-neg­a­tive AIHA, a bone mar­row exam­i­na­tion helps doc­
u­ment accel­er­ated eryth­ro­poi­e­sis and may exclude other pri­
mary bone mar­row dis­or­ders. Measurement of lev­els of nutri­ents
and eryth­ro­poi­e­tin assess for ade­quacy of bone mar­row com­
pen­sa­tion. Although their role in predicting throm­botic risk has
not been dem­on­strated, this reviewer rou­tinely checks D-dimer
and a pro­throm­bin frag­ment.
of Hgb, no evi­dence of hemo­ly­sis (nor­mal bil­i­ru­bin, LDH, hap­to­
glo­bin, and retic­u­lo­cytes), and the absence of trans­fu­sions.
What is the ini­tial ther­apy for wAIHA?
Corticosteroids are the ini­
tial ther­
apy for almost all­patients
(Table 4). In patients with sig­nif­i­cant symp­tom­atic ane­mia, RBC
trans­fu­sion needs to be con­sid­ered. All patients should be sup­
plemented with folic acid of at least 0.4mg/d and repleted with
B12 and iron if so defi­cient.
Corticosteroids
Prednisone at 1mg/kg/d for 2 to 4 weeks until the Hgb level
is over 10 is a com­mon starting dose.20,21 Although dexa­meth­a­
sone is some­times used, this reviewer finds its effects to be tran­
sient and can­not rec­om­mend its use. Over 85% of patients have
a rise in Hgb.22,23 Slow ste­roid taper­ing over 5 to 12 weeks to the
low­est effec­tive dose is then under­taken, with close mon­i­tor­ing
(every 1-2 weeks) of Hgb, retic­u­lo­cyte count, LDH, and hap­to­
glo­bin. Pneu­mo­cys­tis cari­nii pneu­mo­nia pro­phy­laxis should be
con­sid­ered in patients who are maintained at pred­ni­sone doses
over 20mg/d for over 1 month. Unfortunately, with ste­roid taper­
Table 4. Treatments for chronic wAIHA
RBC trans­fu­sion
Adequate nutri­tional sup­port
Folate, B12, iron
Reduce anti­body pro­duc­tion
Corticosteroids 1mg/kg/d
Rituximab
MMF
Azathioprine
Cyclophosphamide
Cyclosporine
Danazol
Alemtuzumab
Daratumumab
Reduce mac­ro­phage func­tion
Corticosteroids
Splenectomy
Intravenous immu­no­glob­u­lin
Vincristine
Increase RBC pro­duc­tion
ESAs
110 | Hematology 2022 | ASH Education Program
Subsequent treat­ments for wAIHA
Although most patients have a sig­nif­i­cant response to cor­ti­co­
ste­roids, with taper­ing most patients lose their response and
require addi­tional ther­apy. The many cur­rent ther­apy options
are listed in Table 4, and the most com­mon are discussed sep­
a­rately next.
Rituximab
Rituximab is highly effec­tive in patients resis­tant to or depen­
dent upon cor­ti­co­ste­roids. Standard doses of 375mg/m2 weekly
for 4 weeks or 2 doses of 1000mg 2 weeks apart have been stud­
ied. When added to cor­ti­co­ste­roids, the response rate is much
higher than with cor­ti­co­ste­roids alone. In one study, response
rates of 75% to 80% at 12 months with mostly CRs (defined as a
nor­mal Hgb level on no treat­ment) were reported; the CR rate
with cor­ti­co­ste­roids was only 36%.20 The response dura­tion was
more than 36 months in 70% of patients. In a sec­ond smaller
phase 3 study, over­all responses at 1 year were 31% (95% CI, 11.058.7; 5 CR) for pred­ni­sone alone ver­sus 75% (95% CI, 47.6-92.7;
11 CR, 1 par­tial remis­sion [PR]) for pred­ni­sone plus rituximab.21
After 2 years, 10 in 16 rituximab patients but only 3 in 16 pla­cebo
patients maintained a CR (P = 0.011).21 Eighteen wAIHA patients
in a third study were treated with a brief course of cor­ti­co­ste­
roids plus 100mg of rituximab weekly for 4 weeks. All patients
responded by 6 and 12 months with approx­i­ma­tely 70% CR24;
relapse-free sur­vival at 36 months was 76%.25
Splenectomy
Although often declined by patients, sple­nec­tomy is an effec­
tive treat­ment for wAIHA.26,27 Laparoscopic approaches are pre­
ferred. Presplenectomy vac­ci­na­tion is recommended if there is
ade­quate time. Deep vein throm­bo­sis (DVT) pro­phy­laxis for up
to a month fol­low­ing treat­ment should be con­sid­ered. CR rates
of 75% have been reported in one study and of 81% at 3 years
in another. 26,27 Long-term com­pli­ca­tions of infec­tion and throm­
bo­sis need to be con­sid­ered. Shared deci­sion-mak­ing is encour­
aged.
Intravenous immu­no­glob­u­lin
As with most treat­ments for wAIHA, there are lit­tle data. Intra­
venous immu­no­glob­u­lin doses of 0.4g/kg/d for 5 days have
been stud­ied. In 73 patients the over­all response rate was 29
in 73 (40%).28 A 2-g/dL Hgb rise in 2 weeks was seen in 25 of 62
(40%), and a Hgb rise over 2g/dL in 2 weeks with a Hgb higher
than 10g/dL was seen in 9 of 61 (15%) patients. The treat­ment
response was short-lived.
Mycophenolate
Doses of 500 to 1000mg of MMF twice a day have been stud­
ied. This drug may take 4 to 8 weeks before reaching its full
effect. Pooled data from 2 short series show that all 8 patients
responded with a sig­nif­i­cant rise in Hgb.29,30 MMF is well tol­er­
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/105/2021916/105kuter.pdf by guest on 09 December 2022
RBC trans­fu­sions
In med­i­cally unsta­ble, symp­tom­atic patients, par­tic­u­larly those
with Hgb lev­els lower than 6g/dL, RBC trans­fu­sion is indi­cated.
Appropriately crossmatched blood may not be imme­
di­
ately
avail­­able because of the panagglutinin, and there­fore type-spe­
cific blood is used. In gen­eral, wors­en­ing of hemo­ly­sis is more
likely to occur with alloimmune antibodies than with autoanti­
bodies.
ing, the ini­tial high response is not maintained: 50% maintained
response at 3 months but only 31% to 38% at 12 months. Overall,
nearly two-thirds of cor­ti­co­ste­roid respond­ers require addi­tional
ther­apy.22
In patients fail­ing to respond rap­idly or for those presenting
with severe dis­ease, the addi­tion of rituximab should be con­sid­
ered (see below).
ated except for mild nau­sea and remains the favor­ite immu­no­
sup­pres­sive drug for this reviewer.
Cyclophosphamide
A 50-mg/m2 dose of cyclo­phos­pha­mide daily for 4 days pro­
duced a CR rate in 6 of 9 patients and a PR rate (absence of trans­
fu­sions) in 3 of 9 patients.31 Another study reported a response
rate of 59% of 17 patients.32 Single intra­ve­nous doses of 600 to
750mg/m2 have also been reported as effec­tive.
Erythropoiesis-stim­u­lat­ing agents
Approximately 20% to 40% of patients with AIHA have a
reduced retic­
u­
lo­
cyte response due to antibodies against
retic­u­lo­cytes or reduced eryth­ro­poi­e­sis; this cor­re­lates with a
poor response. In a recent mul­ti­cen­ter pro­spec­tive study, 51
patients with wAIHA had severely reduced eryth­ro­poi­e­tin lev­els
com­pared with other forms of ane­mia; they were treated with
an eryth­ro­poi­e­sis-stim­u­lat­ing agent (ESA) begin­ning at a median
(range) of 24 (0.03-187) months from diag­
no­
sis.35 The over­all
response rate (Hgb >10g/dL or ≥2g/dL increase in the absence
of trans­fu­sions) was 55% at 15 days, 71% at 30 days, 73% at
3 months, 76% at 6 months, and 78% at 12 months. The median
(range) Hgb rise at 6 months was 2.1 (0-2.18) g/dL.
Other agents
Individual case reports and small ret­ro­spec­tive ana­ly­ses have
reported ther­a­peu­tic responses to danazol, cyclo­spor­ine, borte­
zomib, daratumumab, vin­cris­tine, eculizumab, and alemtuzumab.
Putting it all­together
A pro­posed treat­ment algo­rithm can be found in (Figure 4). As
in this case, the ini­tial ther­apy for most patients should be cor­
ti­co­ste­roids, usu­ally in the form of pred­ni­sone at 1mg/kg/d. If
the patient is severely affected or shows no response within
2 weeks, rituximab should be given at a dose of either 375mg/m2
for 4 weeks or 2 sep­
a­
rate weekly doses of 1000mg each.
Corticosteroids are weaned slowly over 8 to 12 weeks, and if an
ade­quate ther­a­peu­tic response (eg, Hgb over 9-10 g/dL) can­
not be maintained at a dose of pred­ni­sone less than or equal to
5mg/d, rituximab (if not already given) is admin­is­tered. If an ade­
quate ther­a­peu­tic response is not obtained after 2 or 3 months,
our pre­ferred approach is 1000mg of MMF twice a day in those
who have dem­on­strated a strong cor­ti­co­ste­roid response. In
those hav­ing never had an ade­quate cor­ti­co­ste­roid response or
fail­ing MMF, an ESA, such as 200 mcg of darbepoetin alfa every
2 weeks, is admin­is­tered for sev­eral months in those with an
endog­e­nous eryth­ro­poi­e­tin level under 500 mIU/mL. Through­
out this treat­ment algo­rithm, the low­est effec­tive dose of pred­
ni­sone is admin­is­tered to main­tain an ade­quate Hgb response;
pred­ni­sone taper­ing is under­taken when­ever another ther­a­
peu­tic modal­ity dem­on­strates a ben­e­fi­cial effect. In patients
refrac­tory to low-dose cor­ti­co­ste­roids, rituximab, immu­no­sup­
pres­sive agents, and ESAs, we reassess the diag­no­sis and indi­
Complications of wAIHA and their treat­ment
Aside from the well-known com­pli­ca­tions of long-term cor­ti­co­ste­
roid treat­ment, 3 other aspects of wAIHA war­rant brief com­ment:
•
•
•
Patients face an increased risk of devel­op­ing LPDs. An LPD
devel­oped in 7 of 71 (10%) of patients with wAIHA after a
median (range) of 26.5 (9-76) months, suggesting that wAIHA
might be the first man­i­fes­ta­tion of a nascent LPD.36
Patients’ risk of throm­bo­em­bo­lism (venous more than arte­rial)
is increased. In a recent study, 15 of 126 (12%) patients devel­
oped throm­bo­sis.37 Increased LDH and trans­
fu­
sions were
asso­ci­ated with increased risk. Thrombosis is of par­tic­u­lar
con­cern fol­low­ing sple­nec­tomy38; appro­pri­ate post­op­er­a­tive
DVT pro­phy­laxis should be pro­vided. Hospitalized patients
should receive rou­tine DVT pro­phy­laxis.
Monitoring for dia­be­tes is impaired. Given the reduced red
cell sur­vival, hemo­glo­bin A1c lev­els do not pro­vide accu­rate
glu­cose mon­i­tor­ing.39
Newer ther­a­pies in devel­op­ment for wAIHA
wAIHA is a frus­trat­ing dis­or­der, and many patients remain on
inap­pro­pri­ately high doses of cor­ti­co­ste­roids for years. Newer
ther­a­pies are needed.40
Figure 4. Proposed treatment algorithm for wAIHA.
Warm auto­im­mune hemo­lytic ane­mia: diag­no­sis and treat­ment | 111
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/105/2021916/105kuter.pdf by guest on 09 December 2022
Azathioprine
wAIHA is not uncom­mon in dogs; 87 of 93 dogs responded to
aza­thi­o­prine and sur­vived.33 In humans, doses of 100 to 200mg
daily have been used, with reported response rates of 26 of
33 (79%) in 2.2 months in one study and 7 of 8 (1 CR, 6 PR) in
another.23,32 With dis­con­tin­u­a­tion of the drug, 42% relapsed.32
To avoid severe hema­
to­
log­
i­
cal tox­
ic­
ity, patients need to be
screened for deficiencies of thiopurine methyltransferase.34
Immunoproteasome inhib­i­tors such as bortezomib and KZR–
616 may reduce auto­an­ti­body for­ma­tion, as might CD38 inhib­
i­tors such as daratumumab. Drugs that block the neo­na­tal Fc
recep­tor shorten IgG half-life and reduce anti-RBC anti­body.40
Inhibitors of medi­a­tors of phago­cy­to­sis such as syk kinase (fos­
tamatinib) and Bruton kinase (rilzabrutinib) are cur­rently being
stud­ied.41,42 Fostamatinib increased Hgb to a level above 10g/dL,
with an increase of 2g/dL in 11 of 24 (46%) refrac­tory wAIHA
patients; Hgb improve­ment was often seen after 2 weeks and
was maintained with few adverse events.43 Clinical tri­als are also
under­way to assess the effi­cacy of inhib­i­tors of com­ple­ment fac­
tors C1 and C3.
ca­tions for treat­ment. Our sub­se­quent pref­er­ence is for clin­i­cal
tri­als, such as those with fostamatinib, rilzabrutinib, or a neo­na­tal
Fc recep­tor inhib­i­tor. While not averse to sple­nec­tomy, this is
usu­ally reserved for those highly refrac­tory to the ther­a­peu­tic
modal­i­ties described above and for whom clin­i­cal tri­als are not
appro­pri­ate.
Acknowledgment
I am pleased to thank Dr Drew Provan for the art­istry in Figure 2
and the Visual Abstract.
Conflict-of-inter­est dis­clo­sure
Off-label drug use
David J. Kuter: none of the pharmacologic therapies listed have
received regulatory approval for the treatment of wAIHA.
Correspondence
David J. Kuter, Hematology Division, Mas­sa­chu­setts General Hos­
pital, Ste 118, Rm 110, Zero Emerson Pl, Bos­ton, MA 02114; e-mail:
dkuter@MGH​­.harvard​­.edu.
References
1.
Jäger U, Barcellini W, Broome CM, et al. Diagnosis and treat­ment of auto­
im­mune hemo­lytic ane­mia in adults: rec­om­men­da­tions from the First Inter­
national Consensus Meeting. Blood Rev. 2020;41:100648.
2. Wagner KH, Shiels RG, Lang CA, Seyed Khoei N, Bulmer AC. Diagnostic
cri­te­ria and con­trib­u­tors to Gilbert’s syn­drome. Crit Rev Clin Lab Sci.
2018;55(2):129-139.
3. Trudel G, Shahin N, Ramsay T, Laneuville O, Louati H. Hemolysis con­trib­utes
to ane­mia dur­ing long-dura­tion space flight. Nat Med. 2022;28(1):59-62.
4. Marchand A, Galen RS, Van Lente F. The pre­dic­tive value of serum hap­to­
glo­bin in hemo­lytic dis­ease. JAMA. 1980;243(19):1909-1911.
5. Galen RS. Application of the pre­dic­tive value model in the anal­y­sis of test
effec­tive­ness. Clin Lab Med. 1982;2(4):685-699.
6. Coombs RR, Mourant AE, Race RR. A new test for the detec­tion of weak and
incom­plete Rh agglu­ti­nins. Br J Exp Pathol. 1945;26:255-266. https:​­/​­/www​
­.ncbi​­.nlm​­.nih​­.gov​­/pmc​­/articles​­/PMC2065689​­/pdf​­/brjexppathol00160​
­-0040​­.pdf. Accessed 26 August 2022.
7. Coombs RR, Edin B, Mourant AE, Race RR. Detection of weak and “incom­
plete” Rh agglu­ti­nins: a new test. Lancet. 1945;246(6358):15-16.
8. Zarandona JM, Yazer MH. The role of the Coombs test in eval­u­at­ing hemo­
ly­sis in adults. CMAJ. 2006;174(3):305-307.
9. Mauro FR, Foa R, Cerretti R, et al. Autoimmune hemo­lytic ane­mia in chronic
lym­pho­cytic leu­ke­mia: clin­i­cal, ther­a­peu­tic, and prog­nos­tic fea­tures.
Blood. 2000;95(9):2786-2792.
10. Barcellini W, Fattizzo B. The chang­ing land­scape of auto­im­mune hemo­lytic
ane­mia. Front Immunol. 2020;11:946.
11. Hill QA, Horan M, Charlton A, et al. Developing the evi­dence base for the
man­age­ment of auto­im­mune haemolytic anae­mia (AIHA): the UK expe­ri­
ence. Br J Haematol. 2021;192(2):e54-e57.
12. Hansen DL, Möller S, Andersen K, Gaist D, Frederiksen H. Increasing inci­
dence and prev­a­lence of acquired hemo­lytic ane­mias in Denmark, 19802016. Clin Epidemiol. 2020;12:497-508.
13. Maquet J, Lafaurie M, Walter O, et al. Epidemiology of auto­im­mune hemo­
lytic ane­mia: a nation­wide pop­u­la­tion-based study in France. Am J Hematol. 2021;96(8):E291-E293.
112 | Hematology 2022 | ASH Education Program
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/105/2021916/105kuter.pdf by guest on 09 December 2022
David J. Kuter: research funding: Argenx, Biocryst, Immunovant,
Rigel, Sanofi (Principia), Takeda (Bioverativ), UCB; con­sul­tancy:
Alexion (Syntimmune), Amgen, Argenx, BioCryst, Bristol Myers
Squibb, Caremark, Cellphire, Cellularity, CRICO, Daiichi Sankyo,
Hengrui, Immunovant, Incyte, Kyowa-Kirin, Merck Sharp Dohme,
Momenta, Novartis, Pfizer, Platelet Biogenesis, Platelet Disorder
Support Association, Rigel, Sanofi (Bioveratif), Sanofi (Genzyme),
Sanofi (Principia), Sobi (Dova), Takeda, UCB, Up-To-Date; stock
own­er­ship: Rubius.
14. Tranekær S, Hansen DL, Frederiksen H. Epidemiology of sec­ond­ary warm
auto­im­mune haemolytic anae­mia—a sys­tem­atic review and meta-anal­y­sis.
J Clin Med. 2021;10(6).
15. Sokol RJ, Hewitt S, Stamps BK. Autoimmune haemolysis: an 18-year study
of 865 cases referred to a regional trans­fu­sion cen­tre. Br Med J (Clin Res Ed)
1981;282(6281):2023-2027.
16. Engelfriet CP, Overbeeke MA, von dem Borne AE. Autoimmune hemo­lytic
ane­mia. Semin Hematol. 1992;29(1):3-12.
17. Barcellini W, Fattizzo B, Zaninoni A. Management of refrac­tory auto­im­mune
hemo­lytic ane­mia after allo­ge­neic hema­to­poi­etic stem cell trans­plan­ta­
tion: cur­rent per­spec­tives. J Blood Med. 2019;10(August):265-278.
18. Hollenhorst MA, Al-Samkari H, Kuter DJ. Markers of auto­im­mu­nity in immune
throm­bo­cy­to­pe­nia: prev­a­lence and prog­nos­tic sig­nif­i­cance. Blood Adv.
2019;3(22):3515-3521.
19. Hansen DL, Möller S, Andersen K, Gaist D, Frederiksen H. Evans syn­drome
in adults—inci­dence, prev­a­lence, and sur­vival in a nation­wide cohort. Am
J Hematol. 2019;94(10):1081-1090.
20. Birgens H, Frederiksen H, Hasselbalch HC, et al. A phase III ran­dom­ized
trial com­par­ing glu­co­cor­ti­coid monotherapy ver­sus glu­co­cor­ti­coid and rit­
uximab in patients with auto­im­mune haemolytic anae­mia. Br J Haematol.
2013;163(3):393-399.
21. Michel M, Terriou L, Roudot-Thoraval F, et al. A ran­dom­ized and dou­bleblind con­trolled trial eval­u­at­ing the safety and effi­cacy of rituximab for
warm auto-immune hemo­lytic ane­mia in adults (the RAIHA study). Am J
Hematol. 2017;92(1):23-27.
22. Roumier M, Loustau V, Guillaud C, et al. Characteristics and out­
come of warm auto­
im­
mune hemo­
lytic ane­
mia in adults: new insights
based on a sin­
gle-cen­
ter expe­
ri­
ence with 60 patients. Am J Hematol.
2014;89(9):E150-E155.
23. Prabhu R, Bhaskaran R, Shenoy V, Reema G, Sidharthan N. Clinical char­ac­
ter­is­tics and treat­ment out­comes of pri­mary auto­im­mune hemo­lytic ane­
mia: a sin­gle cen­ter study from South India. Blood Res. 2016;51(2):88-94.
24. Barcellini W, Zaja F, Zaninoni A, et al. Low-dose rituximab in adult patients
with idi­o­pathic auto­im­mune hemo­lytic ane­mia: clin­i­cal effi­cacy and bio­
logic stud­ies. Blood. 2012;119(16):3691-3697.
25. Barcellini W, Zaja F, Zaninoni A, et al. Sustained response to low-dose
rituximab in idi­o­pathic auto­im­mune hemo­lytic ane­mia. Eur J Haematol.
2013;91(6):546-551.
26. Barcellini W, Fattizzo B, Zaninoni A, et al. Clinical het­ero­ge­ne­ity and pre­
dic­tors of out­come in pri­mary auto­im­mune hemo­lytic ane­mia: a GIMEMA
study of 308 patients. Blood. 2014;124(19):2930-2936.
27. Giudice V, Rosamilio R, Ferrara I, Sen­eca E, Serio B, Selleri C. Efficacy and
safety of sple­nec­tomy in adult auto­im­mune hemo­lytic ane­mia. Open Med
(Wars). 2016;11(1):374-380.
28. Flores G, Cunningham-Rundles C, Newland AC, Bussel JB. Efficacy of intra­
ve­nous immu­no­glob­u­lin in the treat­ment of auto­im­mune hemo­lytic ane­
mia: results in 73 patients. Am J Hematol. 1993;44(4):237-242.
29. Howard J, Hoffbrand AV, Prentice HG, Mehta A. Mycophenolate mofetil for
the treat­ment of refrac­tory auto-immune haemolytic anae­mia and autoimmune throm­bo­cy­to­pe­nia pur­pura. Br J Haematol. 2002;117(3):712-715.
30. Kotb R, Pinganaud C, Trichet C, et al. Efficacy of mycophenolate mofetil
in adult refrac­tory auto-immune cytopenias: a sin­gle cen­ter pre­lim­i­nary
study. Eur J Haematol. 2005;75(1):60-64.
31. Moyo VM, Smith D, Brodsky I, Crilley P, Jones RJ, Brodsky RA. High-dose
cyclo­phos­pha­mide for refrac­tory auto­im­mune hemo­lytic ane­mia. Blood.
2002;100(2):704-706.
32. Hantaweepant C, Pairattanakorn P, Karaketklang K, Owattanapanich
W, Chinthammitr Y. Efficacy and safety of sec­ond-line treat­ment in Thai
patients with pri­mary warm-type auto­im­mune hemo­lytic ane­mia. Hematology. 2019;24(1):720-726.
33. Piek CJ, Junius G, Dekker A, Schrauwen E, Slappendel RJ, Teske E. Idiopathic
immune-medi­ated hemo­lytic ane­mia: treat­ment out­come and prog­nos­tic
fac­tors in 149 dogs. J Vet Intern Med. 2008;22(2):366-373.
34. Lennard L. Implementation of TPMT test­
ing. Br J Clin Pharmacol.
2014;77(4):704-714.
35. Fattizzo B, Michel M, Zaninoni A, et al. Efficacy of recom­bi­nant eryth­ro­poi­
e­tin in auto­im­mune hemo­lytic ane­mia: a mul­ti­cen­ter inter­na­tional study.
Haematologica. 2021;106(2):622-625.
36. Sallah S, Wan JY, Hanrahan LR. Future devel­op­ment of lymphoproliferative
dis­or­ders in patients with auto­im­mune hemo­lytic ane­mia. Clin Cancer Res.
2001;7(4):791-794.
37. Fattizzo B, Barcellini W. Autoimmune hemo­lytic ane­mia: causes and con­se­
quences. Expert Rev Clin Immunol. 2022;18(7):731-745.
38. Crary SE, Buchanan GR. Vascular com­
pli­
ca­
tions after sple­
nec­
tomy for
hema­to­logic dis­or­ders. Blood. 2009;114(14):2861-2868.
39. Lum G. Artefactually low hemo­glo­bin A1C in a patient with hemo­lytic ane­
mia. Lab Med. 2010;41(5):267-270.
40. Kuter DJ. Novel ther­a­pies for immune throm­bo­cy­to­pe­nia. Br J Haematol.
2022;196(6):1311-1328.
41. Podolanczuk A, Laz­a­rus AH, Crow AR, Grossbard E, Bussel JB. Of mice and
men: an open-label pilot study for treat­ment of immune throm­bo­cy­to­pe­
nic pur­pura by an inhib­i­tor of Syk. Blood. 2009;113(14):3154-3160.
42. Langrish CL, Bradshaw JM, Francesco MR, et al. Preclinical effi­cacy and antiinflam­ma­tory mech­a­nisms of action of the Bruton tyro­sine kinase inhib­i­tor
rilzabrutinib for immune-medi­ated dis­ease. J Immunol. 2021;206(7):14541468.
43. Kuter DJ, Rogers KA, Boxer MA, et al. Fostamatinib for the treat­ment of
warm anti­body auto­im­mune hemo­lytic ane­mia: phase 2, mul­ti­cen­ter,
open-label study. Am J Hematol. 2022;97(6):691-699.
© 2022 by The Amer­i­can Society of Hematology
DOI 10.1182/hema­tol­ogy.2022000405
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/105/2021916/105kuter.pdf by guest on 09 December 2022
Warm auto­im­mune hemo­lytic ane­mia: diag­no­sis and treat­ment | 113
Dr Prakash Singh Shekhawat
BEYOND ROUTINE FRONTLINE THERAPY OF CML
Christian Niederwieser and Nicolaus Kröger
Department of Stem Cell Transplantation, University Medical Center Hamburg-Eppendorf, Germany
Molecular therapy with tyrosine kinase inhibitors (TKIs) has significantly reduced the indication for allogeneic hematopoietic stem cell transplantation (allo-HSCT) in chronic myeloid leukemia (CML). Treatment-free remission can be obtained
in about 50% of patients with an optimal response. However, cure rates up to 90% are restricted to patients receiving
HSCT. Timing is essential since HSCT in the early stages of the disease has the best outcome. Patients in a more advanced
phase (AdP) than chronic-phase (chP) CML undergo HSCT with suboptimal outcomes, and the gap between chP and AdP
disease is widening. First-line therapy should start with first- or second-generation (G) TKIs. Patients failing treatment
(BCR-ABL1 transcripts of greater than 10% at 3 and 6 months and greater than 1% at 12 months) should be switched to
second-line TKIs, and HSCT should be considered. Patients not responding to 2G-TKI therapy as well as patients in an
accelerated phase (AP) or blast crisis (BC) are candidates for HSCT. Therapy resistant BCR-ABL1 mutations, high-risk additional cytogenetic abnormalities, and molecular signs of leukemia progression should trigger the indication for HSCT.
Patients who, despite dose adjustments, do not tolerate or develop severe adverse events, including vascular events, to
multiple TKIs are also candidates for HSCT. In AdP CML, TKIs do not show long-lasting results, and the outcome of HSCT
is less optimal without pretransplant therapy. In these patients the induction of chP2 with TKIs, either alone (AP) or in
combination with intensive chemotherapy (BC), followed by HSCT should be pursued.
LEARNING OBJECTIVES
• Understand the criteria for considering allogeneic stem cell transplantation for CML patients
• Define the risk factors for outcomes after allogeneic stem cell transplantation for CML patients
• Acquire the knowledge of how to use TKIs before and after transplantation in CML patients
Introduction
Chronic myeloid leukemia (CML), a disease of predominantly older adult (age >60 years) and male patients, is in
many aspects a model for malignant diseases. Described
by Virchow and Bennett in 1845 as leucocythemia, it was
the first malignancy with a common chromosomal alteration.1,2 After molecular identification of the breakpoints,
the main molecular pathomechanisms were unraveled.3,4
Hematopoietic stem cell transplantation (HSCT) was
shown to be the only curative option.5 Almost at the same
time, treatment with interferon alfa was shown to induce
cytogenetic remissions (CyR) in a small proportion of
patients.6 The use of BCR-ABL1 transcripts for disease monitoring and donor lymphocyte infusion (DLI) to treat relapse
after HSCT were the next hallmarks in the treatment of this
disease.7,8 Finally, in 2001 a specific inhibitor of BCR-ABL
tyrosine kinase revolutionized the treatment of CML.9 In
2010, the discontinuation of TKIs in patients with optimal
114 | Hematology 2022 | ASH Education Program
response was described and the term “operational cure”
used.10
After decades of development, HSCT was increasingly
used worldwide, and more than 1.5 million transplants (now
90 000 annually) were reported up to 2019.11 Patients with
early chronic-phase (chP) CML were considered an ideal
indication for HSCT.12 Interval-diagnosis HSCT (<12 months
and >12 months), disease stage (chP, accelerated phase
[AP], and blast crisis [BC]), donor (related and unrelated),
and gender differences were recognized as major prognostic factors influencing transplant-related mortality (TRM;
nonrelapse mortality caused by graft-versus-host disease
[GVHD], infections, or toxicities), and relapse incidence
(European Society for Blood and Marrow Transplantation
[EBMT] risk score).12 Alternative donors became available in 1989 using cord blood and haploidentical donors
by ex vivo or in vivo T-cell depletion.13,14 In 1998 reducedintensity preparative regimens for older adults and patients
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/114/2021884/114niederwieser.pdf by guest on 09 December 2022
Transplantation in CML in the TKI era:
who, when, and how?
with comorbidities were devel­oped.15 Today, TRM of 5% to 10%
in chP CML and 20% to 25% in advanced phase (AdP) CML
remain, and GVHD rep­re­sents a major com­pli­ca­tion after HSCT.
After the avail­abil­ity of TKIs, HSCT activ­ity in the EBMT reg­is­try
con­tin­u­ously decreased for chP but remained sta­ble for AP/BC
(Figure 1). Nowadays, more patients in non-chP than in chP are
transplanted world­wide,11 which under­lines the impor­tance of
opti­miz­ing out­come by improv­ing the tim­ing of HSCT.
Oral TKI ther­apy in patients with chP CML resulted in an over­
all sur­vival (OS) of 80% at 10 years, but only approx­i­ma­tely 50%
remained on first-gen­er­a­tion (1G) treat­ment.9,16 After first, sec­
ond, and third G-TKIs became avail­­able, major molec­u­lar remis­
sions (MMR) and even treat­ment-free remis­sions were observed.17
Despite this suc­cess, a sig­nif­i­cant pro­por­tion of patients become
refrac­tory or prog­ress, become intol­er­ant, or develop cytopenia. Approximately 31% of patients switch from first-line to sec­
ond-line ther­apy after 11 months.18 Reasons for switching include
fail­ure in 32%, intol­er­ance in 57%, and other issues in 12% of
patients. Recently, asciminib has dem­on­strated a higher MMR
rate in combination with a more favorable safety profile as bosutinib.19 Despite being well tol­er­ated over­all, side effects on TKIs,
espe­cially in higher G-TKIs, occur (eg, vasculopathy up to 53%
of patients on nilotinib, vas­cu­lar events in 37% on ponatinib, and
pleu­ral effu­sion in approx­i­ma­tely 20% on dasatinib).20,21 Other
stud­ies showed an intol­er­ance of 16% or 7% on treat­ment with
dasatinib (5 years) and/or imatinib (5 and 10 years).16,22
Despite HSCT being the only cura­
tive treat­
ment, tim­
ing
assumes an essen­tial role in this pro­cess. Decisions should not
be taken too early and incur unnec­es­sary risks or too late, which
could jeop­ar­dize the out­come after HSCT. Results in AdP CML
(espe­cially in BC) are poor with sin­gle TKI ther­apy and sin­gle HSCT
(see below). Recommendations from experts in the field and published results are of fun­da­men­tal impor­tance in this pro­cess.
CLINICAL CASE 1
A 38-year-old con­
struc­
tion worker presented with sud­
den
fatigue, nose­bleeds, and pneu­mo­nia. He had leu­ko­cy­to­sis
(112 × 109/L), throm­bo­cy­to­pe­nia (15 × 109/L), and ane­mia (hemo­
glo­
bin level of 8.4 g/dL). Initially, an acute mye­
loid leu­
ke­
mia (AML) was suspected after the bone mar­row aspi­ra­tion
revealed 97% blasts with a mye­loid phe­no­type. In addi­tion, a
Philadelphia chro­mo­some (Ph) in 20 out of 20 meta­phases and
a com­plex aber­rant kar­yo­type (includ­ing mono­somy 7) were
detected. A T315I muta­tion was described in the molec­u­lar ana­
ly­ses and a blast cri­sis (BC) of CML diag­nosed.
CLINICAL CASE 2
At a rou­tine checkup, a 60-year-old woman had 30×109/L leu­
ko­cytes, 400×109/L plate­lets, and a hemo­glo­bin level within a
nor­mal range. She cur­rently expe­ri­enced short­ness of breath but
oth­er­wise felt healthy. After refer­ral to a hema­tol­o­gist, a left shift
was found in the dif­fer­en­tial but no enlarge­ment of the lymph
nodes or spleen. Since the Ph chro­mo­some was detected in all­
meta­phases upon bone mar­row aspi­ra­tion, chP CML and a lowrisk EUTOS long-term-sur­vival-score was diag­nosed.
Who should undergo trans­plan­ta­tion
Outcomes of HSCT have improved sub­stan­tially dur­ing the last
decades. In a real-world anal­y­sis, the Swed­ish reg­is­try reported
an OS of 96% at 5 years in chP1 in patients with a median age
spec­
tive
of 43 years using matched donors.23 In another pro­
study, the Ger­man CML study group reported 3-year OS rates
Dr Prakash Singh Shekhawat
HSCT in CML | 115
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/114/2021884/114niederwieser.pdf by guest on 09 December 2022
Figure 1. Development of allogeneic stem cell transplantation for CML in Europe from 1990 to 2020 (EBMT registry) showing chP
and AdP CML (AP and BC).
con­sid­ered for HSCT, while de novo AP CML might become eli­gi­
ble for HSCT if the response to TKIs is not opti­mal.
Recommendations according to the ELN and to the NCCN
regard­ing allo­ge­neic stem cell trans­plan­ta­tion are outlined in
Table 1.
Although AdP CML occurs in a minor­ity of patients (de novo
10%; 5% on dasatinib and 7% on imatinib develop AdP after
5 years), out­
comes are infe­
rior to chP CML after HSCT and
TKIs alone. Outcomes of BC HSCT in the pre-imatinib era were
reported to be only 21% OS at 2 years.28 1G-TKIs in patients with
BC resulted in a median OS of 7 to 10 months, while treat­ment
with 2G-TKIs (nilotinib or dasatinib) resulted in an OS of 32%
and 30% at 2 years, respec­tively. In a ret­ro­spec­tive study with
104 patients, 1-3G-TKIs plus inten­sive che­mo­ther­apy (IC) and
TKIs plus hypomethylating agents led to a higher rate of CRi
(57.5% vs 33%), a higher com­plete CyR rate (45% vs 10.7%), and
more patients pro­ceed­ing to HSCT (32.5% vs 10.7%) than TKIs
or IC alone. Long-term results were sim­i­lar in the com­bi­na­tions
and clearly infe­rior to TKIs or IC alone (OS, 30%-28% vs 13%-0%
at 5 years).31 HSCT resulted in long-term OS in patients with
advanced CML (34% CI, 23-46, at 15 years). OS was improved in
non-BC patients at HSCT with donors 36 years of age or youn­
ger and with a higher CD34+ cell dose in the graft.32 The ELN
and NCCN pro­vide infor­ma­tion on induc­tion che­mo­ther­apy
according to AML-based mor­phol­ogy and acute lym­pho­cytic
leu­ke­mia treat­ment. The ELN empha­sizes the need to attempt
to return to chP CML with sub­se­quent HSCT with­out delay and
that patients with untreated BC should not undergo HSCT. A
study found that in patients in remis­sion for BC, con­ven­tional
risk fac­tors such as advanced age, poor per­for­mance sta­tus,
a lon­ger inter­val from diag­no­sis to HSCT, myeloablative con­
di­
tion­
ing (MAC), and unre­
lated donors remained the major
deter­
mi­
nants of out­
come, whereas in those with active BC
at trans­plant, unre­lated donor trans­plan­ta­tion was asso­ci­ated
with prolonged leu­ke­mia-free sur­vival (LFS).33 Similar results for
advanced CML in the AP or BC or pretreated with TKIs beyond
Table 1. Recommendations for allo­ge­neic stem cell trans­plan­ta­tion in CML according to the ELN and the NCCN
Chronic phase (chP)
Accelerated phase (AP)
Blast phase (BC)
ELN 2020
- Disease resis­tant or intol­er­ant (sub­op­ti­mal
response to 2 or more TKIs)
- For the very rare patient with inad­e­quate
recov­ery of nor­mal hema­to­poi­e­sis
- Resistance to 2G-TKIs (first or sec­ond line)
ponatinib or exper­i­men­tal agent
- Failure to respond to ponatinib after
3 months’ treat­ment
- Emergence of high-risk cyto­ge­net­ics:
observe closely, con­sider inten­si­fi­ca­tion of
treat­ment (ponatinib, early allo-SCT)
ELN 2020
- A patient presenting in AP should be treated
as a high-risk patient, becom­ing eli­gi­ble for
HSCT if the response is not opti­mal
- A patient progressing to AP dur­ing treat­ment
should imme­di­ately be con­sid­ered for HSCT
ELN 202025
- Attempt at return to chP2
- Addition of che­mo­ther­apy based on AML
reg­i­mens for mye­loid BP (such as dasatinib
or ponatinib + FLAG-IDA) or ALL reg­i­mens for
lym­phoid BP (such as imatinib or dasatinib +
hyperfractionated CVAD) recommended
- After CP2 is achieved pro­ceed to allo-SCT
with­out delay
- Transplantation in active BP is not
recommended
NCCN guide­lines26
- If TKI-resis­tant dis­ease BCR-ABL1 (IS) >10%
at >3 months, switch to alter­nate TKI and
eval­u­ate for HSCT
NCCN guide­lines26
- Disease pro­gres­sion to AP while on TKI
ther­apy should be con­sid­ered for HSCT
- Patients who pres­ent with AP at diag­no­sis
should be treated with a TKI, followed by
eval­u­a­tion for allo­ge­neic HSCT based on
response to ther­apy after 3, 6, or 12 months
NCCN guide­lines26
- Recommendation does not depend on
response
- After ther­apy with mor­phol­ogy-based
induc­tion che­mo­ther­apy + TKI in lym­phoid
and mye­loid blast cri­sis or TKI plus ste­roids in
lym­phoid blast cri­sis and sole TKI in mye­loid
blast cri­sis
25
25
CVAD, chemotherapy combination used to treat some types of acute lymphoblastic leukemia (ALL) and non-Hodgkin lymphoma (NHL). Hyper-CVAD
includes the drugs cyclophosphamide, vincristine sulfate, doxorubicin hydrochloride (Adriamycin), and dexamethasone; FLAG-IDA, fludarabine, highdose cytosine arabinoside (AraC), idarubicin, and granulocyte colony-stimulating factor (G-CSF); IS, international scale.
116 | Hematology 2022 | ASH Education Program
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/114/2021884/114niederwieser.pdf by guest on 09 December 2022
of 88% and 94% after elec­tive HSCT in high-risk chP (but a 0-1
EBMT score) and in imatinib-fail­ure chP CML patients, respec­
tively. TRM of only 8% was described in this mul­ti­cen­ter study,
which com­pares favor­ably to TRM of 26% in an ear­lier ran­dom­
ized study after IFN-based treat­ment in chP CML.24 TKI ther­apy
with only imatinib led approx­i­ma­tely 96% OS at 3 years in the
same study.24 The Euro­pean LeukemiaNet (ELN) cri­te­ria and the
US National Comprehensive Cancer Network (NCCN) guide­
lines are updated on a reg­u­lar basis and guide the treat­ment
not only in chP CML. The dif­fer­ence between the 2 involves the
def­i­ni­tion of response to ther­apy. Transcripts higher than 10%
at 3 months define patients with pos­si­ble TKI-resis­tant dis­ease
and after 3 months with def­i­nite TKI-resis­tant dis­ease (NCCN
guide­lines). In the ELN guide­lines, tran­scripts higher than 10%
are defined as a warn­ing at 3 months and if con­firmed within the
next 3 months as fail­ure. Failure means chang­ing to another TKI
if ther­apy started with 1G-TKI but to assess­ment for HSCT only
if ther­apy started with 2G-TKI. In NCCN guide­lines, pos­si­ble TKI
resis­tance auto­mat­i­cally leads to eval­u­a­tion for HSCT.25,26 On the
other hand, the ELN rec­om­men­da­tions spec­ify resis­tance to 2GTKI as an indi­ca­tion for HSCT.
HSCT results in patients with AP CML are clearly infe­rior to
those in chP1. Before the TKI era, out­comes of HSCT in AP CML
were 35% at 2 years and in the imatinib era, 59% at 2 years (com­
pared to 88% and 94% in early chP CML).27,28 In a pro­spec­tive
study, Jiang et al dem­on­strated an advan­tage for HSCT over TKI
in AP CML (6-year OS, 83.3% vs 51.4%).29 Similar results of 50% to
60% OS at 5 years after imatinib, nilotinib, and dasatinib (60%70% OS and 10% MMR at 2 years) and bosutinib (60% OS and 11%
MMR at 4 years) were reported with TKI monotherapy.30 Ponatinib had slightly higher response rates (84% OS and 34% MMR at
1 year), but ran­dom­ized com­par­i­sons are lacking. Patients with
de novo AP CML treated with nilotinib or dasatinib (70% MMR
and 90% OS at 3 years, respec­tively) have been reported to have
supe­rior results than AP CML that devel­ops while on ther­apy.30
Patients with AP CML under treat­ment should imme­di­ately be
The patient was treated with an AML-like induc­tion che­mo­ther­
apy in com­bi­na­tion with ponatinib, a 3G-TKI to over­come the
resis­tance to the T315I muta­tion. After reaching chP, the patient
was referred to an HSCT pro­gram.
at 3 and 6 months of treat­ment and higher than 1% at 12 months
are con­sid­ered treat­ment fail­ures and should lead to sec­ond-line
TKI treat­ment (either 2G- or 3G-TKI or a TKI respon­sive to the
detected muta­tion). A donor search should be ini­ti­ated upon
resis­tance to 2G-TKI treat­ment.25 Approximately 65% respond
to sec­ond-line treat­ment with BCR-ABL1 tran­scripts of less than
or equal to 10%. Nonresponding patients (>10% tran­scripts at
3 months) have dura­ble CyR of only 50% at 4 years.51
HSCT in early chP CML has been shown to have the best out­
come, even after first-line TKI resis­tance.12 In addi­tion, TKI ther­
apy before HSCT has been asso­ci­ated with bet­ter posttransplant
out­comes.32 A Center for International Blood and mar­row Transplant Research (CIBMTR) study con­firmed the ben­e­fi­cial effect of
pretransplant TKIs in chP CML for posttransplant sur­vival.53 Delaying HSCT to late chP CML incurs the risk of pro­gres­sion to AdP
dis­ease, with unfa­vor­able results. Patients not eli­gi­ble for HSCT
(>80 years of age) should con­tinue on 3G-TKIs or new inhib­i­tors
in the test­ing phase. In pedi­at­ric patients, life­long TKI ther­apy,
includ­ing side effects, needs to be bal­anced with an increased
HSCT TRM at the begin­ning and a pos­si­ble increased mor­bid­ity
with chronic GVHD.54
The goal in advanced CML is a return to a chP followed by
HSCT. The choice of pretransplant TKIs for AdP CML is not well
stan­
dard­
ized, but dasatinib and nilotinib have at least safely
been admin­
is­
tered before HSCT with­
out increased TRM.50,55
Thus, it is appro­pri­ate to use TKIs to reduce the dis­ease bur­den
before HSCT for AP CML. The search for a donor should be ini­ti­
ated at diag­no­sis and HSCT planned after a response (Table 2). In
BC, a down­grade to chP2 should be con­sid­ered a clin­i­cal con­di­
tion requir­ing, when­ever pos­si­ble, treat­ment with a com­bi­na­tion
of TKIs with IC or TKIs with hypomethylating agents followed
by HSCT.31 The com­bi­na­tion of TKIs plus IC with HSCT has been
shown to result in 54% OS at 2 years in AdP CML.27 HSCT remains
a unique ther­a­peu­tic option for patients in chP CML after the fail­
ure of 2 TKIs or in those poten­tially har­bor­ing the T315I muta­tion
(after a trial of ponatinib ther­apy).56
CLINICAL CASE 2 (Con­t in­u ed)
CLINICAL CASE 1 (Con­tin­ued)
CLINICAL CASE 1 (Con­t in­u ed)
After ther­apy with hydroxy­urea, the patient was started on the
1G-TKI imatinib. Her response according to ELN cri­te­ria was
opti­mal after 3 months. After she discussed the pos­si­bil­ity of
HSCT with her hema­tol­o­gist, it was decided to con­tinue TKI
ther­apy under con­tin­u­ous BCR-ABL1 mon­i­tor­ing.
Timing: when to per­form a trans­plant
It is gen­er­ally accept­able to start first-line ther­apy with TKIs
(first or sec­ond gen­er­a­tion) in chP CML with­out affect­ing the
out­
come of HSCT.25 The response and tol­
er­
ance to the TKI
deter­mine fur­ther treat­ment. The response to the TKI depends
on the EUTOS long-term sur­vival score and the pres­ence of
high-risk cyto­ge­net­ics and TKI muta­tions.51 The pres­ence of the
T315I muta­tion, which is respon­sive to ponatinib, is a trig­ger
for HSCT.
The ELN pro­vi­des reg­u­lar rec­om­men­da­tions on how to man­
age TKI treat­ment according to BCR-ABL1 tran­script reduc­tion
at defined time points.52 BCR-ABL1 tran­scripts higher than 10%
After weighing the pros and cons, the patient decided to
undergo HSCT from the HLA-matched brother at a JACIE-cer­ti­fied
trans­plant cen­ter. Three months after diag­no­sis, he under­went
periph­eral blood HSCT.
CLINICAL CASE 2 (Con­tin­ued)
After 2 years of opti­mal response, the patient showed increased
BCR-ABL1 tran­scripts. Molecular ana­ly­ses detected V299L resis­
tant to dasatinib and Y253H resis­tant to nilotinib, but both
were sen­si­tive only to ponatinib. Treatment with ponatinib was
started. Since her 70-year-old brother, her only sib­ling, had
CLL, an unre­lated donor search was ini­ti­ated. Following intol­er­
ance to ponatinib (caus­ing head­aches and vomiting) but after
achiev­ing BCR-ABL1 neg­a­tiv­ity and lacking access to asciminib,
an HSCT was con­sid­ered.
Dr Prakash Singh Shekhawat
HSCT in CML | 117
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/114/2021884/114niederwieser.pdf by guest on 09 December 2022
front­line ther­apy have been reported by other inves­ti­ga­tors
and are listed in Table 2.34-41 The emer­gence of high-risk addi­
tional chro­mo­somal aber­ra­tions (ACAs; high-risk ACAs include
+8, a sec­ond Ph chro­mo­some, i(17q), +19, −7/7q-, 11q23, 3q26.2,
and com­plex aber­rant kar­yo­types) pre­dict a poorer response
to TKIs and a higher risk of pro­gres­sion. According to ELN cri­
te­
ria, patients with high-risk ACAs are con­
sid­
ered high-risk
patients,25 and they should be observed and con­sid­ered for
inten­si­fi­ca­tion of treat­ment, includ­ing early HSCT. Gene muta­
tions (RUNX1, ASXL1, IKZF1, WT1, TET2, IDH1/2, CBFB/MYH11,
TP53) are found in CML and might be asso­ci­ated with pro­gres­
sion to BC.42,43 Such muta­tions may lead to a genet­i­cally based
risk clas­si­fi­ca­tion in the future with the poten­tial for com­bi­na­
tion with non–BCR-ABL1 targeted ther­apy. The molec­u­lar land­
scape of muta­
tions, espe­
cially concerning epi­
ge­
net­
ics, has
been fea­tured in recent pub­li­ca­tions.44
Optimal pre­con­di­tions for HSCT are a low CI Sorror score and
good per­for­mance sta­tus, in addi­tion to the require­ment for a
suit­
able donor, pref­
er­
a­
bly a human leu­
ko­
cyte antigen (HLA)–
com­pat­i­ble sib­ling followed by an unre­lated donor (10 out of 10
matches).45 If no matched donor is avail­­able, alter­na­tive donors,
such as haploidentical HSCT with post cyclo­phos­pha­mide or
mismatched donors or cord blood, should be taken into con­
sid­er­ation.46-49
In addi­tion, HSCT is fea­si­ble in patients pre­vi­ously treated
with 2G-TKIs with a posttransplant com­pli­ca­tion rate com­pa­ra­
ble to that of TKI-naive or imatinib-treated patients.50
Table 2. Selection of CML stud­ies using HSCT includ­ing patients beyond front­line ther­apy
Authors
Number of
patients
HSCT y
Disease stage
(patient %)
TKI used
before HSCT
(patient %)
TRM %
DFS/LFS/
PFS/RI %
OS %
Study type
chP
449
177 (<18y)
272 (18-29y)
2001-2010
chP1 (100)
TKI (60) <18y
TKI (48) 18-29y
13 (at 1y),
18 (at 3y),
20 (at 5y)
LFS 59%
57 (<18y)
60 (18-29y)
n.s.
75% OS
(at 5y)
76% (<18y)
74% (18-29y)
n.s.
Retrospective,
mul­ti­cen­ter,
chil­dren and
young adults
(CIBMTR)
Yassine et al35
199 (adults)
97 (chil­dren/
adults)
Metaanal­y­sis
chP (100)
Meta-anal­y­sis
20 (adults)
28 (chil­dren/
adults)
Adult DFS
66%, adults/
chil­dren DFS
47%/PFS
82%
Adults 84%,
chil­dren 91%,
adults/
chil­dren 76%
Meta-anal­y­sis
resis­tant/
intol­er­ant chP
to >1 TKI
Jiang et al29
132
2001-2008
AP (100)
Imatinib (66%)
Imatinib + HSCT
(34)
11
Low risk:
imatinib 85%
HSCT 95%
(PFS at 6y)
High risk:
imatinib 19%
HSCT 100%
(PFS at 5y)
Low risk:
imatinib 100%
HSCT 81% (OS
at 6y)
High risk:
imatinib 18%
HSCT 100%
(OS at 5y)
Prospective
sin­gle-cen­ter
study
(imatinib vs
HSCT)
Khoury et al57
449
1999-2004
chP2 (41),
AP (41),
BC (18)
Imatinib (50)
chP2 33, AP
34, BC 46
(at 1y)
chP2 27%,
AP 37%,
BC 10%
(LFS at 3y)
chP2 36%,
AP 43%,
BC 14% (OS
at 3y)
Retrospective
mul­ti­cen­ter
study
(CIBMTR)
Zheng et al36
32
2002-2011
AP (59),
BC (41)
Imatinib (53)
Cord 38,
sib 12
(at 0.5y)
Cord 50%
sib 40%
(LFS at 5y)
Cord 62%,
Sib 49%
(OS at 5y)
Retrospective
sin­gle-cen­ter
study (cord
vs sib)
Radujkovic et al33
170
2004-2016
BC in 2
chP (56)
BC active (44)
1G-TKI (59),
2G-TKI (33),
3G-TKI (8)
19.7 (at 1y)
23.3 (at 3y)
Active BC:
27.1% 3y
BC in
remis­sion:
20.2% 3y
LFS 34.6, RI
45.7 (at 1y).
LFS 26.1, RI
50.7 (at 3y)
LFS: active
BC: 11.6% 3y
BC in
remis­sion
33.8% 3y
RI: active BC:
56.4.6% 3y
BC in
remis­sion
45.9% 3y
57.5%(at 1y)
38.5 (at 3y)
Active BC:
23.8% (3y)
BC in
remis­sion 51%
(3y)
Retrospective
EBMT study
HSCT in
treated vs
untreated BC
Yang et al41
278
2002-2021
de novo
AP (100)
Imatinib (67)
or 2G-TKI (33;
nilotinib 24,
dasatanib 7,
flumatinib 1)
Censored
at HSCT
TFS 89%
(at 6y)
Low risk 95
(5y)
Interm 76
(5y)
High-risk 19
(5y)
No
sig­nif­i­cance
between
1G-TKI ver­sus
2G for TFS
OS 90%
(at 6y)
Low risk 54
(5y)
Interm. 36
(5y)
High-risk 10
(5y)
No ­
sig­nif­i­cance
between
1G-TKI vs 2G
for OS
Comparison
between
imatinib
and 2G-TKI
before HSCT
AP and BC
118 | Hematology 2022 | ASH Education Program
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/114/2021884/114niederwieser.pdf by guest on 09 December 2022
Chaudhury et al34
Table 2. Selection of CML studies using HSCT including patients beyond frontline therapy (Continued)
Authors
Number of
patients
HSCT y
Disease stage
(patient %)
TKI used
before HSCT
(patient %)
TRM %
DFS/LFS/
PFS/RI %
OS %
Study type
Different dis­ease phase
84
2003-2008
chP1 (TKI)
(23), chP1 (TKI
fail­ure) (44),
AP (4), BC
(30)
Imatinib
8 in chP
(at 3y)
18 in AdP
(at 3y)
CMR at last
PCR 88%
chP1 (elec­tive)
88%,
chP1 (imatinib
fail­ure) 94%,
AP 59%
(OS at 3y)
Prospective
mul­ti­cen­ter
study (CML IV
study)
Jabbour et al55
47
2004-2007
chP1 (34),
chP2 (21),
AP (25),
BC (19)
Imatinib fail­ure
(100)
+ 2G-TKI (62)
13% (at 2y)
49% (EFS at
2 y) (muta­
tion 36% vs
no muta­tion
58%)
63% (OS at
2y) (muta­tion
44% vs no
muta­tion 76%)
Retrospective
sin­gle-cen­ter
study
Topcuoglu et al37
84
1989-2007
chP1 (79),
chP2 (6),
AP (15)
NA
7% RIC
14% MAC
48% (LFS at
5y) No
dif­fer­ence
RIC vs MAC
56% (OS at
5y) No
dif­fer­ence RIC
vs MAC
Retrospective
sin­gle-cen­ter
study (RIC vs
MAC)
Oyekunle et al38
68
2002 –
2009
chP1 (40),
>chP1 (60)
Pre-HSCT TKI
(71),
post-HSCT TKI
(29)
NA
54% (LFS at
2y)
63% (OS at
2y)
Single
insti­tute HSCT
in TKI era
Milojkovic et al51
5732
2000-2011
Prior TKI: chP
(51), chP >1
2(59), AP (14),
BC (10),
no-TKI: NR
Prior TKI (22),
no TKI (78)
NA
Non-TKI 46%
Prior TKI 42%
(PFS at 5y)
Non-TKI 61%
Prior TKI 59%
(OS at 5y)
Retrospective
mul­ti­cen­ter
study (EBMT)
Piekarska et al39
25
2008-2013
chP1 (50),
chP2/AP (29),
BC (21)
Dasatinib (53),
nilotinib (18), or
both (29)
7.1 (chP1),
12.5
(chP2/AP);
50 (BC)
RI: 29.6%
RI: chP1
21.4%
RI: chP2/AP
12.5%
RI: BP 50%
chP1 92.9%,
chP2/AP
85.7%
BP 0%
(at 1 or 3y)
Prospective
Lubking et al23
118
2002-2017
chP1 (47.5),
chP >1 (40.7),
AP/BC (11.9)
Imatinib
(39.8), imatinib
+2G-TKI (33.1),
imatinib +2GTKI +3G-TKI
(5.1), 2G-TKI
(13.6), 2G-TKI
+3G-TKI (5.1),
no TKI (3.4)
11.6 in chP ≥1
(at 5y)
23.1 in AP/BC
(at 5y)
chP: 66%
molec
relapse
(at 2y)
AP/BC: 71.4%
prog­ress to
AP/BC
50% to AdP
in chP >1
chP 96,3%
70.1% >
chP1/AP
36.9% BP
AP/BP to chP
70.1%
chP TKI resist.
96.8%
(all­at 5y)
Swed­ish
reg­is­try
study
Hu et al40
1223
2001-2013
chP1 (60),
chP2 (21), AP
(12), BC (7),
pro­gres­sion
to AP/BC (19)
TKI 1 (median,
range 0–3)
10–20 (at 1y)
NA
chP 1 HSCT
(vs non-HSCT
infe­rior OS,
HR 2.4)
No dif­fer­ence
chP2 and
AP; BC trend
favor­ing HSCT
Life
expec­tancy
cal­cu­la­tion in
com­par­i­son to
no HSCT
(CIBMTR)
Niederwieser
et al27
147
1990-2018
Non-BC (75)
BC (25)
Prior TKI:
1G (27.2); 1G +
2G (13.6); 1G +
3G (0.7); 2G±3G
(38.1); 3G (0.7)
28 (at 15y)
24 in BC
(at 5y)
24 non-BC
(at 5y)
30% at 10y
and 26% at
15y
BP 24% 5y
Non-BP 31%
5y
OS 15y
34%
BP 30% 10y;
30% 5y
non-BP 41%
10y; 44% 5y
Bicentric
ret­ro­spec­tive
study
Long-term
fol­low-up
Masouridi-Levrat
et al50
383
2009-2013
chP1 (38)
AP > chP1 (45)
BC (16)
Prior TKI:
dasatinib (40)
or nilotinib (17)
or sequen­tial ±
bosutinib/
ponatinib (43)
18 (at 1y)
24% (at 5y)
No
dif­fer­ence
between
2G-TKIs
RFS 40% (at
5y)
RI 29% (at
2y)
RI 36% (at
5y)
65% (at 2y)
56% (at 5y)
chP1 67%
AP/chP >1
57%
BC 37%
EBMT study,
ret­ro­spec­tive
study
CMR, com­plete molec­u­lar response; cord, cord blood; NA, not avail­­able; n.s., not sig­nif­i­cantly dif­fer­ent; PCR, poly­mer­ase chain reac­tion;
PFS, pro­gres­sion-free sur­vival; RI, relapse inci­dence; sib, sib­ling; TFS, trans­for­ma­tion-free sur­vival.
Dr Prakash Singh Shekhawat
HSCT in CML | 119
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/114/2021884/114niederwieser.pdf by guest on 09 December 2022
Saussele et al24
How to man­age trans­plant ther­a­pies
detected 6 months after HSCT. Because of a high risk of relapse,
ponatinib was restarted, and the BCR-ABL1 tran­scripts became
neg­a­tive in the blood. After con­sul­ta­tion with his phy­si­cian fol­
low­ing con­tin­u­ous neg­a­tive BCR-ABL1 tran­
scripts, the patient
opted to ter­mi­nate TKI treat­ment 5 years post trans­plant.
CLINICAL CASE 2 (Con­tin­ued)
A matched unre­
lated donor was found, and after RIC with
busul­fan/fludarabine, HSCT was performed with­out com­pli­ca­
tions. BCR-ABL1 tran­scripts remained unde­tect­able post HSCT.
No TKI pro­phy­laxis was given, and the patient remained BCRABL1-neg­a­tive at 5 years after HSCT.
Summary
Rapid devel­op­ments in the treat­ment of CML require con­tin­ual
adap­ta­tion regard­ing indi­ca­tions for cura­tive stem cell trans­
plan­ta­tion. Patients with chP CML should be started with 1G- or
2G-TKIs and mon­i­tored according to ELN or NCCN guide­lines.
The emer­gence of high-risk ACAs pre­dict a poorer response to
TKIs and a higher risk of pro­gres­sion and are trig­gers for HSCT.
Patients with resis­tant dis­ease (NCCN) or fail­ure (ELN) should be
regarded as can­di­dates for HSCT. Patients with resis­tance to G2TKIs should be con­sid­ered for HSCT if they are not responding
to G3 after 3 months. An HLA-iden­ti­cal related or fully matched
unre­lated donor should be selected, and the inten­sity of the
con­di­tion­ing reg­i­men should be adjusted according to age and
comorbidities. Patients in AP and BC CML should be started
on IC with the pos­si­ble addi­tion of TKIs and are can­di­dates for
HSCT. Posttransplant mon­i­tor­ing with quan­ti­ta­tive poly­mer­ase
chain reac­tion test­ing is indi­cated to guide pro­phy­lac­tic or pre­
emp­tive TKI treat­ment with or with­out DLIs.
Conflict-of-inter­est dis­clo­sure
Chris­
tian Niederwieser: no com­
pet­
ing finan­
cial inter­
ests to
declare.
Nicolaus Kröger: research funding: Bristol Myers Squibb, Jazz,
Neovii, Novartis, Riemser; hon­o­raria: Bristol Myers Squibb, Gilead,
Jazz, Neovii, Novartis, Riemser, Sanofi.
Off-label drug use
Chris­tian Niederwieser: nothing to disclose.
Nicolaus Kröger: nothing to disclose.
Correspondence
Nicolaus Kröger, Department of Stem Cell Transplantation, University Medical Center Hamburg-Eppendorf, Martinistr 52, 20246
Hamburg, Germany; e-mail: nkroeger@uke​­.uni-ham­burg​­.de.
References
1.
CLINICAL CASE 1 (Con­t in­u ed)
The patient under­went a trans­plant, after con­di­tion­ing with fludarabine and myeloablative busul­fan, from an HLA-iden­ti­cal brother
and, except for neutropenic fever, had an unevent­ful posttransplant period. During immu­no­sup­pres­sion taper­ing, he devel­oped
mild chronic skin GVHD. However, BCR-ABL1 tran­scripts were still
120 | Hematology 2022 | ASH Education Program
Nowell P, Hungerford D. A min­ute chro­mo­some in human chronic gran­u­lo­
cytic leu­ke­mia. Blut. 1962;8(April):65-66.
2. Rowley JD. Letter: a new con­sis­tent chro­mo­somal abnor­mal­ity in chronic
mye­log­e­nous leu­kae­mia iden­ti­fied by quin­a­crine fluo­res­cence and Giemsa
staining. Nature. 1973;243(5405):290-293.
3. Groffen J, Stephenson J, Heisterkamp N, Deklein A, Bartram C, Grosveld
G. Philadelphia chro­mo­somal breakpoints are clus­tered within a lim­ited
region, bcr, on chro­mo­some 22. Cell. 1984;36(1):93-99.
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/114/2021884/114niederwieser.pdf by guest on 09 December 2022
In a CIBMTR study, the prog­nos­tic favor­able fac­tors for LFS and
OS after HSCT in advanced CML were the time from diag­no­sis to
trans­plan­ta­tion, good per­for­mance sta­tus, and access to an HLA
fully matched donor.57
In an EBMT study, risk fac­tors for poorer sur­vival were active
BC at trans­plant, advanced age (>45 years), low per­for­mance
sta­tus, a lon­ger inter­val from diag­no­sis to trans­plant, MAC, and
the use of an unre­lated donor.33
Other stud­ies reported that a low CD34+ cell count in the
graft and a higher donor age were adverse risk fac­tors for OS.27
Although no pro­
spec­
tive ran­
dom­
ized study exists, reduced
inten­sity con­di­tion­ing (RIC) seems to be a fea­si­ble approach that
increases acces­si­bil­ity to the trans­plant pro­ce­dure.58 Whether
youn­ger patients and patients with­out comorbidities should be
transplanted after RIC instead of MAC is still a mat­ter of debate.
Similar OS but lower chronic GVHD rates were observed in RIC
when com­pared to MAC.59 RIC may be a rea­son­able alter­na­tive
to MAC in the TKI era, but results should be con­firmed in a ran­
dom­ized study. However, due to the high risk of relapse in AdP
CML, patients who can tol­er­ate MAC should receive it.
Ex vivo T-cell deple­tion is asso­ci­ated with a high relapse risk,
but it has been observed that using anti–T-lym­pho­cyte glob­u­lin
within the con­di­tion­ing reg­i­men reduces the inci­dence of GVHD
with­out increas­ing the risk of relapse.60
The detec­
tion of BCR-ABL1 tran­
scripts early after HSCT has
no adverse prog­nos­tic sig­nif­i­cance. However, BCR-ABL1 tran­
scripts (per­sis­tently neg­a­tive, fluc­tu­at­ing, or per­sis­tently pos­i­tive)
6 months post trans­
plant pre­
dict a risk of relapse.61 ­Several
reports sug­
gest that early posttransplant TKIs (includ­
ing 2GTKIs) are safe to admin­is­ter effec­tively in chP CML but are less effec­
tive in AdP CML.62,63 Maintenance ther­apy with TKIs with or with­out
DLIs appears to be safe and has been asso­ci­ated with a lower inci­
dence of exten­sive chronic GVHD.27,61 Continued reg­u­lar long-term
mon­i­tor­ing of BCR-ABL1 tran­scripts is required to antic­i­pate the
occa­sional late relaps­ing patient; how­ever, both the opti­mal fre­
quency of mon­i­tor­ing and the thresh­old of BCR-ABL1 tran­scripts for
pre­emp­tive ther­apy with TKIs or DLIs need to be established in the
con­texts of con­di­tion­ing reg­i­men and graft manip­u­la­tion. Although
the use of posttransplant TKI ther­apy is wide­spread, pro­spec­tive
stud­ies are needed to explore the best TKI dose, the treat­ment
dura­tion, and pos­si­ble coad­min­is­tra­tion with DLIs.
A relapse of CML can occur as late as the sec­ond decade after
HSCT and involves molec­u­lar, cyto­ge­netic, and hema­to­log­i­cal
relapse.5 Relapsed CML has been treated with DLIs, TKIs, che­
mo­ther­apy, and sec­ond HSCTs.27,64-67 Combinations of TKI ther­
apy and DLIs can achieve a 5-year postrelapse sur­vival rate of
62%, and some patients became TKI-free, suggesting a per­sis­
tent graft-ver­sus-leu­ke­mia effect.67 Molecular remis­sion with­out
GVHD can be achieved, espe­cially when DLIs were given beyond
1 year from HSCT for molec­u­lar or cyto­ge­netic relapse.68
26. Deininger MW, Shah NP, Altman JK, et al. Chronic mye­loid leu­ke­mia, ver­sion
2.2021, NCCN clin­i­cal prac­tice guide­lines in oncol­ogy. J Natl Compr Canc
Netw. 2020;18(10):1385-1415.
27. Niederwieser C, Morozova E, Zubarovskaya L, et al. Risk fac­tors for out­
come after allo­ge­neic stem cell trans­plan­ta­tion in patients with advanced
phase CML. Bone Marrow Transplant. 2021;56(11):2834-2841.
28. Gratwohl A, Brand R, Apperley J, et al; Chronic Leukemia Working Party of
the Euro­pean Group for Blood and Marrow Transplantation. Allogeneic hema­
to­poi­etic stem cell trans­plan­ta­tion for chronic mye­loid leu­ke­mia in Europe
2006: trans­plant activ­ity, long-term data and cur­rent results. An anal­y­sis by
the Chronic Leukemia Working Party of the Euro­pean Group for Blood and
Marrow Transplantation (EBMT). Haematologica. 2006;91(4):513-521.
29. Jiang Q , Xu LP, Liu DH, et al. Imatinib mesylate ver­sus allo­ge­neic hema­to­
poi­etic stem cell trans­plan­ta­tion for patients with chronic mye­log­e­nous
leu­ke­mia in the accel­er­ated phase. Blood. 2011;117(11):3032-3040.
30. How J, Venkataraman V, Soriano Hobbs, G. Blast and accel­er­ated phase
CML: room for improve­ment. ASH Educational Program. 2021;2021(1):122128.
31. Saxena K, Jabbour E, Issa G, et al. Impact of front­line treat­ment approach
on out­comes of mye­loid blast phase CML. J Hematol Oncol. 2021;14(1):
94.
32. Oehler VG, Gooley T, Snyder DS, et al. The effects of imatinib mesylate
treat­ment before allo­ge­neic trans­plan­ta­tion for chronic mye­loid leu­ke­mia.
Blood. 2007;109(4):1782-1789.
33. Radujkovic A, Dietrich S, Blok HJ, et al. Allogeneic stem cell trans­plan­ta­tion
for blast cri­sis chronic mye­loid leu­ke­mia in the era of tyro­sine kinase inhib­
i­tors: a ret­ro­spec­tive study by the EBMT Chronic Malignancies Working
Party. Biol Blood Marrow Transplant. 2019;25(10):2008-2016.
34. Chaudhury S, Sparapani R, Hu ZH, et al. Outcomes of allo­ge­neic hema­to­
poi­etic cell trans­plan­ta­tion in chil­dren and young adults with chronic mye­
loid leu­ke­mia: a CIBMTR cohort anal­y­sis. Biol Blood Marrow Transplant.
2016;22(6):1056-1064.
35. Yassine F, Reljic T, Moustafa MA, et al. Efficacy of allo­ge­neic hema­to­poi­etic
cell trans­plan­ta­tion in patients with chronic phase CML resis­tant or intol­er­
ant to tyro­sine kinase inhib­i­tors [published online ahead of print 11 March
2022]. Hematol Oncol Stem Cell Ther.
36. Zheng C, Tang B, Yao W, et al. Comparison of unre­lated cord blood trans­
plan­ta­tion and HLA-matched sib­ling hema­to­poi­etic stem cell trans­plan­ta­
tion for patients with chronic mye­loid leu­ke­mia in advanced stage. Biol
Blood Marrow Transplant. 2013;19(12):1708-1712.
37. Topcuoglu P, Arat M, Ozcan M, et al. Case-matched com­par­i­son with stan­
dard ver­sus reduced inten­sity con­di­tion­ing reg­i­men in chronic mye­loid
leu­ke­mia patients. Ann Hematol. 2012;91(4):577-586.
38. Oyekunle A, Zander AR, Binder M, et al. Outcome of allo­ge­neic SCT in
patients with chronic mye­loid leu­ke­mia in the era of tyro­sine kinase inhib­i­
tor ther­apy. Ann Hematol. 2013;92(4):487-496.
39. Piekarska A, Gil L, Prejzner W, et al. Pretransplantation use of the sec­ondgen­er­a­tion tyro­sine kinase inhib­i­tors has no neg­a­tive impact on the HCT
out­come. Ann Hematol. 2015;94(11):1891-1897.
40. Hu B, Lin X, Lee HC, et al. Timing of allo­ge­neic hema­to­poi­etic cell trans­
plan­ta­tion (alloHCT) for chronic mye­loid leu­ke­mia (CML) patients. Leuk
Lymphoma. 2020;61(12):2811-2820.
41. Yang S, Zhang XS, Gale RP, et al. Co-var­i­ates asso­ci­ated with out­comes
of tyro­sine kinase-inhib­i­tor ther­apy in per­sons with chronic mye­loid leu­kae­
mia ini­
tially presenting in accel­
er­
ated phase. Leukemia. 2022;36(7):18181824.
42. Hehlmann R. The new ELN rec­om­men­da­tions for treating CML. J Clin Med.
2020;9(11):3671.
43. McKenney AS, Levine RL. Isocitrate dehy­dro­ge­nase muta­tions in leu­ke­mia.
J Clin Invest. 2013;123(9):3672-3677.
44. Adnan-Awad S, Kankainen M, Mustjoki S. Mutational land­scape of chronic
mye­loid leu­ke­mia: more than a sin­gle onco­gene leu­ke­mia. Leuk ­Lymphoma.
2021;62(9):2064-2078.
45. Sorror ML, Maris MB, Storer B, et al. Comparing mor­bid­ity and mor­tal­ity
of HLA-matched unre­lated donor hema­to­poi­etic cell trans­plan­ta­tion after
nonmyeloablative and myeloablative con­di­tion­ing: influ­ence of pretransplantation comorbidities. Blood. 2004;104(4):961-968.
46. Mielcarek M, Furlong T, O’Donnell PV, et al. Posttransplantation cyclo­phos­
pha­mide for pre­ven­tion of graft-ver­sus-host dis­ease after HLA-matched
mobi­lized blood cell trans­plan­ta­tion. Blood. 2016;127(11):1502-1508.
47. Chen H, Si Y, He X, et al. HLA-mismatched hema­
to­
poi­
etic stem cell
trans­plan­ta­tion for treat­ment of chronic mye­log­e­nous leu­ke­mia. Blood.
2008;112(11):1113-1113.
Dr Prakash Singh Shekhawat
HSCT in CML | 121
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/114/2021884/114niederwieser.pdf by guest on 09 December 2022
4. Shtivelman E, Lifshitz B, Gale RP, Canaani E. Fused tran­script of abl and bcr
genes in chronic mye­log­e­nous leu­kae­mia. Nature. 1985;315(6020):550-554.
5. Goldman JM, Majhail NS, Klein JP, et al. Relapse and late mor­
tal­
ity in
5-year sur­vi­vors of myeloablative allo­ge­neic hema­to­poi­etic cell trans­plan­
ta­tion for chronic mye­loid leu­ke­mia in first chronic phase. J Clin Oncol.
2010;28(11):1888-1895.
6. Talpaz M, Kantarjian HM, McCredie K, Trujillo JM, Keating MJ, Gutterman JU.
Hematologic remis­sion and cyto­ge­netic improve­ment induced by recom­
bi­nant human inter­feron alpha A in chronic mye­log­e­nous leu­ke­mia. N Engl
J Med. 1986;314(17):1065-1069.
7. Lee MS, Chang KS, Freireich EJ, et al. Detection of min­i­mal resid­ual bcr/abl
tran­scripts by a mod­i­fied poly­mer­ase chain reac­tion. Blood. 1988;72(3):893897.
8. Kolb HJ, Mittermüller J, Clemm C, et al. Donor leu­ko­cyte trans­fu­sions for
treat­ment of recur­rent chronic mye­log­e­nous leu­ke­mia in mar­row trans­
plant patients. Blood. 1990;76(12):2462-2465.
9. Druker BJ, Talpaz M, Resta DJ, et al. Efficacy and safety of a spe­cific inhib­i­tor
of the BCR-ABL tyro­sine kinase in chronic mye­loid leu­ke­mia. N Engl J Med.
2001;344(14):1031-1037.
10. Mahon FX, Réa D, Guilhot J, et al; Intergroupe Français des Leucémies
Myéloïdes Chroniques. Discontinuation of imatinib in patients with chronic
mye­loid leu­kae­mia who have maintained com­plete molec­u­lar remis­sion for
at least 2 years: the pro­spec­tive, multicentre Stop Imatinib (STIM) trial.
Lancet Oncol. 2010;11(11):1029-1035.
11. Niederwieser D, Baldomero H, Bazuaye N, et al. One and a half mil­lion
hema­to­poi­etic stem cell trans­plants: con­tin­u­ous and dif­fer­en­tial improve­
ment in world­wide access with the use of non-iden­ti­cal fam­ily donors.
Haematologica. 2022;107(5):1045-1053.
12. Gratwohl A, Hermans J, Goldman JM, et al. Risk assess­ment for patients
with chronic mye­loid leu­kae­mia before allo­ge­neic blood or mar­row trans­
plan­ta­tion. Chronic Leukemia Working Party of the Euro­pean Group for
Blood and Marrow Transplantation. Lancet. 1998;352(9134):1087-1092.
13. Gluckman E, Broxmeyer HA, Auerbach AD, et al. Hematopoietic recon­sti­tu­
tion in a patient with Fanconi’s ane­mia by means of umbil­i­cal-cord blood
from an HLA-iden­ti­cal sib­ling. N Engl J Med. 1989;321(17):1174-1178.
14. Luznik L, O’Donnell PV, Fuchs EJ. Post-trans­plan­ta­tion cyclo­phos­pha­mide
for tol­er­ance induc­tion in HLA-haploidentical bone mar­row trans­plan­ta­
tion. Semin Oncol. 2012;39(6):683-693.
15. Slavin S, Nagler A, Naparstek E, et al. Nonmyeloablative stem cell trans­
plan­ta­tion and cell ther­apy as an alter­na­tive to con­ven­tional bone mar­row
trans­plan­ta­tion with lethal cytoreduction for the treat­ment of malig­nant
and non­ma­lig­nant hema­to­logic dis­eases. Blood. 1998;91(3):756-763.
16. Hochhaus A, Larson RA, Guilhot F, et al; IRIS Investigators. Long-term out­
comes of imatinib treat­ment for chronic mye­loid leu­ke­mia. N Engl J Med.
2017;376(10):917-927.
17. Mahon FX, Delbrel X, Cony-Makhoul P, et al. Follow-up of com­plete cyto­ge­
netic remis­sion in patients with chronic mye­loid leu­ke­mia after ces­sa­tion of
inter­feron alfa. J Clin Oncol. 2002;20(1):214-220.
18. Kohlbrenner K, Fabarius A, Reiser M, et al. Therapy in patients with chronic
mye­loid leu­ke­mia out­side of clin­i­cal tri­als: results of the Ger­man CMLreg­is­try (CML-VI). Paper presented at: Euro­pean Society of Hypertension
31st Euro­pean Meeting; 17-20 June 2022; Athens, Greece.
19. Réa D, Mauro MJ, Boquimpani C, et al. A phase 3, open-label, ran­dom­ized
study of asciminib, a STAMP inhib­i­tor, vs bosutinib in CML after 2 or more
prior TKIs. Blood. 2021;138(21):2031-2041.
20. Pasvolsky O, Leader A, Iakobishvili Z, Wasserstrum Y, Kornowski R, Raanani
P. Tyrosine kinase inhib­i­tor asso­ci­ated vas­cu­lar tox­ic­ity in chronic mye­loid
leu­ke­mia. Cardiooncology. 2015;1(1):5.
21. Brixey AG, Light RW. Pleural effu­sions due to dasatinib. Curr Opin Pulm
Med. 2010;16(4):351-356.
22. Cortes JE, Saglio G, Kantarjian HM, et al. Final 5-year study results of DASISION: the Dasatinib ver­
sus Imatinib Study in Treatment-Naïve Chronic
Myeloid Leukemia Patients trial. J Clin Oncol. 2016;34(20):2333-2340.
23. Lübking A, Dreimane A, Sandin F, et al. Allogeneic stem cell trans­plan­ta­tion
for chronic mye­loid leu­ke­mia in the TKI era: pop­u­la­tion-based data from
the Swed­ish CML reg­is­try. Bone Marrow Transplant. 2019;54(11):1764-1774.
24. Saussele S, Lauseker M, Gratwohl A, et al; Ger­man CML Study Group. Allogeneic hema­to­poi­etic stem cell trans­plan­ta­tion (allo SCT) for chronic mye­
loid leu­ke­mia in the imatinib era: eval­u­a­tion of its impact within a sub­group
of the ran­dom­ized Ger­man CML Study IV. Blood. 2010;115(10):1880-1885.
25. Hochhaus A, Baccarani M, Silver RT, et al. Euro­pean LeukemiaNet 2020
rec­om­men­da­tions for treating chronic mye­loid leu­ke­mia. Leukemia.
2020;34(4):966-984.
122 | Hematology 2022 | ASH Education Program
60. Zander AR, Kröger N, Schleuning M, et al. ATG as part of the con­di­tion­ing
reg­i­men reduces trans­plant-related mor­tal­ity (TRM) and improves over­all
sur­vival after unre­lated stem cell trans­plan­ta­tion in patients with chronic
mye­log­e­nous leu­ke­mia (CML). Bone Marrow Transplant. 2003;32(4):355361.
61. Kaeda J, O’Shea D, Szydlo RM, et al. Serial mea­sure­ment of BCR-ABL tran­
scripts in the periph­eral blood after allo­ge­neic stem cell trans­plan­ta­tion
for chronic mye­loid leu­ke­mia: an attempt to define patients who may not
require fur­ther ther­apy. Blood. 2006;107(10):4171-4176.
62. Carpenter PA, Snyder DS, Flowers ME, et al. Prophylactic admin­is­tra­tion of
imatinib after hema­to­poi­etic cell trans­plan­ta­tion for high-risk Philadelphia
chro­mo­some-pos­i­tive leu­ke­mia. Blood. 2007;109(7):2791-2793.
63. Nakasone H, Kanda Y, Takasaki H, et al; Kanto Study Group for Cell Therapy. Prophylactic impact of imatinib admin­is­tra­tion after allo­ge­neic stem
cell trans­plan­ta­tion on the inci­dence and sever­ity of chronic graft ver­sus
host dis­ease in patients with Philadelphia chro­mo­some-pos­i­tive leu­ke­mia.
Leukemia. 2010;24(6):1236-1239.
64. Goldman JM, Melo JV. Chronic mye­loid leu­ke­mia—advances in biol­ogy and
new approaches to treat­ment. N Engl J Med. 2003;349(15):1451-1464.
65. Kim YJ, Kim DW, Lee S, et al. Cytogenetic clonal evo­
lu­
tion alone in
CML relapse post-trans­plan­ta­tion does not adversely affect response to
imatinib mesylate treat­
ment. Bone Marrow Transplant. 2004;33(2):237242.
66. DeAngelo DJ, Hochberg EP, Alyea EP, et al. Extended fol­low-up of patients
treated with imatinib mesylate (Gleevec) for chronic mye­log­e­nous leu­ke­
mia relapse after allo­ge­neic trans­plan­ta­tion: dura­ble cyto­ge­netic remis­sion
and con­ver­sion to com­plete donor chi­me­rism with­out graft-ver­sus-host
dis­ease. Clin Cancer Res. 2004;10(15):5065-5071.
67. Savani BN, Montero A, Kurlander R, Childs R, Hensel N, Barrett AJ. Imatinib
synergizes with donor lym­pho­cyte infu­sions to achieve rapid molec­u­lar
remis­sion of CML relaps­ing after allo­ge­neic stem cell trans­plan­ta­tion. Bone
Marrow Transplant. 2005;36(11):1009-1015.
68. Radujkovic A, Guglielmi C, Bergantini S, et al; Chronic Malignancies Working Party of the Euro­
pean Society for Blood and Marrow Transplantation. Donor lym­pho­cyte infu­sions for chronic mye­loid leu­ke­mia relaps­ing
after allo­ge­neic stem cell trans­plan­ta­tion: may we pre­dict graft-ver­susleu­ke­mia with­out graft-ver­sus-host dis­ease? Biol Blood Marrow Transplant.
2015;21(7):1230-1236.
© 2022 by The Amer­i­can Society of Hematology
DOI 10.1182/hema­tol­ogy.2022000329
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/114/2021884/114niederwieser.pdf by guest on 09 December 2022
48. Xiao-Jun H, Lan-Ping X, Kai-Yan L, et al. HLA-mismatched/haploidentical
hema­to­poi­etic stem cell trans­plan­ta­tion with­out in vitro T cell deple­tion for
chronic mye­loid leu­ke­mia: improved out­comes in patients in accel­er­ated
phase and blast cri­sis phase. Ann Med. 2008;40(6):444-455.
49. Sanz J, Sanz GF. Umbilical cord blood trans­
plan­
ta­
tion from unre­
lated
donors in adult patients with chronic mye­loid leu­kae­mia. Best Pract Res
Clin Haematol. 2010;23(2):217-222.
50. Masouridi-Levrat S, Olavarria E, Iacobelli S, et al. Outcomes and tox­ic­ity of
allo­ge­neic hema­to­poi­etic cell trans­plan­ta­tion in chronic mye­loid leu­ke­mia
patients pre­vi­ously treated with sec­ond-gen­er­a­tion tyro­sine kinase inhib­i­
tors: a pro­spec­tive non-interventional study from the Chronic Malignancy
Working Party of the EBMT. Bone Marrow Transplant. 2022;57(1):23-30.
51. Milojkovic D, Apperley JF, Gerrard G, et al. Responses to sec­ond-line tyro­
sine kinase inhib­i­tors are dura­ble: an inten­tion-to-treat anal­y­sis in chronic
mye­loid leu­ke­mia patients. Blood. 2012;119(8):1838-1843.
52. Baccarani M, Castagnetti F, Gugliotta G, Rosti G. A review of the Euro­pean
LeukemiaNet rec­om­men­da­tions for the man­age­ment of CML. Ann Hematol. 2015;94(suppl 2):S141-S147.
53. Lee SJ, Kukreja M, Wang T, et al. Impact of prior imatinib mesylate on the
out­come of hema­to­poi­etic cell trans­plan­ta­tion for chronic mye­loid leu­ke­
mia. Blood. 2008;112(8):3500-3507.
54. Hafez HA, Abdallah A, Hammad M, et al. Outcomes of allo­genic hema­to­
poi­etic cell trans­plan­ta­tion for child­hood chronic mye­loid leu­ke­mia: sin­glecen­ter expe­ri­ence. Pediatr Transplant. 2020;24(4):e13664.
55. Jabbour E, Cortes J, Santos FP, et al. Results of allo­ge­neic hema­to­poi­etic
stem cell trans­plan­ta­tion for chronic mye­log­e­nous leu­ke­mia patients who
failed tyro­sine kinase inhib­i­tors after devel­op­ing BCR-ABL1 kinase domain
muta­tions. Blood. 2011;117(13):3641-3647.
56. Nicolini FE, Basak GW, Kim DW, et al. Overall sur­
vival with ponatinib
ver­sus allo­ge­neic stem cell trans­plan­ta­tion in Philadelphia chro­mo­somepos­i­tive leu­ke­mias with the T315I muta­tion. Cancer. 2017;123(15):28752880.
57. Khoury HJ, Kukreja M, Goldman JM, et al. Prognostic fac­tors for out­comes
in allo­ge­neic trans­plan­ta­tion for CML in the imatinib era: a CIBMTR anal­y­sis.
Bone Marrow Transplant. 2012;47(6):810-816.
58. Crawley C, Szydlo R, Lalancette M, et al; Chronic Leukemia Working Party
of the EBMT. Outcomes of reduced-inten­sity trans­plan­ta­tion for chronic
mye­loid leu­ke­mia: an anal­y­sis of prog­nos­tic fac­tors from the Chronic Leukemia Working Party of the EBMT. Blood. 2005;106(9):2969-2976.
59. Chhabra S, Ahn KW, Hu ZH, et al. Myeloablative vs reduced-inten­sity con­di­
tion­ing allo­ge­neic hema­to­poi­etic cell trans­plan­ta­tion for chronic mye­loid
leu­ke­mia. Blood Adv. 2018;2(21):2922-2936.
BEYOND ROUTINE FRONTLINE THERAPY OF CML
Treatment of CML in pregnancy
Center for Hematology, Imperial College London, London, UK; and Department of Clinical Hematology, Hammersmith Hospital, Imperial College
Healthcare NHS Trust, London, UK
Since the introduction of tyrosine kinase inhibitors (TKIs) at the beginning of the millennium, the outlook for patients
with chronic myeloid leukemia (CML) has improved remarkably. As such, the question of life expectancy and survival has
become less problematic while quality of life and family planning have become more so. While TKIs are the cornerstone
of CML management, their teratogenicity renders them contraindicated during pregnancy. In recent years, patients who
satisfy standardized criteria can stop TKI therapy altogether, and indeed, in eligible patients who wish to become pregnant, these objectives overlap. However, not all patients satisfy these criteria. Some pregnancies are unplanned, and a
number of patients are pregnant when diagnosed with CML. In these patients the way forward is less clear, and there
remains a paucity of good evidence available to guide treatment. In this article, we summarize the relevant literature and
provide a framework for clinicians faced with the challenge of managing CML and pregnancy.
LEARNING OBJECTIVES
• Manage CML diagnosed in pregnancy
• Manage established CML in unplanned pregnancies
• Facilitate planned childbirth in established CML
CLINICAL CASE
A 35-year-old woman was diagnosed with chronic myeloid leukemia (CML) in chronic phase in 2014. There were
no additional chromosomal abnormalities, and the European Treatment and Outcome Study (EUTOS) long-term
survival score was intermediate. She was started on imatinib and achieved complete cytogenetic remission (CCyR)
at 12 months but was not in major molecular remission
(MMR) at 2 years. She has no children but indicates that
having a child is very important to her.
Introduction
Targeted tyrosine kinase inhibitors (TKIs) have revolutionized the course of CML. An individual diagnosed with CML,
previously a fatal disease, now has an excellent chance of a
normal life expectancy.1 Furthermore, for many patients the
aim is to achieve deep and sustained molecular responses
so that TKI therapy can be discontinued indefinitely.
In Western countries the median age at diagnosis is
57 to 60 years, so many patients are women of childbearing age.2 This is even more relevant in emerging nations,
where the median age at diagnosis is 30 to 40 years.3
Before the advent of TKIs, pregnancy and CML were largely
incompatible, but the ability to induce profound responses
and an increasing prevalence in young patients mean that
family planning is an increasingly important discussion
topic. There are now 6 TKIs in clinical practice and welldefined sets of hematological, cytogenetic, and molecular landmarks to guide management.4 All produce excellent
responses in the majority, but their teratogenicity hinders
their use in pregnancy.
This therefore poses a therapeutic and ethical
dilemma for the clinician and patient, and decisions
must be made with the disease, mother, and fetus in
mind. Poorly controlled CML in pregnancy can lead to
adverse outcomes for both parties. Persistently raised
white blood cell (WBC) counts can result in leukastasis,
which may subsequently result in placental insufficiency,
intrauterine growth retardation, or fetal mortality.5 Thus,
2 equally important considerations must be balanced: (1)
adequately controlling the disease to avoid progression,
and (2) minimizing exposure of the fetus to potentially
harmful systemic medications.
Despite increasing attention over the last decade, standardized guidance for the management of CML in pregnancy is lacking. Appropriately stratifying patients by
Dr Prakash Singh Shekhawat
Treatment of CML in pregnancy | 123
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/123/2021708/123robertson.pdf by guest on 09 December 2022
Harry F. Robertson and Jane F. Apperley
dis­ease sta­tus at diag­no­sis and the risk of relapse/pro­gres­sion is
help­ful when approaching man­age­ment. They can be gen­er­ally
defined by the fol­low­ing:
a) CML diag­nosed dur­ing preg­nancy
b) Established CML when the patient is a can­di­date for treat­mentfree remis­sion (TFR)
c) Established CML when the patient is not a can­di­date for TFR
CML diag­nosed dur­ing preg­nancy
Table 1. Summary of stud­ies describ­ing the man­age­ment of CML diag­nosed in preg­nancy
Treatment dur­ing
preg­nancy (n, %)
Outcome (n, %)
15
HC (5, 33)
IFN (3, 20)
Leukapheresis (1, 7)
Nilotinib (1, 7)
Observation (5, 33)
Live birth (12, 80)
Spontaneous
ter­mi­na­tion (2, 13)
Elective ter­mi­na­tion
(1, 7)
Imatinib (4, 27)
Dasatinib (6, 40)
Nilotinib (5, 33)
14 (93)
Blast cri­sis and
death (1, 7)
48
Imatinib (10, 30)
IFN only (2, 6)
IFN + imatinib (3, 10)
HC (4, 12)
Observation (14, 42)a
Live birth (33, 71)
Spontaneous
ter­mi­na­tion (1, 2)
Elective ter­mi­na­tion
(14, 29)
Imatinib (30, 94)
Nilotinib (1, 2)
Dasatinib (1, 2)a,b
17 (52)a
Blast cri­sis and
death (2, 6)a
Study
N
Assi et al9
Chelysheva et al10
TKI after preg­nancy
end (n, %)
Denominator derived from the num­ber of patients who con­tin­ued preg­nancy to term (n=33).
One patient had a set of twins, ana­lyzed as 2 preg­nan­cies.
a
b
124 | Hematology 2022 | ASH Education Program
Dr Prakash Singh Shekhawat
≥MMR, n (%)
Adverse dis­ease
out­comes (n, %)
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/123/2021708/123robertson.pdf by guest on 09 December 2022
A diag­
no­
sis of CML dur­
ing preg­
nancy is not an uncom­
mon
occur­rence; approx­i­ma­tely 10% of preg­nancy-related leu­ke­mias
are CML, and 20% of preg­nancy-related CML is in patients diag­
nosed dur­ing preg­nancy.6,7 Unsurprisingly, patients often meet
the deliv­ery of such grave news at this time with pro­found shock
and anx­i­ety surrounding their own health and that of the unborn
child. For the cli­ni­cian, it is impor­tant to be empathic, sen­si­tive,
and to draw upon the mul­ti­fac­eted skill set of the surrounding
health care team, includ­ing but not lim­ited to obste­tri­cians, spe­
cial­ist nurses, psy­chol­o­gists, and ther­a­pists.
At diag­
no­
sis there are 3 three main aspects to con­
sider,
which guide treat­ment deci­sions: dis­ease phase, dis­ease bur­den
and time to deliv­ery. If the patient is presenting in blast cri­sis or
chronic phase with mul­ti­ple high-risk fea­tures, seri­ous con­sid­er­
ation should be given to ter­mi­na­tion of the preg­nancy in order
to com­mence treat­ment. Blast cri­sis is now increas­ingly rare,
but the out­look remains bleak, with median sur­viv­als around
12 months.8 Fortunately, the vast major­ity pres­ent in the chronic
phase. Upon first con­sul­ta­tion, a full out­line of the treat­ment
options, the asso­ci­ated risks, and an expected preg­nancy road
map should be discussed with the patient.
In addi­tion to iso­lated case reports, 2 stud­ies pro­vided data
on mul­ti­ple patients presenting with CML in preg­nancy (Table 1).
Assi et al described 15 patients, the major­ity of whom achieved at
least MMR fol­low­ing the start of TKIs.9 One patient who was man­
aged with inter­feron alfa (IFN-α) dur­ing preg­nancy was resis­tant
to dasatinib, progressed to blast cri­sis, and died. Chelysheva
et al presented 48 patients, of which 33 went to term.10 Imatinib
was given in 13 cases at a median of 18 weeks’ ges­ta­tion (range,
16-35). No con­gen­i­tal abnor­mal­i­ties occurred, but the moth­ers of
6 of 7 infants born with low birth weights had received imatinib.
Two patients dem­on­strated resis­tance to mul­ti­ple TKIs and died
sec­ond­ary to blast cri­sis. At our cen­ter we man­aged 11 cases of
CML diag­nosed dur­ing preg­nancy (Table 2). All patients were
alive at last fol­low-up.
Many patients diag­nosed in preg­nancy pres­ent with man­
age­able cell counts and low-risk dis­ease such that a “watch and
wait” strat­egy is a com­pletely rea­son­able approach. For patients
with rap­idly ris­ing WBC counts and/or trou­ble­some symp­toms,
we dis­cuss below the var­i­ous options for treat­ment. Pregnancy
itself does not alter the dis­ease course of CML, with the caveat
that the true effect of treat­ment delay on long-term dis­ease con­
trol is uncer­tain.11
TKIs should be avoided in preg­nancy, par­tic­u­larly dur­ing the
first tri­mes­ter, when the major­ity of organ­o­gen­e­sis occurs, and
the fetus is most sen­si­tive to mater­nal sys­temic med­i­ca­tions.12
In both clin­
i­
cal and pre­
clin­
i­
cal stud­
ies, the most com­
monly
observed abnor­
mal­
i­
ties in fetuses exposed to TKIs are bony
skull defects and exomphalos.13 Significant “off-tar­
get” inhi­
bi­
tion, most prob­a­bly of plate­let-derived growth fac­tor recep­tor
alpha, is likely respon­si­ble for these malformations.14 Pye et al
reviewed the larg­est cohort to date of 125 patients treated with
imatinib dur­ing preg­nancy, with 103 exposed in the first tri­mes­
ter. Fifty per­cent of the preg­nan­cies resulted in live births, and
12 fetal malformations were observed. Moreover, the sim­
i­
lar­
ity of rare malformations in 3 infants sug­gests that they were
drug induced.13 These find­ings were cor­rob­o­rated by Abruzzese
et al, who con­cluded that although the major­ity of preg­nan­cies
exposed to imatinib result in nor­mal live births, there is a sig­
nif­i­cant risk of fetal abnor­mal­ity.15 Dasatinib exhib­its the highest
inci­dence of fetal abnor­mal­i­ties and should not be used at any
point dur­ing preg­nancy.16 Nilotinib should not be used in the first
tri­mes­ter, and the spar­sity of evi­dence for bosutinib cur­rently
sup­ports avoid­ance through­out preg­nancy.17,18 There have been
no published reports on the effect of the newest-gen­er­a­tion
TKIs, ponatinib and asciminib, on human preg­nancy out­comes;
how­ever; both dem­on­strated embryo-fetal tox­ic­ity in pre­clin­i­cal
stud­ies (Ariad and Novartis inves­ti­ga­tor bro­chures, respec­tively).
While there is no doubt that TKIs should be avoided in the
first tri­mes­ter, there is grow­ing evi­dence of their safety when
used in later preg­nancy. The preg­nancy reg­is­try of the Euro­pean
LeukemiaNet described 41 women who started on a TKI late in
ges­ta­tion (imatinib=33, nilotinib=8), and while a small num­ber
had malformations, these were deemed unre­lated to the TKI by
Table 2. Summary of patients diag­nosed with CML dur­ing
preg­nancy at Hammersmith Hospital, London
Characteristics (N=11)
Median (range); N [%]
Age at diag­no­sis (years)
33 (25-40)
Trimester
1
6 [55]
2
5 [45]
CML phase
Chronic
11 [100]
80 (14.7-215.4)
Sokal score
Low
6 [55]
Intermediate
2 [18]
No data
3 [27]
Pregnancy out­come
Live birth
10 [91]
Miscarriage
1 [9]
Management in preg­nancy
Observation
3 [27]
IFNa
2 [18]
Leukapheresis
3 [27]
IFN + leukapheresis
3 [27]
TKI started after preg­nancy end
Imatinib
8 (73)
Dasatinib
3 (27)
Duration of TKI delay (months)
7 (1-10)
≥MR1 at 3 mo
11 [100]
≥CCyR at 6 mo
9 [82]
≥MMR at 12 mo
8 [73]
One patient was switched from IFN-α to PEG-IFN-α due to an adverse
reac­tion to the drug.
a
the treating phy­si­cians.19 Abruzzese and col­leagues con­firmed
these find­
ings and reported preg­
nan­
cies in 17 patients, 3 of
whom started a TKI dur­ing preg­nancy (imatinib=1, nilotinib=2).
All 3 preg­nan­cies resulted in live healthy infants.20 Therefore,
when nec­es­sary in those with rap­idly increas­ing BCR-ABL1 tran­
scripts, we rec­om­mend con­sid­er­ation of TKIs in later preg­nancy
(sec­ond or third tri­mes­ter). Patients should be aware of the man­
u­fac­turer’s advice to avoid TKIs dur­ing preg­nancy as well as the
poten­tial risks to starting or delaying TKI ther­apy.
A cru­
cial fac­
tor in avoiding com­
pli­
ca­
tions is the man­
age­
ment of thrombocytosis and leu­ko­cy­to­sis. Those with sig­nif­i­cant
thrombocytosis should start low-dose aspi­rin (75-100mg) and
low-molec­u­lar-weight hep­a­rin, both of which are con­sid­ered
safe in preg­nancy, to reduce the risk of throm­bo­sis.21,22 While
there is no evi­dence of a thresh­old for starting antithrombotic
drugs, recent expert rec­om­men­da­tions sug­gest ini­ti­a­tion when
the plate­lets exceed 600×109/L.7 Clearly, the ben­e­fits of avoiding throm­bo­sis must be care­fully bal­anced against the risk of
Established CML: patient is a can­di­date for TFR
Regardless of preg­nancy, a num­ber of patients are eli­gi­ble for
tri­als of treat­ment dis­con­tin­u­a­tion. Indeed, the numer­ous tri­
als of treat­ment dis­con­tin­u­a­tion in non­preg­nant patients can
guide man­age­ment of the preg­nant pop­u­la­tion. Stopping TKIs
offers myr­iad ben­e­fits, includ­ing a res­o­lu­tion of side effects and
a reduc­tion in health care costs and the risk of fetal abnor­mal­i­ties
in those attempting preg­nancy. Guidelines indi­cate that patients
in the chronic phase who have been on first- or sec­ond-line
(intol­er­ance only) TKIs for over 5 years and have sustained deep
Dr Prakash Singh Shekhawat
Treatment of CML in preg­nancy | 125
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/123/2021708/123robertson.pdf by guest on 09 December 2022
WBC (×109/L)
preg­nancy-related bleed­ing. There is lit­tle evi­dence to sug­gest
the best time to stop treat­ment prior to labor, and the deci­sion is
ulti­mately led by the obste­tri­cian based on the expected mode
of deliv­ery and asso­ci­ated pro­ce­dures.
There are a num­ber of reports of safe hydroxycarbamide (HC)
use dur­ing preg­nancy, albeit mostly in mye­lo­pro­lif­er­a­tive neo­
plasms.23,24 The National Toxicology Program gath­ered infor­ma­
tion from 33 patients exposed to HC dur­ing preg­nancy. They
reported major malformations in 8% of those exposed to HC in
the first tri­mes­ter and 14% in the sec­ond and/or third tri­mes­ter.
Two still­births were noted.25 In the recent expert rec­om­men­da­
tions, HC was not recommended for use dur­ing preg­nancy.7
Leukapheresis is effec­tive in con­trol­ling high WBC counts and
can be used safely in preg­nancy.26,27 It is, how­ever, not with­out
draw­backs, includ­ing avail­abil­ity, cost, and poten­tial prob­lems
with repeated intra­
ve­
nous access with risks of infec­
tion and
bleed­ing. Our per­sonal prac­tice is to com­mence leukapheresis
when the WBC count is higher than 100×109/L.
IFN-α is the only ther­a­peu­tic option that exhib­its anti­can­cer
prop­er­ties and is safe to use through­out preg­nancy.28,29 Polyethylene gly­col (PEG)–IFN-α is more com­monly used now due
to its faster and lon­
ger-last­
ing treat­
ment response. Initially,
PEG–IFN-α was contraindicated in preg­nancy due to con­cerns
around the pre­ser­va­tive. Multiple case reports from patients
with mye­lo­pro­lif­er­a­tive neo­plasms and viral hep­a­ti­tis and expe­
ri­ence from our own cen­ter have not dem­on­strated any harm­ful
effects.30 The abil­ity of IFN-α to pro­duce mean­ing­ful dis­ease
con­trol over the remaining period of ges­ta­tion is var­i­able, and
it is asso­ci­ated with unpleas­ant side effects such as pyrexia and
myalgia.31
TKI ther­apy should begin as soon as pos­si­ble after preg­nancy
has con­cluded. Imatinib and nilotinib are pres­ent in breast milk,
but the effects of these low drug con­cen­tra­tions on a devel­op­
ing infant are unknown.32-36 There is lit­tle infor­ma­tion on whether
the other TKIs are pres­ent in breast milk, but it is safer to avoid
their use when breastfeeding. Many women wish to breastfeed,
but the deci­sion to delay effec­tive ther­apy in order to facil­i­tate
this must be on a case-by-case basis after a thor­ough dis­cus­sion
of the risks involved.
It is clear that a “one-size-fits-all­” approach is not pos­si­ble,
and treat­ment should be tai­lored based on the patient’s per­
sonal, cul­tural, and reli­gious wishes. A mul­ti­dis­ci­plin­ary approach
is par­a­mount. A treat­ment-free first tri­mes­ter is the objec­tive,
with man­age­ment in the sec­ond and third tri­mes­ters remaining
largely the same, with reas­sur­ance that any treat­ment admin­
is­tered has a lower risk of caus­ing fetal abnor­mal­i­ties. TKI ther­
apy should be avoided dur­ing breastfeeding, but if the mother
is insis­tent, then it should be lim­ited to 2 to 3 weeks to min­i­mize
the risk of dis­ease pro­gres­sion.
Established CML: patient is not a can­di­date for TFR
This is the major­ity of women who rep­re­sent a spec­trum of sce­
nar­ios rang­ing from planned preg­nan­cies in those with min­i­mal
TKI expo­sure and a lack of ade­quate response, to those on longterm treat­ment who have sustained MMR with­out sat­is­fy­ing TFR
cri­te­ria, to those who con­ceive while on TKI. The evi­dence for
TKI dis­con­tin­u­a­tion in the non-TFR pop­u­la­tion is grow­ing and
shows that con­cep­tion and preg­nancy can be safely nav­i­gated.
In order to appro­pri­ately strat­ify these patients, the dura­tion of
TKI expo­sure and the molec­u­lar response are impor­tant con­sid­
er­ations.
Women who have received TKI ther­apy for less than 3 years
or who have failed to achieve CCyR or MMR even after 3 years
of treat­ment, such as our patient, have a higher risk of loss of
CHR upon TKI dis­con­tin­u­a­tion. The prime con­sid­er­ation in these
patients should be the leu­
ke­
mic bur­
den. The most straight­
for­ward man­age­ment is to delay preg­nancy and con­tinue TKI
ther­apy in the pur­suit of deeper responses. If the patient is on
imatinib, this could be switched to a more potent TKI. If this is
not fea­si­ble due to patient pref­er­ence, then once preg­nancy is
con­firmed the TKI must be stopped and the patient man­aged
sim­i­larly to women diag­nosed in preg­nancy.
A third option is assisted con­cep­tion, which may be par­tic­
u­
larly rel­
e­
vant in older women concerned about age-related
fer­til­ity loss. Depending on per­sonal cir­cum­stances and the con­
sent of their part­ner, these women can be con­sid­ered for ovar­
ian hyper­stim­u­la­tion, embryo cre­a­tion, and sub­se­quent embryo
implan­ta­tion in order to reduce the time off TKIs. Women with­out
part­ners can be offered oocyte or ovar­ian cryo­pres­er­va­tion. Strategies such as TKI hol­i­days, in which highly moti­vated patients
with reg­
u­
lar cycles discontinue their TKI after men­
stru­
a­
tion,
per­form a preg­nancy test 5 days post the expected ovu­la­tion,
and resume their TKI if neg­a­tive, are the­o­ret­i­cally pos­si­ble but
carry a risk of dis­
ease pro­
gres­
sion and can­
not be rou­
tinely
recommended.
For women in MMR but not deep molec­u­lar remis­sion, pre­
vi­ous guide­lines recommended that those who had sustained
MMR for more than 12 months could safely discontinue their TKI
and sub­se­quently regain the same dis­ease con­trol after restarting.7 Table 3 sum­ma­rizes the stud­ies reporting dis­ease con­trol
fol­low­ing TKI dis­con­tin­u­a­tion for preg­nancy. These all­ dem­on­
strated that while a sig­nif­i­cant pro­por­tion of patients lose MMR
upon TKI dis­con­tin­u­a­tion, their orig­i­nal response can be regained
in the major­ity fol­low­ing TKI resump­tion.40-43 Furthermore, Lee
et al dem­on­strated, impor­tantly, that almost half of those who
lost MMR retained CCyR and could be man­aged with­out inter­
ven­tion through­out preg­nancy.44
Table 3. Summary of recent stud­ies reporting dis­ease out­comes in patients who discontinue TKI while in MMR or deeper
N (patients,
preg­nan­cies)
TKI ther­apy prior
to stop, n (%)
Time on TKI in
months, median
(range)
≥MMR at time of
stop­ping TKI, n (%)
Loss of MMR, n (%)
≥MMR recov­ery after
TKI restart in those
who lost MMR, n (%)
28, 38
ND
ND
29 (76)
20 (69)
18 (90)a
Chelysheva et al
87, 87
ND
ND
87 (100)
6 mo—57%
12 mo—66%
6 mo—50%
12 mo—75%
Lee et al44
39, 50
Imatinib 24 (48)
Dasatinib 14 (28)
Nilotinib 10 (20)
Bosutinib 1 (2)
Radotinib 1 (2)
77 (6-194)
44 (88)
20 (45.5)
ND
Lasica et al42
16, 27
Imatinib 13 (81)
Dasatinib 2 (13)
Nilotinib 1 (6)
32 (3-84)
12 (75)
11 (92)
12 (100)
Dou et al43
17, 17
Imatinib 13 (76)
Nilotinib 4 (24)
49 (6-102)
17 (100)
10 (59)
10 (100)
Study
Caldwell et al40
41
The remaining 2 patients were too early for eval­u­a­tion.
ND, no data.
a
126 | Hematology 2022 | ASH Education Program
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/123/2021708/123robertson.pdf by guest on 09 December 2022
molec­u­lar remis­sion for more than 2 years are eli­gi­ble for treat­
ment dis­con­tin­u­a­tion.4,37 Approximately 50% sub­se­quently lose
their MMR, requir­ing TKI resump­tion, 80% of which occurs within
4 to 6 months.38,39 Women who wish to become preg­nant who
are TFR eli­gi­ble are the most straight­for­ward because they can
be man­aged sim­i­larly to their non­preg­nant coun­ter­parts.4 Assi
et al reported planned preg­nan­cies in 7 patients, 6 of whom had
a sustained MR4.5 for at least 18 months.9 Five of these patients
sustained at least an MMR through­out preg­nancy. The 1 remaining patient rap­idly lost the response after stop­ping dasatinib
prior to con­cep­tion but, reas­sur­ingly, regained MR4 6 months
after restarting.
TKIs can be discontinued and the patient mon­
i­
tored for
increases in tran­script lev­els at 4-to-6-week inter­vals: a patient
remaining in MMR can stay off treat­ment indef­i­nitely, regard­less
of preg­nancy. In the event of MMR loss, man­age­ment depends
on whether the woman is preg­nant. If she is not, then she should
resume her orig­i­nal or a more potent TKI and reestablish a deep
response, with the pos­si­bil­ity of attempting preg­nancy in the
future. This of course depends on indi­
vid­
ual cir­
cum­
stances,
includ­ing, impor­tantly, her age. If this is a con­sid­er­ation, then
var­i­ous pos­si­bil­i­ties can be con­sid­ered, includ­ing refer­ral for
oocyte, ovar­ian, or embryo stor­age. If she is preg­nant, then man­
age­ment depends on the level of tumor bur­den. Many women
reach the end of their preg­nancy with­out los­ing CCyR or com­
plete hema­to­log­i­cal remis­sion (CHR) and do not require treat­
ment. If inter­ven­tion is nec­es­sary, then the choices are iden­ti­cal
to those for women presenting in preg­nancy.
In any of these sce­nar­ios, once con­cep­tion is con­firmed, TKI
ther­apy should be discontinued imme­di­ately. There is no need
for increased imag­ing sur­veil­lance if no abnor­mal­i­ties are noted
at the first rou­tine ultra­sound. Those in CCyR should have tran­
script lev­els eval­u­ated every 4 to 6 weeks, whereas those with
less deep responses should be mon­i­tored as appro­pri­ate for
their indi­vid­ual level of dis­ease con­trol. Management is sim­i­lar to
that of patients diag­nosed in preg­nancy.
A loss of dis­ease con­trol can be man­aged in a man­ner sim­i­lar
to those described pre­vi­ously: uti­li­za­tion of IFN-α, ther­a­peu­tic
leukapheresis, and the con­sid­er­ation of TKI use in later preg­
nancy.
Pregnancy and CML can now be suc­cess­fully man­aged in uni­son
with min­i­mal risk to both mother and child. A sig­nif­i­cant pro­
por­tion of patients can be man­aged by obser­va­tion only, and
dis­ease con­trol can be achieved after starting or restarting TKI
ther­apy. If active treat­ment is required, IFN-α can be used in
any tri­mes­ter with min­i­mal risk to the fetus. More recently, the
safe admin­is­tra­tion of TKIs in the later stages of preg­nancy was
reported. Most impor­tantly, a mul­ti­dis­ci­plin­ary approach should
be employed to pro­vide com­pre­hen­si­ble and suc­cinct advice to
empower both mother and fam­ily in the shared deci­sion-mak­ing
pro­cess.
Acknowledgment
Jane F. Apperley and Harry F. Robertson acknowl­
edge the
sup­
port of the National Institute of Health Research Imperial
Biomedical Research Center.
Conflict-of-inter­est dis­clo­sure
Harry F. Robertson: no com­pet­ing finan­cial inter­ests to declare.
Jane F. Apperley: advi­sory board mem­ber: Incyte, Novartis; research funding: Incyte, Pfizer; speak­ers bureau: Incyte, Novartis.
Off-label drug use
Harry F. Robertson: nothing to disclose.
Jane F. Apperley: nothing to disclose.
Correspondence
Jane F. Apperley, Center for Hematology, Imperial College London, Hammersmith Hospital, Du Cane Rd, London W12 0NN,
United Kingdom; e-mail: j​­.apperley@impe­rial​­.ac​­.uk.
References
1.
2.
3.
4.
5.
6.
Hehlmann R, Lauseker M, Saußele S, et al. Assessment of imatinib as firstline treat­ment of chronic mye­loid leu­ke­mia: 10-year sur­vival results of the
ran­dom­ized CML study IV and impact of non-CML deter­mi­nants. Leukemia.
2017;31(11):2398-2406.
Höglund M, Sandin F, Simonsson B. Epidemiology of chronic mye­loid leu­
kae­mia: an update. Ann Hematol. 2015;94(suppl 2):S241-S247.
Malhotra P, Varma S. Chronic mye­
loid leu­
kae­
mia in India. Lancet.
2007;370(9593):1127.
Hochhaus A, Baccarani M, Silver RT, et al. Euro­pean LeukemiaNet 2020
rec­om­men­da­tions for treating chronic mye­loid leu­ke­mia. Leukemia.
2020;34(4):966-984.
Rohilla M, Rai R, Yanamandra U, et al. Obstetric com­pli­ca­tions and man­
age­ment in chronic mye­loid leu­ke­mia. Indian J Hematol Blood Transfus.
2016;32(1):62-66.
Lichtman M. Acute Myelogenous Leukemia. Williams Hematology. 6th ed.
New York: McGraw Hill; 2001:1047.
Abruzzese E, Mauro M, Apperley J, Chelysheva E. Tyrosine kinase inhib­i­tors
and preg­nancy in chronic mye­loid leu­ke­mia: opin­ion, evi­dence, and rec­
om­men­da­tions. Ther Adv Hematol. 2020;11(Octo­ber):2040620720966120.
8. Jain P, Kantarjian HM, Ghorab A, et al. Prognostic fac­
tors and sur­
vival
out­
comes in patients with chronic mye­
loid leu­
ke­
mia in blast phase in
the tyro­sine kinase inhib­i­tor era: cohort study of 477 patients. Cancer.
2017;123(22):4391-4402.
9. Assi R, Kantarjian H, Keating M, et al. Management of chronic mye­
loid leu­ke­mia dur­ing preg­nancy among patients treated with a tyro­sine
kinase inhib­i­tor: a sin­gle-cen­ter expe­ri­ence. Leuk Lymphoma. 2021;62(4):
909-917.
10. Chelysheva E, Abruzzese E, Rea D, et al. Chronic mye­loid leu­ke­mia diag­
nosed dur­
ing preg­
nancy: ther­
apy, out­
comes and fol­
low-up. Blood.
2018;132(suppl 1):4255.
11. Lishner M, Avivi I, Apperley JF, et al. Hematologic malig­nan­cies in preg­
nancy: man­age­ment guide­lines from an inter­na­tional con­sen­sus meet­ing. J
Clin Oncol. 2016;34(5):501-508.
12. Vandenbroucke T, Verheecke M, Fumagalli M, Lok C, Amant F. Effects of
can­cer treat­ment dur­ing preg­nancy on fetal and child devel­op­ment. Lancet Child Adolesc Health. 2017;1(4):302-310.
13. Pye SM, Cortes J, Ault P, et al. The effects of imatinib on preg­nancy out­
come. Blood. 2008;111(12):5505-5508.
14. Ishida Y, Murai K, Yamaguchi K, et al; Inter-Michinoku Dasatinib Study
Group. Pharmacokinetics and phar­
ma­
co­
dy­
nam­
ics of dasatinib in the
chronic phase of newly diag­nosed chronic mye­loid leu­ke­mia. Eur J Clin
Pharmacol. 2016;72(2):185-193.
15. Abruzzese E, Trawinska MM, Perrotti AP, De Fabritiis P. Tyrosine kinase inhib­
i­tors and preg­nancy. Mediterr J Hematol Infect Dis. 2014;6(1):e2014028.
16. Cortes JE, Abruzzese E, Chelysheva E, Guha M, Wallis N, Apperley
JF. The impact of dasatinib on preg­
nancy out­
comes. Am J Hematol.
2015;90(12):1111-1115.
17. Barkoulas T, Hall PD. Experience with dasatinib and nilotinib use in preg­
nancy. J Oncol Pharm Pract. 2018;24(2):121-128.
18. Cortes JE, Gambacorti-Passerini C, Deininger M, Abruzzese E, DeAnnuntis
L, Brümmendorf TH. Pregnancy out­comes in patients treated with bosutinib. Int J Hematol Oncol. 2020;9(2):IJH26.
19. Chelysheva E, Turkina A, Rea D, et al. Pregnancy out­
come in female
patients with chronic mye­loid leu­ke­mia world­wide: anal­y­sis of 305 cases
of the Euro­pean LeukemiaNet reg­is­try: S881. HemaSphere. 2019;3(suppl 1):
395-396.
20. Abruzzese E, Aureli S, Bondanini F, et al. Chronic mye­loid leu­ke­mia and
preg­nancy: when dreams meet real­ity. State of the art, man­age­ment and
out­come of 41 cases, nilotinib pla­cen­tal trans­fer. J Clin Med. 2022;11(7):
1801.
21. James AH, Brancazio LR, Price T. Aspirin and repro­
duc­
tive out­
comes.
Obstet Gynecol Surv. 2008;63(1):49-57.
22. Deruelle P, Coulon C. The use of low-molec­
u­
lar-weight hep­
a­
rins in
preg­nancy—how safe are they? Curr Opin Obstet Gynecol. 2007;19(6):573577.
23. Saleh AJ, Alhejazi A, Ahmed SO, et al. Leukemia dur­ing preg­nancy: long
term fol­low up of 32 cases from a sin­gle insti­tu­tion. Hematol Oncol Stem
Cell Ther. 2014;7(2):63-68.
24. Moura AC, Delamain MT, Duarte GBO, Lorand-Metze I, Souza CA, Pagnano KBB. Management of chronic mye­loid leu­ke­mia dur­ing preg­nancy:
a ret­ro­spec­tive anal­y­sis at a sin­gle cen­ter. Hematol Transfus Cell Ther.
2019;41(2):125-128.
25. National Toxicology Program, US Department of Health and Human Services. NTP mono­graph: devel­op­men­tal effects and preg­nancy out­comes
asso­ci­ated with can­cer che­mo­ther­apy use dur­ing preg­nancy. https:​­/​­/ntp​
­.niehs​­.nih​­.gov​­/ntp​­/ohat​­/cancer_chemo_preg​­/chemopregnancy_monofinal_508​­.pdf.
26. Galera P, Haynes S, Sulmasy P, et al. Physiological mea­
sure­ments cor­
rob­o­rate symp­tom­atic improve­ment after ther­a­peu­tic leukapheresis in
a preg­nant woman with chronic mye­log­e­nous leu­ke­mia. J Clin Apher.
2016;31(4):393-397.
27. Yellu M, Pinkard S, Ghose A, Medlin S. CML in preg­nancy: a case report using
leukapheresis and lit­er­a­ture review. Transfus Apher Sci. 2015;53(3):289292.
28. Balsat M, Etienne M, Elhamri M, Hayette S, Salles G, Thomas X. Successful preg­nan­cies in patients with BCR-ABL-pos­i­tive leu­ke­mias treated with
inter­feron-alpha ther­apy dur­ing the tyro­sine kinase inhib­i­tors era. Eur J
Haematol. 2018;101(6):774-780.
Dr Prakash Singh Shekhawat
Treatment of CML in preg­nancy | 127
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/123/2021708/123robertson.pdf by guest on 09 December 2022
Conclusion
7.
128 | Hematology 2022 | ASH Education Program
38. Campiotti L, Suter MB, Guasti L, et al. Imatinib dis­con­tin­u­a­tion in chronic
mye­loid leu­kae­mia patients with unde­tect­able BCR-ABL tran­script level: a
sys­tem­atic review and a meta-anal­y­sis. Eur J Cancer. 2017;77(May):48-56.
39. Shah NP, García-Gutiérrez V, Jiménez-Velasco A, et al. Dasatinib dis­con­
tin­u­a­tion in patients with chronic-phase chronic mye­loid leu­ke­mia and
sta­ble deep molec­u­lar response: the DASFREE study. Leuk Lymphoma.
2020;61(3):650-659.
40. Caldwell L, Claudiani S, Milojkovic D, et al. Changes in RT-qPCR lev­els dur­
ing and after TKI inter­rup­tions for attempted con­cep­tions in women with
CML. Blood. 2017;130(suppl 1):1612.
41. Chelysheva E, Apperley J, Abruzzese E, et al. Kinetics of the leu­
ke­
mic
clone in patients with chronic mye­loid leu­ke­mia dur­ing preg­nancy. Blood.
2018;132(suppl 1):4254.
42. Lasica M, Willcox A, Burbury K, et al. The effect of tyro­sine kinase inhib­
i­tor inter­rup­tion and inter­feron use on preg­nancy out­comes and longterm dis­
ease con­
trol in chronic mye­
loid leu­
ke­
mia. Leuk Lymphoma.
2019;60(7):1796-1802.
43. Dou X, Qin Y, Huang X, Jiang Q . Planned preg­nancy in female patients
with chronic mye­loid leu­ke­mia receiv­ing tyro­sine kinase inhib­i­tor ther­apy.
Oncologist. 2019;24(11):e1141-e1147.
44. Lee JO, Kim DW, Abruzzese E, Apperley J, Caldwell L, Mauro MJ. Kinetics
of BCR-ABL after TKI inter­rup­tion dur­ing preg­nancy in CML: a mul­ti­na­tional
ret­ro­spec­tive anal­y­sis. Blood. 2018;132(suppl 1):4263.
© 2022 by The Amer­i­can Society of Hematology
DOI 10.1182/hema­tol­ogy.2022000330
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/123/2021708/123robertson.pdf by guest on 09 December 2022
29. Yazdani Brojeni P, Matok I, Garcia Bournissen F, Koren G. A sys­tem­atic
review of the fetal safety of inter­feron alpha. Reprod Toxicol. 2012;33(3):
265-268.
30. Beauverd Y, Radia D, Cargo C, et al. Pegylated inter­feron alpha-2a for
essen­tial thrombocythemia dur­ing preg­nancy: out­come and safety. A case
series. Haematologica. 2016;101(5):e182-e184.
31. Talpaz M, Hehlmann R, Quintás-Cardama A, Mercer J, Cortes J. Re-emer­gence
of inter­feron-α in the treat­ment of chronic mye­loid leu­ke­mia. Leukemia.
2013;27(4):803-812.
32. Ali R, Ozkalemkas F, Kimya Y, et al. Imatinib use dur­ing preg­nancy and
breast feed­ing: a case report and review of the lit­er­a­ture. Arch Gynecol
Obstet. 2009;280(2):169-175.
33. Russell MA, Carpenter MW, Akhtar MS, Lagattuta TF, Egorin MJ. Imatinib
mesylate and metab­o­lite con­cen­tra­tions in mater­nal blood, umbil­i­cal cord
blood, pla­centa and breast milk. J Perinatol. 2007;27(4):241-243.
34. Kronenberger R, Schleyer E, Bornhäuser M, Ehninger G, Gattermann N,
Blum S. Imatinib in breast milk. Ann Hematol. 2009;88(12):1265-1266.
35. Burwick RM, Kuo K, Brewer D, Druker BJ. Maternal, fetal, and neo­na­tal
imatinib lev­els with treat­ment of chronic mye­loid leu­ke­mia in preg­nancy.
Obstet Gynecol. 2017;129(5):831-834.
36. Chelysheva E, Aleshin S, Polushkina E, et al. Breastfeeding in patients with
chronic mye­loid leu­kae­mia: case series with mea­sure­ments of drug con­
cen­tra­tions in mater­nal milk and lit­er­a­ture review. Mediterr J Hematol
Infect Dis. 2018;10(1):e2018027.
37. Deininger MW, Shah NP, Altman JK, et al. Chronic mye­loid leu­ke­mia, ver­sion
2.2021, NCCN clin­i­cal prac­tice guide­lines in oncol­ogy. J Natl Compr Canc
Netw. 2020;18(10):1385-1415.
BEYOND ROUTINE FRONTLINE THERAPY OF CML
Management of TKI-resistant chronic phase CML
Precision Cancer Medicine Theme, South Australian Health and Medical Research Institute, Adelaide, Australia; 2Department of Haematology, Royal
Adelaide Hospital and SA Pathology, Adelaide, Australia; 3Adelaide Medical School, University of Adelaide, Adelaide, Australia; 4Department of
Genetics and Molecular Pathology and Centre for Cancer Biology, SA Pathology, Adelaide, Australia; and 5Clinical and Health Sciences, University of
South Australia, Adelaide, Australia
1
Chronic phase CML (CP-CML) patients who are resistant to 2 or more tyrosine kinase inhibitors (TKIs) have limited therapeutic options and are at significant risk for progression to the blast phase. Ponatinib has been the drug of choice
in this setting for the past decade, but when given at full dose (45 mg/d), the risk of serious vascular occlusive events
is substantial. Lower doses mitigate this risk but also reduce the efficacy. Emerging data suggest that a high dose of
ponatinib is important to achieve response, but a lower dose is usually sufficient to maintain response, introducing a
safer therapeutic pathway for many patients. The recent development and approval of the novel allosteric ABL1 inhibitor, asciminib, for CP-CML patients with resistant disease provides another potentially safe and effective option in this
setting. These recent therapeutic advances mean that for most resistant CP-CML patients who have failed 2 or more
TKIs, 2 excellent options are available for consideration—dose modified ponatinib and asciminib. Patients harboring
the T315I mutation are also candidates for either ponatinib or asciminib, but in this setting, higher doses are critical to
success. Lacking randomized comparisons of ponatinib and asciminib, the best choice for each clinical circumstance
is often difficult to determine. Here we review emerging evidence from recent trials and make some tentative suggestions about which drug is preferable and at what dose in different clinical settings using case studies to illustrate the
key issues to consider.
LEARNING OBJECTIVES
• Determine appropriate assessments for CP-CML patients being considered for a change in therapeutic approach
because of TKI resistance
• Recommend the optimal TKI and the dosing regimen for a CP-CML patient with TKI resistance
What disease and patient assessments are indicated
for the following 3 patients with resistant chronic phase
chronic myeloid leukemia (CP-CML), and what is the best
therapeutic approach in each case?
CLINICAL CASE 1
A 32-year-old woman with CP-CML and a high-risk EUTOS
Long-Term Survival (ELTS) score received frontline dasatinib
at 100 mg/d for 4 months, then switched therapy after confirmed early molecular response failure (>10% BCR::ABL1IS).
After 4 months of nilotinib at a dose of 400 mg twice daily,
there was no improvement in her molecular response.
CLINICAL CASE 2
A 55-year-old woman with CP-CML and a low-risk ELTS
score was started on nilotinib and achieved a major molecular remission (MMR; <0.1% BCR::ABL1IS) by 12 months. After
18 months, she sequentially switched to bosutinib, then
dasatinib, both for intolerance issues, but gradually lost
MR2 (>1% BCR::ABL1IS).
CLINICAL CASE 3
A 72-year-old man with an intermediate-risk ELTS score
failed to achieve MR2 by 12 months on imatinib then
received dasatinib at 100 mg/d for 6 months, with an ini-
Dr Prakash Singh Shekhawat
TKI-resistant CP-CML | 129
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/129/2021810/129hughes.pdf by guest on 09 December 2022
Timothy P. Hughes1–3 and Naranie Shanmuganathan1–5
tial tran­sient improve­ment in response (0.5% BCR::ABL1IS). He
then rap­idly lost molec­u­lar response (15% BCR::ABL1IS).
Introduction
The OPTIC trial
Modeling based on the PACE study and other stud­ies suggested
the risk of AOEs was sig­nif­i­cantly lower at lower doses,11,16 lead­
ing to the approach adopted for the OPTIC trial.13 In OPTIC,
patients were ran­dom­ized to receive either 45 mg/d, 30 mg/d,
or 15 mg/d as a starting dose.13 In the case of patients in the
higher-dos­ing cohorts, the dose was reduced to 15 mg/d once
MR2 (BCR::ABL1IS < 1%) was achieved.13 As predicted, the dos­ing
strat­egy of 45 mg/d with dose reduc­tion proved to be sig­nif­i­
cantly less toxic in terms of AOEs com­pared to the PACE data,
where doses were often maintained at 45 mg/d through­out.13
Remarkably, this approach using high doses ini­tially and deescalated doses in main­te­nance also proved to be highly effi­ca­cious
and sig­nif­i­cantly supe­rior to the other 2 arms, with an MR2 rate
at 12 months of 52% in the 45-mg cohort vs 36% in the 30-mg
cohort and 25% in the 15-mg cohort.13 Furthermore, the OPTIC
study pro­vi­des fur­ther insights into the del­i­cate bal­ance of effi­
cacy and safety that needs to be con­sid­ered when selecting the
ini­tial dose (Figure 1). In resis­tant cases with the T315I muta­tion,
Figure 1. Overall safety and efficacy of ponatinib based on starting dose. The analysis is a descriptive clinical summary of the data to
illustrate the relationship between the efficacy and the AOE rate. TE-AOE, treatment-emergent arterial occlusive events. Reproduced
with permission from Cortes et al.13
130 | Hematology 2022 | ASH Education Program
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/129/2021810/129hughes.pdf by guest on 09 December 2022
For most patients with CP-CML who have achieved and maintained an opti­mal molec­u­lar response, the ongo­ing man­age­
ment is fairly straight­
for­
ward. Ensuring opti­
mal molec­
u­
lar
tar­gets are met and maintained, man­ag­ing drug-related side
effects, mon­i­tor­ing drug adher­ence, and pay­ing care­ful atten­
tion to min­i­mize drug-induced organ toxicities are the main
pri­or­i­ties.1,2 At the extremes of response, man­
age­
ment can
be more com­plex. Patients achiev­ing and maintaining deep
molec­u­lar responses over sev­eral years are con­sid­ered can­di­
dates for treat­ment-free remis­sion, although embarking on a
treat­ment-free remis­sion attempt requires exten­sive dis­cus­sion
and patient edu­ca­tion with a care­ful review of eli­gi­bil­ity. Furthermore, once ther­apy has been ceased, fre­quent mon­i­tor­ing
is essen­tial, as well as a high degree of vig­i­lance looking for
molec­u­lar relapse.3-6 At the oppo­site end of the response spec­
trum, treat­ment fail­ure to at least 2 tyro­sine kinase inhib­i­tors
(TKIs) or to just 1 sec­ond-gen­er­a­tion TKI (2G-TKI) requires expert
man­age­ment while still aiming to max­i­mize opti­mal responses
and also to pre­vent dis­ease pro­gres­sion and TKI-induced tox­
ic­
ity. Treatment fail­
ure to one or more 2G-TKIs infre­
quently
can be sal­vaged by switching to another 2G-TKI,7-10 usu­ally in
cases in which spe­
cific ABL1 kinase domain (KD) muta­
tions
have emerged. The clin­i­cal devel­op­ment of ponatinib has pro­
vided an effec­tive option for many of these resis­tant patients.
The PACE study dem­on­strated its effec­tive­ness in the set­ting
of resis­tant dis­ease, both in the case of resis­tance asso­ci­ated
with a KD muta­tion, includ­ing the gate­keeper T315I muta­tion,
and in cases with­out clear iden­ti­fi­able resis­tance mech­a­nisms.11
The key lim­i­ta­tion with ponatinib ther­apy that emerged in the
PACE study and other stud­ies was the very high rate of arte­rial
occlu­sive events (AOEs) observed, which neces­si­tated the pre­
ma­ture ter­mi­na­tion of some ponatinib-based clin­i­cal tri­als and
a tran­sient with­drawal from the mar­ket.11,12 In the CP-CML arm of
the PACE study, the reported rate of AOEs was 31%.11
The ther­a­peu­tic chal­lenges related to man­ag­ing these TKIresis­tant CP-CML patients have become more com­plex recently
because we now have safer and argu­­
ably bet­
ter ther­
a­
peu­
tic
options to con­sider. The recent dem­on­stra­tion from the OPTIC
trial of a less toxic but still effi­ca­cious dos­ing reg­i­men for ponatinib has changed the risk/ben­e­fit equa­tion, mak­ing ponatinib a
more attrac­tive option in this set­ting.13 Equally prom­is­ing has been
the approval of the allo­ste­ric ABL1 inhib­i­tor asciminib, which offers
good tol­er­a­bil­ity and high effi­cacy in sim­i­lar high-risk set­tings.14,15
Clinicians now have sev­eral options for patients with 2G-TKI resis­
tance, but it is often not clear which is the best drug choice and
the opti­mal dose reg­i­men. The pros­pects of good dis­ease con­
trol and the risks of tox­ic­ity with each option need to be care­fully
assessed. To pro­vide some clar­ity about how to make the best
deci­sion for a par­tic­u­lar patient, we explore the emerg­ing find­ings
from recent tri­als for resis­tant patients. We also pres­ent and dis­
cuss sev­eral clin­i­cal cases to illus­trate the value of con­sid­er­ing all­
patient- and drug-spe­cific param­e­ters before selecting the opti­
mal ther­a­peu­tic approach for each indi­vid­ual patient.
The ASCEMBL trial
The ASCEMBL trial ran­dom­ized CP-CML patients who were intol­
er­ant or resis­tant to 2 or more TKIs to either asciminib or bosutinib in a 2:1 ratio.15 Asciminib dos­ing was set at 40 mg twice daily
based on the tol­er­a­bil­ity assess­ments of a num­ber of dos­ing lev­
els in the phase 1 study and phar­ma­co­ki­netic mod­el­ing dem­on­
strat­ing that either 40 mg twice daily or 80 mg/d of asciminib
exceeded the 90% thresh­
old for mean inhib­
i­
tory con­
cen­
tra­
tions.14 It dem­on­strated a clearly supe­rior rate of MMR (BCR::ABL1IS
< 0.1%) by 6 months for asciminib (26% vs 13%).15 However, MR2
may be a more real­is­tic tar­get in this multiresistant set­ting. The
dif­fer­ence was sig­nif­i­cant here as well—by 12 months, 42% vs
19% had achieved MR2.9 Another major dif­fer­ence between the 2
arms was the rate of dis­con­tin­u­a­tion. In the most recent update,17
43% of patients on the asciminib arm discontinued ther­
apy,
includ­ing 24% for lack of effi­cacy and 6% for adverse events
(median fol­low-up, 19 months). In con­trast, 78% were no lon­ger
tak­ing bosutinib, with 36% stop­ping due to lack of effi­cacy and
24% because of an adverse event.17 There were 7 patients on the
asciminib arm who had AOEs at the most recent anal­y­sis, which
rep­re­sents a rate of 3.4 per 100 patient-years.17 All 7 patients had
received nilotinib pre­
vi­
ously, and 3 had received ponatinib.17
Nonetheless, the observed rate of AOEs was higher with asciminib com­
pared with bosutinib. ASCEMBL dem­
on­
strated the
clear supe­ri­or­ity of asciminib over bosutinib in the resis­tant and
intol­er­ant third-line set­ting.
Comparing ponatinib and asciminib trial data
The best way to deter­mine whether ponatinib or asciminib is
supe­rior as a third-line option in dif­fer­ent clin­i­cal set­tings would
be to con­
duct head-to-head ran­
dom­
ized stud­
ies. However,
as these have not been ini­ti­ated at this stage, we pres­ent our
opin­ions based on the data avail­­able from the phase 1 asciminib
trial, the ASCEMBL trial, and the OPTIC trial to for­mu­late an ini­
tial assess­
ment regard­
ing which drug may be pref­
er­
en­
tial in
spe­cific cases (Table 1). OPTIC recruited a higher pro­por­tion of
patients with KD muta­tions and a higher pro­por­tion of patients
who failed to achieve any molec­u­lar response to their prior TKI.13
In com­par­i­son, ASCEMBL had a higher pro­por­tion of intol­er­ant
patients com­pared to OPTIC. Overall response rates look sim­i­lar
when com­par­ing the 45-mg cohort of OPTIC to the asciminib
arm of ASCEMBL.13,17 In patients with­out T315I, the 12-month rates
of MR2 were sim­i­lar; 42% for asciminib and 49% for the ponatinib
45-mg cohort.13,15 However, response rates appear to be more
var­
i­
able according to the base­
line response in the asciminib
cohort. Notably, only 17% of the patients not in a major cyto­ge­
netic response (MCyR) at base­line achieved MMR at 6 months,
com­pared to 40% of patients who were in MCyR at base­line
(Figure 2).15 Moreover, the cur­rent accepted stan­dard dose of
asciminib of 40 mg twice daily may be inad­e­quate to com­bat
highly resis­tant CML, as evidenced by the higher doses required
in T315I-mutated patients.14 Comparable data from OPTIC are not
avail­­able, but in the 45-mg ponatinib cohort, the MR2 rates were
50% by 12 months regard­less of prior response on his­tor­i­cal TKI
ther­apy.13 On this (ten­ta­tive) basis, ponatinib may be a bet­ter
choice in highly resis­tant cases, and asciminib, with its supe­rior
safety pro­file, may be pref­er­a­ble in patients with less pro­found
resis­tance.
Another impor­tant com­par­i­son is the risk of AOEs with each
agent. Many var­i­ables make this com­par­i­son prob­lem­atic. These
include dif­fer­ent eli­gi­bil­ity cri­te­ria regard­ing car­diac risk fac­tors
and pre­vi­ous car­dio­vas­cu­lar his­tory, dif­fer­ent dura­tion of fol­lowup, and dif­fer­ent cri­te­ria for deter­min­ing a treat­ment-emer­gent
AOE. Acknowledging these lim­i­ta­tions seems to show an advan­
tage for asciminib, but this needs care­ful assess­ment with ongo­
ing data from these and other stud­ies. Rates of AOEs per 100
patient-years were 9.6 for the ponatinib 45-mg cohort in OPTIC
com­pared to 3.4 for asciminib in the ASCEMBL trial.
Optimal ther­apy for patients with T315I muta­tion
Both ponatinib and asciminib are active against the T315I muta­
tion, but in both cases higher doses are impor­tant for effi­cacy.
In the OPTIC trial, 15 of 25 (60%) patients with T315I receiv­ing
the 45-mg ini­tial dose achieved MR2 by 12 months, com­pared
to 5 of 20 (25%) who started at 30 mg/d.13 Asciminib effi­cacy
against the T315I mutant was assessed in the phase 1 trial, but
response rates were low until doses above 150 mg twice daily
were given.14 For that rea­son an expanded cohort of 48 CP-CML
and 4 acute-phase CML patients with the T315I muta­tion were
given asciminib at 200 mg twice daily.18 At this dose, molec­
u­lar responses were fre­quent, and tox­ic­ity appeared sim­i­lar
to stan­dard -ose asciminib (Table 2). With a median fol­low-up
of 17 months, 35 of 52 (67%) remained on ther­apy, and 23 of
49 (47%) who were not in MMR at base­line achieved MMR by
24 months.18 This is impres­sive, espe­cially con­sid­er­ing that 28 of
52 (54%) had prior ther­apy with ponatinib and had devel­oped
either intol­
er­
ance or resis­
tance. In ponatinib-naive patients,
the rate of MMR was 57% by 6 months, com­pared to 29% in
ponatinib-resis­tant/intol­er­ant patients.18
Thus, both drugs are highly active when used at the cor­rect
dose for patients with the T315I muta­tion. When patients on
ponatinib 45 mg/d deescalated dos­ing as per the OPTIC pro­
to­col, the over­all rate of loss of response exceeded 30% (the
actual rate was not spec­i­fied in the OPTIC paper).13 Therefore, a
risk assess­ment should be under­taken before dose deescalating
in the set­ting of ponatinib ther­apy for patients with the T315I
muta­tion. In sup­port of asciminib as the first choice in the T315I
set­ting, tol­er­ance appears supe­rior to ponatinib 45 mg (OPTIC
dos­
ing), while effi­
cacy appears sim­
i­
lar (Table 2).18 In sup­
port
of ponatinib as the first choice, we have more data and lon­ger
fol­low-up to reas­sure us that responses are sta­ble and dura­ble in
T315I patients. Since there is no clear “win­ner,” the choice should
Dr Prakash Singh Shekhawat
TKI-resis­tant CP-CML | 131
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/129/2021810/129hughes.pdf by guest on 09 December 2022
the advan­tage of starting at 45 mg/d over 30 mg/d is sig­nif­i­cant,
with MR2 rates by 12 months at 60% and 25%, respec­tively.13
There was also an advan­tage with the 45-mg starting dose com­
pared to the 30-mg cohort in patients with a non-T315I KD muta­
tion (56% vs 40% MR2 by 12 months).13 In patients with resis­tance
or intol­er­ance with­out a KD muta­tion, the advan­tage was less
marked (46% vs 38%).12
The higher effi­cacy achieved with higher starting doses needs
to be weighed against the higher tox­ic­ity. In patients with recent
AOEs on a prior TKI, even a brief expo­sure to full-dose ponatinib
may prove very haz­ard­ous. Treatment-emer­gent AOEs per 100
patient-years in OPTIC were 3.2 for the 15-mg cohort, 5.3 for the
30-mg cohort, and 9.6 for the 45-mg cohort.13 To dis­tin­guish this
dos­ing strat­egy from the con­ven­tional strat­egy of maintaining
the starting dose if it is well tol­er­ated, we refer to this dos­ing
reg­i­men as ponatinib 45 mg (OPTIC dos­ing) and ponatinib 30 mg
(OPTIC dos­ing) for clar­ity.
Table 1. Comparison of asciminib (ASCEMBL) to ponatinib (OPTIC, 45-mg-dose cohort)
Asciminib 40 mg twice daily, arm
of ASCEMBL study15,17
Ponatinib 45 mg daily, dose
adjusted from OPTIC trial13
Number of patients
157
92
Eligibility
ELN201324 treat­ment fail­ure (61%)
or intol­er­ance (38%)
Intolerant or resis­tant to >2 TKIs
and not in MR2
Exclusion cri­te­ria
Within 6 mo prior to starting study
treat­ment of myo­car­dial infarc­tion,
angina pectoris, cor­o­nary artery
bypass graft, or clin­i­cally
sig­nif­i­cant car­diac arrhyth­mias
Any his­tory of myo­car­dial
infarc­tion, unsta­ble angina,
cere­bro­vas­cu­lar acci­dent,
tran­sient ische­mic attack,
periph­eral vas­cu­lar dis­ease, NYHA
class III or IV con­ges­tive heart
fail­ure, or reduced left ven­tric­u­lar
ejec­tion frac­tion within 6 mo prior
to enroll­ment, clin­i­cally sig­nif­i­cant
car­diac arrhyth­mias within 6 mo
≥3 prior TKI lines
48%
53%
Resistant to last TKI
61%
98%
Disparity in favor of ponatinib in
highly resis­tant cases
Intolerant patients
38%
2%
Disparity in favor of asciminib in
intol­er­ant cases
MCyR at base­line
29%
35%
CCyR at base­line
10%
0%
13%
44%
>10% BCR::ABL1
62%
NA
Cardiac risk fac­tors
Not spec­i­fied
29% had 1, 5% had >1
Follow-up
19 mo
32 months
Still on study ther­apy
57%
47%
Discontinuation due to tox­ic­ity
6%
17%
Discontinuation due to lack of
effi­cacy
24%
19%
42%
44% (49% in non-T315I group)
MR2 by 12 mo in patients in CHR
or worse on prior TKI
NA
27/54 (50%)
MR2 by 12 mo in patients who
had a bet­ter response than CHR to
prior TKI
NA
14/28 (50%)
29%
NA – 34% “over­all rate with lat­est
fol­low-up”
MMR by 6 mo in patients in
MCyR at base­line
40%
NA
MMR by 6 mo in patients NOT in
MCyR at base­line
17%
NA
Suggestion from this data that
asciminib response may be
more depen­dent on the level of
response to prior ther­apy than is
the case for ponatinib
AOEs per 100 patient-years
3.4
7.6 (year 1)
5.9 (year 2)
In the ponatinib cohort
receiv­ing 30 mg ini­tially, AOE
rates/100 patient-years were
3 (year 1) and 2 (year 2)
Comment
KD muta­tions at base­line
(non-T315I)
Lower rate of patients with KD
muta­tions in ASCEMBL
Results
MR2 by 12 mo
MMR by 12 mo
Consistent with bet­ter tol­er­ance
to asciminib than ponatinib
Note high response rate on
ponatinib in highly resis­tant cohort
Data reproduced with per­mis­sion from Cortes et al,13 Réa et al,15 and Mauro et al.17
CCyR, com­plete cyto­ge­netic response; CHR, com­plete hema­to­logic response; NA, not avail­­able; NYHA, New York Heart Association.
132 | Hematology 2022 | ASH Education Program
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/129/2021810/129hughes.pdf by guest on 09 December 2022
Baseline fea­tures
be indi­vid­u­al­ized by weighing the dis­ease-related fac­tors and
the car­dio­vas­cu­lar risk.
Optimal ther­apy in the case of other KD muta­tions
Both ponatinib and asciminib are active against most aden­o­sine
tri­phos­phate-site KD muta­tions,19,20 although sen­si­tiv­ity is gen­
er­ally lower than it is for unmutated BCR::ABL1. KD muta­tions of
con­cern for asciminib include F359V/C/I, which either emerged
or persisted with­out evi­dence of molec­u­lar response in sev­eral
cases in ASCEMBL and the phase 1 study.15,21 This was predicted
to be a par­tially resis­tant muta­tion in in vitro stud­ies.21 More data
are needed to deter­mine asciminib activ­ity against the F317L
muta­tion, which appears to induce a degree of asciminib resis­
tance based on in vitro stud­ies (Figure 3).21 However, in all­presented sce­nar­ios, higher doses of asciminib may be required for
opti­mal effi­cacy, anal­o­gous to the higher doses nec­es­sary to
com­bat the T315I muta­tion.
Any role for 2G-TKIs in this set­ting?
In most cases the opti­
mal choice of next-line ther­
apy is
between ponatinib and asciminib. There are spe­cific cases in
which a 2G-TKI may still have a role. This would include cases
in which resis­tance is driven by a muta­tion that is well cov­ered
by another 2G-TKI. For instance, the devel­op­ment of V299L or
F317L muta­tions is highly sen­si­tive to nilotinib and bosutinib,
and F359V/I/C muta­tions are highly sen­si­tive to both dasatinib
and bosutinib (Figure 3).2
Additionally, in some countries eli­gi­bil­ity rules pre­vent access
to sequen­tial ther­apy with either asciminib or ponatinib in the
set­ting of 2G-TKI fail­ure with­out devel­op­ing resis­tance or intol­
er­ance to an alter­na­tive 2G-TKI. In this set­ting, a 3-to-6 month
trial is gen­er­ally suf­fi­cient to deter­mine if the newly com­menced
2G-TKI is likely to be effec­tive and sal­vage the sit­u­a­tion. It is rare
for patients who have no improve­ment in molec­u­lar response
after 6 months of a sal­vage ther­apy to even­tu­ally achieve MR2 or
bet­ter with lon­ger-term expo­sure.
The intol­er­ant patient
The focus of this chap­ter is TKI resis­tance, but it is worth commenting on the chal­leng­ing patient who is intol­er­ant to 2 or
more TKIs. In this set­ting the effi­cacy and tol­er­a­bil­ity of asciminib
look very favor­able based on the phase 1 asciminib study,14 and in
most cases, ponatinib would not be the pri­mary drug of choice
due to its higher rates of intol­er­ance and tox­ic­ity. Whether lower
doses of ponatinib could achieve com­pa­ra­ble response and tol­
er­ance to asciminib in these patients, where the pre­dom­i­nant
chal­lenge is intol­er­ance, remains to be established.
Future options
Targeting 2 dis­tinct regions on ABL1, the aden­o­sine tri­phos­
phate- and myristate-bind­ing sites, with asciminib and con­ven­
tional TKIs may also be a future option to com­bat resis­tance,
with the phase 1 clin­i­cal trial yet to for­mally report out­comes
for the rel­e­vant treat­ment arms that inves­ti­gated com­bi­na­tion
Dr Prakash Singh Shekhawat
TKI-resis­tant CP-CML | 133
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/129/2021810/129hughes.pdf by guest on 09 December 2022
Figure 2. Subgroup analysis of the ASCEMBL study: risk difference with 95% CIs for MR3 achievement by 24 weeks of either asciminib or bosutinib. CRF, case report form. Reproduced with permission from Rea et al.15
Table 2. Comparing base­line var­i­ables and out­comes for CP-CML patients receiv­ing asciminib at 200 mg twice daily
(phase 1 study) or dose-mod­i­fied ponatinib (OPTIC)
Ponatinib 45 mg (OPTIC)13
Ponatinib 30 mg (OPTIC)13
No. of patients
52
94
94
Follow-up
17 mo
32 mo
32 mo
Molecular sta­tus at
base­line
BCR::ABL1IS
54%> 10%
25% 1-10%
15% <1%
Prior ponatinib ther­apy
28/52 (54%)
NA
NA
Discontinued ther­apy
17/52 (33%)
44/94 (47%)
For all­patients on 45 mg,
not just T315I cohort
53/95 (56%)
For all­patients on 30 mg,
not just T315I cohort
MMR rate at
12 mo
12/21 (57%) in ponatinibnaive cohort
NA
NA
MMR rate at 24 mo
14/21 (66%) in ponatinibnaive cohort
NA
NA
MR2 at 12 mo
NA
15/25 (60%)
5/20 (25%)
MR2 at 24 mo
NA
AOEs
2/52 (4%)
9/94 (9.6%)
5/94 (5.3%)
Comments
Results
Asciminib appears to be
bet­ter tol­er­ated
AOE risk with asciminib
may be sim­i­lar to ponatinib
30 mg (OPTIC dos­ing)
Data reproduced with per­mis­sion from Cortes et al,13 Hughes et al14, and Cortes et al.18
CCyR, com­plete cyto­ge­netic response; CHR, com­plete hema­to­logic response; NA, not avail­­able.
ther­apy.20 Furthermore, murine mod­els have dem­on­strated that
com­bin­ing asciminib with ponatinib can inhibit the growth of
T315I-containing com­pound muta­tions.21 A num­
ber of novel
BCR::ABL1–targeting agents are under inves­ti­ga­tion for resis­tant
CML. One such drug is vodobatinib, a novel 3G-TKI with lim­ited
off-tar­get activ­ity.22 The phase 1 dose esca­la­tion study recently
reported prom­
is­
ing safety and effi­
cacy results gen­
er­
ated in
both ponatinib-naive and ponatinib-treated patients.22 Furthermore, car­dio­vas­cu­lar tox­ic­ity did not appear to be a prominent
fea­
ture at this early stage, which may dis­
tin­
guish vodobatinib from the other avail­­able agents for resis­tant CML. Similarly, another 3G-TKI, olverembatinib, has also had early phase
1 results reported in first- and sec­
ond-gen­
er­
a­
tion-resis­
tant
CP-CML and acute-phase CML.23 Efficacy was observed regard­
less of the study entry response, pre­
dom­
i­
nantly in T315Imutated patients.23 Responses were less impres­
sive in the
absence of T315I. While lon­ger-term fol­low-up in larger data
sets is required, these agents may offer future options for this
resis­tant patient cohort.
CLINICAL CASE 1 (Con­t in­u ed)
This 32-year-old woman with CP-CML and a high-risk ELTS
score received front­line dasatinib at 100 mg/d for 4 months,
then switched to nilotinib at 400 mg twice daily for 4 months
134 | Hematology 2022 | ASH Education Program
after con­firmed early molec­u­lar response fail­ure. After a total of
8 months, her BCR::ABL1IS remained over 10%, hav­ing never
fallen below 10%.
Cytogenetics at 8 months did not show evi­dence of clonal
evo­lu­tion, but 18 of 20 meta­phases were Ph+. No KD muta­tion
was iden­ti­fied at 4 or 8 months. She had no car­dio­vas­cu­lar risk
fac­tors or sig­nif­i­cant comorbidities.
Options
1. Continue nilotinib and review at 12 months.
2. Switch to bosutinib.
3. Switch to ponatinib 45 mg/d (OPTIC dos­ing).
4. Switch to asciminib.
5. Proceed to an allo­graft.
Discussion of options
1. Continue nilotinib: Unless there are sig­nif­i­cant inter­rup­tions to
ther­apy due to tox­ic­ity issues or poor adher­ence that resolve
or are over­come, there is very lit­tle pros­pect that the molec­u­lar
response will improve by con­tinu­ing the cur­rent ther­apy. In this
set­ting, the risk of dis­ease pro­gres­sion is high. One impor­tant
con­sid­er­ation is the need to iden­tify those patients who will not
achieve response with any TKI as soon as pos­si­ble so that an allo­
graft can still be sched­uled in the chronic phase, where there is
still a good pros­pect of cure.
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/129/2021810/129hughes.pdf by guest on 09 December 2022
Asciminib 200 mg twice
daily18
2. Switch to bosutinib: The like­li­hood of a response if a third
2G-TKI is given is very low. In ASCEMBL, only 19% achieved MR2
by 12 months on the bosutinib arm, and the response rate was
even poorer in patients who were not in an MCyR at base­line, as
in this case.
3. Switch to ponatinib 45 mg/d (OPTIC dos­ing): Given the
com­plete lack of response after 8 months of potent TKI ther­apy,
this may be the best option. In OPTIC, 50% of patients who
were not in MCyR at base­line who received 45 mg/d of ponatinib achieved MR2 by 12 months. If a patient is given ponatinib
and does not achieve MR2 after a fur­ther 6 months, the pros­
pect of achiev­ing this tar­get with ongo­ing ponatinib is low,
and hence switching to asciminib or pro­ceed­ing to an allo­graft
would be appro­pri­ate.
4. Switch to asciminib: As discussed ear­lier, responses to asciminib are prob­a­bly infe­rior to ponatinib 45 mg/d (OPTIC dos­
ing) in patients with lit­tle or no response to prior ther­apy. The
excep­tion would be in cases in which the risk of a vas­cu­lar event
was high, where asciminib (or lower-dose ponatinib) may be
pre­ferred.
5. Proceed to an allo­graft: While an allo­graft is not directly
recommended at this stage, it would be pru­dent to tis­sue type
the patient and iden­tify an allo­ge­neic donor in case of fail­ure to
respond to third-line ther­apy. In the event of per­sis­tent ther­apy
fail­ure, the patient can pro­ceed to an allo­ge­neic stem cell trans­
plant promptly with­out delay.
Best option: ponatinib 45 mg daily with a dose reduc­tion to
15 mg daily if MR2 is achieved.
CLINICAL CASE 2 (Con­tin­ued)
The 55-year-old woman with CP-CML and a low-risk ELTS score
was started on nilotinib and responded well, achiev­ing MMR by
12 months. After 18 months, she devel­oped periph­eral vas­cu­lar
dis­ease and switched to bosutinib and then dasatinib due to
fur­ther tox­ic­ity but sequen­tially lost MMR and MR2 after 3 years.
Further assess­ments: There were no KD muta­tions or evi­
dence of clonal evo­lu­tion on mar­row exam­i­na­tion. There were
sub­stan­tial inter­rup­tions sec­ond­ary to adverse events as well
as peri­ods of poor drug adher­ence. She had periph­eral vas­cu­lar
dis­ease and well-con­trolled hyper­ten­sion.
Options
1. Switch to imatinib.
2. Switch to ponatinib 45 mg/d with dose reduc­tion once MR2
is achieved.
3. Switch to ponatinib 30 mg/d with dose reduc­tion once MR2
is achieved.
4. Switch to asciminib 40 mg twice daily.
Discussion of options
1. Switch to imatinib: It is pos­si­ble that her loss of response is
more related to exten­
sive inter­
rup­
tions to ther­
apy and poor
adher­ence rather than drug resis­tance, so imatinib may still be
an effec­tive drug in this set­ting. However, given her intol­er­ance
Dr Prakash Singh Shekhawat
TKI-resis­tant CP-CML | 135
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/129/2021810/129hughes.pdf by guest on 09 December 2022
Figure 3. Heat map summary of IC50 values for TKIs against a panel of Ba/F3 cell lines expressing KD mutations. ATP, adenosine
triphosphate. Reproduced with permission from Eide et al.21
of resis­tance to imatinib that was not driven by a KD muta­tion.
Bosutinib may still be a rea­son­able choice in this set­ting, espe­
cially given its lower risk of AOEs. However, the expe­ri­ence in
ASCEMBL would sug­gest that it may be poorly tol­er­ated.
2. Switch to nilotinib: The F317L muta­tion is sen­si­tive to nilotinib, but the risk of vas­cu­lar events is prob­a­bly too high, espe­
cially since a higher dose is likely to be needed (400 mg twice
daily).
3. Switch to ponatinib 45 mg/d (OPTIC dos­ing): Patients in
the OPTIC study with a non-T315I muta­tion who received the
higher ini­tial dose of ponatinib had a rate of MR2 by 12 months of
56%. The main con­cern would be the AOE risk given the patient’s
his­tory and poorly con­trolled hyper­ten­sion. This patient would
not have been eli­gi­ble for the OPTIC trial, so the rate of AOEs
would likely be sig­nif­i­cantly higher than the rate observed on
this arm of the OPTIC study.
4. Switch to ponatinib 30 mg/d (OPTIC dos­ing): Given the
con­cern about the risk of AOEs in this patient, even this mod­i­
fied dose rep­re­sents a sig­nif­i­cant risk. The rate of AOEs per 100
patient-years was 5.3 for this dose in OPTIC, but this patient
would not have been eli­gi­ble for OPTIC, so the AOE risk is prob­
a­bly higher.
5. Switch to asciminib 40 mg twice daily: There are very lim­
ited data regard­ing the sen­si­tiv­ity of the F317L muta­tion to asciminib. In vitro stud­ies sug­gest a level of resis­tance greater than
that seen with the T315I muta­tion. This muta­tion was only found
in 2 patients at base­line in the asciminib arm of ASCEMBL; nei­ther
had a molec­u­lar response.15 Along with the F359C/I/V muta­tion,
we need more clin­i­cal data before we can deter­mine whether
asciminib is effec­tive in this set­ting and, if so, at what dose.
Best option: bosutinib may be the saf­est option, with a switch
to ponatinib at a mod­i­fied dose if no molec­u­lar response occurs
after 3 to 6 months.
Acknowledgments
CLINICAL CASE 3 (Con­t in­u ed)
A 72-year-old man with inter­
me­
di­
ate-risk ELTS failed to
achieve MR2 by 12 months on imatinib. He then received
dasatinib at 100 mg/d for 6 months with an ini­tial improve­
ment in response (best response, 0.5% BCR::ABL1IS) but then
lost molec­u­lar MR2 (15% BCR::ABL1IS).Further assess­ments: He
had no muta­tion detected when switching to dasatinib, but
when he lost MR2, there was evi­dence of emer­gence of the
F317L muta­tion. From a car­dio­vas­cu­lar risk per­spec­tive, he had
a sin­gle cor­o­nary stent inserted 10 years ago and has poorly
con­trolled hyper­ten­sion.
Conflict-of-inter­est dis­clo­sure
Tim­
o­
thy P. Hughes: research funding: Novartis, Bristol Myers
Squibb; hon­o­raria: Takeda, Novartis, Bristol Myers Squibb.
Naranie Shanmuganathan: research funding: Novartis; hon­o­raria:
Takeda.
Off-label drug use
Tim­o­thy P. Hughes: nothing to disclose.
Naranie Shanmuganathan: nothing to disclose.
Options
5. Switch to bosutinib.
6. Switch to nilotinib.
7. Switch to ponatinib 45 mg/d (OPTIC dos­ing).
8. Switch to ponatinib 30 mg/d (OPTIC dos­ing).
9. Switch to asciminib 40 mg twice daily.
Correspondence
Naranie Shanmuganathan, Centre for Cancer Biology and Precision Cancer Medicine Theme, South Aus­tra­lian Health and Medical Research Institute, Royal Adelaide Hospital, Port Rd, 5000
Adelaide, Australia; e-mail: naranie​­.shanmuganathan@sa​­.gov​­.au.
Discussion of options
1. Switch to bosutinib: The F317L muta­tion would appear to be
sen­si­tive to bosutinib, but in this patient there was also evi­dence
136 | Hematology 2022 | ASH Education Program
Tim­o­thy P. Hughes received sup­port from the National Health
and Medical Research Council of Australia (APP1135949) and has
the finan­cial sup­port of Cancer Council SA’s Beat Cancer Project
on behalf of its donors and the state gov­ern­ment of South Australia through the Department of Health.
References
1.
NCCN Clinical Practice Guidelines in Oncology: Chronic Myelogenous Leukemia. Version 3.2022. Fort Washington, PA; 2022.
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/129/2021810/129hughes.pdf by guest on 09 December 2022
to all 3 2G-TKIs, imatinib may not be well tol­er­ated, and her
adher­ence may be poor once again.
2. Switch to ponatinib 45mg/d (OPTIC dos­ing) with dose
reduc­tion once MR2 is achieved: While she has lost MR2, it is
still very pos­
si­
ble that she still has TKI-sen­
si­
tive dis­
ease, so
she would likely have an excel­lent response to full-dose ponatinib. However, the risk of AOEs is still sig­nif­i­cant at this dose,
espe­cially given her his­tory, and other options may be equally or
more effi­ca­cious with­out the same level of risk.
3. Switch to ponatinib 30 mg/d (OPTIC dos­ing) with dose
reduc­tion once MR2 is achieved: This dose pro­vi­des a lower
risk of AOEs (5.3 events per 100 patient-years com­pared to 9.6
for the 45-mg ponatinib cohort in OPTIC). However, the rate of
molec­u­lar response is lower than with the higher dose of ponatinib. In patients with no KD muta­tions, 46% and 38% achieved
MR2 by 12 months in the ponatinib 45-mg and 30-mg dose
cohort, respec­tively. This may not be a major dif­fer­ence as pro­
gres­sion-free and over­all sur­vival were not sub­stan­tially infe­rior
with the mod­i­fied dose. This would be an accept­able option
except for the fact that the patient has a his­tory of periph­eral
vas­cu­lar dis­ease.
4. Switch to asciminib at 40 mg twice daily: This would be
a rel­a­tively safe and effec­tive option in this set­ting. Patients
in MCyR in ASCEMBL who received asciminib had a 40% rate
of MMR by 6 months. The risk of AOEs (per 100 patient-years)
in this cohort was 3.4.17 Given her con­sis­tent tol­er­a­bil­ity issues
with all 3 2G-TKIs, asciminib would be the option most likely to
be well tol­er­ated.
Best option: Asciminib at 40 mg twice daily and con­sider a
switch to ponatinib if MR2 is not achieved after 6 to 12 months.
Imatinib may be another option, but given the loss of responses
on pre­vi­ous ther­apy, asciminib would be pref­er­en­tially recommended.
14. Hughes TP, Mauro MJ, Cortes JE, et al. Asciminib in chronic mye­loid leu­ke­
mia after ABL kinase inhib­i­tor fail­ure. N Engl J Med. 2019;381(24):2315-2326.
15. Réa D, Mauro MJ, Boquimpani C, et al. A phase 3, open-label, ran­dom­ized
study of asciminib, a STAMP inhib­i­tor, vs bosutinib in CML after 2 or more
prior TKIs. Blood. 2021;138(21):2031-2041.
16. Dorer DJ, Knickerbocker RK, Baccarani M, et al. Impact of dose inten­sity of
ponatinib on selected adverse events: mul­ti­var­i­ate ana­ly­ses from a pooled
pop­u­la­tion of clin­i­cal trial patients. Leuk Res. 2016;48(Sep­tem­ber):84-91.
17. Mauro MJ, Minami Y, Rea D, et al. Efficacy and safety results from ASCEMBL,
a mul­ti­cen­ter, open-label, phase 3 study of asciminib, a first-in-class STAMP
inhib­i­tor, vs bosutinib in patients with chronic mye­loid leu­ke­mia in chronic
phase after ≥2 prior tyro­
sine kinase inhib­
i­
tors: update after 48 weeks
[ab­stract]. Blood. 2021;138(suppl 1). Abstract 310.
18. Cortes JE, Hughes TP, Mauro MJ, et al. Asciminib, a first-in-class STAMP
inhib­i­tor, pro­vi­des dura­ble molec­u­lar response in patients (pts) with
chronic mye­loid leu­ke­mia (CML) har­bor­ing the T315I muta­tion: pri­mary effi­
cacy and safety results from a phase 1 trial. Blood. 2020;136(suppl 1):47-50.
19. Soverini S, Branford S, Nicolini FE, et al. Implications of BCR::ABL1 kinase
domain-medi­ated resis­tance in chronic mye­loid leu­ke­mia. Leuk Res.
2014;38(1):10-20.
20. Wylie AA, Schoepfer J, Jahnke W, et al. The allo­ste­ric inhib­i­tor ABL001
enables dual targeting of BCR::ABL1. Nature. 2017;543(7647):733-737.
21. Eide CA, Zabriskie MS, Savage Stevens SL, et al. Combining the allo­
ste­ric inhib­i­tor asciminib with ponatinib suppresses emer­gence of and
restores effi­cacy against highly resis­tant BCR::ABL1 mutants. Cancer Cell.
2019;36(4):431-443.e5443e5.
22. Cortes JE, Saikia T, Kim DW, et al. Phase 1 trial of vodobatinib, a novel oral
BCR::ABL1 tyro­sine kinase inhib­i­tor (TKI): activ­ity in CML chronic phase
patients fail­ing TKI ther­a­pies includ­ing ponatinib. Blood. 2020;136(suppl
1):51-52.
23. Qian J, Shi D, Li Z, et al. Updated safety and effi­cacy results of phase 1
study of olverembatinib (HQP1351), a novel third-gen­er­a­tion BCR-ABL tyro­
sine kinase inhib­i­tor (TKI), in patients with TKI-resis­tant chronic mye­loid
leu­ke­mia (CML). Blood. 2021;138(suppl 1):311.
24. Baccarani M, Deininger MW, Rosti G, et al. Euro­pean LeukemiaNet rec­om­
men­da­tions for the man­age­ment of chronic mye­loid leu­ke­mia: 2013. Blood.
2013;122(6):872-884.
© 2022 by The Amer­i­can Society of Hematology
DOI 10.1182/hema­tol­ogy.2022000328
Dr Prakash Singh Shekhawat
TKI-resis­tant CP-CML | 137
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/129/2021810/129hughes.pdf by guest on 09 December 2022
2. Hochhaus A, Baccarani M, Silver RT, et al. Euro­pean LeukemiaNet 2020
rec­om­men­da­tions for treating chronic mye­loid leu­ke­mia. Leukemia. 2020;
34(4):966-984.
3. Mahon FX, Réa D, Guilhot J, et al; Intergroupe Français des Leucémies
Myéloïdes Chroniques. Discontinuation of imatinib in patients with chronic
mye­loid leu­kae­mia who have maintained com­plete molec­u­lar remis­sion for
at least 2 years: the pro­spec­tive, multicentre Stop Imatinib (STIM) trial.
Lancet Oncol. 2010;11(11):1029-1035.
4. Ross DM, Branford S, Seymour JF, et al. Safety and effi­cacy of imatinib ces­
sa­tion for CML patients with sta­ble unde­tect­able min­i­mal resid­ual dis­ease:
results from the TWISTER study. Blood. 2013;122(4):515-522.
5. Saussele S, Richter J, Guilhot J, et al; EURO-SKI Investigators. Discontinuation of tyro­sine kinase inhib­i­tor ther­apy in chronic mye­loid leu­kae­mia
(EURO-SKI): a prespecified interim anal­y­sis of a pro­spec­tive, multicentre,
non-randomised, trial. Lancet Oncol. 2018;19(6):747-757.
6. Hughes TP, Ross DM. Moving treat­ment-free remis­sion into main­stream
clin­i­cal prac­tice in CML. Blood. 2016;128(1):17-23.
7. Garg RJ, Kantarjian H, O’Brien S, et al. The use of nilotinib or dasatinib after
fail­ure to 2 prior tyro­sine kinase inhib­i­tors: long-term fol­low-up. Blood.
2009;114(20):4361-4368.
8. Lipton JH, Bryden P, Sidhu MK, et al. Comparative effi­cacy of tyro­sine
kinase inhib­
i­
tor treat­
ments in the third-line set­
ting, for chronic-phase
chronic mye­log­e­nous leu­ke­mia after fail­ure of sec­ond-gen­er­a­tion tyro­sine
kinase inhib­i­tors. Leuk Res. 2015;39(1):58-64.
9. Ibrahim AR, Paliompeis C, Bua M, et al. Efficacy of tyro­sine kinase inhib­
i­tors (TKIs) as third-line ther­apy in patients with chronic mye­loid leu­ke­
mia in chronic phase who have failed 2 prior lines of TKI ther­apy. Blood.
2010;116(25):5497-5500.
10. Hochhaus A, Gambacorti-Passerini C, Abboud C, et al; BYOND Study Investigators. Bosutinib for pretreated patients with chronic phase chronic
mye­loid leu­ke­mia: pri­mary results of the phase 4 BYOND study. Leukemia.
2020;34(8):2125-2137.
11. Cortes JE, Kim DW, Pinilla-Ibarz J, et al. Ponatinib effi­cacy and safety in Philadelphia chro­mo­some-pos­i­tive leu­ke­mia: final 5-year results of the phase 2
PACE trial. Blood. 2018;132(4):393-404.
12. Lipton JH, Chuah C, Guerci-Bresler A, et al; EPIC Investigators. Ponatinib
ver­sus imatinib for newly diag­nosed chronic mye­loid leu­kae­mia: an inter­na­
tional, randomised, open-label, phase 3 trial. Lancet Oncol. 2016;17(5):612621.
13. Cortes J, Apperley J, Lomaia E, et al. Ponatinib dose-rang­ing study in chronic-phase chronic mye­loid leu­ke­mia: a ran­dom­ized, open-label phase 2 clin­
i­cal trial. Blood. 2021;138(21):2042-2050.
CONTROVERSIES IN AGGRESSIVE NHL
CNS prophylaxis in aggressive B-cell lymphoma
Department of Haematology, Beatson West of Scotland Cancer Centre, Glasgow, United Kingdom; 2Department of Haematology, Vall d’Hebron
Institute of Oncology, University Hospital Vall d’Hebron, Barcelona, Spain; and 3Department of Haematology, University College London Hospital,
London, United Kingdom
1
The prevention of central nervous system (CNS) relapse in diffuse large B-cell lymphoma (DLBCL) continues to be one of
the most contentious areas of lymphoma management. Outcomes for patients with secondary CNS lymphoma (SCNSL)
have historically been very poor. However, in recent years improved responses have been reported with intensive immunochemotherapy approaches, and there is a growing interest in potential novel/cellular therapies. Traditional methods
for selecting patients for CNS prophylaxis, including the CNS International Prognostic Index, are hampered by a lack of
specificity, and there is accumulating evidence to question the efficacy of widely employed prophylactic interventions,
including intrathecal and high-dose methotrexate (HD-MTX). Given the potential toxicity of HD-MTX in particular and
the ongoing need to prioritize systemic disease control in high-risk patients, there is an urgent need to develop more
robust methods for identifying patients at highest risk of CNS relapse, as well as investigating prophylactic interventions
with greater efficacy. Here we review new evidence in this field from the last 5 years, focusing on the potential use of
molecular diagnostics to improve the identification of high-risk patients, recent large data sets questioning the efficacy
of HD-MTX, and the current approach to management of patients with SCNSL. We provide a suggested algorithm for
approaching this very challenging clinical scenario.
LEARNING OBJECTIVES
• Understand the currently available methods for identifying patients at high risk of CNS relapse and the potential
for novel molecular diagnostics to improve patient selection in the future
• Review recent evidence to question the efficacy of traditional methods for delivering CNS prophylaxis and to
evaluate the increasing focus on alternative interventions for this important clinical problem
CLINICAL CASE
A 62-year-old man with no previous medical history presented in January 2020 with a short history of weight loss,
night sweats, hip pain, and bilateral groin lymphadenopathy. His lactate dehydrogenase (LDH) was elevated (>3
times the upper limit of normal). Fluorodeoxyglucosepositron emission tomographic (FDG-PET) imaging
revealed widespread hypermetabolic lymphadenopathy
(largest lesion, 4 cm diameter) as well as pathological
FDG uptake in multiple areas of bone (scapula, L3 vertebra, left hemipelvis) and in the left kidney. A core biopsy
from a left inguinal lymph node demonstrated a diagnosis
of diffuse large B-cell lymphoma (DLBCL), non-germinal
center subtype (Hans algorithm), with overexpression of
MYC (>90%) and BCL2 (>60%) by immunohistochemistry—
that is, the double-expressor subtype. Fluorescence in
situ hybridization (FISH) studies showed no evidence
of MYC, BCL2, or BCL6 rearrangements. His Eastern
138 | Hematology 2022 | ASH Education Program
Cooperative Oncology Group (ECOG) performance status (PS) was 1, resulting in an international prognostic
index (IPI) of 4 (age >60, stage IVB, raised LDH, ≥2 extranodal (EN) sites) and a central nervous system (CNS) IPI
of 5 (aforementioned IPI factors plus renal involvement).
Six cycles of rituximab, cyclophosphamide, doxorubicin,
vincristine, and prednisolone (R-CHOP) therapy were
planned at 21-day intervals. Consideration was given as
to whether CNS prophylaxis should be incorporated to
reduce risk of CNS relapse.
Introduction
CNS relapse (otherwise referred to as secondary CNS lymphoma [SCNSL]) is a relatively rare but often devastating
complication for patients with DLBCL. Estimates of CNS
relapse incidence vary, occurring overall in approximately
5% of DLBCL patients but with subgroups in which the risk
is significantly higher.1,2 Most CNS relapse events occur
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/138/2021790/138wilson.pdf by guest on 09 December 2022
Matthew R. Wilson,1 Sabela Bobillo,2 and Kate Cwynarski3
Table 1. Summary of con­sen­sus guide­line rec­om­men­da­tions for CNS pro­phy­laxis in DLBCL
Patient selec­tion
Method for CNS pro­phy­laxis suggested
Brit­ish Society for Haematology (2021)8
Offer to:
• High (4-6) CNS-IPI
• ≥3 EN sites
• High-risk EN site involve­ment—tes­tic­u­lar,
renal/adre­nal, intra­vas­cu­lar
Consider in:
• Breast involve­ment
• Uterine involve­ment
• HD-MTX (≥3g/m2 for 2-3 cycles) as early as pos­si­ble
as part of first-line ther­apy with­out com­pro­mis­ing
dose and time inten­sity of R-CHOP-like treat­ment
• IT pro­phy­laxis not recommended if HD-MTX
suc­cess­fully deliv­ered
• Consider IT as well as sys­temic pro­phy­laxis in
tes­tic­u­lar DLBCL
NCCN (2022)48
Consider in:
• High (4-6) CNS-IPI
• Double/tri­ple-hit HGBL
• High-risk EN site involve­ment—tes­tic­u­lar,
breast, pri­mary cuta­ne­ous, renal/adre­nal
• HD-MTX (3-3.5g/m2 for 2-4 cycles) dur­ing or after the
course of treat­ment and/or
• IT meth­o­trex­ate and/or cytarabine (4-8 doses)
dur­ing or after the course of treat­ment
ESMO (2018)49
Consider in:
• High IPI
• High-risk EN site involve­ment—tes­tic­u­lar,
renal/adre­nal, breast, bone mar­row, bone
• HD-MTX is “an option . . . ​even though the level of
supporting evi­dence is low”
• “Little or no role” for IT ther­apy
ESMO, Euro­pean Society for Medical Oncology; HGBL, high-grade B-cell lym­phoma; NCCN, National Comprehensive Cancer Network.
either dur­ing or closely fol­low­ing front­line immunochemotherapy, with a median time in recent pro­spec­tive clin­i­cal tri­als of
6 to 8 months.1,3 Management of SCNSL is often chal­leng­ing,
with his­tor­i­cally poor out­comes. As a result, much atten­tion has
focused on both the iden­ti­fi­ca­tion of patients at highest risk for
this com­pli­ca­tion, as well as pro­phy­lac­tic treat­ments aimed at
abro­gat­ing risk as much as pos­si­ble. Although our under­stand­
ing of which patients are at highest risk of SCNSL has improved,
par­tic­u­larly with the intro­duc­tion of the CNS-IPI and increased
under­
stand­
ing of the molec­
u­
lar biol­
ogy of DLBCL,4 deci­sionmak­ing around pro­phy­lac­tic inter­ven­tions con­tin­ues to be based
either on ret­ro­spec­tive ana­ly­ses or data extrapolated from other
disease subtypes, with no prospective randomized trials performed aimed at addressing CNS prophylaxis efficacy directly.
Clinicians often are faced with the dilemma of trying to prevent such a feared complication whilst ensuring that the patient
is not exposed to additional therapy with associated risk of toxicity and a limited evidence-base to demonstrate its efficacy.
The lim­i­ta­tions of the evi­dence to inform deci­sion-mak­ing are
reflected in the var­i­a­tion between national guide­lines on the
topic (Table 1), as well as the sig­nif­i­cant dis­par­ity in prac­tice
between cen­ters within the same health care sys­tem.
A 2017 Amer­i­can Society of Hematology Educational Program
review gave a com­pre­hen­sive over­view of the risk fac­tors for CNS
relapse and evi­dence to guide pro­phy­lac­tic inter­ven­tions at that
time.5 In this arti­cle we focus on updates in the field in the last
5 years, with par­tic­u­lar atten­tion to the advances in molec­u­lar
diag­nos­tics and impli­ca­tions for SCNSL, as well as new evi­dence
to ques­tion the effi­cacy of high-dose meth­o­trex­ate (HD-MTX).
How do we iden­tify patients at high risk of CNS relapse?
Clinical risk fac­tors
Numerous stud­
ies have inves­
ti­
gated poten­
tial risk fac­
tors for
CNS relapse in DLBCL.5 In 2016 the Ger­man High-Grade NonHodgkin Lymphoma Study Group (DSHNHL) devel­oped a prog­
nos­tic model (CNS-IPI) incor­po­rat­ing the 5 stan­dard IPI fac­tors as
well as involve­ment of the kid­neys or adre­nal glands, strat­i­fy­ing
patients into 3 risk categories (Table 2).4 Notably, patients with 5
or 6 risk fac­tors had a much higher risk of CNS relapse of 15% and
32.5%, respec­tively. Although the CNS-IPI is a robust model and
has been val­i­dated in sub­se­quent stud­ies, it lacks spec­i­fic­ity,
and half of events occur among patients with low to inter­me­di­
ate scores. It should also be noted that although a small num­ber
of patients in the DSHNHL tri­als used to for­mu­late the CNS-IPI
had Bur­kitt lym­phoma, the final model is val­i­dated for patients
with DLBCL only, and CNS pro­phy­laxis strat­e­gies for Bur­kitt lym­
phoma should be con­sid­ered sep­a­rately.
Certain EN sites have been asso­ci­ated with a higher risk of
CNS recur­rence, with kid­ney/adre­nal involve­ment included in
the CNS-IPI model and intra­vas­cu­lar lym­phoma a dis­tinct entity
with a well-established risk of CNS involve­ment at base­line or
at relapse. Testicular involve­ment has long been rec­og­nized as
a risk fac­tor, in the con­text of both lim­ited and advanced stage,
with a 10-year CNS relapse risk of 10% to 25% (see section Testicular DLBCL).6 Breast involve­ment has been asso­ci­ated with a
higher risk of CNS relapse (~15%) in ret­ro­spec­tive series,7 whereas
other EN sites such as the uterus, blood, bone mar­row, or epi­
du­ral area showed more incon­sis­tent results and are unlikely to
be inde­pen­dently pre­dic­tive of CNS relapse.8 Finally, a large ret­
ro­spec­tive study reported that the involve­ment of 3 or more EN
sites as deter­mined by PET-com­puted tomog­ra­phy con­ferred a
3-year cumu­la­tive risk of CNS relapse of 15%.9
Biological risk fac­tors
The dual overexpression of MYC and BCL2, deter­
mined by
immu­no­his­to­chem­is­try (dou­ble-expressor DLBCL), has not been
con­sis­tently asso­ci­ated with a high risk of CNS relapse.3,10 However, most dou­ble-expressor cases are clas­si­fied as the acti­vated
B-cell (ABC) sub­type, which, when deter­mined by gene expres­
sion pro­fil­ing, has been asso­ci­ated with a CNS relapse risk of 7%
to 9% and 15% when com­bined with a high CNS-IPI.3,10
Recently, multiplatform anal­y­sis defined new molec­u­lar sub­
groups, or clus­ters.11,12 The MCD and C5 clus­ters, char­ac­ter­ized
by a high fre­quency of MYD88L265P and CD79 muta­tions, occur
almost exclu­sively in the ABC sub­type. Genetic alter­ations defin­
ing these sub­
types are also recur­
rently mutated in pri­
mary
Dr Prakash Singh Shekhawat
CNS pro­phy­laxis in aggres­sive B-cell lym­phoma | 139
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/138/2021790/138wilson.pdf by guest on 09 December 2022
Guideline
Table 2. CNS-IPI risk categories with cor­re­spond­ing 2 year rates of CNS relapse and pro­por­tion of patients in each cat­e­gory from
train­ing (DSHNHL) and val­i­da­tion (BCCA) cohorts
DSHNHL cohort
Risk group
Risk fac­tors
Low
Intermediate
BCCA cohort
N (%)
2-year risk
of CNS relapse
N (%)
2-year risk of CNS relapse
0-1
1002 (46)
0.6%
463 (31)
0.8%
2-3
896 (41)
3.4%
694 (46)
3.9%
High*
4
188 (9)
7.4%
344 (23)
12%
High
5
62 (3)
15%
High
6
13 (1)
32.5%
EN lym­
pho­
mas orig­
i­
nat­
ing in the CNS, tes­
tes, breasts, skin,
and intra­vas­cu­lar spaces. Interestingly, a recent series of SCNSL
(n=13) con­firmed a higher prev­a­lence of the MCD sub­type than
a ref­er­ence cohort of relapsed DLBCL with no CNS involve­ment
(38% vs 8%; P=.003).13 Furthermore, the hcMCD sub­type defined
by MYD88L265P muta­tion or more than 3 muta­tions in CD79, PIM1,
ETV6, BTG1, PRDM1, or PBL1XR1 con­sti­tuted almost half of the
patients with CNS recur­rence (46%). The remaining cases were
either dou­
ble-hit lym­
phoma (DHL) or asso­
ci­
ated with TP53
muta­tions. Although these data need to be val­i­dated, there is
clear poten­tial for next gen­er­a­tion sequenc­ing anal­y­sis to help
iden­tify patients at risk of CNS relapse.
High-grade B-cell lym­pho­mas har­bor­ing MYC trans­lo­ca­tion
along with BCL2 and/or BCL6 trans­lo­ca­tion (DHL or tri­ple-hit
lym­pho­mas] have his­tor­i­cally been asso­ci­ated with a high risk
of CNS involve­ment). However, there is accu­mu­lat­ing evi­dence
to sug­gest that early data overestimated this risk, as FISH was
not performed con­sis­tently,10 and such patients often meet other
clin­i­cal cri­te­ria.
Baseline screen­ing
Baseline screen­ing with brain imag­ing and lum­bar punc­ture/
cere­bro­spi­nal fluid (CSF) anal­y­sis is increas­ingly used to iden­
tify high-risk patients with CNS involve­ment who may ben­e­fit
from CNS-directed ther­a­pies. Several stud­ies have shown that
CSF anal­y­sis with flow cytom­e­try is more sen­si­tive than cytol­
ogy for the detec­tion of occult CNS involve­ment.14 However, a
pro­por­tion of patients with a neg­a­tive flow cytom­e­try result
relapse in the CNS shortly after treat­ment, suggesting the need
for more sen­si­tive tech­niques. Cell-free cir­cu­lat­ing tumor DNA
(ctDNA) has recently appeared as a prog­
nos­
tic bio­
marker in
patients with CNS lym­phoma, with good cor­re­la­tion between
ctDNA lev­els (MYD88L265P muta­tion) and treat­ment response and
out­comes.15-17 Two stud­ies have assessed the role of CSF ctDNA
anal­y­sis in patients with high-risk B-cell lym­phoma.17,18 The first
ana­
lyzed spe­
cific tumor-derived muta­
tions in sequen­
tial CSF
sam­
ples from 12 patients receiv­
ing front­
line treat­
ment, and
CSF ctDNA was detected 3 months before CNS relapse in 1 of 2
patients in whom this occurred.17 More recently, Olszewski et al
ana­lyzed CSF from 22 patients with aggres­sive B-cell lym­phoma
140 | Hematology 2022 | ASH Education Program
using a next gen­er­a­tion sequenc­ing-min­i­mal resid­ual assay.18 At
diag­no­sis, CSF ctDNA was iden­ti­fied in 8 patients, 2 of whom
relapsed in the CNS, with a 12-month cumu­la­tive risk of CNS
recur­rence of 29% in patients with a pos­i­tive anal­y­sis vs a 0% risk
for patients with neg­a­tive CSF.18 Taken together, acknowl­edg­ing
the lim­i­ta­tion of the small num­ber of patients, these results sug­
gest the poten­tial util­ity of CSF ctDNA to iden­tify patients with
a higher risk of CNS relapse. Further stud­ies are ongo­ing to val­i­
date these find­ings before the tech­nol­ogy can be incor­po­rated
into rou­tine clin­i­cal prac­tice.
How do we man­age patients with CNS relapse/SCNSL?
Historically, SCNSL has been asso­ci­ated with a dis­mal prog­no­
sis and median over­all sur­vival (OS) of approx­i­ma­tely 6 months.19
Identifying effec­tive ther­a­peu­tic approaches remains chal­leng­
ing, and the major­ity of patients still prog­ress or relapse shortly
after treat­ment.
In recent years, com­bi­na­tions of inten­sive chemotherapies
includ­ing HD-MTX followed by autol­o­gous stem cell trans­plan­
ta­tion (ASCT) have been adopted, with 2-year OS of 25% to 68%
reported.20,21 Recently, the MARIETTA phase 2 study exam­ined the
effi­cacy of 3 courses of MATRix (rituximab, MTX, cytarabine, thio­
tepa) plus 3 courses of RICE (rituximab, ifosfamide, etoposide,
carboplatin) followed by carmustine and thio­tepa-con­di­tioned
ASCT in 75 patients with CNS involve­
ment at diag­
no­
sis or
relapse. Two-year OS for the inten­tion-to-treat pop­u­la­tion was
46%,22 while those under­go­ing ASCT (37/75 patients) had 2-year
OS of 83%. Two-year pro­gres­sion-free sur­vival (PFS) of 71% was
reported in patients with SCNSL at ini­tial diag­no­sis, but PFS was
only 28% in patients pre­vi­ously treated with R-CHOP. Although
this study was restricted to patients under the age of 70, an
ECOG PS of 3 or lower, and ade­quate organ func­tion, thio­tepabased ASCT is increas­ingly uti­lized in older patients.23 However,
induc­tion che­mo­ther­apy reg­i­mens are intense and less well tol­
er­ated in older patients in the real-world set­ting.24
Chimeric anti­gen recep­tor (CAR) T-cell ther­apy has shown
prom­
is­
ing results in relapsed/refrac­
tory DLBCL, includ­
ing in
older and unfit patients. There is accu­mu­lat­ing data dem­on­strat­
ing the effi­cacy and safety of CAR T cells in CNS lym­phoma.25
In SCNSL, the TRANSCEND study dem­
on­
strated com­
plete
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/138/2021790/138wilson.pdf by guest on 09 December 2022
One point is scored for any of the fol­low­ing: age >60 years, LDH > nor­mal, ECOG per­for­mances sta­tus >1, stage III/IV dis­ease, extranodal involve­ment
≥2 sites, kid­ney and/or adre­nal involve­ment.
BCCA, Brit­ish Colombia Cancer Agency.
*High risk group (4-6 fac­tors) over­all 2-year risk of CNS relapse of 10.2% in DSHNHL cohort
remis­sion (CR) in 3 of 6 patients, with grade 3 neu­ro­log­i­cal events
in 2 cases (27).26 Similarly, small (≤8 patients) series of patients
with highly refrac­tory SCNSL treated with com­mer­cial agents
have been reported as hav­ing CR rates of approx­i­ma­tely 50%
with no sig­nif­i­cant tox­ic­ity.27,28 These find­ings sug­gest that CAR T
cells may be a via­ble sal­vage treat­ment for this chal­leng­ing pop­
u­la­tion, espe­cially for patients not fit enough to receive inten­
sive immunochemotherapy. Furthermore, a num­ber of phase 1/2
stud­ies eval­u­at­ing CAR T cells in CNS lym­phoma are cur­rently
ongo­ing (NCT04608487, NCT04464200, NCT03484702).
deliv­ery of IT ther­apy can be chal­leng­ing and uncom­fort­able
for the patient, with some evi­dence to sug­gest an asso­ci­a­tion
with infec­tion-related hos­pi­tal­i­za­tion in older patients.30 With
an increas­ing rec­og­ni­tion that the major­ity of CNS relapses in
DLBCL involve the brain paren­chyma, an area not pen­e­trated
by IT ther­apy alone, IT use has dimin­ished in this set­ting, with
a move toward sys­temic anti­me­tab­o­lite ther­apy instead. An
excep­tion to this is in tes­tic­u­lar DLBCL, where IT ther­apy may
con­tinue to have a role based on data from pro­spec­tive IELSG
tri­als (see below).
Methods for deliv­ery of CNS pro­phy­laxis
Intrathecal che­mo­ther­apy
HD-MTX
Approximately 70% to 80% of CNS relapses in DLBCL involve the
brain paren­chyma,31 and there­fore there is a ratio­nale for pro­phy­
lac­tic ther­a­pies that cross the blood-brain bar­rier and pen­e­trate
all­CNS com­
part­
ments. Intravenous HD-MTX has increas­
ingly
been used over the last 10 years as CNS pro­phy­laxis in DLBCL,
with ini­tial supporting evi­dence mainly derived from its effi­cacy
in pri­mary CNS lym­phoma. Over the years sev­eral stud­ies that
have the com­mon theme of being nonrandomized, ret­ro­spec­
tive ana­ly­ses have addressed this area, but they have been of
var­i­able size and have pro­duced dis­crep­ant results (Table 3).32-40
While there has been wide­spread incor­po­ra­tion of HD-MTX as
Table 3. Summary of recent stud­ies eval­u­at­ing use of HD-MTX in DLBCL
Study (year)
n
Design
Risk fac­tors
Lewis et al32
(2022)
2300
Multicenter,
ret­ro­spec­tive
CNS-IPI ≥4
Testicular, breast
involve­ment
DHL
Wilson et al33
(2022)
1384
Multicenter,
ret­ro­spec­tive
Orellana-Noia et al34
(2022)
1030
Puckrin et al35
(2021)
Systemic
treat­ment
CNS Prophylaxis
CNS relapse
Comments
R-CHOP (94%)
R-EPOCH (6%)
1. HD-MTX (18%)
2. No HD-MTX
(82%)
1. 9.2% (5y)
2. 8.1% (5y)
No ben­e­fit
HD-MTX
High-risk EN sites
CNS-IPI ≥4
≥2 EN and LDH ↑
R-CHOP
1. HD-MTX (all­,
inter­ca­lated, or
EOT)
1. 5.7% (3y)
2. 5.8% (3y)
No dif­fer­ence
between EOT
and inter­ca­lated
HD-MTX
Multicenter,
ret­ro­spec­tive
Not described
R-CHOP (48%)
R-EPOCH (45%)
Other (7%)
1. HD-MTX (20%)
2. IT (77%)
1. 6.8%
2. 5.4%
No ben­e­fit
HD-MTX vs IT
326
Multicenter,
ret­ro­spec­tive
CNS-IPI ≥4 Testicular
DHL
LDH ↑ + ECOG >1 +
>1 EN
R-CHOP (85%)
Intensive che­mo­
ther­apy (15%)
1. HD-MTX (35%)
2. No HD-MTX
(65%)
1. 12.2%
2. 11.2%
No ben­e­fit
HD-MTX
Bobillo et al36
(2021)
585
Single-cen­ter,
ret­ro­spec­tive
CNS-IPI ≥4
High-risk EN sites
DHL
R-CHOP (68%)
R-EPOCH (15%)
Other (17%)
1. HD-MTX (7%)
2. IT MTX (43%)
3. None (50%)
1. 7.5% (5y)
2. 5.5% (3y)
3. 5%
No ben­e­fit (IT
or HD-MTX)
Ong et al37
(2021)
226
Multicenter,
ret­ro­spec­tive
High-risk EN sites
CNS-IPI ≥4
R-CHOP
1. HD-MTX (29%)
2. No HD-MTX
(71%)
1. 3.1% (3y,
iso­lated)
2. 14.6% (3y,
iso­lated)
HD-MTX sig­nif­i­
cantly reduced
risk of iso­lated
CNS relapse
Wilson et al38
(2020)
334
Multicenter,
ret­ro­spec­tive
CNS-IPI ≥4
High-risk EN sites
≥2 EN sites and
LDH ↑
R-CHOP
1. HD-MTX (all­,
inter­ca­lated, or
EOT)
1. 6.8% (3y)
2. 4.7% (3y)
No dif­fer­ence
between EOT
and inter­ca­lated
HD-MTX
Lee et al39
(2019)
130
Single-cen­ter,
ret­ro­spec­tive
CNS-IPI ≥4
High-risk EN sites
≥2 EN and LDH ↑
R-CHOP
1. HD-MTX (49%)
2. None (51%)
1. 6.9% (2y)
2. 8.1% (2y)
No ben­e­fit
HD-MTX
Goldschmidt et al40
(2019)
480
Multicenter,
ret­ro­spec­tive
High-risk EN sites
Stage IV, LDH ↑,
≥1 EN
CHOP +/−R (80%)
1. HD-MTX (27%)
2. None (73%)
1. 6.9%
2. 6.3%
No ben­e­fit
HD-MTX
Dr Prakash Singh Shekhawat
CNS pro­phy­laxis in aggres­sive B-cell lym­phoma | 141
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/138/2021790/138wilson.pdf by guest on 09 December 2022
For many years, intra­the­cal (IT) cyto­toxic che­mo­ther­apy was
used as CNS pro­phy­laxis in DLBCL, with supporting evi­dence
derived mainly from nonrandomized, ret­ro­spec­tive ana­ly­ses as
well as data extrap­o­lated from other B-cell malig­nan­cies. More
recently, a sys­tem­atic review of stand-alone IT pro­phy­laxis ana­
lyzed a total of 7357 patients treated with anti-CD20 mono­clo­
nal anti­body–based immunochemotherapy, incor­po­rat­ing 3 post
hoc trial ana­ly­ses and 10 ret­ro­spec­tive stud­ies.29 Overall, IT pro­
phy­laxis was not found to be asso­ci­ated with a reduc­tion in
CNS relapse rate on univariable or mul­ti­var­i­able ana­ly­ses. The
142 | Hematology 2022 | ASH Education Program
gether or, at the very least, deliv­er­ing at end of treat­ment (EOT).
An alter­na­tive approach for some very high-risk patients may be
to use inten­si­fied sys­temic reg­i­mens that incor­po­rate HD-MTX—
for exam­ple, R-CODOX-M/IVAC, which has prom­is­ing data in a
phase 2 trial but has not been dem­on­strated to be supe­rior to
R-CHOP in a ran­dom­ized trial and is asso­ci­ated with sig­nif­i­cant
tox­ic­ity.42
Testicular DLBCL
Testicular lym­phoma has been asso­ci­ated with a high risk of longterm CNS relapse, with a 5-year risk of 10% and 25% for lim­ited
(pri­mary testicular lymphoma [PTL]) and advanced dis­ease in the
rituximab era, respec­tively.6 Biologically, over 75% of tes­tic­u­lar
lym­pho­mas resem­ble the ABC sub­type and are enriched for the
somatic muta­tions com­monly seen in CNS lym­phoma, such as
MYD88L265P and CD79, which are pres­ent in up to 70% of cases.
Two pro­spec­tive stud­ies explored the role of CNS pro­phy­laxis
in the rituximab era. The IELSG-10 phase 2 study dem­on­strated
that patients with PTL treated with R-CHOP and con­tra­lat­eral
radi­a­tion ther­apy plus 4 doses of IT MTX had a lower risk of CNS
relapse com­pared to his­tor­i­cal series (5-year cumu­la­tive risk of
6% vs 20%).43 More recently, the IELSG-30 trial included a total
of 54 patients with PTL receiv­ing R-CHOP, con­tra­lat­eral radi­a­tion
ther­apy, and 2 courses of HD-MTX (dose, 1.5g/m2), along with
4 doses of IT lipo­so­mal cytarabine. Preliminary results showed
no CNS relapses after a median fol­low-up of 5 years.44 According
to these stud­ies, patients with tes­tic­u­lar lym­phoma may ben­e­fit
from CNS pro­phy­laxis incor­po­rat­ing HD-MTX and/or IT che­mo­
ther­apy.
CLINICAL CASE (Con­tin­ued)
Due to the pres­ence of high-risk fea­tures for SCNSL, the patient
had a base­line MRI head/spine and lum­bar punc­ture with CSF
anal­
y­
sis (flow cytom­
e­
try) with no CNS disease evident. He
went on to receive 6 cycles of R-CHOP-21, with inter­ca­lated
HD-MTX (3 g/m2) planned on day 8 of R-CHOP cycle 2 and cycle
4. No IT pro­phy­laxis was admin­is­tered. Following the first HDMTX treat­ment after R-CHOP cycle 2, the patient expe­ri­enced
a 7-day hos­pi­tal admis­sion with grade 2 renal tox­ic­ity and line
infec­tion. As a result, cycle 3 R-CHOP was delayed by 10 days.
No fur­ther HD-MTX was given, and he received the remaining
cycles of R-CHOP on sched­
ule, with end-of-treat­
ment PETcom­puted tomog­ra­phy dem­on­strat­ing a com­plete met­a­bolic
response. At 12 months’ fol­low-up, he remains well with no evi­
dence of sys­temic or CNS dis­ease relapse.
How do we approach CNS pro­phy­laxis in 2022?
The above case, treated prior to the pub­li­ca­tion of the most
recent large HD-MTX ana­ly­ses,32,33 dem­on­strates the dif­fi­cul­
ties in deci­sion-mak­ing in this area. The patient had a CNS-IPI
score of 5, cor­re­spond­ing to a 2-year risk of CNS relapse of
15% according to the trial data sets used in the for­mu­la­tion of
the score. Had his ECOG per­for­mance sta­tus been 2 instead
of 1, the CNS-IPI score would have been 6, con­fer­ring an esti­
mated risk of 33%, although it should be empha­sized that only
13 patients were in this cat­e­gory in the CNS-IPI trial data set.
However, the patient had an inher­ently greater risk of sys­temic
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/138/2021790/138wilson.pdf by guest on 09 December 2022
pro­phy­laxis for high-risk patients, dis­agree­ment has arisen about
the saf­est and most effec­tive way to incor­po­rate it into front­
line ther­
apy. The prac­
tice of “inter­
ca­
lated” HD-MTX was first
described in a sin­gle-cen­ter ret­ro­spec­tive study of 65 patients
in which HD-MTX was deliv­ered at days 10 to 15 in between
cycles of R-CHOP, resulting in a CNS relapse rate of 3%.41 While
this approach deliv­ers early CNS-directed ther­apy, poten­tially
advan­
ta­
geous given the often early onset of CNS relapse, it
intro­
duces poten­
tial tox­
ic­
ity and delays to sys­
temic R-CHOP
ther­apy, with many choos­ing to wait and deliver HD-MTX after
R-CHOP com­ple­tion instead.
In the last year, a num­ber of stud­ies in this area have been
published to date, argu­­ably pro­vid­ing the most robust data to
inform our prac­tice in the absence of pro­spec­tive clin­i­cal tri­als.
These stud­ies have addressed 2 sep­a­rate ques­tions: (1) Is HDMTX effec­tive?, and (2) How should it be incor­po­rated into front­
line (R-CHOP/R-CHOP-like) ther­apy? Lewis et al car­ried out an
inter­na­tional mul­ti­cen­ter anal­y­sis of 2300 patients deemed at
high risk for SCNSL on the basis of high CNS-IPI, the pres­ence
of dou­ble-hit FISH abnor­mal­i­ties, or the involve­ment of high risk
sites (breast or tes­tic­u­lar).32 Patients received either HD-MTX
(n=410) with or with­out con­cur­rent IT ther­apy, IT ther­apy alone,
or no CNS pro­phy­laxis. There was no sig­nif­i­cant dif­fer­ence in
the 5-year cumu­la­tive inci­dence of CNS relapse between the
HD-MTX and no–HD-MTX arms (9.1% vs 8.4%, respec­tively), with
results unchanged when ana­ly­ses were restricted to patients
achiev­ing CR at the end of sys­temic treat­ment (5.0% vs 6.0%)
and in subanalyses of patients with “ultra”–high-risk char­ac­ter­
is­tics. These find­ings are con­sis­tent with those of another large
ret­ro­spec­tive study by Orellana-Noia et al, in which no reduc­tion
in CNS relapse was seen in patients receiv­ing HD-MTX (n = 236)
com­pared to those receiv­ing IT pro­phy­laxis alone (n = 894).
Wilson et al reported an inter­na­tional mul­ti­cen­ter anal­y­sis of
1384 patients, all­of whom received HD-MTX CNS pro­phy­laxis
deliv­ered either inter­ca­lated (n = 749) or at the end of R-CHOP
ther­apy (n = 635).33 There was no dif­
fer­
ence in CNS relapse
between the 2 deliv­ery approaches (3-year rate of 5.7% vs 5.8%,
respec­tively), with inter­ca­lated deliv­ery caus­ing sig­nif­i­cantly
higher rates of R-CHOP delay. Notably, in ana­
ly­ses restricted
to patients with high CNS-IPI (n = 600), the 3-year rate of CNS
relapse was 9.1%, very sim­i­lar to the rates reported in the orig­i­nal
CNS-IPI study in which min­i­mal CNS pro­phy­laxis was used.4
Both these stud­ies carry inher­ent cave­ats asso­ci­ated with ret­
ro­spec­tive data col­lec­tion. Notably, the Lewis et al study had a
rel­a­tively low num­ber of patients in the HD-MTX arm, with poten­
tial for a sig­nal toward ben­e­fit in very high-risk patients being
missed as a result. The Wilson et al study had wide var­i­a­tion in
the cri­te­ria used for selec­tion for CNS pro­phy­laxis, and both data
sets contained sig­nif­i­cant num­bers of patients receiv­ing con­cur­
rent IT pro­phy­laxis. However, both stud­ies add com­pel­ling data
to the argu­ment that HD-MTX may not sig­nif­i­cantly reduce rates
of CNS relapse for the major­ity of patients deemed to be “high
risk” by tra­di­tional cri­te­ria. If the abso­lute risk reduc­tion of 1%
with HD-MTX from the Lewis et al study is accu­rate, 100 high-risk
patients would need to be treated to avoid 1 CNS relapse. HDMTX car­ries a sig­nif­i­cant risk of tox­ic­ity, includ­ing acute kid­ney
injury, mucositis, and hep­a­to­tox­ic­ity.38 Considering that sys­temic
treat­ment fail­ure is a greater risk than CNS relapse, it appears
likely that the bal­ance of risks for the vast major­ity of patients
favors pri­or­i­tiz­ing sys­temic ther­apy and for­go­ing HD-MTX alto­
treat­ment fail­ure. He expe­ri­enced sig­nif­i­cant tox­ic­ity fol­low­
ing the first inter­ca­lated HD-MTX, resulting in delayed R-CHOP
ther­apy that could poten­tially have had det­ri­men­tal effect on
sys­temic dis­ease con­trol.
If this patient were to pres­ent now for treat­ment, suggested
approaches based on recent data are outlined in Figure 1. The
deci­
sion-mak­
ing involved essen­
tially places greater empha­
sis
on base­line screen­ing for occult CNS involve­ment in high-risk
patients, as well as more judi­cious use of HD-MTX.
Future direc­tions and con­clu­sions
Although we have seen advances in this extremely con­ten­tious
area of DLBCL man­age­ment in the last 5 years, it is clear that we
need to con­tinue to develop more spe­cific meth­ods of iden­ti­
fy­ing patients at highest risk of CNS relapse and to inves­ti­gate
more effec­tive pro­phy­lac­tic inter­ven­tions for those at highest
risk. As outlined above, the use of ctDNA as a base­line screen­ing
tool car­ries much poten­tial for improv­ing patient selec­tion for
pro­phy­laxis. The incor­po­ra­tion of novel agents a
­ ble to cross the
blood-brain bar­rier is likely to be an area of ongo­ing research.
Although tri­als thus far have not dem­on­strated an over­all ben­e­
fit with the addi­tion of ibrutinib or lenalidomide to R-CHOP,45,46
stud­
ies such as REMoDL-A (NCT04546620) inves­
ti­
gat­
ing the
addi­tion of acalabrutinib are ongo­ing, and results with regard
to CNS relapse rates will be of inter­est. Improving sys­temic dis­
ease con­trol may be an effec­tive way to reduce CNS relapses,
par­tic­u­larly those that occur con­cur­rent with sys­temic relapse.
The recent POLARIX trial dem­
on­
strated an addi­
tional agent
(polatuzumab vedotin) that can improve PFS over R-CHOP alone
for the first time, rais­ing the ques­tion of whether broad adop­tion
of such a front­line reg­i­men could have an impact on CNS events
over time.47 Until then, we must use cur­rently avail­­able risk-strat­
i­fi­ca­tion mod­els to care­fully select patients for more strin­gent
base­line screen­ing for CNS dis­ease and exer­cise greater cau­tion
in the use of pro­phy­lac­tic HD-MTX in light of recently published
data.
Conflict-of-inter­est dis­clo­sure
Mat­
thew R. Wilson: speakers’ bureau: Kite/Gilead, Janssen;
consultancy/advisor: Veriton; conference/travel support: Takeda,
Janssen, Kite/Gilead; research funding: Abbvie.
Sabela Bobillo: speakers’ bureau: Janssen, Roche, Gilead; travel
support: Gilead.
Kate Cwynarski: con­sul­tancy/advi­sor: Roche, Takeda, Celgene,
Atara, Gilead, KITE, Janssen, Incyte; speak­ers’ bureau: Roche,
Takeda, KITE, Gilead, Incyte; research funding: Roche, Takeda,
KITE, Janssen, Bristol Myers Squibb.
Off-label drug use
Mat­thew R. Wilson: nothing to disclose.
Sabela Bobillo: nothing to disclose.
Kate Cwynarski: nothing to disclose.
Dr Prakash Singh Shekhawat
CNS pro­phy­laxis in aggres­sive B-cell lym­phoma | 143
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/138/2021790/138wilson.pdf by guest on 09 December 2022
Figure 1. Proposed algorithm for CNS prophylaxis in DLBCL in 2022. CMR, complete metabolic response; CT, computed tomography; PD, progressive disease; PR, partial response; SD, stable disease.
Correspondence
Kate Cwynarski, University College London Hospital, 3rd Fl,
West, 250 Euston Rd, London NW1 2PG, United Kingdom;
e-mail: kate​­.cwynarski@nhs​­.net.
References
1.
144 | Hematology 2022 | ASH Education Program
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/138/2021790/138wilson.pdf by guest on 09 December 2022
Gleeson M, Counsell N, Cunningham D, et al. Central ner­
vous sys­
tem
relapse of dif­fuse large B-cell lym­phoma in the rituximab era: results of the
UK NCRI R-CHOP-14 ver­sus 21 trial. Ann Oncol. 2017;28(10):2511-2516.
2. Ghose A, Elias HK, Guha G, Yellu M, Kundu R, Latif T. Influence of rituximab
on cen­tral ner­vous sys­tem relapse in dif­fuse large B-cell lym­phoma and
role of pro­phy­laxis—a sys­tem­atic review of pro­spec­tive stud­ies. Clin Lymphoma Myeloma Leuk. 2015;15(8):451-457.
3. Klanova M, Sehn LH, Bence-Bruckler I, et al. Integration of cell of ori­gin into
the clin­i­cal CNS International Prognostic Index improves CNS relapse pre­
dic­tion in DLBCL. Blood. 2019;133(9):919-926.
4. Schmitz N, Zeynalova S, Nickelsen M, et al. CNS International Prognostic
Index: a risk model for CNS relapse in patients with dif­fuse large B-cell lym­
phoma treated with R-CHOP. J Clin Oncol. 2016;34(26):3150-3156.
5. Savage KJ. Secondary CNS relapse in dif­fuse large B-cell lym­phoma: defin­
ing high-risk patients and opti­mi­za­tion of pro­phy­laxis strat­e­gies. Hematology Am Soc Hematol Educ Program. 2017;2017(1):578-586.
6. Kridel R, Telio D, Villa D, et al. Diffuse large B-cell lym­phoma with tes­tic­u­lar
involve­ment: out­come and risk of CNS relapse in the rituximab era. Br J
Haematol. 2017;176(2):210-221.
7. Hu S, Song Y, Li Y, et al. Primary breast dif­fuse large B cell lym­phoma in
the rituximab era: out­comes of a mul­ti­cen­ter ret­ro­spec­tive study by the
Lymphoma and Leukemia Committee of Chi­nese Geriatric Oncology Society (LLC-CGOS). Paper presented at: 58th Annual Meeting of the Amer­i­can
Society of Hematology; 3-6 Decem­ber 2016; San Diego, CA.
8. McKay P, Wilson MR, Chaganti S, et al; Brit­ish Society of Haematology. The
pre­ven­tion of cen­tral ner­vous sys­tem relapse in dif­fuse large B-cell lym­
phoma: a Brit­ish Society for Haematology Good Practice Paper. Br J Haematol. 2020;190(5):708-714.
9. El-Galaly TC, Villa D, Michaelsen TY, et al. The num­ber of extranodal sites
assessed by PET/CT scan is a pow­
er­
ful pre­
dic­
tor of CNS relapse for
patients with dif­fuse large B-cell lym­phoma: an inter­na­tional mul­ti­cen­ter
study of 1532 patients treated with chemoimmunotherapy. Eur J Cancer.
2017;75(April):195-203.
10. Savage KJ, Slack GW, Mottok A, et al. Impact of dual expres­sion of MYC
and BCL2 by immu­no­his­to­chem­is­try on the risk of CNS relapse in DLBCL.
Blood. 2016;127(18):2182-2188.
11. Schmitz R, Wright GW, Huang DW, et al. Genetics and path­o­gen­e­sis of dif­
fuse large B-cell lym­phoma. N Engl J Med. 2018;378(15):1396-1407.
12. Chapuy B, Stewart C, Dunford AJ, et al. Molecular sub­types of dif­fuse large
B cell lym­phoma are asso­ci­ated with dis­tinct path­o­genic mech­a­nisms and
out­comes. Nat Med. 2018;24(5):679-690.
13. Ollila TA, Kurt H, Waroich J, et al. Genomic sub­types may pre­dict the risk of
cen­tral ner­vous sys­tem recur­rence in dif­fuse large B-cell lym­phoma. Blood.
2021;137(8):1120-1124.
14. Wilson WH, Bromberg JE, Stetler-Stevenson M, et al. Detection and out­
come of occult leptomeningeal dis­ease in dif­fuse large B-cell lym­phoma
and Bur­kitt lym­phoma. Haematologica. 2014;99(7):1228-1235.
15. Hiemcke-Jiwa LS, Leguit RJ, Snijders TJ, et al. MYD88 p.(L265P) detec­tion
on cell-free DNA in liq­uid biop­sies of patients with pri­mary cen­tral ner­vous
sys­tem lym­phoma. Br J Haematol. 2019;185(5):974-977.
16. Mutter JA, Alig S, Lauer EM, et al. Profiling of cir­cu­lat­ing tumor DNA for
non­in­va­sive dis­ease detec­tion, risk strat­i­fi­ca­tion, and MRD mon­i­tor­ing in
patients with CNS lym­phoma. Blood. 2021;138(suppl 1):6.
17. Bobillo S, Crespo M, Escudero L, et al. Cell free cir­cu­lat­ing tumor DNA in
cere­bro­spi­nal fluid detects and mon­i­tors cen­tral ner­vous sys­tem involve­
ment of B-cell lym­pho­mas. Haematologica. 2021;106(2):513-521.
18. Olszewski AJ, Chorzalska AD, Petersen M, et al. Detection of clonotypic
DNA in the cere­bro­spi­nal fluid as a marker of cen­tral ner­vous sys­tem inva­
sion in lym­phoma. Blood Adv. 2021;5(24):5525-5535.
19. El-Galaly TC, Cheah CY, Bendtsen MD, et al. Treatment strat­e­gies, out­
comes and prog­nos­tic fac­tors in 291 patients with sec­ond­ary CNS involve­
ment by dif­fuse large B-cell lym­phoma. Eur J Cancer. 2018;93(April):57-68.
20. Ferreri AJ, Donadoni G, Cabras MG, et al. High doses of anti­me­tab­o­lites
followed by high-dose sequen­tial chemoimmunotherapy and autol­o­gous
stem-cell trans­plan­ta­tion in patients with sys­temic B-cell lym­phoma and
sec­ond­ary CNS involve­ment: final results of a mul­ti­cen­ter phase II trial.
J Clin Oncol. 2015;33(33):3903-3910.
21. Doorduijn JK, van Imhoff GW, van der Holt B, et al. Treatment of sec­
ond­ary cen­tral ner­vous sys­tem lym­phoma with intra­the­cal rituximab,
high-dose meth­o­trex­ate, and R-DHAP followed by autol­o­gous stem cell
trans­plan­ta­tion: results of the HOVON 80 phase 2 study. Hematol Oncol.
2017;35(4):497-503.
22. Ferreri AJM, Doorduijn JK, Re A, et al; International Extranodal Lymphoma
Study Group. MATRix-RICE ther­apy and autol­o­gous haematopoietic stemcell trans­plan­ta­tion in dif­fuse large B-cell lym­phoma with sec­ond­ary CNS
involve­ment (MARIETTA): an inter­na­tional, sin­gle-arm, phase 2 trial. Lancet
Haematol. 2021;8(2):e110-e121.
23. Khwaja J, Schorb E, Goradia H, et al. International multi-cen­
tre ret­
ro­
spec­
tive anal­
y­
sis of out­
comes of thio­
tepa-based autol­
o­
gous stem cell
trans­plan­ta­tion for sec­ond­ary CNS lym­phoma. Euro­pean Hematology
Association Open Access Library [poster abstract]. 12 June 2020. Abstract
EP1376.
24. Schorb E, Fox CP, Kasenda B, et al. Induction ther­apy with the MATRix reg­
i­men in patients with newly diag­nosed pri­mary dif­fuse large B-cell lym­
phoma of the cen­tral ner­vous sys­tem—an inter­na­tional study of fea­si­bil­ity
and effi­cacy in rou­tine clin­i­cal prac­tice. Br J Haematol. 2020;189(5):879887.
25. Frigault MJ, Dietrich J, Gallagher K, et al. Safety and effi­cacy of tisagenlecleucel in pri­
mary CNS lym­
phoma: a phase 1/2 clin­
i­
cal trial. Blood.
2022;139(15):2306-2315.
26. Abramson JS, Palomba ML, Gordon LI, et al. Lisocabtagene maraleucel for patients with relapsed or refrac­
tory large B-cell lym­
pho­
mas
(TRANSCEND NHL 001): a multicentre seam­
less design study. Lancet.
2020;396(10254):839-852.
27. Frigault MJ, Dietrich J, Martinez-Lage M, et al. Tisagenlecleucel CAR T-cell
ther­apy in sec­ond­ary CNS lym­phoma. Blood. 2019;134(11):860-866.
28. Ahmed G, Hamadani M, Shah NN. CAR T-cell ther­apy for sec­ond­ary CNS
DLBCL. Blood Adv. 2021;5(24):5626-5630.
29. Eyre TA, Djebbari F, Kirkwood AA, Collins GP. Efficacy of cen­tral ner­vous
sys­tem pro­phy­laxis with stand-alone intra­the­cal che­mo­ther­apy in dif­fuse
large B-cell lym­
phoma patients treated with anthracycline-based che­
mo­
ther­
apy in the rituximab era: a sys­
tem­
atic review. Haematologica.
2020;105(7):1914-1924.
30. Eyre TA, Kirkwood AA, Wolf J, et al. Stand-alone intra­the­cal cen­tral ner­
vous sys­tem (CNS) pro­phy­laxis pro­vide unclear ben­e­fit in reduc­ing CNS
relapse risk in elderly DLBCL patients treated with R-CHOP and is asso­ci­
ated increased infec­tion-related tox­ic­ity. Br J Haematol. 2019;187(2):185194.
31. Kansara R, Villa D, Gerrie AS, et al. Site of cen­tral ner­vous sys­tem (CNS)
relapse in patients with dif­fuse large B-cell lym­phoma (DLBCL) by the CNSIPI risk model. Br J Haematol. 2017;179(3):508-510.
32. Lewis KL, Jakobsen LH, Villa D, et al. High-dose meth­o­trex­ate is not asso­
ci­ated with reduc­tion in CNS relapse in patients with aggres­sive B-cell
lym­phoma: an inter­na­tional ret­ro­spec­tive study of 2300 high-risk patients.
Blood. 2021;138(suppl 1):181.
33. Wilson MR, Eyre TA, Kirkwood AA, et al. Timing of high-dose meth­o­trex­
ate CNS pro­phy­laxis in DLBCL: a mul­ti­cen­ter inter­na­tional anal­y­sis of 1384
patients. Blood. 2022;139(16):2499-2511.
34. Orellana-Noia VM, Reed DR, McCook AA, et al. Single-route CNS pro­phy­
laxis for aggres­sive non-Hodgkin lym­pho­mas: real-world out­comes from 21
US aca­demic insti­tu­tions. Blood. 2022;139(3):413-423.
35. Puckrin R, El Darsa H, Ghosh S, Peters A, Owen C, Stewart D. Ineffectiveness of high-dose meth­o­trex­ate for pre­ven­tion of CNS relapse in dif­fuse
large B-cell lym­phoma. Am J Hematol. 2021;96(7):764-771.
36. Bobillo S, Joffe E, Sermer D, et al. Prophylaxis with intra­the­cal or high-dose
meth­o­trex­ate in dif­fuse large B-cell lym­phoma and high risk of CNS relapse.
Blood Cancer J. 2021;11(6):113.
37. Ong SY, de Mel S, Grigoropoulos NF, et al. High-dose meth­o­trex­ate is effec­
tive for pre­ven­tion of iso­lated CNS relapse in dif­fuse large B cell lym­phoma.
Blood Cancer J. 2021;11(8):143.
38. Wilson MR, Eyre TA, Martinez-Calle N, et al. Timing of high-dose meth­o­
trex­ate CNS pro­phy­laxis in DLBCL: an anal­y­sis of tox­ic­ity and impact on
R-CHOP deliv­ery. Blood Adv. 2020;4(15):3586-3593.
39. Lee K, Yoon DH, Hong JY, et al. Systemic HD-MTX for CNS pro­phy­laxis in
high-risk DLBCL patients: a pro­spec­tively col­lected, sin­gle-cen­ter cohort
anal­y­sis. Int J Hematol. 2019;110(1):86-94.
40. Goldschmidt N, Horowitz NA, Heffes V, et al. Addition of high-dose meth­o­
trex­ate to stan­dard treat­ment for patients with high-risk dif­fuse large B-cell
lym­phoma con­trib­utes to improved free­dom from pro­gres­sion and sur­vival
but does not pre­vent cen­tral ner­vous sys­tem relapse. Leuk Lymphoma.
2019;60(8):1890-1898.
cin, vin­cris­tine, and pred­ni­sone in non-ger­mi­nal cen­ter B-cell dif­fuse large
B-cell lym­phoma. J Clin Oncol. 2019;37(15):1285-1295.
46. Nowakowski GS, Chiappella A, Gascoyne RD, et al. ROBUST: a phase
III study of lenalidomide plus R-CHOP ver­
sus pla­
cebo plus R-CHOP in
­pre­vi­ously untreated patients with ABC-type dif­fuse large B-cell lym­phoma.
J Clin Oncol. 2021;39(12):1317-1328.
47. Tilly H, Morschhauser F, Sehn LH, et al. Polatuzumab vedotin in pre­vi­ously
untreated dif­fuse large B-cell lym­phoma. N Engl J Med. 2022;386(4):351-363.
48. National Comprehensive Cancer Network. Clinical prac­
tice guide­
lines
in oncol­ogy: B-cell lym­pho­mas. Version 3.2022. https:​­/​­/www​­.nccn​­.org​­
/professionals​­/physician_gls​­/pdf​­/b​­-cell​­.pdf. Accessed 25 April 2022.
49. Tilly H, Gomes da Silva M, Vitolo U, et al; ESMO Guidelines Committee.
Diffuse large B-cell lym­
phoma (DLBCL): ESMO Clinical Practice Guidelines for diag­no­sis, treat­ment and fol­low-up. Ann Oncol. 2015;26 (suppl
5):v116-v125.
© 2022 by The Amer­i­can Society of Hematology
DOI 10.1182/hema­tol­ogy.2022000331
Dr Prakash Singh Shekhawat
CNS pro­phy­laxis in aggres­sive B-cell lym­phoma | 145
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/138/2021790/138wilson.pdf by guest on 09 December 2022
41. Abramson JS, Hellmann M, Barnes JA, et al. Intravenous meth­o­trex­ate as
cen­tral ner­vous sys­tem (CNS) pro­phy­laxis is asso­ci­ated with a low risk of
CNS recur­rence in high-risk patients with dif­fuse large B-cell lym­phoma.
Cancer. 2010;116(18):4283-4290.
42. McMillan AK, Phil­lips EH, Kirkwood AA, et al. Favourable out­comes for highrisk dif­fuse large B-cell lym­phoma (IPI 3-5) treated with front-line R-CODOXM/R-IVAC che­mo­ther­apy: results of a phase 2 UK NCRI trial. Ann Oncol.
2020;31(9):1251-1259.
43. Vitolo U, Chiappella A, Ferreri AJ, et al. First-line treat­ment for pri­mary
tes­tic­u­lar dif­fuse large B-cell lym­phoma with rituximab-CHOP, CNS pro­
phy­laxis, and con­tra­lat­eral tes­tis irra­di­a­tion: final results of an inter­na­tional
phase II trial. J Clin Oncol. 2011;29(20):2766-2772.
44. Conconi A, Chiappella A, Orsucci L, et al. Intensified (intra­ve­nous and
intra­the­cal) CNS pro­phy­laxis in pri­mary tes­tic­u­lar dif­fuse large B-cell
lym­
phoma: 5-year results of the IELSG30 trial. Hematol Oncol. 2021;
39(suppl S2).
45. Younes A, Sehn LH, Johnson P, et al; PHOENIX Investigators. Randomized
phase III trial of ibrutinib and rituximab plus cyclo­phos­pha­mide, doxo­ru­bi­
CONTROVERSIES IN AGGRESSIVE NHL
Sequencing therapy in relapsed DLBCL
Department of Lymphoma/Myeloma, The University of Texas MD Anderson Cancer Center, Houston, TX; and 2Lymphoma and Myeloma,
Dana-Farber Cancer Institute, Boston, MA
1
Diffuse large B-cell lymphoma (DLBCL) is the most common lymphoid malignancy worldwide, comprising approximately
30% of all lymphomas. Currently, 50% to 60% of patients diagnosed with DLBCL are alive at 5 years and cured with modern therapy, but about 10% to 15% of patients are refractory to first-line therapy, and an additional 20% to 30% relapse
following a complete response. Patients who have relapses beyond 2 years may experience more favorable outcomes
and have forms of DLBCL that can be distinguished biologically. Patients who experience early relapse or who have primary refractory disease (less than a complete response or relapse within 3 to 6 months of initial therapy) have worse
outcomes. For decades, the standard of care treatment strategy for fit patients with relapsed DLBCL has been salvage
therapy with non–cross-resistant combination chemoimmunotherapy regimens followed by high-dose chemotherapy
and autologous stem cell transplantation (ASCT) as stem cell rescue for patients with chemosensitive disease. Recent
data suggest that certain patients may benefit from chimeric antigen receptor T-cell therapy (CAR T) in the second-line
setting. Additional novel therapies exist for patients who are ineligible, who are unable to access these therapies, or who
fail ASCT and/or CAR T. Despite the advent of new therapies for DLBCL and improved outcomes, DLBCL remains a lifethreatening illness. Thus, it is essential for clinicians to engage in serious illness conversations with their patients. Goalsof-care communication can be improved through skills-based training and has been demonstrated to have an impact on
patient experiences.
LEARNING OBJECTIVES
• Recognize the therapeutic options available for patients with relapsed/refractory DLBCL
• Understand the triggers and process for goals-of-care discussions for patients with relapsed/refractory DLBCL
Introduction
Diffuse large B-cell lymphoma (DLBCL) is the most common
lymphoid malignancy worldwide, comprising approximately 30% of all lymphomas. Currently, 55% to 60% of patients
diagnosed with DLBCL are alive at 5 years and cured with
modern therapy,1 but approximately 10% to 15% of patients
are refractory to first-line therapy, and an additional 20%
to 30% relapse following a complete response (CR).2,3
This high-risk DLBCL population constitutes approximately
11 000 individuals annually in the United States.3,4 Evidence
suggests that molecular subgroups of DLBCL exhibit significantly different biology, response to standard therapies,
and overall survival (OS) based on the cell of origin.5,6 Standard of care first-line therapy for DLBCL has been based on
French and US randomized controlled trials (RCTs) completed in 2001 demonstrating that the addition of rituximab
(R) to cyclophosphamide, doxorubicin, vincristine, and
prednisone (CHOP) combination chemotherapy improved
outcomes over CHOP.7,8 Since then, numerous RCTs have
failed to improve upon R-CHOP, including trials targeting
146 | Hematology 2022 | ASH Education Program
patients with poor-risk clinical features by cell-of-origin
subtype.9-16 Recently, a double-blind, placebo-controlled,
international RCT compared a modified regimen of R-CHOP
(pola-R-CHP), in which vincristine was replaced with
polatuzumab vedotin as a first-line therapy for patients
with intermediate-risk or high-risk DLBCL according to the
International Prognostic Index (IPI).17 Progression-free survival (PFS) was significantly higher in the pola-R-CHP group
than in the R-CHOP group (stratified hazard ratio [HR],
0.73; 95% CI, 0.57-0.95) with 2-year PFS of 76.7% (95% CI,
72.7-80.8) vs 70.2% (95% CI, 65.8-74.6) for pola-R-CHP and
R-CHOP, respectively. The safety profile was similar in the 2
regimens. Changes in the frontline management of DLBCL
based upon these results would likely have an impact on
therapeutic decisions in the relapse setting (Table 1).
DLBCL prognosis and social determinants of health
The hypothetical stories of Alvin and Wilma are based
on real patient scenarios and illustrate some of the key
challenges faced by patients with newly diagnosed and
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/146/2021680/146flowers.pdf by guest on 09 December 2022
Christopher R. Flowers1 and Oreofe O. Odejide2
Table 1. A tale of 2 patients
Alvin was born in 1955 in a large city that bor­ders the
ocean. In 2017 he was diag­nosed with DLBCL, not oth­er­wise
spec­i­fied. His diag­nos­tic tumor block under­went IHC ­
anal­y­sis, which revealed this to be a non-GCB sub­type
with­out evi­dence of MYC or BCL2 trans­lo­ca­tion. He had
ele­vated LDH greater than 2 times the upper limit of nor­mal
and kid­ney and bone mar­row involve­ment.
Wilma was born in a rural town in the south­ern United
States in the same year as Alvin. She remem­bers watching
how her small town changed over time after the Fair
Housing Act. She was diag­nosed with non-GCB DLBCL in
the same year as Alvin. She also had ele­vated LDH and bone
mar­row but no kid­ney involve­ment.
First-line ther­apy
He had excel­lent per­for­mance sta­tus (ECOG PS 0), was
offered a ran­dom­ized con­trolled trial as ini­tial ther­apy, was
ran­dom­ized to an arm where he received R-CHOP, and
started treat­ment 18 days after diag­no­sis.
She was self-employed and did not have insur­ance at the
time of diag­no­sis. She was unable to con­duct any work
activ­i­ties (ECOG PS 2) and was admit­ted to the hos­pi­tal to
start R-CHOP at 13 days after diag­no­sis.
Treatment at relapse
In early 2019 (11 months fol­low­ing R-CHOP), Alvin noticed
a lymph node in his left axilla. A biopsy showed relapse of
non-GCB DLBCL. He was ­able to pur­chase geno­mic test­ing
using an avail­­able solid-tumor panel, but it did not pro­vide
infor­ma­tion that his treating phy­si­cian thought rel­e­vant to
guide ther­apy. His phy­si­cian performed HLA test­ing for Alvin
and his sib­lings, and Alvin under­went sal­vage ther­apy with
RICE (achiev­ing a par­tial response) followed by ASCT based
on the best avail­­able evi­dence.
Last year fol­low-up imag­ing revealed evi­dence of
relapsed dis­ease with renal involve­ment (serum cre­at­i­nine
2.1), and chemoimmunotherapy was restarted. Alvin was
referred for CAR T ther­apy. He is now in your office awaiting
scan results 6 months fol­low­ing com­ple­tion of treat­ment.
Wilma had dif­fi­culty with fatigue, alo­pe­cia, and periph­eral
neu­rop­a­thy dur­ing her treat­ment but recov­ered to feel­ing
as good as she had sev­eral years prior to her diag­no­sis.
She returned to work for 34 months and then began to feel
pro­gres­sively more tired and unable to per­form her usual
activ­i­ties. She went back to her treating phy­si­cian and had
repeat scans that showed evi­dence of DLBCL relapse. Her
phy­si­cian referred her to a cen­ter more than 250 miles
away to con­sider ASCT. She connected Wilma with a social
worker who aided in gar­ner­ing state-based insur­ance for
can­cer patients. A patient nav­i­ga­tor at the cen­ter
coor­di­nated vis­its and recommended places for Wilma
and her care­giver to stay around the cen­ter for the period
imme­di­ately fol­low­ing ASCT. She now returns to see you
2.5 years after ASCT feel­ing well.
HLA, human leu­ko­cyte anti­gen; RICE, rituximab, ifosfamide, carboplatin, etoposide.
relapsed DLBCL. Both patients were diag­nosed with stage IV
non–ger­mi­nal cen­ter B-cell (GCB) DLBCL with poor prog­nos­tic
fea­tures and an IPI score of 4, plac­ing both in the high-risk cat­e­
gory based on sim­i­lar fea­tures. A recent anal­y­sis com­par­ing the
IPI, revised IPI, and National Comprehensive Cancer Network IPI
involved 2124 DLBCL patients treated with R-CHOP or a var­i­ant
in 1 of 7 mul­ti­cen­ter RCTs. These data dem­on­strated that DLBCL
patients in the high-IPI cat­e­gory had 5-year PFS of 45.8% (95%
CI, 41.1%-51.0%) and 5-year OS of 53.9% (95% CI, 49.3%-59.1%).18
Although immu­no­his­to­chem­is­try (IHC) algo­rithms have been
avail­­able for more than a decade and gene expres­sion–based
approaches uti­liz­ing for­ma­lin-fixed par­af­fin-embed­ded tis­sue have
been dem­on­strated to reli­ably clas­sify DLBCL into bio­log­i­cally and
clin­i­cally dis­tinct sub­groups,19-23 these tests have had lit­tle impact
on clin­
i­cal man­age­ment for patients with newly diag­
nosed or
relapsed DLBCL. Based on gene expres­sion, patients with acti­
vated B-cell (ABC) DLBCL (the major­ity of non-GCB DLBCL) had
5-year PFS of 48% and 5-year OS of 56%, and the cell of ori­gin
was iden­ti­fied in this study to have prog­nos­tic sig­nif­i­cance inde­
pen­dent of IPI.23 Additional stud­ies found that sig­nif­i­cant geno­mic
diver­sity exists within the ABC and GCB sub­groups and defined
geno­mic sub­types with prog­nos­tic value and poten­tial ther­a­peu­
tic impli­ca­tions.24-26 Currently, there are no clin­i­cal assays that can
com­pre­hen­sively and accu­rately eval­u­ate the geno­mic pro­file of
DLBCLs with a short turn­around so that this molec­u­lar infor­ma­tion
can be incor­po­rated into clin­i­cal workflow and prac­tice.
Additionally, stud­
ies indi­
cate that patients diag­
nosed with
DLBCL from rural areas of the United States and that unin­sured
patients (HR, 1.39; 95% CI, 1.14-1.70) and Med­ic­aid-insured patients
(HR, 1.48; 95% CI, 1.23-1.78) with DLBCL had lower sur­vival com­
pared with patients who had pri­vate insur­ance.27,28 Moreover, rural
patients, less com­monly, had pri­vate insur­ance and high socio­
eco­nomic sta­tus and were more likely to receive treat­ment within
14 days of diag­no­sis.27 Accounting for these social deter­mi­nants
of health is crit­i­cal in under­stand­ing the prog­no­sis for indi­vid­u­als
with DLBCL and devel­op­ing suit­able man­age­ment strat­e­gies for
first-line ther­apy and relapse dis­ease. As dem­on­strated by these
patient sit­u­a­tions, these sociodemographic fac­tors influ­ence the
means to access ther­a­pies that can be cura­tive. It is also i­mpor­tant
to note that even when social fac­tors make ther­a­pies more dif­fi­
cult to access, patients can be supported using evi­dence-based
approaches like patient nav­
i­
ga­
tion that have a dem­
on­
strated
impact on out­comes in the first-line and relapse set­ting.29
Other impor­tant stud­ies showed that a shorter diag­
no­sisto-treat­ment inter­val (DTI) was asso­ci­ated with poor prog­nos­
tic fac­tors. DLBCL patients ini­ti­at­ing treat­ment within 14 days
had higher lac­tate dehy­dro­ge­nase (LDH) lev­els, worse Eastern
Cooperative Oncology Group Performance Status (ECOG PS),
more fre­quent B symp­toms, more bulky dis­ease, and a worse
age-­adjusted IPI.30 A shorter DTI was asso­ci­ated with worse out­
comes in both ABC and GCB sub­groups, was prog­nos­tic even
after adjusting for the IPI, and was a pre­
dic­
tor of sur­
vival in
cohorts across the United States, Europe, Canada, and Japan.30-33
Interestingly, a recent study showed that higher lev­els of cir­cu­
lat­ing tumor DNA (ctDNA) were asso­ci­ated with shorter DTI and
sim­i­larly predicted such out­comes, pro­vid­ing a bio­log­i­cal path­
way for under­stand­ing these worse out­comes and a poten­tial
bio­marker for inter­ven­tion.31 However, there is lim­ited appli­ca­bil­
ity of ctDNA in rou­tine clin­i­cal prac­tice cur­rently.
Conventional man­age­ment strat­e­gies for patients
with relapsed DLBCL
For the major­ity of patients who relapse, this occurs within the
first 2 years of ini­
tial ther­
apy. Patients who expe­
ri­
ence early
Dr Prakash Singh Shekhawat
Sequencing ther­apy in relapsed DLBCL | 147
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/146/2021680/146flowers.pdf by guest on 09 December 2022
Initial pre­sen­ta­tion
148 | Hematology 2022 | ASH Education Program
Chimeric anti­gen recep­tor T-cell ther­apy
Current approved chimeric anti­gen recep­tor T-cell (CAR T) ther­a­
pies are genet­i­cally engineered cell prod­ucts targeting the B-cell
marker CD19 that are available for patients with R/R DLBCL after
at least 2 lines of ther­apy. These include axicabtagene ciloleucel
(axi-cel), tisagenlecleucel (tisa-cel), and lisocabtagene maraleucel (liso-cel). These cel­lu­lar ther­a­pies may dif­fer in their mech­
a­nisms of viral trans­fec­tion, costimulatory domains, and T-cell
selec­tion, but each has dem­on­strated the abil­ity to pro­duce
dura­ble remis­sions in about 30% of patients, includ­ing patients
who have failed prior ASCT.45-47 Axi-cel uses the CD28 costimulatory domain that inter­acts with the CD3-zeta acti­va­tion domain
to enhance acti­va­tion and pro­lif­er­a­tion of CAR T cells. Tisa-cel
and liso-cel have a sim­i­lar extra­cel­lu­lar CD19 sin­gle-chain var­i­able
frag­ment attached to a 4-1BB intra­cel­lu­lar costimulatory domain
connected to a CD3ζ sig­nal­ing domain. Liso-cel has a dis­tinct
manufactur­ing pro­cess involv­ing the sep­a­ra­tion of CD4 and CD8
T cells after leukapheresis-inde­pen­dent trans­duc­tion, expan­sion,
and admin­is­tra­tion to patients at equal tar­get con­cen­tra­tions.
JULIET was an inter­na­tional phase 2 study of tisa-cel in 93
patients with R/R DLBCL who were inel­i­gi­ble for or had dis­ease
pro­gres­sion after ASCT.47 The best over­all response rate (ORR)
was 52% (95% CI, 41%-62%). Grade 3 or 4 adverse events of
inter­est included cyto­kine release syn­drome (CRS) in 22%, neu­
ro­log­i­cal events in 12%, and cytopenias last­ing more than 28
days in 32%. In a sub­se­quent study of this patient group at a
median fol­low-up of 40.3 months, the ORR was 53% and the PFS
at 3 years was 77.8% for patients in CR at 6 months.48 Among
111 patients enrolled in ZUMA-1, axi-cel was manufactured for
110 patients and admin­is­tered to 101 patients.46 The objec­tive
response rate was 82%, and the CR rate was 54%. Grade 3 or
higher CRS and neu­ro­log­i­cal events occurred in 13% and 28%
of patients, respec­tively. In a sub­se­quent study of this patient
group at a median fol­low-up of 27 months, the median dura­tion
of response for all­101 patients was 11 months and the median
dura­
tion of response for par­
tic­
i­
pants achiev­
ing CR was not
reached.49 In the TRANSCEND study, 344 patients under­went leukapheresis for the man­u­fac­ture of CAR T cells, and 269 patients
received at least 1 dose of liso-cel.45 Of 256 patients evaluable for
response, an objec­tive response was observed in 73% (95% CI,
66.8%–78.0%) and a CR in 53% (95% CI, 46.8%–59.4%). Grade 3
or greater CRS and neu­ro­log­i­cal events occurred in 2% and 10%
of patients, respec­
tively. Together these results dem­
on­
strate
mean­ing­ful and dura­ble responses for patients with R/R DLBCL,
includ­ing those who have failed prior ASCT. This remains a via­
ble alter­na­tive for patients who can access these ther­a­pies. As
described in the patient cases, there may be bar­ri­ers for cer­tain
patient groups, such as those who live in rural areas or who lack
insur­ance cov­er­age for CAR T-cell ther­apy. Additional mea­sures
are nec­es­sary to ensure equi­ta­ble access to poten­tially cura­tive
ther­apy for all­patients.
Common lym­phoma-spe­cific bridg­ing ther­apy admin­is­tered
between leukapheresis and CAR T-cell ther­apy includes ste­roids,
che­mo­ther­apy, targeted ther­apy, or radi­a­tion ther­apy. The opti­
mal approach for bridg­ing ther­apy prior to CAR T-cell ther­apy
remains uncer­tain.3,50-53 Data from 17 aca­demic insti­tu­tions involved
in the US Lymphoma CAR T Consortium showed that bridg­ing
ther­apy was admin­is­tered to 53% of 298 patients who under­went
leukapheresis prior to planned axi-cel. Among these patients
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/146/2021680/146flowers.pdf by guest on 09 December 2022
relapse or who have pri­mary refrac­tory dis­ease (less than a CR
or a relapse within 3 to 6 months to ini­tial ther­apy) have worse
out­comes.1,34,35 Although patients can be cured with treat­ments
fol­
low­
ing these early events, most of these patients die from
DLBCL. An IPI for relapsed/refrac­tory DLBCL (R/R IPI) was devel­
oped using 1234 of 5112 patients treated with front­
line immunochemotherapy in the SEAL con­
sor­
tium who expe­
ri­
enced
pro­gres­sive dis­ease.36 This study found that time to pro­gres­sion
on immunochemotherapy and age at pro­gres­sion were strongly
asso­ci­ated with postprogression OS. A prog­nos­tic model involv­
ing both var­i­ables had good con­cor­dance in the dis­cov­ery (0.67)
and val­i­da­tion sets (cohort 1, c = 0.64; cohort 2, c = 0.68). A cor­
re­
spond­
ing nomo­
gram is avail­­
able to aid in esti­
mat­
ing prog­
no­sis for R/R DLBCL.36 An elec­tronic cal­cu­la­tor of the model is
also avail­­able at https:​­/​­/qxmd​­.com​­/calculate​­/calculator_682​­
/prognosis​­-calculator​­-for​­-relapsed​­-refractory​­-dlbcl and as a
smart­phone app.
Patients who have relapses beyond 2 years may expe­ri­ence
more favor­able out­comes and have forms of DLBCL that can be
dis­tin­guished bio­log­i­cally.37 However, despite pro­vid­ing valu­able
prog­nos­tic and pre­dic­tive infor­ma­tion, molec­u­lar clas­si­fi­ca­tion
assays includ­ing IHC, gene expres­sion pro­fil­ing, and sequenc­
ing pan­els have faced sig­nif­i­cant chal­lenges, pre­clud­ing their
tran­si­tion into rou­tine clin­i­cal workflow and prac­tice.21-23,38,39
Even when such bio­mark­ers are avail­­able, they have had lim­ited
impact on ther­a­peu­tic deci­sion-mak­ing in the relapsed set­ting.
Validated, pre­dic­tive bio­mark­ers will be an inte­gral com­po­nent
of future per­son­al­ized med­i­cine strat­e­gies for DLBCL.39-41
For decades the stan­dard of care treat­ment strat­egy for fit
patients with relapsed DLBCL has been sal­vage ther­apy with non–
cross-resis­tant com­bi­na­tion chemoimmunotherapy reg­i­mens
followed by high-dose che­mo­ther­apy and ASCT as stem cell res­
cue for patients with chemosensitive dis­ease.42 The piv­otal RCTs
supporting the mod­ern prac­tice of stem cell trans­plan­ta­tion for
patients with large B-cell lym­phoma include the Parma trial that
was conducted in the era before the avail­abil­ity of rituximab
and the cur­rent World Health Organization lym­phoma clas­si­fi­
ca­tion. This study ran­dom­ized patients with R/R inter­me­di­ateto high-grade non-Hodgkin lym­phoma with chemosensitivity to
pro­ceed with either autol­o­gous stem cell trans­plan­ta­tion (ASCT)
or addi­tional che­mo­ther­apy with dexa­meth­a­sone, high-dose
cytarabine, and cis­platin (DHAP). It dem­on­strated that high-dose
che­mo­ther­apy followed by ASCT was asso­ci­ated with improved
event-free sur­vival (EFS) at 5 years of 46%, vs 12% with DHAP.42
Importantly, the CORAL trial and the LY.12 trial designed to com­
pare DHAP to alter­na­tive plat­i­num-containing sal­vage reg­i­mens
iden­ti­fied that patients with R/R dis­ease within 12 months of
diag­no­sis and prior expo­sure to rituximab had poor PFS fol­low­
ing sal­vage ther­apy.43,44 These find­ings indi­cate that the prior
dem­on­strated ben­e­fits of ASCT were reduced in the R-CHOP era
due to resis­tance to chemoimmunotherapy, iden­ti­fied by early
first-line treat­ment fail­ure. These poor out­comes asso­ci­ated with
early relapse of DLBCL are evi­dent in Alvin’s sit­u­a­tion. However,
ASCT can pro­vide ben­e­fits, par­tic­u­larly for patients like Wilma
who expe­ri­ence late relapse and have chemosensitive dis­ease.
Recent evi­dence-based guides sug­gest that ASCT remains a pre­
ferred option for patients who expe­ri­ence late relapse.3 Based
on the R/R IPI cal­cu­la­tor, the predicted 2-year sur­vival from pro­
gres­sion for Alvin and Wilma are 34% and 60%, respec­tively.36
bridg­ing reg­i­men as an event if it occurred within the first
12 weeks. Notably, all 3 tri­als dem­on­strated very poor out­
comes for patients with early R/R LBCL who under­went SOC
sal­vage ther­apy and ASCT. A vibrant, detailed, and timely dis­
cus­sion of these tri­als and their impli­ca­tions for man­age­ment
of patients with early relapsed LBCL was recently presented
in Blood by Drs Sehn and Westin.3 They also pro­vide a use­ful
algo­rithm for sec­ond-line ther­apy deci­sion-mak­ing for DLBCL.
Patients inel­i­gi­ble for ASCT or CAR T ther­apy or who
relapse after ASCT and CAR T ther­apy
While ASCT and CAR T have dem­on­strated dura­ble remis­sions for
patients who respond to treat­ment, patients who are ­inel­i­gi­ble,
unable to access these ther­a­pies, or who fail treat­ment often
expe­
ri­
ence poor out­
comes. DLBCL patients in these cir­
cum­
stances have mul­ti­ple options from recently approved ther­a­
pies, includ­
ing loncastuximab tesirine, polatuzumab vedotin,
selinexor, and tafasitamab. However, lim­ited data exist regard­
ing out­comes fol­low­ing CAR T ther­apy. Loncastuximab tesirine
is a CD19-directed anti­body-drug con­ju­gate that was eval­u­ated
in an inter­na­tional, mul­ti­cen­ter phase 2 trial (LOTIS-2) in patients
with R/R DLBCL fol­low­ing 2 or more lines of ther­apy. Among
145 patients who received at least 1 dose of loncastuximab, the
ORR was 48.3% (95% CI, 39.9-56.7), includ­ing 35 patients (24%)
who expe­ri­enced a CR.60 The most com­mon grade 3 or higher
adverse events were neutropenia in 26%, throm­bo­cy­to­pe­nia in
18%, and increased γ-glutamyltransferase in 17%. A matchingadjusted indi­rect com­par­i­son was performed to eval­u­ate the
effi­cacy of loncastuximab com­pared with chemoimmunotherapy in R/R DLBCL.61 In this anal­y­sis, 80 patients from LOTIS-2
were matched by char­ac­ter­is­tics to 278 patients from pooled
exten­sion stud­ies from CORAL, dem­on­strat­ing an ORR of 53.4%
in the loncastuximab cohort com­pared with 40.3% for chemoimmunotherapy. As exem­pli­fied below, real-world evi­dence com­
par­i­sons are becom­ing more com­mon in set­tings like R/R DLBCL
where ran­dom­ized stud­ies are uncom­mon. A detailed dis­cus­sion
of these ana­ly­ses is beyond the scope of this text, but they merit
addi­tional con­sid­er­ation.
The anti­body-drug con­ju­gate polatuzumab vedotin tar­gets
CD79b, a com­po­nent of the B-cell recep­tor that is expressed
on 95% of DLBCLs. Patients with R/R DLBCL inel­i­gi­ble for ASCT
(n = 80) were ran­dom­ized to receive polatuzumab vedotin com­
bined with bendamustine and rituximab (pola-BR) or bendamustine and rituximab (BR).62 Patients who received pola-BR had a
sig­nif­i­cantly higher inde­pen­dent review com­mit­tee–assessed
CR rate (40.0% vs 17.5%; P = .026), PFS (median, 9.5 vs 3.7 months;
P<.001), and OS (median, 12.4 vs 4.7 months), with a median fol­lowup of 22.3 months. Pola-BR was asso­ci­ated with higher rates
of grade 3 to 4 neutropenia (46.2% vs 33.3%), ane­mia (28.2%
vs 17.9%), and throm­bo­cy­to­pe­nia (41% vs 23.1%). In a fol­low-up
sin­gle-arm exten­sion cohort, 106 addi­
tional patients received
pola-BR, dem­on­strat­ing an objec­tive response rate of 41.5% and
a CR rate of 38.7%.63 Median PFS and OS were 6.6 months and
12.5 months, respec­tively. No new safety sig­nals with pola-BR
were iden­ti­fied, and dif­fer­ences between the ran­dom­ized arms
persisted with addi­tional fol­low-up. A bio­marker anal­y­sis led
by Dr Herrera ran a ctDNA assay based on a cus­tom­ized panel
of recur­rently mutated genes in DLBCL at base­line and end of
treat­ment.64 Higher base­line ctDNA lev­els were asso­ci­ated with
poor prog­nos­tic fac­tors and were an inde­pen­dent pre­dic­tor of
Dr Prakash Singh Shekhawat
Sequencing ther­apy in relapsed DLBCL | 149
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/146/2021680/146flowers.pdf by guest on 09 December 2022
23% received ste­
roids alone, 54% received che­
mo­
ther­
apy,
12% received radi­
a­
tion ther­
apy, and 10% received targeted
ther­apy.54 A mul­ti­var­i­able anal­y­sis in this data set revealed that
bridg­ing was asso­ci­ated with infe­rior OS. However, the need for
bridg­ing may be a sur­ro­gate for greater dis­ease bur­den or more
aggres­sive lym­phoma. A study by Pinnix and col­leagues sug­
gests that radi­a­tion may be an effec­tive approach when bridg­
ing is required,50 but addi­tional stud­ies are needed in this area.
Recent data sug­gest that patients such as Alvin may ben­e­
fit from CAR T-cell ther­apy in the sec­ond-line set­ting. Given the
results discussed above and real-world com­par­i­sons to cohorts
of patients with R/R DLBCL,55,56 3 RCTs com­pared CAR T ther­apy
in the sec­ond-line set­ting to the stan­dard of care (SOC) strat­egy
of sal­vage chemoimmunotherapy and ASCT for patients with
chemosensitive dis­ease: ZUMA-7 (axi-cel), Belinda (tisa-cel), and
TRANSFORM (liso-cel).57-59 All tri­als enrolled patients with large
B-cell lym­phoma who relapsed within 12 months from com­ple­
tion of or refrac­tory to first-line ther­apy and were con­sid­ered
can­di­dates for ASCT by the treating phy­si­cian.
In ZUMA-7, 180 patients were ran­dom­ized to receive axi-cel
and 179 to receive SOC.58 The ORR and CR rate were sig­nif­
i­cantly greater in the axi-cel group (83% and 65%) com­pared
with the SOC group (50% and 32%). At a median fol­low-up of
24.9 months, the median EFS was 8.3 months for patients who
received axi-cel and 2.0 months for patients who received SOC,
and the 24-month EFS was 41% and 16%, respec­tively. Estimated
OS at 2 years was 61% in the axi-cel group and 52% for SOC.
Grade 3 or higher adverse events occurred in 91% of the patients
who received axi-cel and in 83% of those who received SOC. In
the TRANSFORM trial, 92 patients were ran­dom­ized to receive
liso-cel and 92 patients to receive SOC.59 The median EFS was
10.1 months for liso-cel vs 2.3 for SOC (HR, 0.349; P<.0001), the
median PFS was 14.8 months and 5.7 months, and the CR rate
was 66% and 39%, respec­
tively. In the group receiv­
ing lisocel, CRS grade 1 occurred in 37%, grade 2 in 11%, and grade 3
in 1 patient. Unlike the other 2 tri­als, the Belinda study found
that CAR T-cell ther­apy was not supe­rior to SOC sal­vage ther­
apy.57 Patients were ran­domly assigned to receive tisa-cel with
optional bridg­ing ther­apy or sal­vage che­mo­ther­apy and ASCT
as SOC. Approximately 26% of the patients in the tisa-cel group
expe­ri­enced pro­gres­sion at week 6, com­pared with approx­i­ma­
tely 14% of the patients who received SOC. Nearly 96% of the
patients in the tisa-cel group received tisa-cel, and 32.5% of the
patients in the SOC group received ASCT. Response occurred for
46.3% of patients receiv­ing tisa-cel vs 42.5% of patients receiv­
ing SOC. The median EFS was 3.0 months in both groups.
Some dif­fer­ences in trial design that may explain these
results include that Belinda allowed for mul­
ti­
ple cycles of
bridg­ing ther­apy, includ­ing switching bridg­ing ther­apy,
whereas TRANSFORM allowed for only 1 cycle of bridg­ing, and
ZUMA-7 allowed for sta­bi­li­za­tion with ste­roids but no bridg­ing
ther­apy. In Belinda bridg­ing che­mo­ther­apy was admin­is­tered
to 83% of the patients receiv­ing tisa-cel; 48% received more
than 1 cycle, and 12% received more than 1 reg­i­men. Patients
enrolled in TRANSFORM and Belinda under­went leukapheresis
before ran­dom­i­za­tion and were allowed to cross over from
SOC. No cross­over was planned in ZUMA-7, but patients were
per­mit­ted to access approved CAR T-cell ther­a­pies out­side
the pro­to­col. As a sta­tis­ti­cal nuance, in the assess­ment of EFS
Belinda did not count the ini­ti­a­tion of a sec­ond sal­vage or
Clinical tri­als involv­ing bispecific antibodies
Bispecific antibodies bind­ing with 2 dif­fer­ent cell sur­face anti­
gens can direct cyto­toxic T cells and other immune effec­tor
cells to areas in prox­im­ity of lym­phoma cells. Several bispecific antibodies bind­ing CD3 and CD20 are in devel­op­ment for
DLBCL, includ­ing mosunetuzumab, glofitamab, e
­ pcoritamab,
plamotamab, and odronextamab, but have some d
­ if­fer­ences in
struc­ture, route of admin­is­tra­tion, and treat­ment sched­ule.71-75
Across agents, com­mon adverse events include neutropenia,
150 | Hematology 2022 | ASH Education Program
throm­bo­cy­to­pe­nia, CRS, hypophosphatemia, fatigue, and
diar­rhea. CRS neu­ro­log­i­cal adverse events were mostly low
grade and con­
fined to cycle 1. Among 129 patients with
R/R aggres­sive B-cell lym­phoma (includ­ing 82 patients with
DLBCL) treated with mosunetuzumab, the ORR was 34.9%,
and the CR rate was 19.4%, with a median dura­tion of response
for all­respond­ers and com­plete respond­ers of 7.6 months
and 22.8 months, respec­tively.71 Among 19 patients who had
pre­vi­ously under­gone CAR T-cell ther­apy, the ORR and CR
rate were 36.8% and 26.3%, respec­tively, but the dura­tion
of response could not be esti­mated due to the small sam­ple
size. In a study of 107 patients who received glofitamab, most
patients were refrac­tory to a prior CD20 anti­body–containing
reg­i­men (85%), many were refrac­tory to their ini­tial ther­apy
(59%), and 32% had prior CAR T. The ORR was 50%, and CR
rates were 32% and 37%, respec­tively, among patients with
and with­
out prior CAR T.76 ­Epcoritamab, a sub­cu­ta­ne­ously
admin­is­tered CD3xCD20 bispecific anti­body,73 odronextamab (REGN1979), a fully human­ized bispecific immu­no­glob­
u­
lin G4 anti­
body with a 14-day half-life,74 and plamotamab
(XmAb13676) also have been exam­ined in early-phase clin­i­cal
tri­als, dem­on­strat­ing mean­ing­ful ORRs and rel­a­tively sim­i­lar
adverse event pro­files.75 For instance, a study of 157 patients
who received epcoritamab dem­on­strated ORRs of 69% for
CAR T–naive patients and 54% for patients who received
prior CAR T.77 Future stud­ies are needed to clar­ify the role and
sequenc­
ing of bispecific antibodies rel­
a­
tive to the cur­
rent
ther­a­pies avail­­able for R/R DLBCL.
Goals-of-care dis­cus­sions for patients with DLBCL
Although the advent of new ther­a­pies for DLBCL has improved
sur­vival out­comes, up to 40% of patients develop R/R dis­ease.2,3
This empha­sizes that DLBCL is still a life-threat­en­ing ill­ness, and
it is thus essen­tial for cli­ni­cians to engage in seri­ous ill­ness con­
ver­sa­tions with their patients. Goals-of-care dis­cus­sions entail
eliciting patients’ goals, val­ues, and pref­er­ences regard­ing their
over­all care as well as end-of-life (EOL) options. Ideally, these
dis­cus­sions should occur early and often in the DLBCL dis­ease
tra­jec­tory and should not be merely lim­ited to EOL time points.78
Patients who have the oppor­tu­nity to engage in goals-of-care
dis­cus­sions with their phy­si­cians are sig­nif­i­cantly more likely to
receive care that is aligned with their pref­er­ences and expe­ri­
ence higher-qual­ity EOL care.79,80 Hematologic oncol­o­gists have
an impor­tant role in ini­ti­at­ing these con­ver­sa­tions with patients
given the trust built through lon­gi­tu­di­nal patient-phy­si­cian rela­
tion­ships. In a ret­ro­spec­tive study of 383 patients with hema­to­
logic malig­nan­cies, of which 37% had lym­phoma, patients who
had their first goals-of-care dis­cus­sion with a hema­to­logic oncol­
o­gist (vs other cli­ni­cians) were sig­nif­i­cantly more likely to enroll
in hos­pice more than 3 days before death and had lower odds
of admis­sion to the inten­sive care unit in the last month of life or
death in the hos­pi­tal.81 Such EOL out­comes are asso­ci­ated with
a bet­ter qual­ity of life for patients and their fam­i­lies, as well as a
lower risk of com­pli­cated grief for bereaved care­giv­ers.82
Goals-of-care dis­cus­sions are crit­i­cal and ben­e­fi­cial for
patients with DLBCL, yet these dis­cus­sions often occur too late
or not at all­. In a study of patients with aggres­sive lym­phoma
relapsed after first- or sec­ond-line treat­ment, while 44.4% had
thought of their care pref­er­ences in the event of becom­ing crit­
i­cally ill, only a quar­ter of these patients had discussed these
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/146/2021680/146flowers.pdf by guest on 09 December 2022
shorter PFS, suggesting that ctDNA may be a tool for iden­ti­fy­ing
patients at higher risk of pro­gres­sion.
Selinexor is an oral selec­tive inhib­i­tor of nuclear export that
works by bind­ing to exportin 1. An inter­na­tional mul­ti­cen­ter
phase 2 trial (SADAL) eval­u­ated selinexor given on days 1 and
3 weekly until dis­ease pro­gres­sion or unac­cept­able tox­ic­ity.
Among 127 patients who received selinexor at 60 mg and
were included in ana­ly­ses of pri­mary out­come and safety, the
ORR was 28%, and the CR rate was 12%.65 The most com­mon
grade 3 to 4 adverse events were throm­bo­cy­to­pe­nia (46%),
neutropenia (24%), ane­
mia (n = 22%), and fatigue (n = 11%).
Additional ana­ly­ses have been performed for bio­log­i­cal and
prior-treat­ment sub­groups.52,66
Tafasitamab is a human­ized, anti-CD19 mono­clo­nal anti­body.
In the inter­na­tional mul­ti­cen­ter phase 2 trial L-MIND, patients with
R/R DLBCL inel­i­gi­ble for ASCT received tafasitamab and lenalidomide for up to 12 cycles followed by tafasitamab monotherapy
(in patients with sta­ble dis­ease or bet­ter) until dis­ease pro­gres­
sion.67,68 Among 80 patients who received tafasitamab plus lenalidomide, 60% (95% CI, 48%-71%) had an objec­tive response and
43% (95% CI, 32%-54%) expe­ri­enced CR. Therapy was gen­er­ally
well tol­er­ated, and the most com­mon grade 3 or higher adverse
events were neutropenia (48%), throm­bo­cy­to­pe­nia (17%), and
febrile neutropenia (12%). The most com­mon nonhematological
adverse effects were rash, diar­rhea, asthe­nia, cough, periph­eral
edema, and fever. Tafasitamab and lenalidomide was more effec­
tive as a sec­ond-line treat­ment than as treat­ment at third line or
beyond (median PFS, 23.5 months vs 7.6 months).68 In addi­tion, it
is impor­tant to note that patients with dou­ble- or tri­ple-hit lym­
phoma (MYC and BCL2 and/or BCL6 rearrangement) or pri­mary
refrac­tory dis­ease were excluded from this study.
In a fol­low-up study, the median PFS was 11.6 months, and
the median OS was 33.5 months.68 A ret­ro­spec­tive obser­va­
tional study (RE-MIND) was performed aligning key eli­gi­bil­ity
cri­te­ria with L-MIND using data from 490 patients with R/R
DLBCL who received lenalidomide monotherapy; 140 qual­i­fied
for matching with the L-MIND cohort, and the pri­mary anal­y­sis
included 76 patients from each cohort.69 The anal­y­sis dem­on­
strated sig­nif­i­cantly higher over­all response from the matched
group from L-MIND. A sec­
ond study performed matchingadjusted indi­
rect com­
par­
i­
sons using data from L-MIND to
match with described clin­i­cal trial cohorts such as pola-BR
and gemcitabine-oxaliplatin-rituximab.70 Such approaches
may increas­ingly be used to inform ther­a­peu­tic deci­sions as
more robust data on treat­ments and out­comes become avail­­
able for patients with R/R DLBCL. For patients under­go­ing
CAR T-cell ther­apy prior to this reg­i­men or con­sid­er­ing future
CAR T-cell ther­apy fol­low­ing tafasitamab, it remains unclear
to what extent anti-CD19 ther­apy or other treat­ments might
impair the effi­cacy of future CD19-targeting ther­a­pies.
Table 2. Triggers for goals-of-care dis­cus­sions for patients
with DLBCL
Refractory dis­ease
Relapsed dis­ease
Declining per­for­mance sta­tus
Organ insuf­fi­ciency
Consideration of ASCT
Consideration of CAR T ther­apy
If you “would not be sur­prised if the patient died
within the next year.”
Table 3. REMAP frame­work for patients with DLBCL
REMAP
Description
Reframe
Assess the patient’s under­stand­ing of his or her dis­ease sta­tus; place details of the ill­ness in a larger con­text, tai­lor­ing
and titrat­ing infor­ma­tion according to patient pref­er­ence.
Expect emo­tion
Acknowledge the patient’s emo­tions.
Map out­pa­tient val­ues
Explore what mat­ters most to the patient in the con­text of DLBCL and the patient’s con­cerns about the future and
goals and the trade-offs they are will­ing to make.
Align with val­ues
Verbally reflect what is heard from the patient to ensure clear under­stand­ing of the patient’s val­ues. This step offers
the oppor­tu­nity for addi­tional clar­i­fi­ca­tion as needed.
Propose a plan
With per­mis­sion from the patient, rec­om­mend a plan of care that has the best chance of max­i­miz­ing the patient’s
goals using a com­bi­na­tion of the patient’s val­ues and your knowl­edge of fea­si­ble med­i­cal treat­ments to help achieve
the patient’s goals.
Adapted from Childers et al90 with per­mis­sion.
Dr Prakash Singh Shekhawat
Sequencing ther­apy in relapsed DLBCL | 151
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/146/2021680/146flowers.pdf by guest on 09 December 2022
pref­er­ences with their cli­ni­cian.83 Most hema­to­logic oncol­o­gists
(56%) in a national sur­vey reported that goals-of-care dis­cus­
sions typ­i­cally occur “too late,”84 and the median time between
the first documented goals-of-care dis­cus­sion and death was
only 15 days in a study of patients who died of DLBCL or other
blood can­cers.81 Several bar­ri­ers con­trib­ute to the cur­rent state
of goals-of-care dis­cus­sions. Given recent treat­ment advances
and the poten­tial for cure in R/R set­tings, the high prog­nos­tic
uncer­tainty in DLBCL may lead some cli­ni­cians to delay dis­cus­
sions until death is very clearly immi­nent. Concerns about tak­ing
away patients’ hope also con­trib­ute to hes­i­tancy in conducting
these dis­cus­sions despite data show­ing that these con­ver­sa­
tions do not pre­clude hope.85 Even when cli­ni­cians grasp the
impor­tance of engag­ing in goals-of-care dis­cus­sions, about 40%
report that they do not know the right thing to say, which may
result in low-qual­ity dis­cus­sions.86 Taken together, these data
sug­gest an urgent need for strat­e­gies to improve goals-of-care
dis­cus­sions for DLBCL patients.
To improve time­
li­
ness of goals-of-care dis­
cus­
sions for
patients with DLBCL, it is impor­tant to start dis­cus­sions before
deci­sions are needed to avoid “cri­sis” com­mu­ni­ca­tion when
patients are very close to death. Having broad con­ver­sa­tions to
under­stand patients’ val­ues and goals for their over­all care and
qual­ity of life dur­ing first-line ther­apy helps to set the foun­da­tion
for future dis­cus­sions. Practical trig­gers to ini­ti­ate or revisit more
spe­cific goals-of-care dis­cus­sions include the pres­ence of refrac­
tory or relapsed dis­ease (Table 2). These dis­ease points were
­iden­ti­fied in a focus group of lym­phoma cli­ni­cians as crit­i­cal sign­
posts for goals-of-care dis­cus­sions.87 Other trig­gers iden­ti­fied
by ­lym­phoma cli­ni­cians include organ insuf­fi­ciency or declin­ing
per­for­mance sta­tus, even in the absence of lym­phoma relapse.
Consideration of inten­sive ther­a­pies such as ASCT or CAR T-cell
ther­apy should also prompt goals-of-care dis­cus­sions. Answering no to the ques­tion “Would you be sur­prised if this patient
died in the next year?” is also an effec­tive trig­ger for engag­ing
in these dis­cus­sions.88 This ques­tion was dem­on­strated to cor­
rectly esti­mate death within 12 months in 68.3% of patients with
hema­to­logic malig­nan­cies.89 Goals-of-care dis­
cus­
sions should
not be con­
sid­
ered one-time con­
ver­
sa­
tions because patients’
pref­er­ences may change as their dis­ease evolves.
Communication skills train­
ing to con­
duct nuanced dis­
cus­
sions that effec­
tively bal­
ance the poten­
tial prom­
ise of
­dis­ease-directed treat­ments with their risks and lim­i­ta­tions are
nec­es­sary to opti­mize goals-of-care dis­cus­sions. Tools such
as the seri­ous ill­ness con­ver­sa­tion guide and REMAP (reframe,
expect emo­tion, map out patient val­ues, align with val­ues, and
pro­pose a plan) frame­work (Table 3) can help cli­ni­cians incor­po­
rate key ele­ments of goals-of-care dis­cus­sions when engag­ing
in these con­ver­sa­tions with DLBCL patients.77,88,90 In using these
tools, cli­ni­cians should assess how much the patient under­
stands and desires to know about the dis­ease tra­jec­tory and
prog­no­sis and then com­pas­sion­ately share infor­ma­tion tai­lored
to the patient’s pref­er­ences. It is also essen­tial to engage the
patient’s loved ones in these con­ver­sa­tions to pro­mote align­
ment of goals across the fam­ily unit. These dis­cus­sions should
be documented in the med­i­cal record to ensure that patients’
pref­er­ences are hon­ored dur­ing care tran­si­tions. Goals-of-care
skills–based train­ing, such as VitalTalk and the Serious Illness
Care Program, have effec­tively incor­po­rated these com­mu­ni­ca­
tion tools.91,92 In a clus­ter ran­dom­ized trial of the Serious Illness
Care Program (a seri­ous ill­ness con­ver­sa­tion guide com­bined
with 2.5 hours of skills-based train­ing) among 278 patients with
advanced can­cer, the pro­gram resulted in more fre­quent, ear­
lier, and higher-qual­ity goals-of-care dis­cus­sions.93 Moreover,
this inter­ven­tion was asso­ci­ated with sig­nif­i­cant reduc­tions in
depres­sion and anx­i­ety symp­toms.92 These find­ings illus­trate
the poten­tial impact of opti­miz­ing goals-of-care dis­cus­sions
through­
out the DLBCL dis­
ease course. Clinicians can apply
this skill set to the dis­cus­sions and care plans for each of our
­exam­ple patients.
Conflict-of-inter­est dis­clo­sure
Off-label drug use
The bispecific antibodies mentioned in this article, mosunetuzumab, glofitamab, epcoritamab, plamotamab, and odronextamab, are not approved agents and their use was discussed
within the context of completed and ongoing clinical trials.
Correspondence
Christopher R. Flowers, Department of Lymphoma/Myeloma,
University of Texas MD Anderson Cancer Center, 1515 Holcombe
Blvd, Houston, TX 77030; e-mail: crflowers@mdanderson​­.org.
References
1.
Sehn LH, Salles G. Diffuse large B-cell lym­
phoma. N Engl J Med.
2021;384(9):842-858.
2. Sarkozy C, Sehn LH. Management of relapsed/refrac­
tory DLBCL. Best
Pract Res Clin Haematol. 2018;31(3):209-216.
3. Westin J, Sehn LH. CAR T cells as a sec­ond-line ther­apy for large B-cell lym­
phoma: a par­a­digm shift? Blood. 2022;139(18):2737-2746.
4. Teras LR, DeSantis CE, Cerhan JR, Morton LM, Jemal A, Flowers CR. 2016 US
lym­phoid malig­nancy sta­tis­tics by World Health Organization sub­types.
CA Cancer J Clin. 2016;66(6):443-459.
5. Alizadeh AA, Eisen MB, Davis RE, et al. Distinct types of dif­
fuse large
B-cell lym­phoma iden­ti­fied by gene expres­sion pro­fil­ing. Nature. 2000;
403(6769):503-511.
6. Wright G, Tan B, Rosenwald A, Hurt EH, Wiestner A, Staudt LM. A gene
expres­sion-based method to diag­nose clin­i­cally dis­tinct sub­groups of dif­
fuse large B cell lym­phoma. Proc Natl Acad Sci USA. 2003;100(17):99919996.
7. Coiffier B, Lepage E, Briere J, et al. CHOP che­mo­ther­apy plus rituximab
com­pared with CHOP alone in elderly patients with dif­fuse large-B-cell
lym­phoma. N Engl J Med. 2002;346(4):235-242.
8. Habermann TM, Weller EA, Morrison VA, et al. Rituximab-CHOP ver­
sus
CHOP alone or with main­te­nance rituximab in older patients with dif­fuse
large B-cell lym­phoma. J Clin Oncol. 2006;24(19):3121-3127.
9. Merli F, Luminari S, Rossi G, et al. Cyclophosphamide, doxo­
ru­
bi­
cin, vin­
cris­tine, pred­ni­sone and rituximab ver­sus epirubicin, cyclo­phos­pha­mide,
­vin­blas­tine, pred­ni­sone and rituximab for the ini­tial treat­ment of elderly “fit”
patients with dif­fuse large B-cell lym­phoma: results from the ANZINTER3 trial
of the Intergruppo Italiano Linfomi. Leuk Lymphoma. 2012;53(4):581-588.
10. Schmitz N, Nickelsen M, Ziepert M, et al; Ger­man High-Grade Lymphoma
Study Group. Conventional che­mo­ther­apy (CHOEP-14) with rituximab or
high-dose che­mo­ther­apy (MegaCHOEP) with rituximab for young, highrisk patients with aggres­sive B-cell lym­phoma: an open-label, randomised,
phase 3 trial (DSHNHL 2002-1). Lancet Oncol. 2012;13(12):1250-1259.
11. Delarue R, Tilly H, Mounier N, et al. Dose-dense rituximab-CHOP com­
pared with stan­dard rituximab-CHOP in elderly patients with dif­fuse large
B-cell lym­phoma (the LNH03-6B study): a randomised phase 3 trial. Lancet
Oncol. 2013;14(6):525-533.
12. Herbrecht R, Cernohous P, Engert A, et al. Comparison of pixantrone-based
reg­i­men (CPOP-R) with doxo­ru­bi­cin-based ther­apy (CHOP-R) for ­treat­ment
of dif­fuse large B-cell lym­phoma. Ann Oncol. 2013;24(10):2618-2623.
152 | Hematology 2022 | ASH Education Program
13. Bartlett NL, Wilson WH, Jung SH, et al. Dose-adjusted EPOCH-R com­pared
with R-CHOP as front­line ther­apy for dif­fuse large B-cell lym­phoma: clin­i­cal
out­comes of the phase III Intergroup Trial Alliance/CALGB 50303. J Clin
Oncol. 2019;37(21):1790-1799.
14. Davies A, Cummin TE, Barrans S, et al. Gene-expres­sion pro­fil­ing of bortezomib added to stan­dard chemoimmunotherapy for dif­fuse large B-cell
lym­phoma (REMoDL-B): an open-label, randomised, phase 3 trial. Lancet
Oncol. 2019;20(5):649-662.
15. Younes A, Sehn LH, Johnson P, et al; PHOENIX Investigators. Randomized
phase III trial of ibrutinib and rituximab plus cyclo­phos­pha­mide, doxo­ru­bi­
cin, vin­cris­tine, and pred­ni­sone in non-ger­mi­nal cen­ter B-cell dif­fuse large
B-cell lym­phoma. J Clin Oncol. 2019;37(15):1285-1295.
16. Nowakowski GS, Chiappella A, Gascoyne RD, et al. ROBUST: a phase III
study of lenalidomide plus R-CHOP ver­sus pla­cebo plus R-CHOP in pre­
vi­ously untreated patients with ABC-type dif­fuse large B-cell lym­phoma.
J Clin Oncol. 2021;39(12):1317-1328.
17. Tilly H, Morschhauser F, Sehn LH, et al. Polatuzumab vedotin in pre­vi­ously
untreated dif­fuse large B-cell lym­phoma. N Engl J Med. 2022;386(4):351-363.
18. Ruppert AS, Dixon JG, Salles G, et al. International prog­nos­tic indi­ces in
dif­fuse large B-cell lym­phoma: a com­par­i­son of IPI, R-IPI, and NCCN-IPI.
Blood. 2020;135(23):2041-2048.
19. Choi WW, Weisenburger DD, Greiner TC, et al. A new immunostain algo­
rithm classifies dif­fuse large B-cell lym­phoma into molec­u­lar sub­types with
high accu­racy. Clin Cancer Res. 2009;15(17):5494-5502.
20. Hans CP, Weisenburger DD, Greiner TC, et al. Confirmation of the molec­u­
lar clas­si­fi­ca­tion of dif­fuse large B-cell lym­phoma by immu­no­his­to­chem­is­
try using a tis­sue microarray. Blood. 2004;103(1):275-282.
21. Ennishi D, Jiang A, Boyle M, et al. Double-hit gene expres­sion sig­na­ture
defi­
nes a dis­
tinct sub­
group of ger­
mi­
nal cen­
ter B-cell-like dif­
fuse large
B-cell lym­phoma. J Clin Oncol. 2019;37(3):190-201.
22. Scott DW, Mottok A, Ennishi D, et al. Prognostic sig­nif­i­cance of dif­fuse
large B-cell lym­phoma cell of ori­gin deter­mined by dig­i­tal gene expres­
sion in for­ma­lin-fixed par­af­fin-embed­ded tis­sue biop­sies. J Clin Oncol.
2015;33(26):2848-2856.
23. Scott DW, Wright GW, Williams PM, et al. Determining cell-of-ori­gin sub­
types of dif­fuse large B-cell lym­phoma using gene expres­sion in for­ma­linfixed par­af­fin-embed­ded tis­sue. Blood. 2014;123(8):1214-1217.
24. Chapuy B, Stewart C, Dunford AJ, et al. Molecular sub­types of dif­fuse large
B cell lym­phoma are asso­ci­ated with dis­tinct path­o­genic mech­a­nisms and
out­comes. Nat Med. 2018;24(5):679-690.
25. Schmitz R, Wright GW, Huang DW, et al. Genetics and path­o­gen­e­sis of dif­
fuse large B-cell lym­phoma. N Engl J Med. 2018;378(15):1396-1407.
26. Wright GW, Huang DW, Phelan JD, et al. A prob­a­bi­lis­tic clas­si­fi­ca­tion tool
for genetic sub­types of dif­fuse large B cell lym­phoma with ther­a­peu­tic
impli­ca­tions. Cancer Cell. 2020;37(4):551-568.e14568e14.
27. Ritter AJ, Goldstein JS, Ayers AA, Flowers CR. Rural and urban patients with
dif­fuse large B-cell and fol­lic­u­lar lym­phoma expe­ri­ence reduced over­all sur­
vival: a National Cancer DataBase study. Leuk Lymphoma. 2019;60(7):16561667.
28. Han X, Jemal A, Flowers CR, Sineshaw H, Nastoupil LJ, Ward E. Insurance sta­tus is related to dif­fuse large B-cell lym­phoma sur­vival. Cancer.
2014;120(8):1220-1227.
29. Hu B, Boselli D, Pye LM, et al. Equal access to care and nurse nav­i­ga­tion
leads to equi­ta­ble out­comes for minor­i­ties with aggres­sive large B-cell
lym­phoma. Cancer. 2021;127(21):3991-3997.
30. Maurer MJ, Ghesquières H, Link BK, et al. Diagnosis-to-treat­ment inter­
val is an impor­tant clin­i­cal fac­tor in newly diag­nosed dif­fuse large B-cell
lym­
phoma and has impli­
ca­
tion for bias in clin­
i­
cal tri­
als. J Clin Oncol.
2018;36(16):1603-1610.
31. Alig S, Macaulay CW, Kurtz DM, et al. Short diag­no­sis-to-treat­ment inter­val
is asso­ci­ated with higher cir­cu­lat­ing tumor DNA lev­els in dif­fuse large B-cell
lym­phoma. J Clin Oncol. 2021;39(23):2605-2616.
32. Blunt DN, Smyth L, Nagamuthu C, et al. Shorter diag­no­sis-to-treat­ment
inter­val in dif­fuse large B-cell lym­phoma is asso­ci­ated with infe­rior over­all
sur­vival in a large, pop­u­la­tion-based reg­is­try. J Natl Compr Canc Netw.
2021;19(6):719-725.
33. Yoshida M, Nakaya Y, Shimizu K, et al. Importance of diag­no­sis-to-treat­
ment inter­val in newly diag­nosed patients with dif­fuse large B-cell lym­
phoma. Sci Rep. 2021;11(1):2837.
34. Maurer MJ, Ghesquières H, Jais JP, et al. Event-free sur­vival at 24 months is
a robust end point for dis­ease-related out­come in dif­fuse large B-cell lym­
phoma treated with immunochemotherapy. J Clin Oncol. 2014;32(10):10661073.
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/146/2021680/146flowers.pdf by guest on 09 December 2022
Christopher R. Flowers: con­sul­tancy: AstraZeneca, Bayer, BeiGene, BioAscend, Bristol Myers Squibb, Celgene, Curio Sciences,
Denovo Biopharma, Epizyme/Incyte, Foresight Diagnostics,
Genentech/Roche, Genmab, MEI Pharmaceuticals, MorphoSys
AG, Pharmacyclics/Janssen, SeaGen; stock options: Foresight
Diagnostics, N Power; research funding: 4D, Abbvie, Acerta,
Adaptimmune, Allogene, Amgen, Bayer, Celgene, Cellectis, EMD,
Gilead, Genentech/Roche, Guardant, Iovance, Janssen Pharmaceutical, Kite, Morphosys, Nektar, Novartis, Pfizer, Pharmacyclics,
Sanofi, Takeda, TGTherapeutics, Xencor, Ziopharm, Burroughs
Wellcome Fund, Eastern Cooperative Oncology Group, National
Cancer Institute, V Foundation, Cancer Prevention and Research
Institute of Texas CPRIT Scholar in Cancer Research.
Oreofe O. Odejide: no com­pet­ing finan­cial inter­ests to declare.
55. Crump M, Neelapu SS, Farooq U, et al. Outcomes in refrac­tory dif­fuse large
B-cell lym­phoma: results from the inter­na­tional SCHOLAR-1 study. Blood.
2017;130(16):1800-1808.
56. Neelapu SS, Locke FL, Bartlett NL, et al. Comparison of 2-year out­comes
with CAR T cells (ZUMA-1) vs sal­vage che­mo­ther­apy in refrac­tory large
B-cell lym­phoma. Blood Adv. 2021;5(20):4149-4155.
57. Bishop MR, Dickinson M, Purtill D, et al. Second-line tisagenlecleucel or stan­
dard care in aggres­sive B-cell lym­phoma. N Engl J Med. 2022;386(7):629639.
58. Locke FL, Miklos DB, Jacobson CA, et al; ZUMA-7 Investigators and Contributing Kite Members. Axicabtagene ciloleucel as sec­ond-line ther­apy for
large B-cell lym­phoma. N Engl J Med. 2022;386(7):640-654.
59. Kamdar M, Sol­o­mon SR, Arnason JE, et al. Lisocabtagene maraleucel (lisocel), a CD19-directed chi­me­ric anti­gen recep­tor (CAR) T cell ther­apy, ver­
sus stan­dard of care (SOC) with sal­vage che­mo­ther­apy (CT) followed by
autol­o­gous stem cell trans­plan­ta­tion (ASCT) as sec­ond-line (2L) treat­ment
in patients (Pts) with relapsed or refrac­tory (R/R) large B-cell lym­phoma
(LBCL): results from the ran­
dom­
ized phase 3 trans­
form study. Blood.
2021;138(suppl 1):91.
60. Caimi PF, Ai W, Alderuccio JP, et al. Loncastuximab tesirine in relapsed or
refrac­tory dif­fuse large B-cell lym­phoma (LOTIS-2): a multicentre, openlabel, sin­gle-arm, phase 2 trial. Lancet Oncol. 2021;22(6):790-800.
61. Hamadani M, Chen L, Song Y, et al. Matching-adjusted indi­rect com­par­i­son
of the effi­cacy of loncastuximab tesirine ver­sus treat­ment in the chemoimmunotherapy era for relapsed/refrac­tory dif­fuse large B-cell lym­phoma.
Clin Lymphoma Myeloma Leuk. 2022;22(8):e738-e744.
62. Sehn LH, Herrera AF, Flowers CR, et al. Polatuzumab vedotin in relapsed
or refrac­tory dif­fuse large B-cell lym­phoma. J Clin Oncol. 2020;38(2):155165.
63. Sehn LH, Hertzberg M, Opat S, et al. Polatuzumab vedotin plus bendamustine and rituximab in relapsed/refrac­tory DLBCL: sur­vival update and new
exten­sion cohort data. Blood Adv. 2022;6(2):533-543.
64. Herrera AF, Tracy S, Croft B, et al. Risk pro­fil­ing of patients with relapsed/
refrac­tory dif­fuse large B-cell lym­phoma by mea­sur­ing cir­cu­lat­ing tumor
DNA. Blood Adv. 2022;6(6):1651-1660.
65. Kalakonda N, Maerevoet M, Cavallo F, et al. Selinexor in patients with
relapsed or refrac­tory dif­fuse large B-cell lym­phoma (SADAL): a sin­gle-arm,
mul­
ti­
na­
tional, multicentre, open-label, phase 2 trial. Lancet Haematol.
2020;7(7):e511-e522.
66. Casasnovas RO, Follows G, Zijlstra JM, et al. Comparison of the effec­tive­
ness and safety of the oral selec­tive inhib­i­tor of nuclear export, selinexor,
in dif­fuse large B cell lym­phoma sub­types. Clin Lymphoma Myeloma Leuk.
2022;22(1):24-33.
67. Salles G, Duell J, González Barca E, et al. Tafasitamab plus lenalidomide in
relapsed or refrac­tory dif­fuse large B-cell lym­phoma (L-MIND): a multicentre,
pro­spec­tive, sin­gle-arm, phase 2 study. Lancet Oncol. 2020;21(7):978-988.
68. Duell J, Maddocks KJ, González-Barca E, et al. Long-term out­comes from
the phase II L-MIND study of tafasitamab (MOR208) plus lenalidomide in
patients with relapsed or refrac­tory dif­fuse large B-cell lym­phoma. Haematologica. 2021;106(9):2417-2426.
69. Zinzani PL, Rodgers T, Marino D, et al. RE-MIND: com­par­ing tafasitamab +
lenalidomide (L-MIND) with a real-world lenalidomide monotherapy cohort
in relapsed or refrac­tory dif­fuse large B-cell lym­phoma. Clin Cancer Res.
2021;27(22):6124-6134.
70. Cordoba R, Prawitz T, Westley T, et al. Tafasitamab plus lenalidomide ver­
sus 3 rituximab-based treat­
ments for non-trans­
plant eli­
gi­
ble relapsed/
refrac­
tory dif­
fuse large B-cell lym­
phoma: a matching-adjusted indi­
rect
com­par­i­son. Adv Ther. 2022;39(6):2668-2687.
71. Budde LE, Assouline S, Sehn LH, et al. Single-agent mosunetuzumab shows
dura­ble com­plete responses in patients with relapsed or refrac­tory B-cell
lym­pho­mas: phase I dose-esca­la­tion study. J Clin Oncol. 2022;40(5):481-491.
72. Hutchings M, Morschhauser F, Iacoboni G, et al. Glofitamab, a novel, biva­
lent CD20-targeting T-cell-engag­ing bispecific anti­body, induces dura­ble
com­plete remis­sions in relapsed or refrac­tory B-cell lym­phoma: a phase I
trial. J Clin Oncol. 2021;39(18):1959-1970.
73. Hutchings M, Mous R, Clausen MR, et al. Dose esca­la­tion of sub­cu­ta­ne­ous
epcoritamab in patients with relapsed or refrac­tory B-cell non-Hodgkin lym­
phoma: an open-label, phase 1/2 study. Lancet. 2021;398(10306):1157-1169.
74. Bannerji R, Arnason JE, Advani RH, et al. Odronextamab, a human
CD20xCD3 bispecific anti­
body in patients with CD20-pos­
i­
tive B-cell
malig­nan­cies (ELM-1): results from the relapsed or refrac­tory non-Hodgkin
lym­phoma cohort in a sin­gle-arm, multicentre, phase 1 trial. Lancet Haematol. 2022;9(5):e327-e339.
Dr Prakash Singh Shekhawat
Sequencing ther­apy in relapsed DLBCL | 153
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/146/2021680/146flowers.pdf by guest on 09 December 2022
35. Maurer MJ, Habermann TM, Shi Q , et al. Progression-free sur­vival at 24
months (PFS24) and sub­se­quent out­come for patients with dif­fuse large
B-cell lym­phoma (DLBCL) enrolled on ran­dom­ized clin­i­cal tri­als. Ann Oncol.
2018;29(8):1822-1827.
36. Maurer MJ, Jakobsen LH, Mwangi R, et al. Relapsed/Refractory International Prognostic Index (R/R-IPI): an inter­na­tional prog­nos­tic cal­cu­la­tor
for relapsed/refrac­
tory dif­
fuse large B-cell lym­
phoma. Am J Hematol.
2021;96(5):599-605.
37. Wang Y, Farooq U, Link BK, et al. Late relapses in patients with dif­fuse
large B-cell lym­phoma treated with immunochemotherapy. J Clin Oncol.
2019;37(21):1819-1827.
38. Robetorye RS, Ramsower CA, Rosenthal AC, et al. Incorporation of dig­i­tal
gene expres­sion pro­fil­ing for cell-of-ori­gin deter­mi­na­tion (Lymph2Cx test­
ing) into the rou­tine work-up of dif­fuse large B cell lym­phoma. J Hematop.
2019;12(1):3-10.
39. Scott DW. Cell-of-ori­
gin in dif­
fuse large B-cell lym­
phoma: are the
assays ready for the clinic? Am Soc Clin Oncol Educ Book. 2015;36(14
May):e458-466.
40. Harkins RA, Chang A, Patel SP, et al. Remaining chal­lenges in predicting
patient out­comes for dif­fuse large B-cell lym­phoma. Expert Rev Hematol.
2019;12(11):959-973.
41. Kurtz DM, Esfahani MS, Scherer F, et al. Dynamic risk pro­
fil­
ing using
serial tumor bio­mark­ers for per­son­al­ized out­come pre­dic­tion. Cell.
2019;178(3):699-713.e19713e19.
42. Philip T, Guglielmi C, Hagenbeek A, et al. Autologous bone mar­
row
trans­plan­ta­tion as com­pared with sal­vage che­mo­ther­apy in relapses
of che­mo­ther­apy-sen­si­tive non-Hodgkin’s lym­phoma. N Engl J Med.
1995;333(23):1540-1545.
43. Gisselbrecht C, Glass B, Mounier N, et al. Salvage reg­i­mens with autol­o­
gous trans­plan­ta­tion for relapsed large B-cell lym­phoma in the rituximab
era. J Clin Oncol. 2010;28(27):4184-4190.
44. Crump M, Kuruvilla J, Couban S, et al. Randomized com­par­i­son of gemcitabine, dexa­meth­a­sone, and cis­platin ver­sus dexa­meth­a­sone, cytarabine,
and cis­platin che­mo­ther­apy before autol­o­gous stem-cell trans­plan­ta­tion
for relapsed and refrac­tory aggres­sive lym­pho­mas: NCIC-CTG LY.12. J Clin
Oncol. 2014;32(31):3490-3496.
45. Abramson JS, Palomba ML, Gordon LI, et al. Lisocabtagene maraleucel for patients with relapsed or refrac­
tory large B-cell lym­
pho­
mas
(TRANSCEND NHL 001): a multicentre seam­
less design study. Lancet.
2020;396(10254):839-852.
46. Neelapu SS, Locke FL, Bartlett NL, et al. Axicabtagene ciloleucel CAR
T-cell ther­
apy in refrac­
tory large B-cell lym­
phoma. N Engl J Med.
2017;377(26):2531-2544.
47. Schuster SJ, Bishop MR, Tam CS, et al; JULIET Investigators. Tisagenlecleucel in adult relapsed or refrac­tory dif­fuse large B-cell lym­phoma. N Engl J
Med. 2019;380(1):45-56.
48. Schuster SJ, Tam CS, Borchmann P, et al. Long-term clin­i­cal out­comes of
tisagenlecleucel in patients with relapsed or refrac­tory aggres­sive B-cell
lym­pho­mas (JULIET): a multicentre, open-label, sin­gle-arm, phase 2 study.
Lancet Oncol. 2021;22(10):1403-1415.
49. Locke FL, Ghobadi A, Jacobson CA, et al. Long-term safety and activ­ity
of axicabtagene ciloleucel in refrac­tory large B-cell lym­phoma (ZUMA-1):
a sin­gle-arm, multicentre, phase 1-2 trial. Lancet Oncol. 2019;20(1):
31-42.
50. Pinnix CC, Gun­ther JR, Dabaja BS, et al. Bridging ther­apy prior to axicabtagene ciloleucel for relapsed/refrac­tory large B-cell lym­phoma. Blood Adv.
2020;4(13):2871-2883.
51. Jain MD, Jacobs MT, Nastoupil LJ, et al. Characteristics and out­
comes
of patients receiv­
ing bridg­
ing ther­
apy while awaiting man­
u­
fac­
ture of
­stan­dard of care axicabtagene ciloleucel CD19 chi­me­ric anti­gen recep­tor
(CAR) T-cell ther­apy for relapsed/refrac­tory large B-cell lym­phoma: results
from the US Lymphoma CAR-T Consortium. Blood. 2019;134(suppl 1):
245.
52. Schuster M, Zijlstra J, Casasnovas RO, et al. Effect of prior ther­apy and dis­
ease refrac­to­ri­ness on the effi­cacy and safety of oral selinexor in patients
with dif­fuse large B-cell lym­phoma (DLBCL): a post-hoc anal­y­sis of the
SADAL study. Clin Lymphoma Myeloma Leuk. 2022;22(7):483-494.
53. Sim AJ, Jain MD, Figura NB, et al. Radiation ther­apy as a bridg­ing strat­egy
for CAR T cell ther­apy with axicabtagene ciloleucel in dif­fuse large B-cell
lym­phoma. Int J Radiat Oncol Biol Phys. 2019;105(5):1012-1021.
54. Nastoupil LJ, Jain MD, Feng L, et al. Standard-of-care axicabtagene ciloleucel for relapsed or refrac­tory large B-cell lym­phoma: results from the US
Lymphoma CAR T Consortium. J Clin Oncol. 2020;38(27):3119-3128.
154 | Hematology 2022 | ASH Education Program
85. Loggers ET, Lee S, Chilson K, Back AL, Block S, Loberiza FR. Advance care
plan­
ning among hema­
to­
poi­
etic cell trans­
plant patients and bereaved
care­giv­ers. Bone Marrow Transplant. 2014;49(10):1317-1322.
86. Odejide OO, Cronin AM, Condron NB, et al. Barriers to qual­ity end-of-life
care for patients with blood can­cers. J Clin Oncol. 2016;34(26):31263132.
87. Odejide OO, Salas Coronado DY, Watts CD, Wright AA, Abel GA. End-of-life
care for blood can­cers: a series of focus groups with hema­to­logic oncol­o­
gists. J Oncol Pract. 2014;10(6):e396-e403.
88. Bernacki RE, Block SD; Amer­i­can College of Physicians High Value Care
Task Force. Communication about seri­ous ill­ness care goals: a review
and syn­the­sis of best prac­
tices. JAMA Intern Med. 2014;174(12):19942003.
89. Hudson KE, Wolf SP, Samsa GP, Kamal AH, Abernethy AP, LeBlanc TW. The
sur­prise ques­tion and iden­ti­fi­ca­tion of pal­li­a­tive care needs among hos­pi­
tal­ized patients with advanced hema­to­logic or solid malig­nan­cies. J Palliat
Med. 2018;21(6):789-795.
90. Childers JW, Back AL, Tulsky JA, Arnold RM. REMAP: a frame­work for goals
of care con­ver­sa­tions. J Oncol Pract. 2017;13(10):e844-e850.
91. VITALtalk. Address goals of care: smooth­ing dis­cus­sions about prog­no­sis
and treat­ment. Vol. 2020. https:​­/​­/www​­.vitaltalk​­.org​­/topics​­/reset​­-goals​­-of​
­-care​­/. Accessed 3 Octo­ber 2022.
92. Bernacki R, Paladino J, Neville BA, et al. Effect of the Serious Illness Care
Program in out­pa­tient oncol­ogy: a clus­ter ran­dom­ized clin­i­cal trial. JAMA
Intern Med. 2019;179(6):751-759.
93. Paladino J, Bernacki R, Neville BA, et al. Evaluating an inter­
ven­
tion to
improve com­mu­ni­ca­tion between oncol­ogy cli­ni­cians and patients with
life-lim­it­ing can­cer: a clus­ter ran­dom­ized clin­i­cal trial of the Serious Illness
Care Program. JAMA Oncol. 2019;5(6):801-809.
© 2022 by The Amer­i­can Society of Hematology
DOI 10.1182/hema­tol­ogy.2022000332
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/146/2021680/146flowers.pdf by guest on 09 December 2022
75. Patel K, Michot JM, Chanan-Khan A, et al. Safety and anti-tumor activ­ity of
plamotamab (XmAb13676), an anti-CD20 x anti-CD3 bispecific anti­body,
in sub­
jects with relapsed/refrac­
tory non-Hodgkin’s lym­
phoma. Blood.
2021;138(suppl 1):2494.
76. Dickinson M, Carlo-Stella C, Morschhauser F, et al. S220: glofitamab
induces dura­ble com­plete remis­sions and has favor­able safety in patients
with relapsed/refrac­tory dif­fuse large B-cell lym­phoma and ≥2 prior ther­
a­
pies: piv­
otal phase II expan­
sion results. HemaSphere. 2022;6(June):
121-122.
77. Thieblemont C, Phil­lips T, Ghesquieres H, et al. Primary results of sub­cu­ta­
ne­ous epcoritamab dose expan­sion in patients with relapsed or refrac­tory
large B-cell lym­phoma: a phase 2 study. Paper presented at: 2022 Euro­
pean Hematology Association Congress; 2022; Vienna, Austria. Abstract
LB2364.
78. Dans M, Smith T, Back A, et al. NCCN guide­lines insights: pal­li­a­tive care,
Version 2.2017. J Natl Compr Canc Netw. 2017;15(8):989-997.
79. Mack JW, Weeks JC, Wright AA, Block SD, Prigerson HG. End-of-life dis­
cus­sions, goal attain­ment, and dis­tress at the end of life: pre­dic­tors and
out­comes of receipt of care con­sis­tent with pref­er­ences. J Clin Oncol.
2010;28(7):1203-1208.
80. Sheridan R, Roman E, Smith AG, et al. Preferred and actual place of death in
haematological malig­nan­cies: a report from the UK Haematological Malignancy Research Network. BMJ Support Palliat Care. 2021;11(1):7-16.
81. Odejide OO, Uno H, Murillo A, Tulsky JA, Abel GA. Goals of care dis­cus­sions
for patients with blood can­cers: asso­ci­a­tion of per­son, place, and time
with end-of-life care uti­li­za­tion. Cancer. 2020;126(3):515-522.
82. Wright AA, Zhang B, Ray A, et al. Associations between end-of-life dis­
cus­sions, patient men­tal health, med­i­cal care near death, and care­giver
bereave­ment adjust­ment. JAMA. 2008;300(14):1665-1673.
83. Trevino KM, Rutherford SC, Marte C, et al. Illness under­
stand­
ing and
advance care plan­ning in patients with advanced lym­phoma. J Palliat Med.
2020;23(6):832-837.
84. Odejide OO, Cronin AM, Condron N, Earle CC, Wolfe J, Abel GA. Timeliness
of end-of-life dis­cus­sions for blood can­cers: a national sur­vey of hema­to­
logic oncol­o­gists. JAMA Intern Med. 2016;176(2):263-265.
CONTROVERSIES IN AGGRESSIVE NHL
Anita Kumar
Department of Medicine, Lymphoma Service, Memorial Sloan Kettering Cancer Center, New York, NY
Up-front autologous stem cell transplantation (ASCT) is the established standard of care for younger, transplant-eligible
MCL patients and is associated with a prolonged progression-free survival (PFS) benefit. However, there is no randomized
controlled trial data, with therapy including rituximab and cytarabine, that has established a PFS and overall survival (OS)
benefit with ASCT in the modern era. Multiple retrospective studies have failed to identify an OS benefit associated with
ASCT in younger MCL patients. The high-risk patient subgroup with evidence of baseline TP53 mutation has a dismal outcome with intensive chemoimmunotherapy followed by ASCT, thus up-front ASCT is not optimal for this patient subset.
Ongoing randomized clinical trials will help to clarify the role of up-front ASCT in the future. For example, the ongoing
European MCL Network Triangle study incorporating ibrutinib into chemoimmunotherapy induction and maintenance
with and without ASCT will help define the role of ASCT in the era of novel biologically targeted agents (ClinicalTrials.gov
identifier: NCT02858258). Additionally, minimal residual disease (MRD) assessment is a powerful prognostic tool in MCL,
and the ongoing Eastern Cooperative Oncology Group-American College of Radiology Imaging Network E4151 study is
comparing maintenance rituximab alone vs ASCT consolidation in MCL patients who achieve remission and MRD-undetectable status post induction (ClinicalTrials.gov identifier: NCT03267433). ASCT remains a highly efficacious initial therapy for younger MCL patients; however, ultimately the decision to pursue ASCT requires discussion of risks vs benefits,
incorporating patient preferences and values.
LEARNING OBJECTIVES
• Understand the data that support and challenge the use of ASCT in first remission in younger, fit MCL patients
• Understand that the presence of TP53 mutation is associated with a poor response to standard chemoimmunotherapy, including ASCT
• Review the use of novel therapies and MRD assessment in clinical trials that will help define the future role of ASCT
in the modern era
CLINICAL CASE 1
A 55-year-old otherwise healthy woman presented with
intermittent left lower quadrant abdominal pain. A computed tomographic scan of the abdomen and pelvis
revealed intussusception of the descending colon, and a
colonoscopy showed a 5-cm partially obstructing ulcerated mass in the descending colon. Biopsy of the colonic
mass was consistent with mantle cell lymphoma (MCL).
Neoplastic cells expressed by immunohistochemistry
(IHC) included CD20, CD5, BCL2, and cyclin D1 but did
not express TP53. Ki-67 was 25% to 40%. Mutational analysis revealed no TP53 mutation. A positron emission tomographic (PET) scan revealed diffuse fluorodeoxyglucose
(FDG)-avid lymphadenopathy and FDG avidity throughout
the gastrointestinal (GI) tract. The CBC and lactate dehydrogenase were normal, consistent with the low-risk clinical Mantle Cell Lymphoma International Prognostic Index
(MIPI) score. The patient underwent 4 cycles of RDHAX
chemotherapy (rituximab, dexamethasone, cytarabine,
and oxaliplatin) and achieved a complete remission (CR).
She then received ASCT consolidation followed by maintenance rituximab.
The standard of care frontline therapy for MCL patients
is chemoimmunotherapy. For younger MCL patients
(≤65 years) who are transplant eligible, a widely accepted
Dr Prakash Singh Shekhawat
Autotransplant in mantle cell lymphoma | 155
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/155/2021743/155kumar.pdf by guest on 09 December 2022
What is the role of up-front autologous stem cell
transplantation in mantle cell lymphoma?
156 | Hematology 2022 | ASH Education Program
12.7 and 8.5 years, respec­
tively.10,11 Although there was a
con­
tin­
u­
ous pat­
tern of relapse over time, for low-risk MIPI
patients the median PFS was over 12 years with the Nor­dic MCL2
treat­ment approach. This prolonged ini­tial dura­tion of response
likely has a greater impact on OS in youn­ger vs older MCL
patients given the fewer treat­ment-related toxicities and com­
pet­ing health risks. Given the suc­ces­sive short­en­ing of remis­
sion length with sub­se­quent lines of ther­apy in MCL, opti­miz­ing
ini­tial treat­ment is of value.17 Another rea­son to uti­lize ASCT in
front­line is that MCL can evolve over time, and the acqui­si­tion
of high-risk muta­tions, geno­mic insta­bil­ity, and/or trans­for­ma­
tion to blastoid or pleo­mor­phic dis­ease are increas­ingly com­
mon in the relapsed set­ting and often asso­ci­ated with dis­ease
refrac­
tory to stan­
dard chemoimmunotherapy.18 An extended
“treat­
ment-free” inter­
val after front­
line ther­
apy also affords
time for advances in the field to develop; for exam­ple, between
2013 and 2022, 6 novel ther­
a­
pies were US Food and Drug
Administration or National Comprehensive Cancer Network
approved for R/R MCL, includ­ing lenalidomide, ibrutinib, acalabrutinib, zanubrutnib, venetoclax, and brexucabtagene autoleucel.19-24 In sum­mary, front­line inten­sive induc­tion followed by
ASCT and rituximab main­te­nance for youn­ger, fit MCL patients
is a highly effi­ca­cious treat­ment approach and asso­ci­ated with
a prolonged PFS ben­e­fit.
CLINICAL CASE 2
A 64-year-old man with a his­tory of well-con­trolled hyper­ten­
sion and hyper­lip­id­emia presented with fatigue and drenching
night sweats. A PET scan revealed dif­fuse lymph­ade­nop­a­thy
and splenic involve­ment. Biopsy of a lymph node revealed MCL.
Neoplastic cells expressed CD20, CD5, BCL2, and cyclin D1 and
did not express TP53 by IHC. Ki-67 was 20%. Mutational anal­y­sis
revealed ATM muta­tion alone. Bone mar­row (BM) biopsy and
aspi­ra­tion showed 50% to 70% involve­ment with MCL. Labs
revealed mild ane­mia and throm­bo­cy­to­pe­nia. The MIPI score
was inter­me­di­ate risk. The patient expressed a strong desire to
receive out­pa­tient chemoimmunotherapy, balked at the idea
of tem­po­rar­ily relocating to New York City for an ASCT, and
wanted to main­tain his qual­ity of life to con­tinue work­ing and
car­ing for his elderly par­ent. The patient received treat­ment
with bendamustine and rituximab for 6 cycles followed by rituximab main­te­nance in his local com­mu­nity.
Although ASCT in CR1 established the stan­dard of care for youn­
ger MCL patients, it is not the only accept­able treat­ment. It is
becom­ing increas­ingly clear based on ret­ro­spec­tive and realworld data that only a frac­tion of youn­ger patients with MCL
receive up-front ASCT despite national and inter­na­tional guide­
lines supporting its use. Recently published data from the Flatiron Health data­
base describ­
ing treat­
ment pat­
terns for MCL
patients treated in pre­dom­i­nantly com­mu­nity-based prac­tices
from 2011 to 2021 showed that only approx­i­ma­tely 1 in 4 patients
under 65 years of age received an up-front ASCT.25 Data from the
Center for International Blood and Marrow Transplant Research
data­base and esti­mated MCL inci­dence rates from the National
Cancer Institute’s Surveillance, Epidemiology, and End Results
data­base between 2000 and 2017 showed that in patients aged
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/155/2021743/155kumar.pdf by guest on 09 December 2022
treat­ment approach con­sists of 3 phases: (1) cytarabine-containing
induc­
tion chemoimmunotherapy, (2) ASCT, and (3) main­
te­
nance rituximab. This treat­ment par­a­digm is supported by var­
i­
ous national and inter­
na­
tional guide­
lines and expert pan­
els,
includ­ing the National Comprehensive Cancer Network, Amer­
i­can Society of Transplantation and Cellular Therapy, Center for
International Blood and Marrow Transplant Research, Euro­pean
Society for Blood and Marrow Transplantation, Euro­pean Society
for Medical Oncology, and Euro­pean Hematology Association.1-4
The piv­otal study that established the role of consolidative
ASCT in first remis­sion for fit MCL patients was the open-label,
mul­ti­cen­ter, ran­dom­ized phase 3 trial performed by the Euro­
pean Mantle Cell Lymphoma Network enroll­ing patients between
Sep­tem­ber 1996 and March 2004.5 In this study, patients with
advanced-stage MCL aged 65 years or youn­ger were ran­dom­
ized to receive either ASCT or inter­feron alfa (IFNα) main­te­nance
ther­apy after cyclo­phos­pha­mide, doxo­ru­bi­cin, vin­cris­tine, and
pred­ni­sone (CHOP)-like induc­tion ther­apy with or with­out rituximab (ie, R-CHOP). In the ini­tial report with a median fol­low-up
of 25 months, patients in the ASCT arm vs the IFNα arm had a
sig­nif­i­cantly lon­ger pro­gres­sion-free sur­vival (PFS; median of 39
vs 17 months; P=.0108) but no sig­nif­i­cant dif­fer­ence in over­all sur­
vival (OS). Recently published long-term results with a 14-year
median fol­low-up con­tin­ued to dem­on­strate a PFS ben­e­fit and
also showed an OS ben­e­fit favor­ing ASCT con­sol­i­da­tion vs IFNα
main­te­nance.6 Building upon this approach, the treat­ment par­
a­digm for youn­ger, fit MCL patients has been fur­ther refined to
include rituximab and cytarabine in the ini­tial induc­tion reg­i­men
and to include rituximab main­te­nance post ASCT.7,8 The ther­a­
peu­tic impor­tance of high-dose cytarabine was established by
an open-label, phase 3 trial from the Euro­pean Mantle Cell Lymphoma Network in which MCL patients aged 65 years or youn­ger
were ran­dom­ized to receive R-CHOP vs alter­nat­ing R-CHOP and
R-DHAP (rituximab plus dexa­meth­a­sone, high-dose cytarabine,
and cis­platin) for 6 cycles followed by ASCT.7 After a median
fol­low-up of 6 years, the time to treat­ment fail­ure (TTF) was sig­
nif­i­cantly lon­ger in the R-CHOP/R-DHAP group vs the R-CHOPalone group (5-year TTF was 65% vs 40%, respec­tively), thus
supporting the inclu­sion of cytarabine as part of ini­tial induc­
tion chemoimmunotherapy. In the ini­tial report, there was no
OS dif­fer­ence between the R-DHAP-containing vs R-CHOP-alone
treat­ment arms, but with 5 addi­tional years of fol­low-up, when
adjusted for MIPI score with­out and with Ki-67, OS was sig­nif­
i­
cantly supe­
rior in the R-DHAP arm (haz­
ard ratio [HR], 0.74;
P = .038 and 0.60; P=.0066).9 The LYMA study established rituximab main­te­nance post ASCT as a stan­dard of care.8 In this study,
MCL patients aged 65 years or youn­ger received 4 courses of
R-DHAP (any plat­i­num) followed by ASCT con­sol­i­da­tion and then
were ran­dom­ized to receive rituximab main­te­nance vs obser­va­
tion post trans­plant. The PFS and OS at 4 years were supe­rior in
the rituximab main­te­nance vs obser­va­tion group (PFS, 83% vs
64%; P<.001; OS, 89% vs 80%; P=.04), respec­tively.
Multiple phase 2 stud­ies have dem­on­strated long-term dura­
ble remis­sions with cytarabine-containing induc­tion chemoimmunotherapy followed by ASCT con­sol­i­da­tion (Table 1).7,8,10-16
For exam­ple, with a median fol­low-up of 11.4 years, the sec­
ond Nor­dic MCL trial (MCL2) with patients under 66 years of
age who received alter­nat­ing courses of R-maxi-CHOP (dose-­
intensified rituximab, cyclophosphamide, doxorubicin, vincristine, and prednisone) and high-dose cytarabine (HIDAC) and
ASCT con­sol­i­da­tion was asso­ci­ated with a median OS and PFS of
Table 1. Selected pro­spec­tive stud­ies includ­ing ASCT in youn­ger MCL patients
Study/reg­i­men author
(ref)
Details of reg­i­men
Primary end point
Patients
Outcomes
Euro­pean Mantle Cell
Lymphoma Network
(MCL Younger): phase
3 study of R-CHOP vs
R-CHOP alter­nat­ing with
R-DHAP → ASCT
Hermine et al7
(2016)
Untreated MCL
Age 18-65y, stage II-IV,
trans­plant eli­gi­ble
Arm 1: R-CHOP×6 cycles
followed by HDT/ASCT
Arm 2: R-CHOP alter­nat­ing
with R-DHAP followed by
HDT/ASCT
Investigator-assessed
TTF
n=497
TTF sig­nif­i­cantly lon­ger in
cytarabine group: median
9.1y (95% CI, 6.3-not
reached) vs con­trol group:
median 3.9y
(3.2-4.4); HR, 0.56; P=.038.
5-year TTF 65% vs 40%.
No OS dif­fer­ence.
LYMA phase 3 study of
R-DHAP → ASCT →
rituximab main­te­nance
vs other
Le Gouill et al8
(2017)
Untreated MCL
Age >65y, stages II-IV,
trans­plant eli­gi­ble
Arm 1: R-DHAP (plat­i­num
deriv­a­tive)×4 cycles, if
CR/PR then HDT/ASCT
then rituximab
main­te­nance every
2 mo for 3y
*If <CR/PR, then R-CHOP
Arm 2: R-DHAP×4 cycles,
if CR/PR then HDT/ASCT
then obser­va­tion
EFS after ASCT
n=257
4-year EFS was 79% (95%
CI, 70-86) in the rituximab
group vs 61% (95% CI,
51-70) in the obser­va­tion
group (P=.001)
4-year OS was 89% (95%
CI, 81-94) in the rituximab
group vs 80% (95% CI,
72-88) in the obser­va­tion
group (P=.04).
Second Nor­dic MCL
study (MCL2): phase 2
study of R-maxi-CHOP
alter­nat­ing with RHIDAC → ASCT
Geisler et al10
(2008)
Eskelund et al11
(2016)
Untreated MCL
Age <66y, stages II-IV,
trans­plant eli­gi­ble
R-maxi-CHOP alter­nat­ing
with R-high-dose
cytarabine then HDT/ASCT
*Pts in CR who converted
from PCR-neg­a­tive to
PCR-pos­i­tive were offered
pre­emp­tive rituximab
375mg/m2/wk for 4
weeks to pre­vent clin­i­cal
relapse
OS, EFS, and PFS
n=160
6-year OS, EFS, and PFS
were 70%, 56%, and 66%,
respec­tively.
GELA phase 2 study:
R-CHOP×3 followed by
R-DHAP×3 → ASCT
Delarue et al12
(2013)
Untreated MCL
Stage III-IV, trans­plant
eli­gi­ble
Blastoid var­i­ants
excluded
CHOP×2, R-CHOP×1,
R-DHAP×3 → ASCT
Primary end point EFS
n=60
With median fol­low-up
of 67 mo, median EFS 83
mo and median OS not
reached. 5-y OS is 75%.
Phase 2 study of RB/RC
induc­tion followed by
ASCT
2 sites: DFCI and WUSTL
Merryman et al14
(2020)
DFCI: 18-69y,
untreated MCL,
trans­plant eli­gi­ble
WUSTL: age 18-65y,
untreated MCL,
trans­plant eli­gi­ble
DFCI: BR×3 cycles then
R-high-dose cytarabine×3
cycles → ASCT
WUSTL: BR alter­nat­ing
with R-HIDAC, total 6
cycles → ASCT
Primary end point
ORR
n=88
End-of-induc­tion ORR 97%.
After a median fol­low-up
of 33 mo, 3-y PFS 83% and
3-y OS 92%.
Phase 2 study of RHCVAD reg­i­men
alter­nat­ing with
high-dose meth­o­trex­ate
and cytarabine
Massaro et al16
(2021)
Chihara et al15 (2016)
Age <70y, untreated
MCL, trans­plant
eli­gi­ble
R-HCVAD reg­i­men
alter­nat­ing with highdose meth­o­trex­ate and
cytarabine×4 cycles. If CR
→ no fur­ther tx
PR → consolidative ASCT
Primary end point
com­plete response
rate
n=63
With median fol­low-up of
10.5 years, 10-year PFS and
OS of 35% and 61%
BR, bendamustine-rituximab; EFS, event-free sur­vival; HCT, high-dose che­mo­ther­apy; PR, par­tial response; Pts, patients; RB/RC, rituximab
bendamustine/rituximab cytarabine; tx, treatment.
65 years or youn­ger, ASCT uti­li­za­tion increased from 5.4% in 2001
to a peak of 22.3% in 2013 and plateaued at 18.1% in 2017.26 In
addi­tion, an anal­y­sis from the National Cancer Database esti­
mated ASCT uti­li­za­tion to be approx­i­ma­tely 30% for patients
aged 60 years or youn­ger.27 Rates of ASCT uti­li­za­tion for youn­ger
patients appear higher at US aca­demic med­i­cal cen­ters (67%)
and in cer­tain geog­ra­phies, like Sweden and Denmark.28-30
A num­ber of fac­tors may con­trib­ute to why ASCT uti­li­za­tion
rates are lower than antic­i­pated in youn­ger MCL patients. In the
Surveillance, Epidemiology, and End Results data­base study,
pri­vate health insur­ance, care at an aca­demic/research med­i­cal
cen­ter, and geo­graphic loca­tion were sig­nif­i­cantly asso­ci­ated
with ASCT uti­li­za­tion.27 A Swed­ish reg­is­try study found that MCL
patients who had never mar­ried, were divorced, or had a lower
edu­ca­tional level under­went trans­plan­ta­tion less often.29 In the
mul­ti­cen­ter ret­ro­spec­tive study that char­ac­ter­ized out­comes
and treat­ment pat­terns for youn­ger MCL patients treated at
US aca­demic med­i­cal cen­ters, the rea­sons for not selecting
ASCT con­sol­i­da­tion included phy­si­cian choice, 67%; patient
pref­er­ence, 18%; other rea­son (eg, mobi­li­za­tion fail­ure), 3%;
and unknown, 12%.28 Overall, fac­tors that influ­ence whether
up-front ASCT is pur­sued include patient and/or phy­si­cian pref­
er­ence, aca­demic vs com­mu­nity-based care, and socio­eco­
nomic fac­tors.
Dr Prakash Singh Shekhawat
Autotransplant in man­tle cell lym­phoma | 157
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/155/2021743/155kumar.pdf by guest on 09 December 2022
Key eli­gi­bil­ity cri­te­ria
Table 2. Comparing up-front inten­sive chemoimmunotherapy with ASCT con­sol­i­da­tion vs bendamustine-rituximab
Bendamustine-rituximab followed
by rituximab main­te­nance (2y)
Median pro­gres­sion-free sur­vival
7-10y
*Nordic MCL2 study11
4.4-5.3y
*E1411 and SHINE stud­ies49,50
Median over­all sur­vival
~10-13y
~10-13y
Key toxicities
Alopecia, fatigue, severe myelosuppression, infec­tion,
graft fail­ure (<1%), mucositis, GI tox­ic­ity, B-cell deple­tion,
car­dio­pul­mo­nary tox­ic­ity, liver tox­ic­ity, renal insuf­fi­ciency,
cog­ni­tive impair­ment, sec­ond­ary malig­nancy
Fatigue, myelosuppression, infec­tion, GI tox­ic­ity
(mild), B-cell deple­tion, sec­ond­ary malig­nancy
Quality-of-life con­sid­er­ations
*Hospitalization for trans­plant (or close prox­im­ity to
hos­pi­tal)
*Seek care at a trans­plant cen­ter
*Maintain per­for­mance sta­tus with­out
require­ment for hos­pi­tal­i­za­tion
*Can receive com­mu­nity-based care
One of the rea­sons why phy­si­cians may not rec­om­mend upfront ASCT in MCL is because of emerg­ing data that fail to con­
vinc­ingly dem­on­strate an asso­ci­ated OS ben­e­fit. The sur­vival
advan­
tage pre­
vi­
ously described with up-front ASCT may be
dimin­ished in the con­text of improved front­line induc­tion pro­
grams (includ­ing rituximab and cytarabine) with incor­po­ra­tion
of main­te­nance post induc­tion, as well as with improved sal­vage
ther­
a­
pies for relapsed MCL, such as Bruton’s tyro­
sine kinase
inhib­i­tors, CD19-directed chi­me­ric-anti­gen recep­tor T-cell ther­
apy, and other novel agents.
One crit­i­cal ques­tion that under­lies the con­tro­versy surrounding ASCT in MCL is: Would the same PFS and OS ben­e­fit of ASCT
reported in the only ran­dom­ized piv­otal study from the Euro­pean
Mantle Cell Lymphoma Network remain in the cur­rent era given
access to rituximab, cytarabine, and bet­ter sal­vage ther­a­pies? To
explore this ques­tion, a post hoc sub­group explor­atory anal­y­sis
was performed from the piv­otal study, focus­ing on patients who
received rituximab-containing induc­tion reg­i­mens. Neither PFS
nor OS were sig­nif­i­cantly dif­fer­ent between the ASCT vs IFNα
main­te­nance groups, with a median PFS of 3.4 years vs 1.7 years
(P=.087) and a median OS of 9.6 years vs 5.5 years (P=0.68),
respec­tively.6 However, this post hoc sub­group anal­y­sis included
a small num­ber of patients who received rituximab-containing
induc­tion ther­apy (41 patients who received ASCT con­sol­i­da­tion
and 27 patients who received IFNα main­te­nance) and thus was
under­pow­ered to detect a sig­nif­i­cant dif­fer­ence between the 2
approaches.6 Another impor­tant ret­ro­spec­tive study aiming to
assess the ben­e­fit of ASCT in the rituximab era was a mul­ti­cen­ter
anal­y­sis from sev­eral North Amer­i­can aca­demic med­i­cal cen­ters
that included 1029 patients aged 65 years or youn­ger.28 In this
study, 64% of patients (n=657) under­went ASCT con­sol­i­da­tion,
and after a median fol­low-up of 76 months, ASCT was asso­ci­
ated, using a mul­ti­var­i­able regres­sion anal­y­sis, with a sta­tis­ti­cally
sig­nif­i­cant improve­ment in PFS and a trend toward improved
OS. In a pro­pen­sity score-weighted anal­y­sis aimed to address
con­found­ers, the median PFS was more prolonged with ASCT
(78 vs 48.5 months), but an improve­ment in OS was not observed.
In addi­tion, sev­eral real-world ana­ly­ses also do not show a clear
sur­vival ben­e­fit to be asso­ci­ated with ASCT. For exam­ple, the
Flatiron Health data­base exam­ined ASCT-eli­gi­ble patients
(<65 years who were alive and did not ini­ti­ate sub­se­quent treat­
ment within 6 months of first-line treat­ment) and found no sig­nif­i­
cant improve­ment in real-world time-to-next-treat­ment or OS with
ASCT (vs with­out).25 In a real-world study of patients diag­nosed
158 | Hematology 2022 | ASH Education Program
in Sweden between 2007 and 2017, for patients under 70 years
of age there was no sig­nif­i­cant dif­fer­ence in OS with inten­si­
fied MCL2 pro­
to­
col or R-CHOP com­
pared to bendamustinerituximab.31
Additionally, it may be pos­si­ble to con­sider the defer­ral of
ASCT to the relapsed/refrac­tory (R/R) set­ting. There are data to
sug­gest that ASCT can be effec­tive in CR2, although it is unclear
whether out­comes are equiv­a­lent to up-front ASCT.32,33
In light of these data, for the patient in Clinical Case 2 I rec­
om­
mend discussing the out­
come data that sup­
port up-front
ASCT, the emerg­ing data that chal­lenge this ther­a­peu­tic strat­
egy in the mod­ern era, and the short- and long-term side effects
of an inten­sive approach vs other options. Ultimately, the oncol­
o­gist and the patient can col­lab­o­rate to devise a care plan based
on clin­i­cal evi­dence that bal­ances risks and expected out­comes
with patient pref­er­ences and val­ues (Table 2). For this patient,
the sig­nif­i­cant time com­mit­ment and need for hos­pi­tal­i­za­tion,
the cost of travel to an urban trans­plant cen­ter, and the poten­tial
toxicities asso­ci­ated with ASCT represented an undue bur­den
not off­set by the likely prolonged remis­sion dura­tion post trans­
plant, par­tic­u­larly given the poten­tial lack of OS ben­e­fit asso­ci­
ated with up-front ASCT in the mod­ern era.
CLINICAL CASE 3
A 40-year-old oth­er­wise healthy woman presented with wors­
en­ing left lower quad­rant pain requir­ing opi­oid ther­apy, and
a com­
puted tomo­
graphic scan of the abdo­
men and pel­
vis
revealed a bulky ret­ro­per­i­to­neal mass mea­sur­ing 12 cm, caus­
ing left-sided hydronephrosis. Biopsy of the mass revealed
MCL, blastoid var­i­ant. Neoplastic cells expressed CD20, CD5,
BCL2, and cyclin D1 by IHC. TP53 expres­sion by IHC was not
eval­u­ated. Ki-67 was 60% to 70%. Mutational anal­y­sis was sent
and pend­ing at the time of the ini­tial pre­sen­ta­tion. A PET scan
revealed dif­fuse FDG-avid lymph­ade­nop­a­thy, includ­ing a bulky
lymph node con­
glom­
er­
ate in the retroperitoneum, and BM
biopsy and aspi­ra­tion showed 10% to 15% involve­ment with
MCL. The MIPI score was high risk. The patient was admit­ted
for pain con­trol and the ini­ti­a­tion of RDHAX with tumor lysis
mon­i­tor­ing. She achieved a CR after 4 cycles. During ther­apy,
muta­tional pro­fil­ing results returned and revealed the pres­ence
of a mis­sense muta­tion in the TP53 gene (R249T). Although she
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/155/2021743/155kumar.pdf by guest on 09 December 2022
Cytarabine-containing induc­tion and ASCT con­sol­i­da­tion
followed by rituximab main­te­nance (3y)
had been sched­uled for ASCT con­sol­i­da­tion, this was aborted,
and the patient proceeded to receive rituximab main­te­nance.
CLINICAL CASE 1 (Con­t in­u ed)
We return again to the 55-year-old woman with MCL, Ki-67 25%
to 40%, TP53 wild-type stage IV MIPI low-risk dis­ease involv­ing
the GI tract who received 4 cycles of RDHAX and achieved a
CR. In this case, the patient enrolled in the ECOG-ACRIN EA4151
trial titled Rituximab With or Without Stem Cell Transplant in
Treating Patients With Minimal Residual Disease-Negative Mantle Cell Lymphoma in First Complete Remission (ClinicalTrials​
The lack of clear OS ben­e­fit asso­ci­ated with ASCT in MCL in the
mod­ern era has paved the way for the devel­op­ment of clin­i­cal
tri­als test­ing alter­nate ther­a­peu­tic strat­e­gies omit­ting ASCT.
These clin­i­cal tri­als have tested dif­fer­ent approaches includ­ing
the use of main­te­nance ther­apy post induc­tion, the incor­po­ra­
tion of novel agents, and the use of MRD assess­ment to develop
risk-adapted treat­ment par­a­digms. For exam­ple, the E1405 study
uti­lized a mod­i­fied rituximab, hyperfractionated cyclo­phos­pha­
mide, doxo­ru­bi­cin, vin­cris­tine, and dexa­meth­a­sone (R-HCVAD)
reg­i­men incor­po­rat­ing bortezomib with main­te­nance rituximab,
and the pri­mary end point of the study was over­all response rate
(ORR) to this com­bi­na­tion.41 The ORR was 95%, and a CR was
achieved in 68% of patients. Interestingly, in this study patients
could receive either main­te­nance rituximab or ASCT per phy­si­
cian and patient pref­er­ence, and with a median fol­low-up of 4.5
years, no sub­stan­tial dif­fer­ence in PFS or OS was found between
patients treated with main­
te­
nance rituximab (n=44) vs ASCT
(n=22). Although this was only hypoth­e­sis gen­er­at­ing given the
lack of a ran­dom­ized study design, short fol­low-up time, and
small sam­ple sizes, this study set the stage for future clin­i­cal
tri­als to more defin­i­tively com­pare main­te­nance rituximab and
ASCT post induc­tion in a larger ran­dom­ized con­text.
The prog­nos­tic role of MRD assess­ment, uti­liz­ing dif­fer­ent
assays includ­ing flow cytom­e­try, reverse-tran­scrip­tion poly­
mer­ase chain reac­tion (PCR)–based tech­niques, and immunosequencing plat­forms, has been well established in MCL.42
MRD sta­tus post induc­tion che­mo­ther­apy and pre-ASCT has
been shown to be an impor­tant pre­dic­tor of PFS and a poten­
tial bio­marker for risk-adapted treat­ments.43,44 Thus, the ongo­
ing EA4151 study is com­par­ing main­te­nance rituximab alone vs
ASCT con­sol­i­da­tion in MCL patients who achieve a remis­sion
and an MRD-unde­tect­able sta­tus post induc­tion. The pri­mary
end point of this study is 6-year OS. The study has the poten­
tial to estab­
lish an MRD-guided treat­
ment strat­
egy, allowing for ther­apy de-esca­la­tion in selected MRD-unde­tect­able
patients.
Other stud­
ies have incor­
po­
rated novel agents, such as
ibrutinib, into up-front reg­
i­
mens. The phase 2 WINDOW-1
study min­i­mized che­mo­ther­apy for pre­vi­ously untreated MCL
patients 65 years or youn­ger.45 Patients were ini­tially treated
with ibrutinib and rituximab, and after achiev­ing CR, patients
received R-HCVAD alter­nat­ing with meth­o­trex­ate/cytarabine
for 4 cycles. The ORR was 98% (n = 131 patients treated), the
3-year esti­mated PFS was 79%, and the 3-year OS was 95%.
Additionally, in the ongo­ing Euro­pean MCL Network Triangle
study, trans­plant-eli­gi­ble patients under 65 were ran­dom­ized
to 1 of 3 arms (rituximab main­te­nance can be added to all­
arms): (1) R-CHOP/R-DHAP induc­tion with ASCT, (2) R-CHOP
plus ibrutinib/R-DHAP with ASCT and then 2 years of ibrutinib
main­te­nance, and (3) R-CHOP plus ibrutinib/R-DHAP followed
by ibrutinib main­te­nance for 2 years with­out ASCT (ClinicalTrials​­.gov iden­ti­fier: NCT02858258).46 The pri­mary end point
is supe­ri­or­ity of 1 of the 3 study arms by inves­ti­ga­tor-assessed
fail­
ure-free sur­
vival. This study will deter­
mine the value of
Dr Prakash Singh Shekhawat
Autotransplant in man­tle cell lym­phoma | 159
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/155/2021743/155kumar.pdf by guest on 09 December 2022
The pres­ence of TP53 muta­tion is asso­ci­ated with a very poor
prog­no­sis when treated with con­ven­tional chemoimmunotherapy, includ­ing ASCT con­sol­i­da­tion. In a pooled anal­y­sis from the
Nor­dic MCL2 and MCL3 clin­i­cal tri­als, the prog­nos­tic value of
TP53 muta­tion at base­line was assessed in 183 patients (of the
total 320 enrolled) who had avail­­able DNA from diag­nos­tic BM
sam­ples. In the MCL2 and MCL3 stud­ies, patients under 66 years
of age with stage II to IV dis­ease who were trans­plant eli­gi­ble
received alter­nat­ing R-maxi-CHOP and R-high-dose-cytarabine
followed by ASCT.34 TP53 muta­tions were observed in 11%, del17p
in 16%, NOTCH1 muta­
tions in 4%, and CDKN2A muta­
tions in
20%. All these genetic sub­groups, in addi­tion to the MIPI, MIPI-c,
blastoid mor­phol­ogy, and Ki-67% over 30% groups, were each
asso­ci­ated with infe­rior sur­vival on univariable anal­y­sis. However, on mul­ti­var­i­able anal­y­sis, only the MIPI-c high-risk group
(HR, 2.6; P=.003) and TP53 muta­tion (HR, 6.2; P<.0001) retained
inde­pen­dent prog­nos­tic sig­nif­i­cance for OS. Patients with TP53
muta­tion had a per­cent­age of Ki-67 greater than 30%, blastoid
mor­phol­ogy, and a high-risk MIPI. The out­comes for patients
with TP53 muta­tion were dis­mal, with a median PFS of 0.9 years
and a median OS of 1.8 years. TP53 muta­tion with or with­out
del17p was asso­ci­ated with infe­rior out­comes com­pared to the
pres­ence of del17p alone. Increased TP53 expres­sion by IHC also
has prog­nos­tic value in MCL and is highly cor­re­lated with the
pres­ence of TP53 muta­tion.35-37 Given the min­i­mal ben­e­fit and
poten­tial tox­ic­ity of ASCT in this high-risk sub­set of patients,
ASCT con­sol­i­da­tion is not recommended in MCL patients with
evi­dence of TP53 muta­tion.
There exists no opti­
mal front­
line treat­
ment approach for
this high-risk sub­set of patients, and clin­i­cal trial enroll­ment is
highly recommended. Emerging treat­ment options for high-risk
patients include com­bi­na­tions of bio­log­i­cally targeted ther­
a­pies, such as dual inhi­bi­tion of Bruton’s tyro­sine kinase and
BCL2 in con­junc­tion with anti-CD20 mono­clo­nal anti­body ther­
apy, which has been, for exam­ple, inves­ti­gated in the OASIS and
BOVen stud­ies.38,39 Encouragingly, brexucabtagene autoleucel
(KTE-X19), an anti-CD19 autol­o­gous CAR-T cell US Food and Drug
Administration approved for R/R MCL, has shown prom­
is­
ing
effi­cacy even among high-risk MCL patients with ele­vated pro­
lif­er­a­tive indi­ces, blastic or pleo­mor­phic mor­phol­ogy, and TP53
muta­tion.24,40 In the future, novel treat­ment options, includ­ing
anti-CD19 chi­me­ric-anti­gen recep­tor T-cell ther­apy or CD20/CD3
bispecific anti­body ther­apy, will likely be incor­po­rated into clin­i­
cal tri­als in ear­lier lines, par­tic­u­larly for high-risk patients.
.­gov iden­ti­fier: NCT03267433). At the end of induc­tion che­mo­
ther­apy, she was min­i­mal resid­ual dis­ease (MRD) unde­tect­able
by an immunosequencing assay (clonoSEQ®) and was ran­dom­
ized to receive rituximab main­te­nance alone (ASCT omit­ted).
ibrutinib dur­ing induc­tion and main­te­nance and deter­mine if
ASCT can be omit­ted with the addi­tion of ibrutinib.
advi­
sory board: Celgene, Genentech, Kite Pharmaceuticals,
Loxo Oncology/Lily, Astra Zeneca.
Conclusion
Off-label drug use
All 4 clin­i­cal cases presented high­light the evolv­ing role of upfrontASCTin MCL, and I have sum­ma­rized my approach in Figure 1.
While up-front ASCT remains a well-established and highly
effec­tive ini­tial ther­a­peu­tic strat­egy for youn­ger MCL
patients, there are sub­
groups of patients in whom this
approach is not warranted, such as patients with TP53
muta­tion. Although ASCT con­sol­i­da­tion has been uti­lized
in patients older than 65 years who are fit with­out sig­nif­
i­
cant comorbidity, I do not sup­
port this approach since
the grand major­
ity of clin­
i­
cal trial data supporting upfront ASCT in MCL have focused on youn­ger patients, and
the already uncer­tain OS ben­e­fit of ASCT is even less well
established in older patients who have an increased risk for
trans­plant-asso­ci­ated toxicities and com­pet­ing risks for sur­
vival with advanced age. 47,48 Ultimately, after dis­cus­sion of
ben­e­fits vs risks, it is crit­i­cal to develop indi­vid­u­al­ized treat­
ment plans that incor­po­rate patient pref­er­ences and val­ues.
Key clin­i­cal tri­als, includ­ing EA4151 and the Euro­pean Triangle study, will help to fur­ther clar­ify the role of ASCT in the
future.
Conflict-of-inter­est dis­clo­sure
Anita Kumar: research funding: Abbvie Pharmaceuticals, Adaptive Biotechnologies, Celgene, Pharmacyclics, Seattle Genetics,
Astra Zeneca, Loxo Oncology/Lily; hon­o­raria: Celgene, Genentech, Kite Pharmaceuticals, Loxo Oncology/Lily, Astra Zeneca;
160 | Hematology 2022 | ASH Education Program
Anita Kumar: nothing to disclose.
Correspondence
Anita Kumar, Department of Medicine, Lymphoma Service,
­Memorial Sloan Kettering Cancer Center, 1275 York Ave, Box 468,
New York, NY 10065; e-mail: kumara2@mskcc​­.org.
References
1.
2.
3.
4.
5.
6.
Zelenetz AD, Gordon LI, Chang JE, et al. NCCN Guidelines® insights: B-cell
lym­pho­mas, Version 5.2021. J Natl Compr Canc Netw. 2021;19(11):12181230.
Munshi PN, Hamadani M, Kumar A, et al. ASTCT, CIBMTR, and EBMT clin­i­cal
prac­tice rec­om­men­da­tions for trans­plant and cel­lu­lar ther­a­pies in man­tle
cell lym­phoma. Bone Marrow Transplant. 2021;56(12):2911-2921.
Dreyling M, Campo E, Hermine O, et al. Newly diag­nosed and relapsed
man­tle cell lym­phoma: ESMO Clinical Practice Guidelines for diag­no­sis,
treat­ment and fol­low-up. Ann Oncol. 2017;28(suppl 4):iv62-iv71.
Linton K, Dreyling M. EHA endorse­ment of ESMO Clinical Practice Guidelines
for Newly Diagnosed and Relapsed Mantle Cell Lymphoma. Hemasphere.
2020;4(5):e464.
Dreyling M, Lenz G, Hoster E, et al. Early con­sol­i­da­tion by myeloablative
radiochemotherapy followed by autol­o­gous stem cell trans­plan­ta­tion in
first remis­sion sig­nif­i­cantly pro­longs pro­gres­sion-free sur­vival in man­tlecell lym­phoma: results of a pro­spec­tive ran­dom­ized trial of the Euro­pean
MCL Network. Blood. 2005;105(7):2677-2684.
Zoellner AK, Unterhalt M, Stilgenbauer S, et al; Euro­pean Mantle Cell Lymphoma Network. Long-term sur­vival of patients with man­tle cell lym­phoma
after autol­o­gous haematopoietic stem-cell trans­plan­ta­tion in first remis­
sion: a post-hoc anal­y­sis of an open-label, multicentre, randomised, phase
3 trial. Lancet Haematol. 2021;8(9):e648-e657.
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/155/2021743/155kumar.pdf by guest on 09 December 2022
Figure 1. Treatment algorithm for untreated MCL patients.
7.
26. Riedell PA, Hamadani M, Ahn KW, et al. Outcomes and uti­li­za­tion trends
of front-line autol­o­gous hema­to­poi­etic cell trans­plan­ta­tion for man­tle cell
lym­phoma. Transplant Cell Ther. 2021;27(11):911.e1-911911.e7.
27. Sawalha Y, Radivoyevitch T, Jia X, et al. The impact of socio­eco­nomic disparities on the use of upfront autol­o­gous stem cell trans­plan­ta­tion for man­
tle cell lym­phoma. Leuk Lymphoma. 2022;63(2):335-343.
28. Gerson JN, Handorf E, Villa D, et al. Survival out­comes of youn­ger patients
with man­tle cell lym­phoma treated in the rituximab era. J Clin Oncol.
2019;37(6):471-480.
29. Glimelius I, Smedby KE, Albertsson-Lindblad A, et al. Unmarried or lessedu­cated patients with man­tle cell lym­phoma are less likely to undergo a
trans­plant, lead­ing to lower sur­vival. Blood Adv. 2021;5(6):1638-1647.
30. Abrahamsson A, Albertsson-Lindblad A, Brown PN, et al. Real world data
on pri­
mary treat­
ment for man­
tle cell lym­
phoma: a Nor­
dic Lymphoma
Group obser­va­tional study. Blood. 2014;124(8):1288-1295.
31. Albertsson-Lindblad A, Palsdottir T, Smedby KE, Weibull CE, Glimelius I,
Jerkeman M. Survival in man­tle cell lym­phoma after front­line treat­ment
with R-bendamustine, R-CHOP and the Nor­
dic MCL2 reg­
i­
men—a real
world study on patients diag­nosed in Sweden 2007-2017. Haematologica.
2022;107(3):740-743.
32. Cassaday RD, Guthrie KA, Budde EL, et al. Specific fea­
tures iden­
tify
patients with relapsed or refrac­
tory man­
tle cell lym­
phoma benefit­
ting
from autol­o­gous hema­to­poi­etic cell trans­plan­ta­tion. Biol Blood Marrow
Transplant. 2013;19(9):1403-1406.
33. Fenske TS, Zhang MJ, Carreras J, et al. Autologous or reduced-inten­sity
con­di­tion­ing allo­ge­neic hema­to­poi­etic cell trans­plan­ta­tion for che­mo­ther­
apy-sen­si­tive man­tle-cell lym­phoma: anal­y­sis of trans­plan­ta­tion tim­ing and
modal­ity. J Clin Oncol. 2014;32(4):273-281.
34. Eskelund CW, Dahl C, Hansen JW, et al. TP53 muta­tions iden­tify youn­ger
man­tle cell lym­phoma patients who do not ben­e­fit from inten­sive chemoimmunotherapy. Blood. 2017;130(17):1903-1910.
35. Aukema SM, Hoster E, Rosenwald A, et al. Expression of TP53 is asso­ci­ated
with the out­come of MCL inde­pen­dent of MIPI and Ki-67 in tri­als of the
Euro­pean MCL Network. Blood. 2018;131(4):417-420.
36. Rodrigues JM, Hassan M, Freiburghaus C, et al. p53 is asso­ci­ated with highrisk and pin­points TP53 mis­sense muta­tions in man­tle cell lym­phoma. Br J
Haematol. 2020;191(5):796-805.
37. Ferrero S, Rossi D, Rinaldi A, et al. KMT2D muta­tions and TP53 dis­rup­tions
are poor prog­nos­tic bio­mark­ers in man­tle cell lym­phoma receiv­ing highdose ther­apy: a FIL study. Haematologica. 2020;105(6):1604-1612.
38. Le Gouill S, Morschhauser F, Chiron D, et al. Ibrutinib, obinutuzumab, and
venetoclax in relapsed and untreated patients with man­tle cell lym­phoma:
a phase 1/2 trial. Blood. 2021;137(7):877-887.
39. Kumar A, Soumerai JD, Abramson JS, et al. Preliminary safety and effi­cacy
from a mul­ti­cen­ter, inves­ti­ga­tor-ini­ti­ated phase II study in untreated TP53
mutant man­tle cell lym­phoma with zanubrutinib, obinutuzumab, and venetoclax (BOVen). Blood. 2021;138(suppl 1):3540.
40. Wang M, Munoz J, Goy A, et al. Three-year fol­low-up of KTE-X19 in patients
with relapsed/refrac­tory man­tle cell lym­phoma, includ­ing high-risk sub­
groups, in the ZUMA-2 study [published online 4 June 2022]. J Clin Oncol.
2022;JCO2102370.
41. Chang JE, Li H, Smith MR, et al. Phase 2 study of VcR-CVAD with main­te­
nance rituximab for untreated man­tle cell lym­phoma: an Eastern Cooperative Oncology Group study (E1405). Blood. 2014;123(11):1665-1673.
42. Ladetto M, Tavarozzi R, Pott C. Minimal resid­ual dis­ease in man­tle cell lym­
phoma: meth­ods and clin­i­cal sig­nif­i­cance. Hematol Oncol Clin North Am.
2020;34(5):887-901.
43. Pott C, Hoster E, Delfau-Larue MH, et al. Molecular remis­sion is an inde­pen­
dent pre­dic­tor of clin­i­cal out­come in patients with man­tle cell lym­phoma
after com­bined immunochemotherapy: a Euro­pean MCL inter­group study.
Blood. 2010;115(16):3215-3223.
44. Liu H, Johnson JL, Koval G, et al. Detection of min­i­mal resid­ual dis­ease fol­
low­ing induc­tion immunochemotherapy pre­dicts pro­gres­sion free sur­vival
in man­tle cell lym­phoma: final results of CALGB 59909. Haematologica.
2012;97(4):579-585.
45. Wang ML, Jain P, Zhao S, et al; Mantle Cell Research Group. Ibrutinibrituximab followed by R-HCVAD as front­line treat­ment for young patients
(≤65 years) with man­tle cell lym­phoma (WINDOW-1): a sin­gle-arm, phase 2
trial. Lancet Oncol. 2022;23(3):406-415.
46. Dreyling M, Ladetto M, Doorduijn JK, et al. Triangle: autol­o­gous trans­plan­
ta­tion after a rituximab/ibrutinib/ara-c containing induc­tion in gen­er­al­ized
Dr Prakash Singh Shekhawat
Autotransplant in man­tle cell lym­phoma | 161
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/155/2021743/155kumar.pdf by guest on 09 December 2022
Hermine O, Hoster E, Walewski J, et al; Euro­pean Mantle Cell Lymphoma
Network. Addition of high-dose cytarabine to immunochemotherapy
before autol­
o­
gous stem-cell trans­
plan­
ta­
tion in patients aged 65 years
or youn­
ger with man­
tle cell lym­
phoma (MCL Younger): a randomised,
open-label, phase 3 trial of the Euro­pean Mantle Cell Lymphoma Network.
Lancet. 2016;388(10044):565-575.
8. Le Gouill S, Thieblemont C, Oberic L, et al; LYSA Group. Rituximab after
autol­o­gous stem-cell trans­plan­ta­tion in man­tle-cell lym­phoma. N Engl J
Med. 2017;377(13):1250-1260.
9. Hermine O, Jiang L, Walewski J, et al. Addition of high-dose cytarabine
to immunochemotherapy before autol­o­gous stem-cell trans­plan­ta­tion
in patients aged 65 years or youn­ger with man­tle cell lym­phoma (MCL
Younger): a long-term fol­low-up of the ran­dom­ized, open-label, phase
3 trial of the Euro­pean Mantle Cell Lymphoma Network. Blood. 2021;
138(suppl 1):380.
10. Geisler CH, Kolstad A, Laurell A, et al; Nor­dic Lymphoma Group. Long-term
pro­gres­sion-free sur­vival of man­tle cell lym­phoma after inten­sive frontline immunochemotherapy with in vivo-purged stem cell res­cue: a nonrandomized phase 2 mul­ti­cen­ter study by the Nor­dic Lymphoma Group.
Blood. 2008;112(7):2687-2693.
11. Eskelund CW, Kolstad A, Jerkeman M, et al. 15-year fol­low-up of the Second
Nor­dic Mantle Cell Lymphoma trial (MCL2): prolonged remis­sions with­out
sur­vival pla­teau. Br J Haematol. 2016;175(3):410-418.
12. Delarue R, Haioun C, Ribrag V, et al; Groupe d’Etude des Lymphomes
de l’Adulte (GELA). CHOP and DHAP plus rituximab followed by autol­o­
gous stem cell trans­plan­ta­tion in man­tle cell lym­phoma: a phase 2 study
from the Groupe d’Etude des Lymphomes de l’Adulte. Blood. 2013;121(1):
48-53.
13. Romaguera JE, Fayad LE, Feng L, et al. Ten-year fol­low-up after intense chemoimmunotherapy with rituximab-HyperCVAD alter­
nat­
ing with rituximabhigh dose meth­o­trex­ate/cytarabine (R-MA) and with­out stem cell trans­
plan­ta­tion in patients with untreated aggres­sive man­tle cell lym­phoma.
Br J Haematol. 2010;150(2):200-208.
14. Merryman RW, Edwin N, Redd R, et al. Rituximab/bendamustine and rituximab/cytarabine induc­tion ther­apy for trans­plant-eli­gi­ble man­tle cell lym­
phoma. Blood Adv. 2020;4(5):858-867.
15. Chihara D, Cheah CY, Westin JR, et al. Rituximab plus hyper-CVAD alter­
nat­ing with MTX/Ara-C in patients with newly diag­nosed man­tle cell lym­
phoma: 15-year fol­low-up of a phase II study from the MD Anderson Cancer
Center. Br J Haematol. 2016;172(1):80-88.
16. Massaro F, Stepanishyna Y, Manni M, et al. Long-term results of the MCL01
phase II trial of rituximab plus HyperCVAD alter­
nat­
ing with high-dose
cytarabine and meth­o­trex­ate for the ini­tial treat­ment of patients with man­
tle cell lym­phoma. Br J Haematol. 2021;192(6):1011-1014.
17. Kumar A, Sha F, Toure A, et al. Patterns of sur­vival in patients with recur­rent
man­tle cell lym­phoma in the mod­ern era: pro­gres­sive short­en­ing in response
dura­tion and sur­vival after each relapse. Blood Cancer J. 2019;9(6):50.
18. Jain P, Wang M. Blastoid man­tle cell lym­phoma. Hematol Oncol Clin North
Am. 2020;34(5):941-956.
19. Goy A, Sinha R, Williams ME, et al. Single-agent lenalidomide in patients
with man­tle-cell lym­phoma who relapsed or progressed after or were
refrac­tory to bortezomib: phase II MCL-001 (EMERGE) study. J Clin Oncol.
2013;31(29):3688-3695.
20. Wang ML, Rule S, Martin P, et al. Targeting BTK with ibrutinib in relapsed or
refrac­tory man­tle-cell lym­phoma. N Engl J Med. 2013;369(6):507-516.
21. Wang M, Rule S, Zinzani PL, et al. Acalabrutinib in relapsed or refrac­tory
man­tle cell lym­phoma (ACE-LY-004): a sin­gle-arm, multicentre, phase 2
trial. Lancet. 2018;391(10121):659-667.
22. Tam CS, Opat S, Simpson D, et al. Zanubrutinib for the treat­
ment of
relapsed or refrac­tory man­tle cell lym­phoma. Blood Adv. 2021;5(12):25772585.
23. Davids MS, Roberts AW, Seymour JF, et al. Phase I first-in-human study of
venetoclax in patients with relapsed or refrac­tory non-Hodgkin lym­phoma.
J Clin Oncol. 2017;35(8):826-833.
24. Wang M, Munoz J, Goy A, et al. KTE-X19 CAR T-cell ther­apy in relapsed
or refrac­tory man­tle-cell lym­phoma. N Engl J Med. 2020;382(14):13311342.
25. Martin P, Cohen JB, Wang M, et al. Treatment out­
comes and roles of
trans­plan­ta­tion and main­te­nance rituximab in patients with pre­vi­ously
untreated man­tle cell lym­phoma: results from large real-world cohorts
[published online 28 June 2022]. J Clin Oncol. 2022;JCO2102698.
man­tle cell lym­phoma—a ran­dom­ized Euro­pean MCL Network trial. Blood.
2019;134(suppl 1):2816.
47. Dahi PB, Tamari R, Devlin SM, et al. Favorable out­comes in elderly patients
under­go­ing high-dose ther­apy and autol­o­gous stem cell trans­plan­ta­tion for
non-Hodgkin lym­phoma. Biol Blood Marrow Transplant. 2014;20(12):20042009.
48. Dahi PB, Lee J, Devlin SM, et al. Toxicities of high-dose che­mo­ther­apy and
autol­o­gous hema­to­poi­etic cell trans­plan­ta­tion in older patients with lym­
phoma. Blood Adv. 2021;5(12):2608-2618.
49. Smith MR, Jegede O, Martin P, et al. ECOG-ACRIN E1411 ran­dom­ized phase
2 trial of bendamustine-rituximab (BR)-based induc­tion followed by ritux-
imab (R)±lenalidomide (L) con­sol­i­da­tion for Mantle cell lym­phoma: effect
of adding bortezomib to front-line BR induc­tion on PFS. JCO. 2021;39(suppl
15):7503.
50. Wang ML, Jurczak W, Jerkeman M, et al; SHINE Investigators. Ibrutinib plus
bendamustine and rituximab in untreated man­tle-cell lym­phoma. N Engl J
Med. 2022;386(26):2482-2494.
© 2022 by The Amer­i­can Society of Hematology
DOI 10.1182/hema­tol­ogy.2022000333
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/155/2021743/155kumar.pdf by guest on 09 December 2022
162 | Hematology 2022 | ASH Education Program
Dr Prakash Singh Shekhawat
IMMUNOTHERAPY IN MULTIPLE MYELOMA
Antibodies and bispecifics for multiple myeloma:
effective effector therapy
Department of Medical Oncology and Hematology, Princess Margaret Cancer Centre, Toronto, Canada
The therapeutic landscape in multiple myeloma (MM) has changed dramatically over the last 2 decades. With the introduction of novel immunotherapies, patients with MM can expect deeper responses, longer remissions, and improved
overall survival. Since its approval by the US Food and Drug Administration in 2015, the monoclonal antibody specific for
CD38, daratumumab, has been incorporated into both frontline and relapsed treatment regimens. Its role as a maintenance therapy is currently being explored. Subsequently, a variety of novel antibody therapeutics have evolved from the
success of daratumumab, using similar concepts to target the malignant plasma cell clone. Noteworthy naked monoclonal antibodies include isatuximab, another agent directed against CD38, and elotuzumab, an agent directed against
SLAM family member 7. Antibody-drug conjugates, complex molecules composed of an antibody tethered to a cytotoxic
drug, target malignant cells and deliver a lethal payload. The first to market is belantamab mafodotin, which targets
B-cell maturation antigen (BCMA) on malignant plasma cells and delivers a potent microtubule inhibitor, monomethyl
auristatin F. Additionally, bispecific T-cell antibodies are in development that engage the immune system directly by
simultaneously binding CD3 on T cells and a target epitope—such as BCMA, G-protein coupled receptor family C group
5 member D (GPRC5d), and Fc receptor homologue 5 (FcRH5)—on malignant cells. Currently, teclistamab, an anti-BCMA
bispecific, is closest to approval for commercial use. In this review, we explore the evolving landscape of antibodies in
the treatment of MM, including their role in frontline and relapse settings.
LEARNING OBJECTIVES
• Review the use of MoAbs in NDMM (transplant eligible and ineligible) and RRMM
• Understand the evolving role of ADCs in RRMM
• Explore the current landscape and future directions of BsAbs in RRMM
Introduction
CLINICAL CASE
A 54­year­old woman is diagnosed with immunoglobulin
G kappa (IgGκ) multiple myeloma (MM). She undergoes
induction therapy with triplet combination, followed by
autologous stem cell transplant (ASCT) and lenalidomide
maintenance. She experiences several relapses and is
cycled through various lines of therapy. She has now had
4 lines of therapy and is considered triple­class refractory
(TCR), having become refractory to a proteosome inhibi­
tor, an immunomodulatory drug, and an anti­CD38 mono­
clonal antibody (MoAb). Her treating physician offers to
send her to an academic center for a clinical trial, but she
opts to stay close to home and is started on belantamab
mafodotin.
Despite decades of scientific advancement, MM remains
incurable in the vast majority of patients.1 Currently, the pri­
mary goal of treatment is to increase survival and maintain
a reasonable quality of life. This is accomplished through
therapeutic suppression of the malignant plasma cell (PC)
clone, which subsequently mitigates disease­related com­
plications.1
Since their approval, MoAbs have been widely incor­
porated into both frontline and relapsed settings. Still,
patients who become TCR remain challenging to treat and
require therapies with novel mechanisms of action. In the
recent LocoMMotion study, TCR MM patients who prospec­
tively received widely available standard of care (SOC) ther­
apies demonstrated an overall response rate (ORR) of 25%,
a median duration of response (DOR) of 4.5 months, and
Dr Prakash Singh Shekhawat
MRD testing in AML 2022 | 163
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/163/2021838/163cipkar.pdf by guest on 09 December 2022
Christopher Cipkar, Christine Chen, and Suzanne Trudel
Monoclonal antibodies
In the last decade, three MoAbs were approved for the treat­
ment of MM. Daratumumab (anti-CD38) and elotuzumab (antiSLAM fam­ily mem­ber 7 [SLAMF7]) were approved for com­mer­cial
use in 2015 while isatuximab (anti-CD38) was approved in 2020.
Daratumumab is a human­ized IgG1κ MoAb that binds to CD38
expressed on malig­
nant PCs.4 It effec­tively elim­i­na­tes CD38-­
expressing PCs through sev­eral mech­a­nisms: anti­body-­depen­dent
T-cel­lu­lar cyto­tox­ic­ity, anti­body-depen­dent T-­cel­lu­lar phago­
cy­to­sis, com­ple­ment-depen­dent cyto­tox­ic­ity, induc­tion of apo­p­
to­sis via Fcγ recep­tor–medi­ated cross-linking, and by var­i­ous
­immu­no­mod­u­la­tory effects.4 Several ran­dom­ized con­trolled tri­
als have observed the clin­i­cal ben­e­fits of adding daratumamab
to immunomodulator drugs and proteosome inhib­i­tors in RRMM,
establishing its use as an SOC agent in this set­ting (Table 1).5-8
In trans­plant-eli­gi­ble patients, the addi­tion of daratumamab to
bortezomib-tha­lid­o­mide-dexa­meth­a­sone (phase 3 CASSIOPEIA
trial) improved depth of response and led to a sig­nif­i­cant pro­gres­
sion-free sur­vival (PFS) ben­e­fit.9 Similarly, the addi­tion of daratu­
mumab to lenalidomide-bortezomib-dexa­meth­a­sone (RVd; phase
2 GRIFFIN trial) in trans­plant-eli­gi­ble patients improved depth of
response, includ­ing strin­gent com­plete remis­sion and min­i­mal
resid­ual dis­ease (MRD) (10−5), and showed PFS benefit (HR 0.45;
p = 0.324), while median OS was not reached in either group with
limited follow-up.10 In both CASSIOPEIA and GRIFFIN, there was
a ben­e­fit to adding daratumumab across all­sub­groups except
those with high-risk cyto­ge­net­ics or International Scoring System
stage III dis­ease.
Table 1. Pivotal ran­dom­ized phase 3 tri­als of mono­clo­nal anti­body treat­ment for MM
Arm (n=num­ber
of patients)
Overall response (%)
Median pro­gres­sion-free
sur­vival (months)
Hazard ratio (P value)
Transplant-inel­i­gi­ble
NDMM
D-VMP (n=350)
91
36.4
0.42 (P<.001)
VMP (n=356)
74
19.3
MAIA11
Transplant-inel­i­gi­ble
NDMM
D-Rd (n=368)
92.9
NR
Rd (n=369)
81.6
34.4
ELOQUENT 124
Transplant-inel­i­gi­ble
NDMM
E-Rd (n=374)
83
31.4
Rd (n=374)
79
29.5
CASSIOPEIA9
Transplant-eli­gi­ble
NDMM
D-VTd (n=543)
93
NR
VTd (n=542)
90
NR
GRIFFIN10,a
Transplant-eli­gi­ble
NDMM
D-RVd (n=99)
99
NR
RVd (n=97)
92
NR
GMMG-HD623
Transplant-eli­gi­ble
NDMM
E-RVd (n=555)
B1 81.5%; B2 80.7% (≥VGPR)
B1 66.2%; B2 67.2% (3y PFS)
RVd (n=559)
A1 78.9%; A2 78.2% (≥VGPR)
A1 68.8%; A2 68.5% (3y PFS)
GMMG-HD719
Transplant-eli­gi­ble
NDMM
I-RVd (n=329)
MRD 50.1%
NR
RVd (n=331)
MRD 35.6%
NR
POLLUX6
RRMM, at least 1 prior
line
D-Rd (n=286)
93
44.5
Rd (n=283)
76
17.5
ELOQUENT 220
RRMM, 1-3 prior lines
E-Rd (n=321)
79
19.4
Rd (n=325)
66
14.9
CANDOR8
RRMM, 1-3 prior lines
D-Kd (n=312)
84
NR
Kd (n=154)
75
15.8
IKEMA18
RRMM, 1-3 prior lines
I-Kd (n=179)
87
NR
Kd (n=123)
83
19.2
CASTOR5
RRMM, at least 1 prior
line
D-Vd (n=251)
83
16.7
Vd (n=247)
63
7.1
Clinical trial
Patient pop­u­la­tion
ALCYONE12
164 | Hematology 2022 | ASH Education Program
Dr Prakash Singh Shekhawat
0.53 (P<.001)
0.93 (P =.44)
0.47 (P<.001)
Not avail­­able
Not avail­­able (P =.86;
no sig­nif­i­cant dif­fer­ence)
Not avail­­able (OR=1.82;
P<.001)
0.44 (P<.001)
0.71 (P<.001)
0.62 (P =.003)
0.53 (P<.001)
0.31 (P<.001)
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/163/2021838/163cipkar.pdf by guest on 09 December 2022
over­all sur­vival (OS) of 11.1 months.2 To address this area of unmet
need, the focus has shifted toward new approaches, such as chi­
me­ric anti­gen recep­tor (CAR) T-cell ther­a­pies, anti­body-drug
con­ju­gates (ADCs), and bispecific antibodies (BsAbs).3 Although
CAR T-cell ther­
apy has shown remark­
able over­
all effi­
cacy of
greater than 70% across var­i­ous early-phase stud­ies—lead­ing
to the approv­als of idecabtagene vicleucel and ciltacabtagene
autoleucel—delays in drug admin­is­tra­tion due to manufactur­ing
time and grade 3/4 adverse events (AEs) from cyto­kine release
syn­
drome (CRS), neu­
ro­
tox­
ic­
ity, and cytopenias remain major
con­cerns.3 The desire for safe “off-the-shelf” immunotherapies
has strength­ened inter­est in ADCs and BsAbs, which them­selves
have dem­on­strated prom­is­ing effi­cacy in relapsed/refrac­tory
MM (RRMM).3
In this review we pres­ent the state of devel­op­ment of each
class of anti­body-directed ther­apy in MM, with an empha­sis on the
rap­idly advanc­ing ADC and BsAb ther­a­peu­tic arma­men­tar­ium.
Table 1. Pivotal ran­dom­ized phase 3 tri­als of mono­clo­nal anti­body treat­ment for MM (Continued )
Clinical trial
Patient pop­u­la­tion
ELOQUENT 321
RRMM, at least 2 prior
lines
APOLLO7
ICARIA-MM17
RRMM, at least 1 prior
line
RRMM, at least 2 prior
lines
Arm (n=num­ber
of patients)
Overall response (%)
Median pro­gres­sion-free
sur­vival (months)
Hazard ratio (P value)
E-Pd (n=60)
53
10.3
0.54 (P=.008)
Pd (n=57)
26
4.7
D-Pd (n=151)
69
12.4
Pd (n=153)
46
6.9
I-Pd (n=154)
63
11.5
Pd (n=153)
32
6.5
0.63 (P=.0018)
0.6 (P=.001)
Randomized phase 2 study
Note: This list is not exhaus­tive for all­ran­dom­ized tri­als includ­ing mono­clo­nal antibodies.
D, daratumumab; d, dexa­meth­a­sone; E, elotuzumab; I, isatuximab; K, carfilzomib; M, mel­pha­lan; OR, odds ratio; P, pomalidomide; p, pred­ni­sone;
R, lenalidomide; T, tha­lid­o­mide; V, bortezomib.
a
in trans­plant-eli­gi­ble NDMM patients (phase 2 SWOG-1211 and
phase 3 GMMG-HD6 tri­als) or to Rd (phase 3 ELOQUENT-1 trial)
in trans­plant-inel­i­gi­ble patients, no improve­ments in out­comes
were dem­on­strated.22-24
Antibody-drug con­ju­gates
Despite the suc­cess of naked MoAbs, the vast major­ity of MM
patients ulti­mately relapse and require novel inter­ven­tions. Con­
sequently, research­ers have clev­erly com­bined the spec­i­fic­ity of
MoAbs with a cyto­toxic drug, cre­at­ing a sophis­ti­cated deliv­ery
sys­tem that trans­ports a lethal pay­load directly to the anti­genexpressing cell (Figure 1). This tech­nol­ogy has already shown
suc­cess in other hema­to­logic malig­nan­cies, includ­ing lym­phoma
(brentuximab vedotin) and acute mye­
loid leu­
ke­
mia (gemtu­
zumab ozogamicin).25
Several ADCs have been eval­u­ated in clin­i­cal tri­als in RRMM
(Table 2), the most prom­is­ing of which tar­get B-cell mat­u­ra­tion
anti­gen (BCMA). BCMA has emerged as an attrac­tive tar­get, as
it is expressed at high lev­els on PCs and plasmablasts but not
on other tis­sues.26 This selec­tiv­ity, in addi­tion to the fact that
BCMA undergoes inter­nal­i­za­tion, makes it an ideal tar­get for
ADCs.26
Belantamab mafodotin (GSK2857916), a first-in-class human­
ized IgG1, afucosylated ADC con­ju­gated to monomethyl auri­
statin-F (MMAF), was the first ADC to dem­on­strate sig­nif­i­cant
ther­a­peu­tic ben­e­fit in MM. Preclinical stud­ies dem­on­strate that
belantamab mafodotin elim­i­na­tes mye­loma cells through sev­
eral mech­a­nisms of action: direct cell kill­ing via the inhi­bi­tion
of micro­tu­bule poly­mer­i­za­tion, clas­si­cal IgG effec­tor func­tions
through an enhanced frag­ment crys­tal­liz­able region (Fc) domain,
and immu­no­genic cell death, a pro­cess in which dying cells elicit
an adap­tive immune response.27
Belantamab mafodotin was eval­
u­
ated in the piv­
otal
DREAMM-2 study.28 This phase 2 study explored doses of
2.5 mg/kg or 3.4 mg/kg admin­
is­
tered intra­
ve­
nously every
3 weeks until pro­gres­sion. In patients receiv­ing the US Food
and Drug Administration–approved 2.5-mg/kg dose (n = 97), all­
were TCR and had a median of 7 (3-12) prior lines of ther­apy.
The ORR was 31% (97.5% CI, 20.8–42.6), and the median PFS
and OS were 2.8 (95% CI, 1.6–3.6) and 13.7 months (95% CI,
9.9-not reached), respec­
tively. For responding patients, the
DOR was 11 months (95% CI, 4.2-not reached). Post hoc ana­
Dr Prakash Singh Shekhawat
MRD test­ing in AML 2022 | 165
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/163/2021838/163cipkar.pdf by guest on 09 December 2022
A PFS ben­e­fit is also seen for front­line daratumamab in com­
bi­na­tion with bortezomib-mel­pha­lan-pred­ni­sone (VMP; phase 3
ALCYONE trial) or lenalidomide-dexa­
meth­
a­
sone (Rd; phase 3
MAIA trial) in trans­plant-inel­i­gi­ble patients.11,12 Further, daratu­
mumab has been com­
bined with carfilzomib/Rd (KRd) as an
induc­tion ther­apy, with or with­out ASCT, in the phase 2 MASTER
and MANHATTAN stud­ies. Early results show impres­sive MRDneg­a­tive (10−5) remis­sions; 80% in the MASTER trial and 71% in the
MANHATTAN trial.13,14
The ques­
tion of daratumumab main­
te­
nance was first
addressed in the sec­ond ran­dom­i­za­tion of the CASSIOPEIA trial,
in which patients were fur­ther ran­dom­ized after ASCT/con­sol­i­
da­tion to daratumumab for up to 2 years or active sur­veil­lance.
The median PFS was not reached (NR) in the daratumumab
main­te­nance arm and was 46.7 months in the obser­va­tion arm
(haz­ard ratio [HR], 0.53; 95% CI, 0.42–0.68; P < .0001).15 Unex­
pectedly, those patients who had received daratumumab
dur­ing induc­tion and con­sol­i­da­tion derived no incre­men­tal
ben­e­fit from daratumumab main­te­nance alone vs obser­va­
tion alone (HR, 1.02; 95% CI, 0.71–1.47; P = .91), suggesting that
depth of response, rather than con­tin­u­ous ther­apy, may have
been the deter­min­ing fac­tor.15 Ongoing tri­
als such as GRIF­
FIN (NCT02874742), DRAMMATIC (NCT04071457), PERSEUS
(NCT03710603), and AURIGA (NCT03901963) will help elu­ci­date
the opti­mal use of daratumumab or daratumamab plus lenalid­
omide as main­te­nance ther­apy.
Isatuximab is another IgG1κ MoAb that binds to CD38.
One nota­
ble dif­
fer­
ence from daratumumab is that isatux­
imab can induce direct cyto­
tox­
ic­
ity via caspase-depen­
dent
­apo­p­­to­sis and lyso­some-medi­ated nonapoptotic cell kill­ing.16
­Isatuximab has been suc­cess­fully com­bined with pomalidomidedexa­meth­a­sone (Pd; phase 3 ICARIA-MM trial) and Kd (phase
3 IKEMA trial) in RRMM, with supe­rior PFS dem­on­strated in the
isatuximab arms (Table 1).17,18 In the GMMG-HD7 phase 3 study in
newly diag­nosed mul­ti­ple mye­loma (NDMM), isatuximab added
to RVd improved MRD (10−5) neg­a­tiv­ity rates (50.1% vs 35.6%)
prior to ASCT, fur­ther supporting the incor­po­ra­tion of an antiCD38 MoAb in front­line treat­ment.19
Elotuzumab, a human­
ized IgG1 MoAb directed against
SLAMF7, has shown improved PFS and OS in com­bi­na­tion with
Rd (phase 3 ELOQUENT-2 trial) and Pd (phase 3 ELOQUENT-3
trial) in RRMM.20,21 Disappointingly, when added to front­line RVd
ly­ses dem­on­strated sim­i­lar response and OS in patients with
high-risk cyto­ge­net­ics and those with impaired renal func­tion,
although patients with extramedullary dis­ease did not appear
to derive the same ben­e­fit.28 Consistent with MMAF-containing
ADCs, the most com­mon AEs (any grade/grade ≥3) were ker­
atopathy (72%/46%), change in best corrected visual acu­ity
(54%/31%), throm­bo­cy­to­pe­nia (38%/22%), ane­mia (27%/21%),
and blurred vision (25%/4%). Dose reduc­tions and delays due
to tox­ic­ity occurred in 54% and 35%, respec­tively, and were
less com­mon in the 2.5-mg/kg arm of the study.
Belantamab mafodotin is US Food and Drug Administration
approved for use in RRMM patients who have received 4 or more
lines of ther­apy. The approval came with a boxed warn­ing, indi­
cat­ing that tox­ic­ity to the cor­nea may result in vision loss, cor­neal
ulcers, or dry eyes. Although the mech­a­nism is yet unclear, pre­
clin­i­cal data sug­gest that ocu­lar tox­ic­ity is related to the recep­torinde­pen­dent uptake of the intact ADC into the epi­the­lial limbal
stem cells of the cor­nea.29 In DREAMM-2, 72% of patients in the
2.5-mg/kg cohort dem­
on­
strated keratopathy on oph­
thal­
mo­
logic exam; how­ever, only 56% expe­ri­enced symp­toms (most
com­monly blurred vision and/or dry eyes), and only 3 patients
(3%) discontinued ther­apy for cor­neal AEs.28 Importantly, expe­
ri­ence from DREAMM-2 indi­cates that patients recover from cor­
neal tox­ic­ity with dose holds for grade 2 events or higher, with
the major­ity of patients (88%) maintaining responses despite
prolonged dose delays. Based on this expe­ri­ence, the rec­om­
men­da­tions for man­age­ment of cor­neal events include the fre­
quent use of pre­
ser­
va­
tive-free lubri­
cant eye drops, eye care
166 | Hematology 2022 | ASH Education Program
pro­
fes­
sional assess­
ments before each dose, and timely dose
holds and mod­i­fi­ca­tions.
Numerous stud­ies are now eval­u­at­ing belantamab mafodotin
in dif­fer­ent drug com­bi­na­tions, dos­ing sched­ules, and treat­ment
set­tings (Table 2). This includes stud­ies com­bin­ing belantamab
mafodotin with pembrolizumab (DREAMM-4),30 with novel agents
(DREAMM-5),31 with lenalidomide or bortezomib (DREAMM-6),32
and with RVd in NDMM (DREAMM-9).33 Notably, in the Algon­quin
study, belantamab mafodotin was com­bined with Pd to iden­tify
the opti­mal dose and sched­ule in pomalidomide-naive patients.34
Across cohorts of patients receiv­
ing belantamab mafodotin,
1.92 mg/kg every 4 weeks or 2.5 mg/kg every 4, 8, or 12 weeks
(n=56), the ORR was 88.9% (≥ very good par­tial response [VGPR],
72%) and PFS, 17 months (14.5-not reached); nota­
bly, 48% of
patients were TCR.34 Additionally, stud­ies explor­ing lower doses
and extended sched­ules of belantamab mafodotin as a strat­egy
to reduce the inci­dence and sever­ity of cor­neal tox­ic­ity are ongo­
ing (DREAMM-9, DREAMM-14, and NCT04808037). Emerging data
from these stud­ies are encour­ag­ing, dem­on­strat­ing a reduc­tion
in the inci­dence and sever­ity of cor­neal tox­ic­ity with good clin­i­cal
effi­cacy.34,35
Several other ADCs targeting BCMA have been devel­
oped. AMG 224 is a human­ized IgG1 anti-BCMA mertansine-­­
con­ju­gated ADC. In a phase 1 study, at the expan­sion dose of
3 mg/kg (n = 11) the ORR was 27%.36 Mild ocu­lar events were
observed in 30% of patients with no reports of keratopathy.
MEDI2228 is another human­ized pyrrolobenzodiazepine (PBD)con­ju­gated anti-BCMA ADC. In a first-in-human study, an ORR
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/163/2021838/163cipkar.pdf by guest on 09 December 2022
Figure 1. Characteristics of ADCs. The components of the ADC and its target antigen influence the efficacy and safety profile. Pref­
erably, target antigens should only be found on malignant cells, be abundantly expressed, be capable of internalization, and not be
shed from the cellular membrane. The cytotoxic drug (payload or warhead) is the ultimate effector component, inducing direct cell
killing either by inhibiting microtubule formation or directly damaging cellular DNA. It should be highly potent in the subnanomolar
range and preferably nonpermeable to avoid damage to surrounding tissues. The linker connects the warhead to the antibody. These
should be stable in circulation and cleavable upon lysosomal degradation. The conjugation chemistry of the linker determines the
drug:antibody ratio, which critically influences the ADC potency. ADCC, antibody-dependent T-cellular cytotoxicity, Fab, fragment
antigen-binding region.
Table 2. Summary of anti­body-drug con­ju­gates for MM
Target
Payload
Combination
Trial phase: num­ber of
patients (n)
Response/
activ­ity
Current sta­tus
(ClinicalTrials​­.gov)
Belantamab
mafodotin
(belamaf)
BCMA
MMAF
Monotherapy28
Phase 1: n=35
ORR 60%
DREAMM-1 (com­pleted)
NCT02064387
Phase 2: n=196
ORR 31%
DREAMM-2 (com­pleted)
NCT03525678
B (Q3W) vs Pd
Phase 3: n=380 (E)
N/A
DREAMM-3 (recruiting)
NCT04162210
B (Q3W) + Pemb30
Phase 1/2: n=41
ORR 47%
DREAMM-4 (active, not
recruiting) NCT03848845
B + novel agent31
Phase 1/2: n=464 (E)
ORR 53% with
feladilimab
DREAMM-5 (recruiting)
NCT04126200
B-Vd OR Rd32
Phase 1/2: n=152 (E)
ORR 78% in BVd
arm
DREAMM-6 (active, not
recruiting) NCT03544281
B-Pd34
Phase 1/2: n=96 (E)
ORR 88.9%
ALGONQUIN (recruiting)
NCT03715478
B-Rd (trans­plantinel­i­gi­ble NDMM)
Phase 1/2: n=66 (E)
N/A
NCT04808037 (recruiting)
B-Vd vs D-Vd
Phase 3: n=575 (E)
N/A
DREAMM-7 (active, not
recruiting) NCT04246047
B-Pd vs V-Pd
Phase 3: n=450 (E)
N/A
DREAMM-8 (recruiting)
NCT04484623
B-VRd (trans­plantinel­i­gi­ble NDMM33
Phase 1: n=144 (E)
ORR 100%
(n=12)
DREAMM-9 (recruiting)
NCT04091126
Monotherapy in
renal impair­ment
Phase 1: n=36 (E)
N/A
DREAMM-12 (recruiting)
NCT04398745
Monotherapy in
liver impair­ment
Phase 1: n=28 (E)
N/A
DREAMM-13 (recruiting)
NCT04398680
Monotherapy
(vary­ing doses and
sched­ules)
Phase 2: n=180 (E)
N/A
DREAMM-14 (recruiting)
NCT05064358
AMG 224
BCMA
Mertansine
Monotherapy36
Phase 1: n=42 (E)
ORR 27%
(3 mg/kg)
NCT02561962 (active, not
recruiting)
CC 99712
BCMA
Maytansinoid-like
Monotherapy
Phase 1: n=160 (E)
N/A
NCT04036461 (recruiting)
MEDI2228
BCMA
PBD
Monotherapy
Phase 1: n=82
ORR 66% at
0.14 mg/kg
NCT03489525 (com­pleted)
Indatuximab
ravtansine
CD138
Maytansinoid
DM4
In com­bi­na­tion
with R or P
Phase 1/2: n=64
ORR 71.7% with R
and 70.6% with P
NCT01001442 (com­pleted)
Lorvotuzumab
mertansine
CD56
Mertansine
Monotherapy
Phase 1: n=37
ORR 5.7%
NCT00346255 (com­pleted)
Milatuzumab
CD74
Doxorubicin
Monotherapy
Phase 1: n=25
26% SD
NCT00421525 (com­pleted)
37
STRO-001
CD74
MMAF
Monotherapy
Phase 1: n=N/A
N/A
NCT03424603 (recruiting)
DFRF4539A
FcRH5
MMAE
Monotherapy
Phase 1: n=39
ORR 5%; 49% SD
NCT01432353 (com­pleted)
SGN-CD48A
CD48
MMAE
Monotherapy
Phase 1: n=14
N/A
NCT03379584 (ter­mi­nated)
ABBV-838
SLAMF7
MMAE
Monotherapy
Phase 1/1b: n=75
ORR 10.7%
NCT02462525 (ter­mi­nated)
Note: This list is not exhaus­tive for all­ADCs devel­oped for MM.
B, belantamab; belamaf, belantamab mafodotin; D, daratumumab; d, dexa­meth­a­sone; E, esti­mate; MMAE, monomethyl auristatin E; N/A, not avail­­
able; Pemb, pembrolizumab; P, pomalidomide; Q3W, once every 3 weeks; R, lenalidomide; SD, sta­ble dis­ease; V, bortezomib.
of 65.9% was reported at the max­i­mum tol­er­ated dose (MTD)
of 0.14 mg/kg (n = 41).37 The safety pro­file was con­sis­tent with
PBD-containing ADCs and included rash (31.7%), throm­bo­cy­to­
pe­nia (31.7%), pleu­ral effu­sions (24.4%), and increased gammaglutamyl trans­fer­ase (24.4%). Unexpectedly, ocu­lar AEs in the
form of pho­to­pho­bia were reported in 58.5% of patients. Phase
1 stud­ies of CC-99712, an anti-BCMA ADC con­ju­gated to 4 may­
tansinoid mol­e­cules, and HDP-101, an anti-BCMA ADC con­ju­
gated to an ama­ni­tin deriv­a­tive, are recruiting (NCT04036461
and NCT04879043, respec­tively).
Dr Prakash Singh Shekhawat
MRD test­ing in AML 2022 | 167
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/163/2021838/163cipkar.pdf by guest on 09 December 2022
Name
ADCs targeting anti­gens other than BCMA have shown less
prom­is­ing effi­cacy in MM. Table 2 sum­ma­rizes the cur­rent state
of devel­op­ment of indatuximab ravtansine (anti-CD138), lorvotu­
zumab mertansine (anti-CD56), STRO-001 (anti-CD74), and sev­eral
other ADCs with novel tar­gets, most hav­ing had their devel­op­
ment halted due to dis­ap­point­ing effi­cacy.
Bispecific antibodies
Figure 2. Characteristics of BsAbs. IgG-like BsAbs include an Fc region, while non–IgG-like BsAbs consist of only Fab (fragment
antigen-binding) variable regions and linkers. Since the Fc portion provides stability and longevity to the molecule in circulation,
most non–IgG-like BsAbs require more frequent dosing to maintain therapeutic plasma levels; they also lack the Fc-mediated effector
functions such as antibody-dependent T-cellular cytotoxicity and complement-dependent cytotoxicity. Several BsAbs, both IgG-like
and non–IgG-like, are under development for the treatment of MM. BiAb, bispecific antibodies.
168 | Hematology 2022 | ASH Education Program
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/163/2021838/163cipkar.pdf by guest on 09 December 2022
BsAbs are unique anti­body con­structs that simul­ta­neously bind
2 anti­gens. In can­cer ther­a­peu­tics, this usu­ally involves target­
ing an anti­
gen on the tumor and a mol­
e­
cule on an immune
cell, resulting in immune effec­tor cell acti­va­tion and tumor lysis
(Figure 2). Similar to the ADCs, BCMA has been the selected anti­
gen tar­get for the devel­op­ment of BsAbs, although oth­ers have
emerged (Table 3).
AMG 420, which tar­gets BCMA on PCs and CD3 on T cells, was
the first BsAb to show effi­cacy in human tri­als. Coined a BiTE®
(bispecific T-cell engager) because it lacks an Fc region, AMG
420 jus­ti­fied the devel­op­ment of bispecifics in MM, dem­on­strat­
ing a 70% ORR at the MTD (400 µg/d; n=10).38 The BiTE® for­mat
offers the advan­tage of bet­ter tis­sue pen­e­trance and access to
epi­topes but with the caveat of a short half-life, neces­si­tat­ing
a con­tin­u­ous intra­ve­nous infu­sion. This has been rem­e­died by
the devel­op­ment of an extended half-life ver­sion, AMG 701,39 and
numer­ous Fc-containing BsAbs that can be admin­is­tered every
1 to 3 weeks, with the major­ity mov­ing to sub­cu­ta­ne­ous for­mu­la­
tions (Table 3).40-46 Fc-containing bispecifics are larger and there­
fore more sta­ble in cir­cu­la­tion, while the Fc func­tion may or may
not be silenced depending on the agent. Thus far, BsAbs have
been eval­u­ated in heavily pretreated patients, with the major­
ity being TCR; a good num­ber have also included older patients
(>80 years). All have dem­on­strated encour­ag­ing clin­i­cal activ­
ity (Table 3), although data on dura­bil­ity of responses are still
imma­ture.
Teclistamab, an anti-BCMA/CD3 BsAb, is fur­thest along in clin­
i­cal devel­op­ment, with an antic­i­pated com­mer­cial approval in
2022. In the phase 1/2 MajesTEC-1 study, 165 patients received
teclistamab (77.8% TCR) at the recommended tar­
get dose of
1.5 mg/kg.40 The ORR was 63% (≥com­plete remis­sion, 39.4%),
with 44 patients (26.7%) achiev­ing MRD (10−5). Responses were
maintained across dif­
fer­
ent sub­
groups, includ­
ing poor-risk
groups, with the excep­tion of those with extramedullary dis­ease,
those with stage III dis­ease, or those with PC mar­row involve­
ment of 60% or greater. The median DOR was 18.4 months (95%
CI, 14.9-not esti­ma­ble), while the PFS was 11.3 months (95% CI,
8.8-17.1). CRS, neutropenia, infec­tion, and neu­ro­tox­ic­ity of any
grade/grade higher than or equal to 3 occurred in 72.1%/0.6%,
70.9%/64.2%, 76.4%/44.8%, and 14.5%/0%, respec­
tively; 19
patients died from AEs, includ­ing 12 deaths due to COVID-19.
In the phase 1 MagnetisMM-1 study, elranatamab (PF06863135), another anti-BCMAxCD3 BsAb, dem­
on­
strated an
ORR of 64% among 55 patients (91% TCR) receiv­
ing doses
of 215 µg/kg or higher.41 Remarkably, 7 of 10 patients treated
with prior BCMA-targeted ther­apy achieved a par­tial response
or bet­ter.41 The inci­dence of CRS at the recommended dose
(1000 µg/kg or 76 mg) was 67% (all­grade 1/2).41 Studies of
Table 3. Summary of bispecific T-cell antibodies for MM
Triple-class
refrac­tory
(median LoT)
Trial
phase
Preliminary
response/
activ­ity
Current sta­tus
(ClinicalTrials​
­.gov)
Target
Antibody
con­struct
AMG 42038
BCMA-CD3
BiTE®
N/A (29%
prior antiCD38, median
5 LoT)
Phase 1
Continuous
infu­sion for
4 wk (out
of 6)
ORR=31%
ORR
MTD=70%
38% CRS (6.25% ≥ gr 3)
5% ≥ gr 3
polyneuropathy
24% ≥ gr 3 infec­tion
Active, not
recruiting
NCT03836053
AMG 70139
BCMA-CD3
HLE-BiTE®
68% (median
6 LoT)
Phase 1/2
Weekly IV
ORR=36%
ORR=83% at
9 mg
75% CRS (10.5% ≥ gr 3)
8% neu­ro­tox­ic­ity
(gr 1-2)
13% ≥ gr 3 infec­tion
Recruiting
NCT03287908
Elranatamab41
BCMA-CD3
Humanized
IgG2a Fc
91% (median
6 LoT; 22%
prior antiBCMA)
Phase 1
Weekly or
every 2 wk
Sc
ORR=64%
for doses
≥215 µg/kg
67% CRS (gr 1-2)
MagnetisMM-1
Recruiting
NCT03269136
REGN545842
BCMA-CD3
Fc Fab arms
97.1% (median
5 LoT)
Phase 1/2
Weekly IV
ORR=73.3%
at 96-200-mg
doses
38.2% CRS (gr 1-2)
4% neu­ro­tox­ic­ity
(gr 1-2)
23% pneu­mo­nia
(11% ≥ gr 3)
Recruiting
NCT03761108
Teclistamab40
BCMA-CD3
Humanized
IgG4 Fc
77.8%
(median 5
LoT; prior
anti-BCMA
not per­mit­
ted)
Phase 1/2
Weekly Sc
ORR=63%
72.1% CRS (gr 3, 0.6%;
no gr 4)
14.5% neu­ro­tox­ic­ity
(1 gr 4 event)
44.8% ≥ gr 3 infec­tion
MajestTEC-1
Recruiting
NCT03145181
CC-9326944
BCMA-CD3
Asymmetric
2-arm IgG
66.7%
(median 6
LoT)
Phase 1
Weekly IV
ORR=83.3%
in 10 pts with
doses ≥6 mg
89.5% CRS (1 gr 5
event)
26.3% infec­tion
Recruiting
NCT03486067
TNB-383B44
BCMA-CD3
IgG4 Fc
CD3
acti­vat­ing
T effec­tor
cells
62% (median
5 LoT)
Phase 1
Q21d IV
ORR=79% at
doses ≥40 mg
52% CRS (3% ≥ gr 3
at RP2D)
28% infec­tion
Recruiting
NCT03933735
Cevostamab45
FcRH5CD3
Humanized
IgG1 Fc
85% (median
6 LoT; 33.5%
prior antiBCMA)
Phase 1
Q21d IV
ORR=54.5%
at 160-mgdose level
80.7% CRS (1.3% ≥
gr 3)
18.8% ≥ gr 3 infec­tion
14.3% neu­ro­tox­ic­ity
(0.3% ≥ gr 3)
Recruiting
NCT03275103
Talquetamab46
GPRC5DCD3
Humanized
IgG4 Fc
Weekly: 77%
(median 6
LoT; 30%
prior antiBCMA)
Biweekly:
65% (median
5 LoT; 17%
prior antiBCMA)
Phase 1/2
Weekly or
biweekly Sc
Weekly:
ORR=70%
Biweekly:
ORR=71%
Weekly: 73% CRS
(1 gr 3)
Biweekly: 78% CRS
(gr 1-2)
MonumenTal-1
Recruiting
NCT03399799
Schedule
Safety
Note: This list is not exhaus­tive for all­bispecific T-cell antibodies devel­oped for MM.
Fab, frag­ment anti­gen-bind­ing; gr, grade; HLE, half-life extended; IV, intra­ve­nous; LoT, lines of ther­apy; pts, patients; Q21d, every 21 days;
RP2D, recommended phase 2 dose; Sc, sub­cu­ta­ne­ous; wk, week.
teclistamab and elranatamab in com­bi­na­tion with immu­no­mod­
u­la­tory drugs and anti-CD38s and in ear­lier lines of treat­ment are
ongo­ing (Table 4).
Other nota­
ble anti-BCMA BsAbs include REGN5458, CC93269, and TNB-383B (Table 3). In a phase 1/2 trial of REGN5458,
the ORR was 73.3% among patients treated at the 96-mg- and
200-mg-dose lev­
els; no patients expe­
ri­
enced CRS of grade
3 or higher.42 Of the 12 patients treated with 6 mg or more of
CC-93269, the ORR was 83.3%.43 CRS was reported in 89.5% of
patients, mostly grade 1/2 (57.9%/26.3%); how­ever, one patient
died in the set­ting of CRS, with infec­tion as a poten­tial con­trib­
u­tor.43 In the dose-esca­la­tion cohort of TNB-383B (doses ≥40 mg
every 3 weeks), the ORR was 79% (n=19/24) with CRS reported
as mainly grade 1/2.44
Fc recep­tor homo­log 5 (FcRH5) is a cell sur­face anti­gen of
unknown func­
tion whose expres­
sion is restricted to B cells,
Dr Prakash Singh Shekhawat
MRD test­ing in AML 2022 | 169
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/163/2021838/163cipkar.pdf by guest on 09 December 2022
Name
Table 4. Summary of bispecific T-cell anti­body com­bi­na­tions for MM
Name
Patient pop­u­la­tion
Trial phase
Combination drugs
Elranatamab
RRMM
Phase 1b/2
Arm 1: elranatamab + nirogacestat (GSI)
Arm 2: elranatamab + lenalidomide + dexa­meth­a­sone
Phase 3
Arm 1: elranatamab
Arm 2: elranatamab + daratumumab
Current sta­tus
(ClinicalTrials​­.gov)
MagnetisMM-4 (recruiting)
NCT05090566
MagnetisMM-5 (recruiting)
NCT05020236
Arm 3: earatumumab + pomalidomide + dexa­meth­a­sone
Teclistamab +
talquetamab
RRMM
Phase 3
Arm 1: teclistamab + daratumumab
Arm 2: daratumumab + pomalidomide + dexa­meth­a­sone
MajestTEC-3 (recruiting)
NCT05083169
Phase 1
Arm 1: teclistamab + talquetamab
NCT04586426 (recruiting)
Arm 2: teclistamab + talquetamab + daratumumab
Phase 1b
Arm 1: daratumumab + teclistamab
Arm 2: daratumumab + talquetamab
TRIMM-2 (recruiting)
NCT04108195
Arm 3: daratumumab + talquetamab + pomalidomide
Arm 4: daratumumab + teclistamab + pomalidomide
Cevostamab
RRMM
Phase 1
Arm 1: cevostamab
Recruiting NCT04910568
Arm 2: cevostamab + pomalidomide + dexa­meth­a­sone
Arm 3: cevostamab + daratumumab + dexa­meth­a­sone
Note: This list is not exhaus­tive for all­bispecific T-cell anti­body com­bi­na­tions devel­oped for MM.
GSI, gamma secretase inhib­i­tor.
with the highest expres­sion on PCs. BFCR4350A (cevostamab)
is a human­ized IgG Fc anti­body targeting FcRH5 and CD3. In
an ongoing phase 1 study, cevostamab was admin­is­tered for
a fixed dura­tion of 17 cycles, unlike other BsAb tri­als that treat
to pro­gres­sion.45 Among 160 patients (85% TCR), the ORR was
56.7% and 36.1% for those patients receiv­ing doses of 132 to
198 mg (n=60) or 20 to 90 mg (n=83), respec­tively. The ORR
among patients who had prior CAR T-cell or anti-BCMA ther­
apy was 44.4% and 36.4%, respec­tively. CRS was observed in
80.7% (128/160) of patients (mainly grade 1/2; 2 cases grade 3).
In the sin­gle step-up dose cohorts (n=86), the median DOR was
11.5 months (95% CI, 6.0, 18.4) at a fol­low-up of 14.3 months.
Thus, cevostamab appears to be a ben­e­fi­cial treat­ment option,
with a unique ther­a­peu­tic tar­get, for RRMM patients.
Talquetamab is an IgG4 Fc-containing BsAb targeting Gpro­tein–cou­pled recep­tor fam­ily C group 5 mem­ber D (GPRC5D).46
GPRC5D is highly expressed on PCs but also on keratinized tis­
sues. In the MonumenTal-1 study, patients received sub­cu­ta­ne­ous
talquetamab at doses of 405µg/kg (n=30) weekly or 800 µg/kg
biweekly (n=23).44 Unique AEs, owing to the expres­
sion of
GPRC5D on keratinized tis­sues, include dysgeusia, pal­mar/plan­tar
des­qua­ma­tion, nail dys­tro­phy, and sys­temic rash, which were
reported in 75% of patients (mostly grade 1/2; 7.5% grade 3).46
Mitigation strat­e­gies for these toxicities included the use of emol­
lient creams and, for oral AEs, saliva-sub­sti­tute sprays and rinses
at the onset of symp­toms. The ORR and CRS in patients receiv­ing
the 405-µg/kg or 800-µg/kg dose was 70% and 73% (1 grade
3) and 71% and 78% (all­grade 1/2), respec­tively.46 Trials explor­
ing talquetamab in com­bi­na­tion with teclistamab, as well as with
SOC antimyeloma agents, are actively recruiting (Table 4).
170 | Hematology 2022 | ASH Education Program
CLINICAL CASE (Con­tin­ued)
The patient responds to belantamab mafodotin; how­ever, she
expe­ri­ences mul­ti­ple dose inter­rup­tions for cor­neal AEs and
progresses after 9 months. She is then referred to an aca­demic
cen­ter and is enrolled in a clin­i­cal trial eval­u­at­ing talquetamab
(anti-GPRC5D) monotherapy. In cycle 1 she expe­ri­ences grade
1 CRS but no neu­ro­tox­ic­ity. After 6 months of ther­apy, she con­
tin­ues to have an ongo­ing response with man­age­able grade 1
dysgeusia and der­ma­to­logic symp­toms.
Conclusion
Alongside CAR T-cell ther­apy, anti­body-based immunotherapies
have emerged as impor­tant off-the-shelf ther­a­peu­tic options for
all­patients with RRMM (Table 5). The sequenc­ing and dura­tion
of these ther­a­pies remain ongo­ing clin­i­cal ques­tions, although
emerg­ing data are encour­ag­ing. In a heavily pretreated RRMM
pop­u­la­tion relaps­ing on a BsAb (n=64), the ORR to sub­se­quent
treat­ment (sec­ond BsAb, n=20; CAR T, n=15) was 58%, with an OS
of 17.6 months (95% CI, 21.6-not reached).47 Further, in the Mag­
netisMM-1 study, 7 of 13 patients pre­vi­ously treated with an antiBCMA targeted ther­apy (5/7 for those treated with anti-BCMA
ADCs) responded to the BCMAxCD3 BsAb, elranatamab.41 Mean­
while in the ongoing phase 1 study of cevostamab (FCRH5xCD3
BsAb) that lim­ited treat­ment to 17 cycles, 6 patients con­tin­ued
to main­tain responses for 6 months or lon­ger after the ces­sa­tion
of treat­ment.45
While ADCs and BsAbs dem­on­strate prom­is­ing effi­cacy,
future efforts need to focus on the opti­mal tim­ing and dura­tion
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/163/2021838/163cipkar.pdf by guest on 09 December 2022
Arm 3: daratumumab + bortezomib + dexa­meth­a­sone
Table 5. Comparison of novel immu­no­ther­apy approaches in mul­ti­ple mye­loma
Advantages
Disadvantages
Bispecific T-cell engagers
CAR T-cell ther­apy
Off-the-shelf ther­apy
Off-the-shelf ther­apy
-
Immune and non­im­mune mech­a­nisms of action
-
-
Infrequent dos­ing (every 3 wk-12 wk)
-
One-time ther­apy
Encouraging response rates
Deep responses
Deep responses
No CRS/ICANS
Mostly grade 1-2 CRS/ICANS
-
Outpatient admin­is­tra­tion
Only ini­tial dos­ing as inpa­tient
Vacation from con­tin­u­ous ther­apy
Continuous ther­apy until pro­gres­sion
Continuous ther­apy until pro­gres­sion
-
Frequent dose inter­rup­tions
Weekly or biweekly dos­ing
Administration delays due to
manufactur­ing time
Ocular tox­ic­ity
Significant immu­no­sup­pres­sion
Potential for severe CRS/ICANS;
prolonged cytopenias
Ophthalmic exams prior to dos­ing
Specialized cen­ters required
Complex infra­struc­ture required
Cost ($$)
Cost ($$)
Cost ($$$)
ICANS, immune effec­tor cell-asso­ci­ated neu­ro­tox­ic­ity syn­drome; wk, week.
of ther­apy, on com­bi­na­tional strat­e­gies, on mit­i­ga­tion of the
risk of immu­no­sup­pres­sion/infec­tion (BsAbs) and cor­neal tox­
ic­ity (ADC), on the mech­a­nisms of resis­tance, and on equi­ta­
ble access. Although much work is still to be done, these novel
approaches offer new hope for a yet incur­able dis­ease.
Conflict-of-inter­est dis­clo­sure
Christopher Cipkar: no com­pet­ing finan­cial inter­ests to declare.
Christine Chen: con­sul­tancy: Forus Therapeutics; advi­sory board:
Amgen, Bristol Myers Squibb, Janssen Pharmaceuticals.
Suzanne Trudel: research funding: Amgen, Bristol Myers Squibb,
Genentech, GlaxoSmithKline, Janssen Pharmaceuticals, Pfizer,
Roche; con­
sul­
tancy: Bristol Myers Squibb, Forus, GlaxoSmith­
Kline, K36 Therapeutics, Roche; advi­sory board: Amgen, Bristol
Myers Squibb, GlaxoSmithKline, Janssen Pharmaceuticals, Pfizer,
Sanofi.
Off-label drug use
Christopher Cipkar: nothing to disclose.
Christine Chen: nothing to disclose.
Suzanne Trudel: nothing to disclose.
Correspondence
Suzanne Trudel, Princess Margaret Cancer Centre, Department
of Medical Oncology and Hematology, 700 University Ave, Ste
6-322, Toronto, ON M5G 1Z5, Canada; e-mail: Suzanne​­.trudel@
uhn​­.ca.
References
1.
Cowan AJ, Green DJ, Kwok M, et al. Diagnosis and man­age­ment of mul­ti­ple
mye­loma: a review. JAMA. 2022;327(5):464-477.
2. Mateos MV, Weisel K, De Stefano V, et al. LocoMMotion: a pro­spec­tive,
non-interventional, mul­ti­na­tional study of real-life cur­rent stan­dards of care
in patients with relapsed and/or refrac­tory mul­ti­ple mye­loma. Leukemia.
2022;36(5):1371-1376.
3. Mohammed T, Mailankody S. “Off-the-shelf” immunotherapies for mul­ti­ple
mye­loma. Semin Oncol. 2022;49(1):60-68.
4. Offidani M, Corvatta L, More S, et al. Daratumumab for the man­age­ment of
newly diag­nosed and relapsed/refrac­tory mul­ti­ple mye­loma: cur­rent and
emerg­ing treat­ments. Front Oncol. 2021;17(10):624-661.
5. Palumbo A, Chanan-Khan A, Weisel K, et al; CASTOR Investigators. Daratu­
mumab, bortezomib, and dexa­meth­a­sone for mul­ti­ple mye­loma. N Engl J
Med. 2016;375(8):754-766.
6. Bahlis NJ, Dimopoulos MA, White DJ, et al. Daratumumab plus lenalidomide
and dexa­meth­a­sone in relapsed/refrac­tory mul­ti­ple mye­loma: extended
fol­low-up of POLLUX, a ran­dom­ized, open-label, phase 3 study. Leukemia.
2020;34(7):1875-1884.
7. Dimopoulos MA, Terpos E, Boccadoro M, et al; APOLLO Trial Investigators.
Daratumumab plus pomalidomide and dexa­
meth­
a­
sone ver­
sus poma­
lidomide and dexa­meth­a­sone alone in pre­vi­ously treated mul­ti­ple mye­
loma (APOLLO): an open-label, randomised, phase 3 trial. Lancet Oncol.
2021;22(6):801-812.
8. Usmani SZ, Quach H, Mateos MV, et al. Carfilzomib, dexa­meth­a­sone, and
daratumumab ver­
sus carfilzomib and dexa­
meth­
a­
sone for patients with
relapsed or refrac­tory mul­ti­ple mye­loma (CANDOR): updated out­comes
from a randomised, multicentre, open-label, phase 3 study. Lancet Oncol.
2022;23(1):65-76.
9. Moreau P, Attal M, Hulin C, et al. Bortezomib, tha­lid­o­mide, and dexa­meth­a­
sone with or with­out daratumumab before and after autol­o­gous stem-cell
trans­plan­ta­tion for newly diag­nosed mul­ti­ple mye­loma (CASSIOPEIA): a
randomised, open-label, phase 3 study. Lancet. 2019;394(10192):29-38.
10. Sborov D, et al. Daratumumab (DARA) plus lenalidomide, bortezomib, and
dexamethasone (RVd) in patients (Pts) with transplant-eligible newly diag­
nosed multiple myeloma (NDMM): final analysis of GRIFFIN after ≥ 1y follow
up. Presented at: The International Myeloma Society 2022 Annual Meeting;
August 27, 2022; Los Angeles, CA.
11. Facon T, Kumar S, Plesner T, et al. Daratumumab, lenalidomide and dexa­
meth­a­sone alone in newly diag­nosed mul­ti­ple mye­loma (MAIA): over­all
sur­vival results from a ran­dom­ized, open-label, phase 3 trial. Lancet Oncol.
2021;22(11):1582-1596. doi: 10.1016/S1470-2045(21)00466-6.
12. Mateos MV, Cavo M, Blade J, et al. Overall sur­vival with daratumumab,
bortezomib, mel­pha­lan, and pred­ni­sone in newly diag­nosed mul­ti­ple
mye­
loma (ALCYONE): a randomised, open-label, phase 3 trial. Lancet.
2020;395(10218):132-141.
13. Costa LJ, Chhabra S, Medvedova E, et al. Daratumumab, carfilzomib, lena­
lidomide, and dexamethasone with minimal residual disease responseadapted therapy in newly diagnosed multiple myeloma. J Clin Oncol.
2022;40(25):2901-2912.
14. Landgren O, Hultcrantz M, Diamond B, et al. Safety and effec­tive­ness of
weekly carfilzomib, lenalidomide, dexa­
meth­
a­
sone, and daratumumab
com­bi­na­tion ther­apy for patients with newly diag­nosed mul­ti­ple mye­loma:
the MANHATTAN nonrandomized clin­i­cal trial. JAMA Oncol. 2021;7(6):862868.
15. Moreau P, Hulin C, Perrot A, et al. Maintenance with daratumumab or
obser­va­tion fol­low­ing treat­ment with bortezomib, tha­lid­o­mide, and dexa­
meth­a­sone with or with­out daratumumab and autol­o­gous stem-cell trans­
Dr Prakash Singh Shekhawat
MRD test­ing in AML 2022 | 171
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/163/2021838/163cipkar.pdf by guest on 09 December 2022
Antibody drug con­ju­gates
172 | Hematology 2022 | ASH Education Program
32. Popat R, Nooka A, Stockerl-Goldstein K, et al. DREAMM-6: safety, tol­er­a­bil­
ity and clin­i­cal activ­ity of belantamab mafodotin (belamaf) in com­bi­na­tion
with bortezomib/dexa­meth­a­sone (BorDex) in relapsed/refrac­tory mul­ti­
ple mye­loma (RRMM) [ab­stract]. Blood. 2020;138(suppl 1);1419. Abstract
1419.
33. Usmani S, Alonso A, Quach H, et al. DREAMM-9: phase I study of belantamab
mafodotin plus stan­dard of care in patients with trans­plant-inel­i­gi­ble newly
diag­nosed mul­ti­ple mye­loma. Blood. 2021;138(suppl 1):2738.
34. Trudel S, McCurdy A, Sutherland H, et al. Part 1 results of a dose-find­ing
study of belantamab mafodotin in com­bi­na­tion with pomalidomide and
dexa­meth­a­sone for the treat­ment of relapsed/refrac­tory mul­ti­ple mye­
loma (RRMM) [ab­stract]. Blood. 2021;138(suppl 1):1653. Abstract 725.
35. Terpos E, Gavriatopoulou M, Ntanasis-Stathopoulos I, et al. Safety and
effi­cacy of belantamab mafodotin in com­bi­na­tion with RD in newly diag­
nosed, trans­plant inel­i­gi­ble mul­ti­ple mye­loma patients: a phase 1/2 study
by the Hel­
lenic Society of Hematology [ab­
stract]. HemaSphere. 2022;
6:79-80. Abstract S178.
36. Lee HC, Raje NS, Landgren O, et al. Phase 1 study of the anti-BCMA anti­
body-drug con­ju­gate AMG 224 in patients with relapsed/refrac­tory mul­ti­
ple mye­loma. Leukemia. 2021;35(1):255-258.
37. Kumar S, Migkou M, Bhutani M, et al. Phase 1, first-in-human study of
MEDI2228, a BCMA-targeted ADC in patients with relapsed/refrac­tory mul­
ti­ple mye­loma [ab­stract]. Blood. 2020;138(suppl 1);179. Abstract 179.
38. Topp MS, Duell J, Zugmaier G, et al. Anti-B-cell mat­u­ra­tion anti­gen BiTE
mol­e­cule AMG 420 induces responses in mul­ti­ple mye­loma. J Clin Oncol.
2020;38(8):775-783.
39. Harrison SJ, Minnema MC, Lee HC, et al. A phase 1 first in human (FIH) study
of AMG 701, an anti-B-cell mat­u­ra­tion anti­gen (BCMA) half-life extended
(HLE) BiTE (bispecific T-cell engager) mol­e­cule, in relapsed/refrac­tory (RR)
mul­ti­ple mye­loma (MM). Blood. 2020;136(suppl 1):181.
40. Moreau P, Garfall AL, van de Donk NWCJ, et al. Teclistamab in relapsed or
refrac­tory mul­ti­ple mye­loma. N Engl J Med. 2022;387(6):495-505.
41. Jakubowiak AJ, Bahlis NJ, Raje NS, et al. Elranatamab, a BCMA-targeted
T-cell redirecting immu­
no­
ther­
apy, for patients with relapsed or refrac­
tory mul­ti­ple mye­loma: updated results from MagnetisMM-1. J Clin Oncol.
2022;40(suppl 16):8014.
42. Zonder JA, Richter J, Bumma N, et al. Early, deep and dura­ble responses
and low rates of cyto­kine release syn­drome with REGN5458, a BCMAxCD3
bispecific mono­
clo­
nal anti­
body, in a phase 1/2 first-in-human study in
patients with relapsed/refrac­tory mul­ti­ple mye­loma (RRMM). Blood. 2021;
138(suppl 1);160.
43. Costa LJ, Wong SW, Bermúdez A, et al. First clin­i­cal study of the B-cell
mat­u­ra­tion anti­gen (BCMA) 2+1 T cell engager (TCE) cc-93269 in patients
(pts) with relapsed/refrac­tory mul­ti­ple mye­loma (RRMM): interim results of
a phase 1 mul­ti­cen­ter trial. Blood. 2019;134 (suppl 1):143.
44. Kumar SK, D’Souza A, Shah N, et al. A first-in-human study of TNB383B, a BCMA x CD3 $$. Blood. 2021;138(suppl 1):900. doi:10.1182/bl
ood-2021-150757.45. Trudel S, Cohen A, Krishnan A, et al. Cevostamab
monotherapy con­tin­ues to show clin­i­cally mean­ing­ful activ­ity and man­
age­able safety in patients with heavily pre-treated relapsed/refrac­tory
mul­ti­ple mye­loma (RRMM): updated results from an ongo­ing phase I study.
Blood. 2021;138(suppl 1):157. doi:10.1182/blood-2021-147983.46. Krishnan
AY, Minnema MC, Berdeja JG, et al. Updated phase 1, results from mon­u­
men­tal: first-in-human study of talquetamab, a G pro­tein-cou­pled recep­
tor fam­ily C group 5 mem­ber D x CD3 bispecific anti­body, in patients
with relapsed and/or refrac­tory mul­ti­ple mye­loma (RRMM). Blood. 2021;
138(suppl 1):158. doi:10.1182/blood-2021-146868.47. Mouhieddine TH, Van
Oekelen O, Pan D, et al. Clinical out­comes of relapsed/refrac­tory mul­ti­ple
mye­loma patients fol­low­ing treat­ment with bispecific antibodies (BiAbs).
Blood. 2021;138(suppl 1):821.
© 2022 by The Amer­i­can Society of Hematology
DOI 10.1182/hema­tol­ogy.2022000334
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/163/2021838/163cipkar.pdf by guest on 09 December 2022
plant in patients with newly diag­nosed mul­ti­ple mye­loma (CASSIOPEIA):
an open-label, randomised, phase 3 trial. Lancet Oncol. 2021;22(10):13781390.
16. Richardson PG, Beksaç M, Špička I, Mikhael J. Isatuximab for the treat­
ment of relapsed/refrac­tory mul­ti­ple mye­loma. Expert Opin Biol Ther.
2020;20(12):1395-1404.
17. Attal M, Richardson PG, Rajkumar SV, et al; ICARIA-MM Study Group. Isat­
uximab plus pomalidomide and low-dose dexa­meth­a­sone ver­sus poma­
lidomide and low-dose dexa­
meth­
a­
sone in patients with relapsed and
refrac­
tory mul­
ti­
ple mye­
loma (ICARIA-MM): a randomised, multicentre,
open-label, phase 3 study. Lancet. 2019;394(10214):2096-2107.
18. Moreau P, Dimopoulos MA, Mikhael J, et al; IKEMA Study Group. Isatux­
imab, carfilzomib, and dexa­
meth­
a­
sone in relapsed mul­
ti­
ple mye­
loma
(IKEMA): a multicentre, open-label, randomised phase 3 trial. Lancet.
2021;397(10292):2361-2371.
19. Goldschmidt H, Mai EK, Nievergali E, et al. Addition of isatuximab to lena­
lidomide, bortezomib and dexa­meth­a­sone as induc­tion ther­apy for newlydiag­nosed, trans­plant-eli­gi­ble mul­ti­ple mye­loma patients: the phase III
GMMG-HD7 trial. Blood. 2021;138(suppl 1):463.
20. Dimopoulos MA, Lonial S, Betts KA, et al. Elotuzumab plus lenalidomide
and dexa­meth­a­sone in relapsed/refrac­tory mul­ti­ple mye­loma: extended
4-year fol­low-up and anal­y­sis of rel­a­tive pro­gres­sion-free sur­vival from the
ran­dom­ized ELOQUENT-2 trial. Cancer. 2018;124(20):4032-4043.
21. Dimopoulos MA, Dytfeld D, Grosicki S, et al. Elotuzumab plus pomalidomide
and dexa­meth­a­sone for mul­ti­ple mye­loma. N Engl J Med. 2018;379(19):18111822.
22. Usmani SZ, Hoering A, Ailawadhi S, et al; SWOG1211 Trial Investigators. Bor­
tezomib, lenalidomide, and dexa­meth­a­sone with or with­out elotuzumab in
patients with untreated, high-risk mul­ti­ple mye­loma (SWOG-1211): pri­mary
anal­y­sis of a randomised, phase 2 trial. Lancet Haematol. 2021;8(1):e45e54.
23. Goldschmidt H, Mai EK, Bertsch U, et al. Elotuzumab in com­bi­na­tion with
lenalidomide, bortezomib, dexa­meth­a­sone and autol­o­gous trans­plan­ta­
tion for newly-diag­nosed mul­ti­ple mye­loma: results from the ran­dom­ized
phase III GMMG-HD6 trial. Blood. 2021;138(suppl 1):486.
24. Dimopoulos MA, Richardson PG, Bahlis NJ, et al; ELOQUENT-1 Investigators.
Addition of elotuzumab to lenalidomide and dexa­meth­a­sone for patients
with newly diag­nosed, trans­plan­ta­tion inel­i­gi­ble mul­ti­ple mye­loma (ELO­
QUENT-1): an open-label, multicentre, randomised, phase 3 trial. Lancet
Haematol. 2022;9(6):e403-e414.
25. Rossi C, Chrétien ML, Casasnovas RO. Antibody-drug con­ju­gates for the
treat­ment of hema­to­log­i­cal malig­nan­cies: a com­pre­hen­sive review. Target
Oncol. 2018;13(3):287-308.
26. Kleber M, Ntanasis-Stathopoulos I, Terpos E. BCMA in mul­ti­ple mye­loma—a
prom­is­ing key to ther­apy. J Clin Med. 2021;10(18).
27. Tai YT, Mayes PA, Acharya C, et al. Novel anti-B-cell mat­u­ra­tion anti­gen
anti­body-drug con­ju­gate (GSK2857916) selec­tively induces kill­ing of mul­ti­
ple mye­loma. Blood. 2014;123(20):3128-3138.
28. Lonial S, Lee HC, Badros A, et al. Longer term out­comes with sin­gle-agent
belantamab mafodotin in patients with relapsed or refrac­
tory mul­
ti­
ple
mye­loma: 13-month fol­low-up from the piv­otal DREAMM-2 study. Cancer.
2021;127(22):4198-4212.
29. Zhao H, Atkinson J, Gulesserian S, et al. Modulation of macropinocytosismedi­ated inter­nal­i­za­tion decreases ocu­lar tox­ic­ity of anti­body-drug con­ju­
gates. Cancer Res. 2018;78(8):2115-2126.
30. Suvannasankha A, Bahlis NJ, Trudel S, et al. Safety and clin­
i­
cal activ­
ity of belantamab mafodotin with pembrolizumab in patients with
relapsed/refrac­
tory mul­
ti­
ple mye­
loma (RRMM): DREAMM-4 Study
[ab­stract]. J Clin Oncol. 2022;(suppl 16);8018. Abstract 8018.
31. Callander N, Ribrag V, Richardson P, et al. DREAMM-5 study: inves­ti­gat­
ing the syn­er­getic effects of belantamab mafodotin plus induc­ible T-cell
co-stim­u­la­tor ago­nist (aICOS) com­bi­na­tion ther­apy in patients with
relapsed/refrac­tory mul­ti­ple mye­loma [ab­stract]. Blood. 2021;138(suppl
1);897. Abstract 897.
IMMUNOTHERAPY IN MULTIPLE MYELOMA
Sarah A. Holstein
Division of Oncology and Hematology, Department of Internal Medicine, University of Nebraska Medical Center, Omaha, NE
The development of novel cellular therapies and bispecific T-cell-engaging antibodies is occurring at breakneck speed
in multiple myeloma (MM). While groundbreaking, these agents have their unique logistical and toxicity issues and currently do not represent a curative approach. In this context, there continues to be an urgent need to develop novel,
off-the-shelf immunotherapy approaches to add to the armamentarium. This article explores novel agents being investigated in combination with standard immunomodulatory drugs as well as next-generation cereblon E3 ligase modulators. These novel agents include drugs being repurposed from their use in other diseases as well as novel monoclonal
antibodies. In addition, agents under development such as immunocytokines, immunotoxins, and natural killer-cell
activators/engagers are reviewed. These novel therapeutic strategies hold the promise of countermanding the
immunosuppressive tumor microenvironment, leading to enhanced anti-MM activity.
LEARNING OBJECTIVES
• Learn about novel IMiD-based combinations and CELMoDs being tested in R/R MM
• Recognize the developing therapeutic landscape of immunocytokines, immunotoxins, and NK-cell activators/
engagers in MM
CLINICAL CASE
A 64-year-old woman with immunoglobulin G (IgG) lambda
relapsed/refractory multiple myeloma (R/R MM) received
commercial idecabtagene vicleucel 6 months ago and
is now experiencing biochemical relapse. She had previously received 5 prior lines of therapy, including prior
autologous stem cell transplant (ASCT), and her disease
is refractory to lenalidomide, pomalidomide, bortezomib,
carfilzomib, daratumumab, and belantamab mafodotin.
Her chimeric antigen receptor (CAR) T-cell course was
complicated by cytokine release syndrome and neurotoxicity, and she wishes to avoid any subsequent therapies
that could be associated with those adverse events (AEs).
What novel non-CAR T-cell/non-bispecific antibody
immunotherapy approaches are currently under investigation?
Introduction
Novel immunotherapies such as CAR T cells and bispecific
T-cell-engaging antibodies show remarkable promise for
MM. However, given the current logistical constraints of
commercial CAR T-cell products, as well as the potential
logistical issues with bispecific antibodies that may limit
access to these therapies to certain populations, there
continues to be a need for alternative approaches. With
the remarkable success of small-molecule agents such as
immunomodulatory drugs (IMiDs) and proteasome inhibitors, the ongoing substantial focus on developing novel
small-molecule agents for MM is not surprising. In addition,
the field is exploring the therapeutic potential of novel
antibody-based approaches and natural killer (NK)-cell
activators/engagers. In this educational session, we discuss the most recent research investigating these novel
non-CAR T-cell/bispecific-based approaches for the treatment of R/R MM.
IMiDs plus novel agent combinations
The IMiDs lenalidomide and pomalidomide have helped
transform the management of MM. Lenalidomide is a current standard of care in the newly diagnosed (ND) and
maintenance settings, while pomalidomide is exclusively
used in the R/R setting. The pleiotropic effects of these
agents, which include both direct cytotoxic effects as well
as a myriad of effects on the immune microenvironment,
Dr Prakash Singh Shekhawat
Next generation of novel therapies for myeloma | 173
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/173/2021746/173holstein.pdf by guest on 09 December 2022
Beyond the cell: novel noncellular immunotherapy
approaches to multiple myeloma
Table 1. Summary of ongo­ing tri­als involv­ing IMiDs and novel agents
Phase
Study pop­u­la­tion
Treatment
Novel agent mech­a­nism
Primary end point(s)
NCT03702725
1
R/R
Ibrutinib + LenDex
BTK inhib­i­tor
MTD, ORR
NCT03110822
1
R/R with ≥2 prior lines
Ruxolitinib + Len + ste­roids
Jak inhib­i­tor
MTD, TEAE
NCT04508790
2
R/R
Leflunomide + PomDex
Dihydroorate
dehy­dro­ge­nase inhib­i­tor
ORR
NCT03143985
1
R/R
Vactosertib + Pom
TGFB1 recep­tor antag­o­nist
MTD
NCT04405167
1
R/R
Tasquinimod +/− IRD
S100A9 inhib­i­tor
MTD
NCT04942067
1/2
R/R
Liasftoclax + PomDex or DaraLenDex
BCL2 inhib­i­tor
DLT
NCT04674514
1/2
R/R with ≥1 prior line
Lisaftoclax + LenDex
BCL2 inhib­i­tor
DLT
NCT04025450
3
ND high risk
Chidamide + VRD vs VRD
HDAC inhib­i­tor
CR rate, AE
NCT03605056
2
R/R with ≥1 prior line
Chidamide + LenDex
HDAC inhib­i­tor
ORR
NCT02400242
1
R/R
ACY-241 + PomDex
HDAC inhib­i­tor
MTD
NCT01997840
1/2
R/R with ≥2 prior lines
Ricolinostat + PomDex
HDAC inhib­i­tor
MTD; ORR
NCT03031730
1
R/R with 1-3 prior lines
KRT-232 + KRD
MDM2 inhib­i­tor
Safety, TEAE
BCL2, B-cell lym­phoma 2; CR, com­plete response; Dara, daratumumab; Dex, dexa­meth­a­sone; DLT, dose lim­it­ing tox­ic­ity; HDAC, his­tone deacetylase;
Jak, Janus kinase; Len, lenalidomide; MDM2, mouse dou­ble min­ute 2; MTD, max­i­mum tol­er­ated dose; Pom, pomalidomide.
cou­pled with gen­er­ally tol­er­a­ble side-effect pro­files, have led
to numer­ous com­bi­na­tion-based stud­ies. Substantial research
is focused on deter­
min­
ing the mech­
a­
nisms of resis­
tance to
these agents and whether these resis­tance mech­a­nisms can be
over­come via novel com­bi­na­tions. In this sec­tion an over­view
of some of the most recent IMiD plus novel agent com­bi­na­tion
strat­e­gies are pro­vided, includ­ing those involv­ing small mol­e­
cules (Table 1) and anti­body-based ther­a­pies (Table 2). In sev­eral
cases, agents being used to treat other malig­nan­cies or dis­eases
are being repurposed for the treat­ment of MM. Other ongo­ing
stud­ies are inves­ti­gat­ing IMiD-based com­bi­na­tions with agents
such as lisaftoclax (BCL2 inhib­i­tor), KRT-232 (MDM2 inhib­i­tor), or
novel his­tone deacetylase inhib­i­tors (eg, ricolinostat, ACY-241,
chidamide) (Table 1). Not discussed in this chap­ter but included
in Table 2 are a num­ber of ongo­ing stud­ies eval­u­at­ing IMiDs with
bispecific antibodies.
The Bruton tyro­sine kinase (BTK) inhib­i­tor ibrutinib is now
widely used in the treat­ment of sev­eral B-cell malig­nan­cies. BTK
is expressed in MM cells, and it may mod­u­late the bone mar­
row (BM) micro­en­vi­ron­ment by decreas­ing cyto­kine/chemokine
secre­
tion and suppressing mye­
loid-derived sup­
pres­
sor cell
activ­ity.1 While sin­gle-agent ibrutinib led to an over­all response
rate (ORR) of 0% in patients with R/R MM (64% lenalidomide
refrac­tory),2 sub­
se­
quent efforts focused on com­
bin­
ing the
agent with IMiDs. A phase 1 study of ibrutinib-lenalidomidedexa­meth­a­sone reported an ORR of 7% with a clin­i­cal ben­e­fit
rate (≥ min­i­mal response) of 80%. Notably, 54% of the patients
were refrac­tory to lenalidomide.3 An ongo­ing phase 1 study is
eval­u­at­ing the com­bi­na­tion in patients with at least 1 prior line of
ther­apy (eli­gi­ble patients must not have progressed on full-dose
lenalidomide but could have expe­ri­enced dis­ease pro­gres­sion
on lenalidomide main­te­nance) (Table 1). A phase 1 study of ibrutinib in com­bi­na­tion with pomalidomide-dexa­meth­a­sone was
com­pleted, and the spon­sor chose not to pur­sue the phase 2
com­po­nent (NCT02548962).
The Janus kinase (Jak) inhib­i­tor ruxolitinib has a num­ber of
immu­no­mod­u­la­tory prop­er­ties that appear rel­e­vant for MM.
174 | Hematology 2022 | ASH Education Program
Recent work has shown that ruxolitinib can decrease PD-L1
expres­sion and decrease expres­sion of a vari­ety of genes impli­
cated in resis­tance to lenalidomide.4,5 A phase 1 study of ruxolitinib in com­bi­na­tion with lenalidomide and meth­yl­pred­nis­o­lone
reported an ORR of 38% in a patient pop­u­la­tion in which 93%
were lenalidomide refrac­tory.6
Leflunomide was approved for the treat­
ment of rheu­
ma­
toid arthri­tis by the US Food and Drug Administration in 1998.
Leflunomide is metab­
o­
lized to teriflunomide, which inhib­
its
dihydroorate dehy­dro­ge­nase, an enzyme involved in pyrim­i­dine
bio­syn­the­sis. While ear­lier stud­ies suggested that leflunomide
induced cyto­tox­ic­ity in MM cells in part related to the dis­rup­
tion of pyrim­i­dine bio­syn­the­sis, more recent work has indi­cated
that this agent downregulates c-Myc expres­
sion by directly
inhibiting the PIM fam­ily of ser­ine/thre­o­nine kinases.7 Preclinical
stud­ies showed the anti-MM activ­ity of leflunomide but only in
immune-com­pe­tent mice.7 Increases in T-cell acti­va­tion mark­ers
(LAMP-1, CD69) with a decrease in the T-cell exhaus­tion marker
CTLA4 were observed in treated ani­mals.7 In addi­tion, the com­
bi­na­tion of lenalidomide with leflunomide prolonged sur­vival in
a mouse model of MM.7 Subsequently a phase 1 study of sin­gleagent leflunomide was conducted in patients with R/R MM. No
objec­tive responses were reported; how­ever, 9 in 11 patients
achieved sta­ble dis­ease.8 Currently, this agent is being inves­ti­
gated in patients with smol­der­ing MM as well as in com­bi­na­tion
with pomalidomide-dexa­meth­a­sone in patients with R/R MM.
Vactosertib is a transforming growth fac­tor beta 1 (TGFB1)
recep­tor antag­o­nist, and as there are data connecting increased
TGFB1 secre­tion by MM cells with impaired immune sur­veil­lance,
stud­ies have been conducted eval­u­at­ing the poten­tial ther­a­peu­
tic use of vactosertib in MM. Preclinical stud­ies with vactosertib
dem­on­strated sin­gle-agent activ­ity in a mouse MM model as well
as in pri­mary patient sam­ples.9 A phase 1 study was conducted
eval­u­at­ing the com­bi­na­tion of vactosertib with pomalidomide
(no ste­roids) and dem­on­strated the safety of this com­bi­na­tion
along with a 6-month pro­gres­sion-free sur­vival (PFS) rate of 80%
(n=15).10
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/173/2021746/173holstein.pdf by guest on 09 December 2022
NCT num­ber
Table 2. Summary of ongo­ing tri­als involv­ing IMiDs and novel immune-mod­u­lat­ing antibodies
Phase
Study pop­u­la­tion
Antibody tar­get
Treatment
Primary end point(s)
1
R/R
CD47
Lemzoparlimab +/− PomDex, CarDex, DaraDex
DLT
NCT04892446
2
R/R
CD47
Magrolimab + PomDex, CarDex, BorDex, Dara
DLT, AE, ORR
NCT04150965
1/2
R/R with ≥3 prior lines
LAG-3
TIGIT
Relatlimab + PomDex;
BMS-986207 + PomDex
ORR, AE
NCT04722146
1
R/R
BCMA/CD3
Tec + DaraPom, DaraLenBor, nirogacestat, Len,
DaraLen
DLT, AE
NCT05090566
2
R/R with ≥3 prior lines
BCMA/CD3
Elran + LenDex
DLT, TEAE
NCT05020236
3
R/R with ≥1 prior line
includ­ing Len + PI
BCMA/CD3
Elran vs Elran+ Dara vs Elran +DaraPomDex
PFS
NCT05137054
1
R/R with ≥3 prior lines
BCMA/CD3
REGN5458 + DaraDex, CarDex, LenDex or BorDex
DLT, TEAE
NCT05317416
3
MRD-pos­i­tive
post–up-front ASCT
BCMA/CD3
Len vs Elran main­te­nance
MRD neg­a­tiv­ity, PFS
NCT05243797
3
Post–up-front ASCT
BCMA/CD3
Len vs TecLen main­te­nance
PFS
NCT04910568
1
R/R
FCRH5/CD3
Cevost +/− PomDex, DaraDex
AE
NCT05050097
1
R/R
GPRC5D/CD3
Talq +/− Car, DaraCar, Len, DaraLen, Pom
DLT, AE
Bor, bortezomib; Car, carfilzomib; Cevost, cevostamab; Dara, daratumumab; Dex, dexa­meth­a­sone; DLT, dose-lim­it­ing tox­ic­ity; Elran, elrantamab; MRD,
min­i­mal resid­ual dis­ease; PI, proteasome inhib­i­tor; Pom, pomalidomide; Talq, talquetamab; Tec, teclistamab.
Tasquinimod is an agent that tar­gets mye­loid-derived sup­
pres­sor cells via the inhi­bi­tion of S100A9. While thus far it has
pri­mar­ily been inves­ti­gated in the set­ting of pros­tate can­cer,
pre­clin­i­cal stud­ies in MM revealed sin­gle-agent activ­ity as well
as in com­bi­na­tion with lenalidomide.11 No direct cyto­
toxic
effect of this agent was observed against MM cells in vitro, suggesting that the mech­a­nism of action is via mod­u­la­tion of the
tumor micro­en­vi­ron­ment.11 A phase 1 study of sin­gle-agent tasquinimod in R/R MM was com­pleted and this agent is currently
being evaluated in com­bi­na­tion with ixazomib-lenalidomidedexa­meth­a­sone (NCT04405167).
While sev­eral phase 3 tri­als study­ing IMiDs in com­bi­na­tion with
anti-programmed cell death 1 pro­tein inhib­i­tors such as pembrolizumab and nivolumab failed,12-14 there con­tin­ues to be inter­est in
inves­ti­gat­ing the ther­a­peu­tic poten­tial of alter­na­tive check­point
inhib­i­tors. Several ongo­ing clin­i­cal tri­als are eval­u­at­ing com­bi­
na­tions of IMiDs with novel check­point inhib­i­tors such as antiLAG3 and anti-TIGIT mono­clo­nal antibodies (mAbs) (Table 2).
There is also inter­
est in targeting CD47, often referred to as
the mac­
ro­
phage “don’t eat me” sig­
nal, and recent/ongo­
ing
stud­
ies are explor­
ing magrolimab- or lemzoparlimab-based
com­bi­na­tions in R/R MM. Whether the tox­ic­ity issues encoun­
tered with IMiDs and anti-programmed cell death 1 inhib­i­tors
will also be observed with any of these novel immune-targeting
mAbs remains to be deter­
mined,15 but it is nota­
ble that the
lemzoparlimab-based study was recently ter­
mi­
nated by the
spon­sor for as yet undis­closed rea­sons while unex­pected seri­
ous AEs were reported with magrolimab in other malig­nan­cies.
CELMoDs
Iberdomide (CC-220) is an orally avail­­able agent that, sim­i­lar
to lenalidomide and pomalidomide, binds to the cereblon E3
ubiquitin ligase com­plex and thus is con­sid­ered to be a CELMoD
(cereblon E3 ligase mod­u­la­tor). Iberdomide more potently binds
cereblon, thus lead­ing to greater deg­ra­da­tion of Ikaros and Ailos in
cell cul­ture stud­ies com­pared to lenalidomide or pomalidomide.16
Preclinical stud­ies dem­on­strated the activ­ity of iberdomide in
lenalidomide- or pomalidomide-refrac­tory MM cells.17 A phase
1b/2a dose-esca­
la­
tion study of iberdomide in R/R MM was
conducted.18 Iberdomide (daily on days 1-21 of a 28-day cycle)
was admin­is­tered in com­bi­na­tion with weekly dexa­meth­a­sone.
The median num­ber of prior reg­i­mens was 5 (2-12), with 100%
receiv­ing lenalidomide and 69% receiv­ing prior pomalidomide.
Dosing started at 0.3 mg, and the recommended phase 2 dose
(RP2D) was deter­mined to be 1.6 mg. The most com­mon grade 3
to 4 AEs were neutropenia (26%), throm­bo­cy­to­pe­nia (11%), and
neu­rop­a­thy (2%). The ORR was 31% while 88% achieved sta­ble
dis­ease or bet­ter (dis­ease con­trol rate).18 The results from the
dose-expan­sion phase of this study showed an ORR of 28% with
a clin­i­cal ben­e­fit rate of 36% and a dis­ease con­trol rate of 79%
(n=107). In patients who had pre­vi­ously received anti–B-cell mat­
u­ra­tion anti­gen (BCMA) ther­apy, the ORR was 25% (n=24).19 Mass
cytometry anal­y­sis of BM aspi­rate sam­ples from patients treated
in the phase 1b/2a study revealed marked changes in the tumor
micro­en­vi­ron­ment, includ­ing a reduc­tion in naive and reg­u­la­
tory B cells, a reduc­tion in CD4+ T cells with an increase in CD8+
T-cells, a shift toward the cyto­toxic effec­tor-mem­ory phe­no­type
of T cells, a decrease in CD8+ T cells expressing inhib­i­tory check­
points (TIGIT, KLRG1), and an increase in NK cells (includ­ing acti­
vated NK cells) with a decrease in TIGIT+ NK cells.20
Clinical tri­als inves­ti­gat­ing iberdomide in com­bi­na­tion with
other stan­dard agents are ongo­ing (Table 3). Preliminary results
from a phase 1/2 study eval­u­at­ing iberdomide in com­bi­na­tion
with either daratumumab-dexa­
meth­
a­
sone or bortezomibdexa­meth­a­sone have been reported.21 The RP2D of iberdomide
in both com­bi­na­tions was also deter­mined to be 1.6 mg daily. The
rate of grade 3 to 4 neutropenia in the iberdomide-daratumumabdexa­meth­a­sone cohort was 67% and 26% in the iberdomidebortezomib-dexa­meth­a­sone cohort. The rates of diarrhea (all
grades) were 26% (iberdomide-daratumumab-dexa­meth­a­sone)
Dr Prakash Singh Shekhawat
Next gen­er­a­tion of novel ther­a­pies for mye­loma | 175
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/173/2021746/173holstein.pdf by guest on 09 December 2022
NCT num­ber
NCT04895410
Table 3. Summary of ongo­ing tri­als involv­ing iberdomide or mezigdomide
Phase
Study pop­u­la­tion
Treatment
Primary end point(s)
2
Intermediate or high-risk smol­der­ing
IberDex×4 cycles ⟶ Iber until PD
vs
Iber until PD
ORR
NCT05272826
2
ND TI
IberBorDex
sCR rate after 4 cycles
NCT02773030
1/2
ND TI, R/R
ND, R/R
R/R post-BCMA
R/R
DaraIberDex
IberBorDex
IberDex
IberCarDex
MTD, RP2D, ORR
NCT05199311
1/2
ND TE
IberCarDex×2-4 cycles ⟶ ASCT
Rate of TEAEs; CR/sCR rates
NCT04934475
(IFM 2020-02)
3
ND TE
Isa-KRD induc­tion. Randomization based
on MRD sta­tus to addi­tional Isa-KRD vs
ASCT + IsaKRD con­sol­i­da­tion vs tan­dem
ASCT. Maintenance includes ran­dom­i­za­
tion to Len×3y vs IberIsa×3y
MRD neg­a­tiv­ity rates
NCT05177536
2
Post–up-front ASCT
Iber main­te­nance until PD
1-y com­ple­tion rate
NCT04564703
2
Post–up-front ASCT
Iber main­te­nance until PD
Response improve­ment
rate after 6 mo; rate of dose
reduc­tions/dis­con­tin­u­a­tions
within 6 mo
NCT05354557
2
Post–up-front ASCT with ≤VGPR on Len
main­te­nance or post-salvage ASCT after
2-3 prior ther­a­pies
Iber main­te­nance×12 cycles
6-mo CR rate
NCT04975997
3
R/R with 1-2 prior lines
DaraIberDex vs DaraBorDex
PFS
NCT05289492
1/2
R/R with ≥3 prior lines (prior BCMA
targeted ther­apy allowed)
EOS884448, EOS884448-Iber,
EOS884448-IberDex
Rate of AEs/SAEs,
num­ber of DLTs, RP2D of
EOS884448 +/− Iber(Dex),
ORR
NCT04392037
2
R/R with 2-4 prior lines; Len refrac­tory
IberCyDex
PFS
NCT03374085
1/2
R/R with ≥3 prior lines (prior CAR T ther­apy
allowed)
MezigDex
AEs, PK param­e­ters, MTD,
ORR
NCT03989414
1/2
ND or R/R
R/R
R/R
R/R
R/R
MezigBorDex
MezigCarDex
MezigEloDex
MezigIsaDex
MezigDaraDex
RP2D, DLTs, AEs, ORR
Bor, bortezomib; Car, carfilzomib; CR, com­plete response; Cy, cyclo­phos­pha­mide; Dara, daratumumab; Dex, dexa­meth­a­sone; DLT, dose-lim­it­ing
tox­ic­ity; Elo, elotuzumab; Iber, iberdomide; Isa, isatuximab; KRD, carfilzomib, lenalidomide, dexa­meth­a­sone; Len, lenalidomide; Mezig, mezigdomide;
MRD, min­i­mum resid­ual dis­ease; MTD, max­i­mum tol­er­ated dose; PD, pro­gres­sive dis­ease; PK, phar­ma­co­ki­netic; SAE, seri­ous adverse event; sCR, strin­
gent com­plete response; TE, trans­plant eli­gi­ble; TI, trans­plant inel­i­gi­ble.
and 35% (iberdomide-bortezomib-dexa­
meth­
a­
sone). No pa­
tients devel­oped throm­botic events. Antithrombotic pro­phy­
laxis is man­
da­
tory in all­ongo­
ing stud­
ies with iberdomide.
Other iberdomide-based com­bi­na­tion tri­als include ixazomibiberdomide-dexa­meth­a­sone, iberdomide-cyclo­phos­pha­midedexa­meth­a­sone, and iberdomide-carfilzomib-dexa­meth­a­sone,
as well as iberdomide in com­
bi­
na­
tion with EOS884448GSK4428859A, an anti-TIGIT mono­clo­nal anti­body. In addi­tion,
sev­eral stud­ies are eval­u­at­ing iberdomide as main­te­nance ther­
apy post ASCT (Table 3).
Mezigdomide (CC-92480) is another potent CELMoD. This
agent has a sig­
nif­i­
cantly higher deg­
ra­
da­
tion effi­
ciency com­
pared to either lenalidomide or pomalidomide.22 The phase 1
study of CC-92480 in heavily treated R/R MM patients (median
prior lines=6; half had tri­ple-class refrac­tory dis­ease) in com­bi­
na­tion with dexa­meth­a­sone showed an ORR of 55% at the RP2D
176 | Hematology 2022 | ASH Education Program
(1.0 mg/d for 21 to 28 days).23 Grade 3/4 treat­ment-emer­gent
AEs (TEAEs) were reported in 88% of patients, with the most
fre­quent grade 3 and 4 AEs includ­ing neutropenia (53%), infec­
tions (30%), ane­mia (29%), and throm­bo­cy­to­pe­nia (17%). Correlative stud­ies have shown Ikaros/Aiolos deg­ra­da­tion in periph­eral
T cells in a dose-depen­dent man­ner at doses greater than or equal
to 0.6 mg and showed sub­strate recov­ery dur­ing drug hol­i­days
(full recov­ery with ≥7-day breaks).24 Preliminary results have dem­
on­strated the safety and fea­si­bil­ity of com­bin­ing mezigdomide
with bortezomib-dexa­meth­a­sone,25 and ongo­ing stud­ies include
other mezigdomide-based com­bi­na­tions, includ­ing carfilzomibdexa­meth­a­sone, elotuzumab-dexa­meth­a­sone, and isatuximab
(or daratumumab)-dexa­meth­a­sone (Table 3). In addi­tion, more
novel com­bi­na­tions such as mezigdomide-dexa­meth­a­sone with
tazemetostat (an EZH2 inhib­i­tor), BMS-986158 (a BET inhib­i­tor),
or trametinib (a MEK inhib­i­tor) are also planned (NCT05372354).
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/173/2021746/173holstein.pdf by guest on 09 December 2022
NCT num­ber
NCT04776395
Immunocytokines and immunotoxins
Figure 1. Examples of immunocytokines and immunotoxins
undergoing clinical trial evaluation in MM. Immunocytokines
are antibody-cytokine fusion proteins, while immunotoxins are
fusion proteins containing antibody fragments and a biologic
toxin. SLTA, Shiga-like toxin A subunit; VH, heavy chain variable
region; VL, light chain variable region. Created with BioRender​
.com.
38% in patients with a median of 7 prior lines of ther­apy, as well
as an ORR of 38% in patients refrac­tory to anti-CD38 mAb ther­
apy.30 The most nota­ble AEs included cytopenias (76% throm­bo­
cy­to­pe­nia, 69% neutropenia, 66% ane­mia) and infu­sion-related
reac­tions (31%).30 Currently, a ran­dom­ized phase 2 study is eval­
u­at­ing 2 dif­fer­ent dose lev­els (flat dos­ing of 120 or 240 mg every
4 weeks) in R/R MM.
TAK-169 is con­sid­ered an immunotoxin because it is an engineered pro­
tein containing a deimmunized form of the ribo­
some-inactivating Shiga-like toxin A sub­unit (SLTA) fused to an
anti­body frag­ment that spe­cif­i­cally tar­gets the CD38 cell sur­
face recep­tor.32 This agent is inter­nal­ized by CD38-expressing
cells, lead­ing to cell death via the irre­vers­ible inhi­bi­tion of pro­
tein syn­the­sis. Ex vivo stud­ies using pri­mary BM sam­ples showed
lysis of the MM cells, with the highest amount of lysis in sam­ples
from ND or daratumumab-naive R/R patients.33 A phase 1 study
in patients with R/R MM or non-Hodgkin lym­phoma is ongo­ing
(NCT04017130). An ini­tial report after 4 patients were enrolled at
the starting dose noted 1 patient expe­ri­enc­ing a TEAE of asymp­
tom­atic grade 2 myo­car­di­tis that was revers­ible.34 A num­ber of
other immunotoxins have under­gone pre­clin­i­cal eval­u­a­tion thus
far, targeting a vari­
ety of tumor anti­
gens (eg, CD38, CD138,
BCMA) and involv­ing a num­ber of tox­ins (eg, Shiga-like toxin A
sub­unit SLTA, pseu­do­mo­nas exo­toxin, saporin, ricin).35
NK-cell acti­va­tors and engagers
Determining how to effec­tively har­ness the anti-MM activ­ity of
NK cells has been an active area of research for decades. While
many ongo­ing stud­ies involve genetic engi­neer­ing of NK cells
(eg, CAR NK cells), inter­est in alter­na­tive strat­e­gies by which
to enhance the activ­ity of endog­e­nous NK cells con­tin­ues. Earlier efforts included mAbs targeting killer immu­no­glob­u­lin-like
recep­tors (KIRs; eg, IPH2101, 1-7F9, lirilumab) and NK group 2
mem­ber A ( monalizumab). Currently, no ongo­ing stud­ies are
inves­ti­gat­ing these agents in MM.
An alter­
na­
tive strat­
egy involves agents that enhance IL-15
activ­ity, as IL-15 plays a key role in NK devel­op­ment and sur­vival.
ALT-803 is an IL-15 superagonist fusion pro­tein that was eval­u­
ated in a phase 1 study in R/R MM; how­ever, the results have
not been disclosed (NCT02099539). A study using ALT-803 in
patients relaps­ing after allo­ge­neic stem cell trans­plant did dem­
on­strate that this agent increased NK and CD8+ T-cell num­bers
and func­tion.36 NKTR-255 is a poly­mer-con­ju­gated IL-15 recep­tor
ago­nist reported to have syn­er­gis­tic activ­ity with daratumumab
in pre­clin­i­cal mod­els.37 Currently, this agent is being exam­ined in
a phase 1 study either as a sin­gle agent or in com­bi­na­tion with
daratumumab in R/R MM (NCT04136756).
Finally, an excit­ing emerg­ing group of ther­a­peu­tic agents
are the bispecific and trispecific killer engagers (BiKEs and
TriKEs), as well as the anti­body-recruiting mol­e­cules (ARMs)
(Figure 2). BiKEs include 2 linked sin­gle-chain var­i­able frag­
ments that rec­og­nize the tumor anti­gen (eg, BCMA) and an
NK cell acti­va­tion recep­tor (eg, NKp30, NK group 2 mem­ber
A, CD16). TriKEs also incor­po­rate another com­po­nent, such as
IL-15, to fur­ther enhance NK cell acti­va­tion. Half-life extended
nanobody-based BiKEs are under devel­op­ment.38 ARMs con­
tain 2 linked ter­mini; 1 ter­mi­nus binds to a tumor anti­gen,
while the other binds endog­e­nous IgG antibodies regard­less
of the anti­gen-bind­ing spec­i­fic­ity.39 This leads to the opsonization of tumor cells with endog­e­nous IgG, resulting in tumor
Dr Prakash Singh Shekhawat
Next gen­er­a­tion of novel ther­a­pies for mye­loma | 177
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/173/2021746/173holstein.pdf by guest on 09 December 2022
The US Food and Drug Administration–approved anti-CD38 mAbs
daratumumab and isatuximab have found wide­spread use in the
R/R set­ting as well as increas­ing use in the ND set­ting. The pre­
dom­i­nant mech­a­nisms of action for these agents include direct
induc­tion of apo­pto­sis, anti­body-depen­dent cell-medi­ated
cyto­tox­ic­ity (ADCC), com­ple­ment-depen­dent cyto­tox­ic­ity, anti­
body-depen­dent cel­lu­lar phago­cy­to­sis, and inhi­bi­tion of CD38
ectoenzyme activ­ity, although a vari­ety of immu­no­mod­u­la­
tory effects have been described as well.26,27 As alter­na­tives to
“naked” mAbs, newer drug devel­op­ment efforts have focused
on the mod­i­fi­ca­tion of anti-CD38 mAbs, thus lead­ing to the gen­
er­a­tion of immunocytokines and immunotoxins (Figure 1).
Modakafusp alfa (for­merly known as TAK-573) is an exam­ple
of an immunocytokine. This agent was designed as a deliv­ery
sys­tem of inter­feron alpha (IFNA) 2b to CD38+ cells. Modakafusp
is a recom­bi­nant human­ized IgG4 anti-CD38 mAb that is fused
to an atten­u­ated IFNA pro­tein. While IFNA was inves­ti­gated as
an anti-MM agent decades ago due to its cyto­toxic and immu­
no­mod­u­la­tory prop­er­ties, the sys­temic tox­ic­ity of the agent
lim­ited its use.28 The hypoth­e­sis under­ly­ing the novel fusion pro­
tein is that the sys­temic tox­ic­ity is lim­ited as a con­se­quence of
both reduced bind­ing affin­ity to the IFNA recep­tor as well as
an increased local­ized con­cen­tra­tion at CD38+ tar­get cells. In
vitro stud­ies dem­on­strated that modakafusp displayed 10 000fold increased spec­i­fic­ity for CD38+ cells while being approx­
i­ma­tely 6000-fold less toxic to nor­mal BM cells com­pared to
native IFNA.29 Preclinical stud­ies showed syn­ergy with stan­dard
agents such as lenalidomide and bortezomib.29 Due to the lim­
ited Fc func­tion­al­ity of this fused pro­tein, it is thought to be
unlikely to induce ADCC. This agent acti­vates innate and adap­
tive immune cells, as well as induces apo­pto­tic sig­nals to the
MM cells via sig­nal­ing through the IFNA recep­tor.30 Interestingly,
modakafusp binds to an epi­tope of CD38 that is unique rel­a­tive
to the com­mer­cially avail­­able anti-CD38 mAbs.31 A first-in-human
phase 1 study in patients with R/R MM has shown that a dose of
1.5 mg/kg (intra­ve­nously) every 4 weeks resulted in an ORR of
lysis via NK cell–medi­ated ADCC.39 Preclinical stud­ies have
dem­on­strated the poten­tial of BiKEs, TriKEs, and ARMs,39-41 and
a clin­i­cal trial eval­u­a­tion is the next step.
Correspondence
Sarah A. Hol­
stein, University of Nebraska Medical Center,
987680 Nebraska Medical Ctr, Omaha, NE 68198; e-mail: sarah​
­.holstein@unmc​­.edu.
References
1.
CLINICAL CASE (Continued)
A dis­cus­sion is held with the patient about the many prom­is­
ing approaches cur­rently under inves­ti­ga­tion, includ­ing the use
of IMiDs with novel agents, CELMoDs, immunocytokines, and
NK-cell acti­va­tors/engagers. As all­of these approaches involve
mod­u­la­tion of the immune sys­tem in some man­ner, off-tar­get
effects are pos­si­ble. She expresses inter­est in learn­ing more
about the clin­i­cal trial options avail­­able at your insti­tu­tion.
2.
3.
4.
5.
Conflict-of-inter­est dis­clo­sure
Sarah A. Hol­stein: con­sul­tancy: Bristol Myers Squibb, Celgene,
Janssen, Genentech, Oncopeptides, Sanofi, Secura Bio, Takeda;
research funding: Oncopeptides.
7.
Off-label drug use
Sarah A. Hol­stein: by definition, everything that is discussed in
this chapter is currently experimental and undergoing clinical
trial and is therefore off-label.
178 | Hematology 2022 | ASH Education Program
6.
8.
Tai YT, Anderson KC. Bruton’s tyro­sine kinase: oncotarget in mye­loma.
Oncotarget. 2012;3(9):913-914.
Richardson PG, Bensinger WI, Huff CA, et al. Ibrutinib alone or with dexa­
meth­
a­
sone for relapsed or relapsed and refrac­
tory mul­
ti­
ple mye­
loma:
phase 2 trial results. Br J Haematol. 2018;180(6):821-830.
Ailawadhi S, Parrondo RD, Moustafa MA, et al. Ibrutinib, lenalidomide and
dexa­meth­a­sone in patients with relapsed and/or refrac­tory mul­ti­ple mye­
loma: phase I trial results [published online ahead of print 30 April 2022].
Hematol Oncol.
Chen H, Li M, Ng N, et al. Ruxolitinib reverses check­point inhi­bi­tion by reduc­ing
programmed cell death ligand-1 (PD-L1) expres­sion and increases anti-tumour
effects of T cells in mul­ti­ple mye­loma. Br J Haematol. 2021;192(3):568-576.
Chen H, Li M, Sanchez E, et al. JAK1/2 path­way inhi­bi­tion suppresses M2
polar­i­za­tion and over­comes resis­tance of mye­loma to lenalidomide by
reduc­ing TRIB1, MUC1, CD44, CXCL12, and CXCR4 expres­sion. Br J Haematol. 2020;188(2):283-294.
Berenson JR, To J, Spektor TM, et al. A phase I study of ruxolitinib, lenalidomide, and ste­roids for patients with relapsed/refrac­tory mul­ti­ple mye­
loma. Clin Cancer Res. 2020;26(10):2346-2353.
Buettner R, Morales C, Caserta E, et al. Leflunomide reg­u­lates c-Myc expres­
sion in mye­loma cells through PIM targeting. Blood Adv. 2019;3(7):1027-1032.
Rosenzweig M, Palmer J, Tsai NC, et al. Repurposing leflunomide for
relapsed/refrac­tory mul­ti­ple mye­loma: a phase 1 study. Leuk Lymphoma.
2020;61(7):1669-1677.
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/173/2021746/173holstein.pdf by guest on 09 December 2022
Figure 2. Novel approaches for improving NK-cell–mediated targeting of myeloma cells. Bispecific killer cell engagers (BiKEs) are
composed of 2 antibody fragments that recognize tumor antigens and NK-cell activation receptors, while trispecific killer cell engagers (TriKEs) also include a component such as IL-15 to further enhance NK-cell activation. Antibody-recruiting molecules (ARMs)
contain a terminus that binds endogenous IgG antibodies and is linked to a second terminus that binds tumor antigen, thus leading
to opsonization of tumor cells with endogenous IgG and resulting in NK-cell–mediated ADCC.
Created with BioRender​.com
mul­ti­ple mye­loma (RRMM): pre­lim­i­nary results from the phase 1/2 study
CC-92480-MM-002. Blood. 2021;138(suppl 1):2731.
26. Krejcik J, Casneuf T, Nijhof IS, et al. Daratumumab depletes CD38+ immune
reg­u­la­tory cells, pro­motes T-cell expan­sion, and skews T-cell rep­er­toire in
mul­ti­ple mye­loma. Blood. 2016;128(3):384-394.
27. Deckert J, Wetzel MC, Bartle LM, et al. SAR650984, a novel human­ized
CD38-targeting anti­body, dem­on­strates potent anti­tu­mor activ­ity in mod­
els of mul­ti­ple mye­loma and other CD38+ hema­to­logic malig­nan­cies. Clin
Cancer Res. 2014;20(17):4574-4583.
28. Fritz E, Ludwig H. Interferon-alpha treat­
ment in mul­
ti­
ple mye­
loma:
meta-anal­y­sis of 30 randomised tri­als among 3948 patients. Ann Oncol.
2000;11(11):1427-1436.
29. Pogue SL, Taura T, Bi M, et al. Targeting atten­u­ated inter­feron-α to mye­
loma cells with a CD38 anti­body induces potent tumor regres­sion with
reduced off-tar­get activ­ity. PLoS One. 2016;11(9):e0162472.
30. Vogl DT, Kaufman JL, Hol­stein SA, et al. Modakafusp alfa (TAK-573), an
immunocytokine, shows clin­i­cal activ­ity in patients with relapsed/refrac­
tory mul­
ti­
ple mye­
loma: updated results from a first-in-human phase 1
study. Blood. 2021;138(suppl 1):898.
31. Vogl DT, Kaufman JL, Hol­stein SA, et al. TAK-573, an anti-CD38/atten­u­ated
IFNα fusion pro­tein, has clin­i­cal activ­ity and mod­u­lates the IFNα recep­tor
(IFNAR) path­way in patients with relapsed/refrac­tory mul­ti­ple mye­loma.
Blood. 2020;136(suppl 1):37-38.
32. Willert EK, Robinson GL, Higgins JP, et al. Abstract 2384: TAK-169, an excep­
tion­ally potent CD38 targeted engineered toxin body, as a novel direct
cell kill approach for the treat­ment of mul­ti­ple mye­loma [ab­stract]. Cancer
Res. 2019;79(suppl 13). Abstract 2384.
33. Bruins WSC, Zheng W, Higgins JP, et al. TAK-169, a novel recom­bi­nant
immunotoxin spe­cific for CD38, induces pow­er­ful pre­clin­i­cal activ­ity
against patient-derived mul­
ti­
ple mye­
loma cells. Blood. 2020;136(suppl
1):11-12.
34. Kamur S, Mamuye A, Dabovic K, et al. Interim results of a phase 1 study
of the novel engineered toxin body TAK-169 in patients with relapsed or
refrac­tory mul­ti­ple mye­loma [ab­stract]. J Immunother Cancer. 2021;9.
Abstract 447.
35. Bruins WSC, Zweegman S, Mutis T, van de Donk NWCJ. Targeted ther­
apy with immunoconjugates for mul­ti­ple mye­loma. Front Immunol. 2020;
(20 June)11:1155.
36. Romee R, Cooley S, Berrien-Elliott MM, et al. First-in-human phase 1 clin­i­
cal study of the IL-15 superagonist com­plex ALT-803 to treat relapse after
trans­plan­ta­tion. Blood. 2018;131(23):2515-2527.
37. Fernandez AF, Pierre-Louis L, Xu Y, et al. Restoring NK cell activ­i­ties in mul­
ti­ple mye­loma with IL-15 recep­tor ago­nist NKTR-255. Abstract presented
at: 62nd ASH Annual Meeting and Exposition; 20 Decem­ber 2020; vir­tual.
Abstract 667.
38. Hambach J, Fumey W, Stähler T, et al. Half-life extended nanobody-based
CD38-spe­cific bispecific killercell engagers induce kill­ing of mul­ti­ple mye­
loma cells. Front Immunol. 2022;13(16 May):838406.
39. Bunin A, McGrath K, Rossi AM, et al. A novel class of bifunc­tional immu­
no­ther­a­peu­tic that exploits a uni­ver­sal anti­body bind­ing ter­mi­nus (uABT)
to recruit endog­e­nous antibodies to cell expressing CD38 dem­on­strate
in vivo effi­cacy in three dis­tinct ani­mal mod­els. Blood. 2019;134(suppl 1):
1820.
40. Wang Y, Li H, Xu W, et al. BCMA-targeting bispecific anti­body that simul­ta­
neously stim­u­lates NKG2D-enhanced effi­cacy against mul­ti­ple mye­loma. J
Immunother. 2020;43(6):175-188.
41. Watkins-Yoon J, Guzman W, Oliphant A, et al. CTX-8573, an innate-cell
engager targeting BCMA, is a highly potent multispecific anti­body for the
treat­ment of mul­ti­ple mye­loma. Blood. 2019;134(suppl 1):3182.
© 2022 by The Amer­i­can Society of Hematology
DOI 10.1182/hema­tol­ogy.2022000335
Dr Prakash Singh Shekhawat
Next gen­er­a­tion of novel ther­a­pies for mye­loma | 179
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/173/2021746/173holstein.pdf by guest on 09 December 2022
9. Malek E, Hwang S, de Lima M, et al. Preclinical stud­ies and phase I trial of
vactosertib in com­bi­na­tion with pomalidomide in relapsed mul­ti­ple mye­
loma: a cor­ti­co­ste­roid-free approach by targeting TGF-β sig­nal­ing path­
way. Blood. 2019;134(suppl 1):3232.
10. Malek E, Hwang S, Caimi PF, et al. Phase Ib trial of vactosertib in com­bi­na­
tion with pomalidomide in relapsed mul­ti­ple mye­loma: a cor­ti­co­ste­roidfree approach by targeting TGF-β sig­nal­ing path­way. J Clin Oncol. 2021;
39(suppl 15):8039.
11. Lin C, Leso A, Rosenwasser M, Torngren M, Eriksson H, Nefedova Y. Inhibition of S100A9 with tasquinimod dem­on­strates potent anti-tumor activ­ity
in pre-clin­i­cal mod­els of mul­ti­ple mye­loma. Paper presented at: Euro­pean
Hematology Association; 19-20 Novem­ber 2020; vir­tual meet­ing.
12. Usmani SZ, Schjesvold F, Oriol A, et al; KEYNOTE-185 Investigators. Pembrolizumab plus lenalidomide and dexa­
meth­
a­
sone for patients with
treat­ment-naive mul­ti­ple mye­loma (KEYNOTE-185): a randomised, openlabel, phase 3 trial. Lancet Haematol. 2019;6(9):e448-e458.
13. Mateos MV, Blacklock H, Schjesvold F, et al; KEYNOTE-183 Investigators.
Pembrolizumab plus pomalidomide and dexa­meth­a­sone for patients with
relapsed or refrac­tory mul­ti­ple mye­loma (KEYNOTE-183): a randomised,
open-label, phase 3 trial. Lancet Haematol. 2019;6(9):e459-e469.
14. Gormley NJ, Pazdur R. Immunotherapy com­bi­na­tions in mul­ti­ple mye­loma—
known unknowns. N Engl J Med. 2018;379(19):1791-1795.
15. Oriol A. A crit­i­cal eval­u­a­tion of pembrolizumab in addi­tion to lenalidomide
and dexa­meth­a­sone for the treat­ment of mul­ti­ple mye­loma. Expert Rev
Hematol. 2020;13(5):435-445.
16. Matyskiela ME, Zhang W, Man HW, et al. A cereblon mod­
u­
la­
tor (CC220) with improved deg­
ra­
da­
tion of Ikaros and Aiolos. J Med Chem.
2018;61(2):535-542.
17. Bjorklund CC, Kang J, Lu L, et al. CC-220 is a potent cereblon mod­u­lat­ing
agent that dis­plays anti-pro­lif­er­a­tive, pro-apo­pto­tic and immu­no­mod­u­la­
tory activ­ity on sen­si­tive and resis­tant mul­ti­ple mye­loma cell lines. Blood.
2016;128(22):1591.
18. Lonial S, van de Donk NWCJ, Popat R, et al. First clin­i­cal (phase 1b/2a)
study of iberdomide (CC-220; IBER), a CELMoD, in com­bi­na­tion with dexa­
meth­a­sone (DEX) in patients (pts) with relapsed/refrac­tory mul­ti­ple mye­
loma (RRMM). J Clin Oncol. 2019;37(suppl 15):8006.
19. Lonial S, Popat R, Hulin C, et al. Iberdomide (IBER) in com­bi­na­tion with
dexa­meth­a­sone (DEX) in patients (pts) with relapsed/refrac­tory mul­ti­ple
mye­loma (RRMM): results from the dose-expan­sion phase of the CC-220MM-001 trial. Blood. 2021;138(suppl 1):162.
20. Van Oekelen O, Amatangelo M, Guo M, et al. Large-scale mass cytometry
reveals sig­nif­i­cant acti­va­tion of innate and adap­tive immu­nity in bone mar­
row tumor micro­en­vi­ron­ment of iberdomide-treated mye­loma patients.
Blood. 2021;138(suppl 1):730.
21. Van de Donk N, Popat R, Larsen J, et al. First results of iberdomide (IBER;
CC-220) in com­bi­na­tion with dexa­meth­a­sone (DEX) and daratumumab
(DARA) or bortezomib (BORT) in patients with relapsed/refrac­tory mul­ti­
ple mye­loma (RRMM). Blood. 2020;136(suppl 1):16-17.
22. Hansen JD, Correa M, Nagy MA, et al. Discovery of CRBN E3 ligase mod­u­
la­tor CC-92480 for the treat­ment of relapsed and refrac­tory mul­ti­ple mye­
loma. J Med Chem. 2020;63(13):6648-6676.
23. Richardson PG, Vangsted AJ, Ramasamy K, et al. First-in-human phase I
study of the novel CELMoD agent CC-92480 com­bined with dexa­meth­
a­sone (DEX) in patients (pts) with relapsed/refrac­tory mul­ti­ple mye­loma
(RRMM). J Clin Oncol. 2020;38(suppl 15):8500.
24. Wong L, Lamba M, Jiménez Nuñez MD, et al. Dose- and sched­ule-depen­dent
immu­no­mod­u­la­tory effects of the novel CELMoD agent CC-92480 in
patients with relapsed/refrac­tory mul­ti­ple mye­loma. Blood. 2020;136(suppl
1):47-48.
25. Richardson PG, Ocio E, Raje NS, et al. CC-92480, a potent, novel cereblon
E3 ligase mod­u­la­tor (CELMoD) agent, in com­bi­na­tion with dexa­meth­a­sone
(DEX) and bortezomib (BORT) in patients (pts) with relapsed/refrac­tory
Paula Rodriguez-Otero and Jesús F. San-Miguel
Clínica Universidad de Navarra, Centro de Investigación Médica Aplicada, Instituto de Investigación Sanitaria de Navarra, Centro de Investigación
Biomédica en Red Cáncer, Pamplona, Spain
Despite significant improvement in the treatment of multiple myeloma (MM), a cure remains elusive, and patients failing proteasome inhibitors, immunomodulatory drugs, and anti-CD38 monoclonal antibodies remain a challenge due to
a lack of standard of care treatment and a dismal survival rate. The development of T-cell redirecting therapies, including bispecific T-cell engagers and chimeric antigen receptor (CAR) T cells, have transformed the outcome of triple-class
exposed relapsed and refractory MM (RRMM). B-cell maturation antigen (BCMA) has proven to be an important target
in MM, and BCMA-directed CAR T cells have shown unprecedented efficacy with a prolonged duration of response in a
population with advanced RRMM, leading to the approval of 2 different BCMA CAR T-cell products. Still, and in contrast
to prior experience in the field of CD19-directed CARs, no plateau has been seen in the survival curves, and relapses
continue to occur. Therefore, further improvement is needed. Early use in the course of the disease as well as of nextgeneration CARs may further augment the efficacy of these therapies. In this review we address current state-of-the-art
approved BCMA-directed CAR T-cell therapy in RRMM, as well as potential future developments focused on optimizing
patient care and novel CAR designs.
LEARNING OBJECTIVES
• Summarize the current evidence regarding BCMA-directed CAR T cells for RRMM, focusing on approved
constructs
• Discuss potential limitations of current CAR T-cell therapy in myeloma
• Describe potential strategies for improving current results, with a focus on the clinical perspective
CLINICAL CASE
A 56-year-old male was diagnosed in September 2019 with
immunoglobulin A lambda multiple myeloma (MM). At diagnosis he presented with diffuse osteolytic lesions and a
revised International Staging System (R-ISS) 3 status (high
lactose dehydrogenase + t(4;14)). He received frontline
treatment with daratumumab in combination with bortezomib, lenalidomide, and dexamethasone (D-VRD) in the context of a randomized clinical trial. He completed 4 planned
induction cycles followed by autologous stem cell transplant and 2 additional consolidation cycles. He achieved a
very good partial response after consolidation and started
maintenance treatment with daratumumab plus lenalidomide in June 2020. Unfortunately, 1 month later he progressed with the presence of a lumbar plasmacytoma. As
of this moment, he has been transferred to our department
to be considered for a B-cell maturation antigen (BCMA)
180 | Hematology 2022 | ASH Education Program
chimeric antigen receptor (CAR) T-cell clinical trial enrolling
patients with early relapse after frontline treatment.
Introduction
The treatment landscape of MM has significantly evolved
in recent years, leading to unprecedented survival rates.1
Nevertheless, this improvement has not been uniform,
and there are still patient populations with high-risk features who have not benefited enough from recently
approved combinations and require innovative strategies
to overcome their dismal prognosis. These populations are,
among others, patients with extramedullary disease, a high
number of circulating plasma cells, high-risk cytogenetic
abnormalities (Cas), particularly double-hit myeloma, and
patients who relapse early after optimized frontline therapy.2-4 Regarding the latter, patients with R-ISS 3 status at
diagnosis who fail to achieve undetectable minimal residual
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/180/2021675/180rodriguez-otero.pdf by guest on 09 December 2022
Cellular therapy for multiple myeloma: what’s
now and what’s next
dis­ease (MRD) after front­line treat­ment are at a high risk of an
early relapse with a median pro­gres­sion-free sur­vival (PFS) of
only 14 months and a median over­all sur­vival (OS) of fewer than
2 years.5 In addi­tion, patients who have failed proteasome inhib­
i­tors, immu­no­mod­u­la­tory drugs, and anti-CD38 mono­clo­nal antibodies have very poor out­comes and lim­ited treat­ment options.6
Therefore, we need new treat­ments with novel mech­a­nisms of
action ­able to res­cue dif­fi­cult-to-treat pop­u­la­tions and improve
patient sur­vival.
BCMA CAR T cells approved: what is now
Figure 1. Several strategies are being evaluated to improve the current clinical efficacy of CAR T-cell therapy in the context of
RRMM. These strategies can be divided into patient-related factors, aspects related to the manufacturing process and CAR design,
patient management during bridging time and infusion, and finally, understanding the mechanisms of resistance and CAR failure.
CCR, chimeric costimulatory receptor; EMD, extramedullary disease; HR-CA, high-risk cytogenetic abnormalities.
Dr Prakash Singh Shekhawat
CAR T-cell ther­apy in mul­ti­ple mye­loma | 181
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/180/2021675/180rodriguez-otero.pdf by guest on 09 December 2022
In this sce­nario CAR T-cell ther­apy has shown unprec­e­dented
response rates in heavily pretreated patients, lead­ing to the
approval of 2 BCMA-directed CAR T-cell prod­ucts (ABECMA™
and CARVIKTY™) for the treat­
ment of tri­
ple-class exposed
relapsed/refrac­
tory (RR) MM who have received at least 4
prior lines of treat­ment. In this review we dis­cuss the cur­rent
state-of-the-art of approved CAR T-cell ther­apy in RRMM and
future devel­op­ments, with a focus on the clin­i­cal per­spec­tive
(Figure 1).
Idecabtagene vicleucel (ide-cel, bb2121, ABECMA™) was the
first cel­lu­lar ther­apy approved for the treat­ment of RRMM. Ide-cel
is a sec­ond-gen­er­a­tion CAR with a 4-1BB costimulatory domain
approved for the treat­ment of RRMM based on the data of the
phase 2 sin­gle-arm piv­otal KarMMa trial.7 Overall, 128 patients were
infused with 2 dif­fer­ent doses of ide-cel 150, 300, and 450×106
CAR T+ cells. The median num­ber of prior lines was 6 (3-18), and
84% of patients were tri­ple-class refrac­tory. All patients received
stan­dard lymphodepleting che­mo­ther­apy with fludarabine and
cyclo­
phos­
pha­
mide. The over­
all response rate (ORR) was 73%
among all­treated patients and 81% in those receiv­ing the higher-dose level (450×106), with 33% and 39% of com­plete responses
(CR), respec­tively. Responses were rapid, with a median time
to first response of 1.0 month (range, 0.5-8.8 months). Median
PFS was 8.8 months in the over­all pop­u­la­tion and was lon­ger in
patients receiv­ing the tar­get dose (median PFS, 12.1 months) and
in those achiev­ing CR or bet­ter (median PFS, 20.2 months; 95%
CI, 12.3-NE [not evaluable]). Median OS was 24.8 months (95% CI,
19.9-31.2). Safety over­all was man­age­able, with cyto­kine release
syn­drome (CRS) reported in 84% of patients across the 3 doses
and 96% at the tar­get dose. The major­ity of events were grade
1 or 2 with a CRS grade higher than or equal to 3 in fewer than
6% of patients, with a median time to onset of 1 day (range, 1-12).
Neurological com­pli­ca­tions were less com­mon, were reported in
(95% CI, 57.6-80.9) in sCR patients. Two-year OS was 71.0% (95% CI,
57.6-80.9). Regarding safety, cytopenias and CRS were the most
fre­quent treat­ment-related adverse events reported. Any grade
of neutropenia was pres­ent in 96% of patients, with 95% hav­ing
grade 3 or higher. The inci­dence of CRS was 95%, mostly grade
1 to 2, with a median time to onset of 7 days (range, 1-12). Neurological events were reported in 20 patients. Among these, 16 presented immune-effec­tor cell–asso­ci­ated neu­ro­tox­ic­ity syn­drome
(ICANS), and 12 patients presented with other neurotoxicities. Of
these, 5 expe­ri­enced a move­ment and neurocognitive dis­or­der
with a median time to onset of 27.0 days (IQR 16.0-73.0) (Table 1).8
Other BCMA-directed CARs are under devel­op­ment, and the
sta­tus of those stud­ies are sum­ma­rized in Table 2.
Table 1. Safety and effi­cacy data of the 2 FDA- and EMA-approved CARs
Idecabtagene vicleucel KarMMa trial1-3
Ciltacabtagene autoleucel CARTITUDE 1 trial
Number of patients infused, n
128 (140 apheresed)
97 (120 apheresed)
Phase
2
1b/2
Target/costimulation
BCMA/4-1BB
BCMA/4-1BB
*2 BCMA-targeting heavy-chain anti­body
scFv
Chimeric mouse
Chimeric llama
Specificity
Autologous
Autologous
Follow-up, median (range)
13.3 mo (0.2-21.2)
21.7 mo (not reported)
Prior lines, median (range)
6 (3 to 16)
6 (3-18)
Triple-class refrac­tory, n (%)
108 (84)
85 (87.6)
Penta-exposed, n (%)
77 (60)
81 (83.5)
Bridging ther­apy, n (%)
112 (88)
73 (75)
Response to bridg­ing ther­apy, n (%)
5/112 (4)
33/73 (45)
EMD, n (%)
50 (39)
CAR T-cell dose
150, 300, 450×10 CAR+T
0.75×106 CAR T+/kg
LD che­mo­ther­apy
Fludarabine 30mg/m2×3 d
Cyclo 300mg/m2×3 d
Fludarabine 30mg/m2×3 d
Cyclo 300mg/m2×3 d
ORR, n (%)
94 (73)
At 450×106 (n=54): 44 (81%)
95 (97.9)
CR or sCR, n (%)
42 (33)
At 450×106 (n=54): 21 (39)
82.5% (sCR)
Time to first response, median (range)
1.0 mo (0.5-8.8)
1.0 mo (IQR 0.9-1.0)
MRD in CR, n (%)
33/42 MRD neg (10 )
DOR, median (95% CI)
10.7 mo (9.0-11.3)
Not reported
PFS, median (95% CI)
8.8 mo (5.6-11.6)
At 450×106: 12.1 mo (8.8-12.3)
NR (16.8-NE)
2-y PFS: 60.5% (48.5-70.4)
13 (13)
6
−5
PFS in high-risk, median (95% CI)
61 evaluable
92% MRD neg­a­tive (10−5)
PFS at 2y, % (95% CI)
ISS 3/R-ISS 3
4.9 mo (1.8-8.2)
NE (NE-NE)
High-risk CA
8.2 mo (4.8-11.9)
48.4% (25.1-68.4)
Plasmacytomas
7.9 mo (5.1-10.9)
47.4% (24.4-67.3)
OS, median (95% CI)
24.8 mo (19.9-31.2)
NR (27.2-NE)
107 (84)
92 (95)
a
CRS, n (%)
Overall
7 (5)
4 (4)
Time to CRS onset, median (range)
Grade 3-4
1 d (1-12)
7 d (5-8)
Duration of CRS, median (range)
5 d (1-63)
4 d (3-6)
182 | Hematology 2022 | ASH Education Program
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/180/2021675/180rodriguez-otero.pdf by guest on 09 December 2022
18% of patients, and were mostly grade 1 to 2. On the other hand,
cytopenias were fre­quent and not dose related, with a median
time to recov­ery from grade 3 or higher neutropenia and throm­
bo­cy­to­pe­nia of 2 and 3 months, respec­tively (Table 1).7
Ciltacabtagene autoleucel (cilta-cel; for­
mer LCAR-B38M,
CARVIKTY™), another sec­ond-gen­er­a­tion CAR T cell containing
2 BCMA-bind­ing domains, was inves­ti­gated in the CARTITUDE-1
study (and pre­vi­ously eval­u­ated in the LEGEND-2 phase 1 study).
Overall, 97 patients were infused with a median num­ber of 6 prior
lines and 87.6% were tri­ple class refractory. The ORR was 97.9%
with 82.5% of patients achiev­ing a strin­gent CR (sCR). With a
median fol­low-up of 2 years, out­stand­ing sur­vival rates have been
reported, with a 2-year PFS of 60.5% (95% CI, 48.5-70.4) and 71.0%
Table 1. Safety and effi­cacy data of the 2 FDA- and EMA-approved CARs (Continued)
Idecabtagene vicleucel KarMMa trial1-3
Ciltacabtagene autoleucel CARTITUDE 1 trial
Overall
23 (18)
20 (21)b
Grade 3-4
4 (3)
9 (9)
Time to neu­ro­tox­ic­ity onset, median (range)
2 d (1-10)
8 d (6-8) for ICANs
Other: CAR-T cell neurotoxicites:
26.5 d (11-108)b
Grade 3-4 neutropenia, n (%)
114 (89)
92 (95)
1.9 mo (1.2-5.6)
Not reported
67 (52)
58 (60)
2.1 mo (1.2-13.8)
Not reported
Infections, n (%)
88 (69)
56 (58)
Grade 3-4 infec­tions, n (%)
28 (22)
19 (20)
Death, n (%)
44 (34)
14
Reference
(1) Munshi et al9
(2) Anderson et al.34
(3) Oriol et al.35
(4) Berdeja et al8
(5) Martin et al.36
(6) Jakubowiak et al11
Neurotoxicity, n (%)
Grade 3-4 throm­bo­cy­to­pe­nia, n (%)
Time to recov­ery, median (range)
Data from the KarMMa and CARTITUDE-1 piv­otal tri­als with ide-cel and cilta-cel, respec­tively, are sum­ma­rized.
a
PFS in R-ISS 3.
b
ICANS occurred in 16 (17%) patients. Other neurotoxicities occurred in 12 patients, with a median time to onset of 27 days. Five patients had a clus­ter
of move­ment and neurocognitive treat­ment–emer­gent adverse events.
EMA: Euro­pean Medicines Agency; EMD, extramedullary dis­ease; FDA: US Food and Drug Administration.
What is next and how to improve cur­rent results
Despite the impres­
sive clin­
i­
cal results obtained thus far and
given the short fol­low-up in some stud­ies, no pla­teau has yet
been seen in the sur­vival curves, and relapses con­tinue to occur.
Therefore, fur­ther improve­ment is needed, and sev­eral points
remain to be discussed.
Improving patient selec­tion
Patient selec­
tion is a crit­
i­
cal aspect of CAR T-cell ther­
apy.
Indeed, it is of par­tic­u­lar rel­e­vance if we con­sider the pres­ent
and future avail­abil­ity of other off-the-shelf BCMA-directed ther­
a­pies. Although high and deep responses are achieved with
both approved ther­a­pies, out­comes in selected high-risk pop­
u­la­tions remain infe­rior com­pared to stan­dard-risk patients. In
the KarMMa trial, patients with extramedullary dis­ease, high-risk
CAs, and R-ISS 3 showed lower CR rates and shorter PFS com­
pared to patients with­out high-risk dis­ease (Table 1).9,10 Likewise,
in the CARTITUDE-1 trial, patients with high-risk CAs, ISS 3, and
the pres­ence of plasmacytomas showed a lower PFS at 2 years
com­pared to the inten­tion-to-treat pop­u­la­tion, suggesting that
fur­ther improve­ment is needed to abro­gate the adverse out­
come of patients with high-risk fea­tures even in the con­text of
potent ther­a­pies such as CAR T cell (Table 1).11
On top of this, data from the dif­fer­ent stud­ies sug­gest that
the achieve­ment of deep responses (CR or MRD neg­a­tiv­ity) is
of utmost impor­tance in the con­text of “one-shot” CAR T-cell
ther­apy and is asso­ci­ated with prolonged dura­tion of response
(DoR) and PFS.8,9 Therefore, under­stand­ing the patient char­ac­
ter­is­tics asso­ci­ated with a higher like­li­hood of achiev­ing CR is of
great impor­tance. A recent anal­y­sis of the KarMMa trial showed
that the pres­ence of high tumor bur­den (sol­u­ble BCMA) and
high inflam­ma­tion (D-dimer, fer­ri­tin) along with the mye­loma
sub­type (immu­no­glob­u­lin G heavy chain) had a neg­a­tive asso­
ci­a­tion with CR/sCR achieve­ment, whereas a high vec­tor copy
num­ber in the drug prod­uct showed a pos­i­tive cor­re­la­tion with
the CR/sCR rate.12 These fac­tors sug­gest that opti­mal and timely
patient selec­tion may be rel­e­vant for CAR T-cell out­comes since
it is not an off-the-shelf ther­apy, and rap­idly pro­gres­sive patients
with high-risk fea­tures or lim­ited bridg­ing options may be dif­fi­
cult to man­age given the manufactur­ing time. In fact, between
10% and 20% of the patients who under­went apher­e­sis in the
dif­fer­ent tri­als did not reach infu­sion due to com­pli­ca­tions dur­
ing the manufactur­ing time or dis­ease pro­gres­sion.
In this regard, ade­quate selec­tion of bridg­ing ther­apy is of
utmost impor­tance to main­tain the patient’s con­di­tion dur­
ing manufactur­ing. Whenever pos­si­ble, agents to which the
patient has not been pre­vi­ously exposed should be pri­or­i­tized.
In this sense, and although the data are lim­ited, BCMA-targeted agents are prob­a­bly not the best option dur­ing the
bridg­ing period. Indeed, patients with prior BCMA treat­ment
were excluded from the piv­otal CAR T stud­ies, and lim­ited data
are there­fore avail­­able. Recent evidence from the cohort C
of the CARTITUDE-2 trial that enrolled patients previously
treated with BCMA-antibody drug conjugates (BCMA-ADC)
and BCMA biespecific antibodies (BCMA-BsAb) showed that
responses to cilta-cel were inferior to that of the CARTITUDE-1
trial (ORR: 62% in the BCMA-ADC and 57% in the BCMA-BsAb
groups, respectively). Factors related to response were a
shorter duration of prior BCMA treatment and a longer median
time between last BCMA treatment and apheresis, suggesting a potential negative impact of prior BCMA treatment in
outcome with CAR T-cell therapies.13 On the other hand, in
the recently reported real-world expe­ri­ence, 22% of patients
had prior anti-BCMA ther­apy, and the ORR was con­sis­tent with
that of the KarMMa trial.14 Even so, both the downregulation of
BCMA expres­sion and the biallelic loss of the BCMA locus have
Dr Prakash Singh Shekhawat
CAR T-cell ther­apy in mul­ti­ple mye­loma | 183
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/180/2021675/180rodriguez-otero.pdf by guest on 09 December 2022
Time to recov­ery, median (range)
Dr Prakash Singh Shekhawat
Autologous
150-450M
Specificity
CAR T-cell dose
50 (69)
NR
Penta-refrac­tory, n (%)
36%
4 (1-10)
NCT03915184
NCT03430011
NCT03274219
Identification
Data of safety and effi­cacy are shown.
NR, not reached.
Chen et al.38
Mailankody
et al.37
NR
NR
1 patient
2 (DL0) & 1 (DL1)
≈30/20
4 (DL0) & 3 (DL1)
2 (DL0) & 1 (DL1)
0 (0)
15/18 (83.3)
Raje et al27
Reference
27.8%
NR
Ongoing
4 (1-10)
NR
50%
NCT05066646
Chunrui et al.39
Ongoing
1
10
0
1 (1.3)
20/34.7
5 (1-30)
6 (1-12)
2 (2.5)
72 (92.4)
25.3 mo (3.0-NE)
58.2%
75 (94.9)
13 (16.5)
85%
94%
5 (3-13)
5 (3-11)
Not fur­ther
devel­oped
2d
Not fur­ther
devel­oped
Status
Duration (d), median (range)
4 (1-6)
2 (3)
3 (4)
7 d (2-24)
Grade 3-4, n (%)
8 (13)
76/52
11 (15)
Onset (d), median (range)
All grade, n (%)
ICANS
4d
53/17
Duration (d), median (range)
Tocilizumab/ste­roids %/%
2 (1-4)
2 (3)
3 (2 grade 5)
2 d (1-20)
Grade 3-4, n (%)
55 (89)
54 (75)
9.3 mo (300M)
36%
12.8 mo (7.3-18.6)
92%
30 (48)
69%
Onset (d), median (range)
All grade, n (%)
CRS
PFS (mo), median (95% CI)
CR, n (%)
ORR, n (%)
Efficacy
6 (3-18)
6 (3-17)
Prior lines, median (range)
Triple-class refrac­tory, n (%)
58 (94)
61 (33-77)
56 (39-70)
Autologous
1.0×106/kg
Autologous
1.5-3.0×108 cells
Autologous
300-600M
62 (36-78)
BCMA/4-1BB
Human
BCMA/4-1BB
1/2
25.3 mo (4.1-36.7)
1b/2
CT103A
FUMANBA-1 (n=79)
6 mo (2-11)
CT053 LUMMICAR
(n=18)
Human
BCMA/4-1BB
62 (33-76)
Age, median (range)
Population
Mouse
BCMA/4-1BB
Mouse
Target/costimulation
scFv
1
9.5 mo
1
23 (9-46)
Phase
Follow-up, median (range)
Orva-cel
EVOLVE (n=62)
BB21217
CRB-402 (n=72)
Table 2. Main reported clin­i­cal tri­als of BCMA-directed CAR T cells in RRMM
NCT05066646
Gan An, et al.40
Ongoing
1
8
0 (0)
1 (4.3)
26/9
5 (2-9)
6 (1-11)
1 (4.3)
21 (91.3)
6 mo PFS: 65.1%
43.5%
22 (95.7)
NR
NR
4 (2-12)
60 (45-74)
1.0-6.0×106/kg
Autologous
Human
BCMA/4-1BB
6.2 mo (0.7-16.1)
1
C-CAR088 (n=23)
1
BCMA/4-1BB
Autologous
—
de Larrea et al24
NCT04309981
Mailankody et al.42
NCT04093596
Costello et al.41
NCT03288493
0
—
0 (0)
NR
Ongoing
0
6 (14)
NR
4 (1-12)
76/12
0 (0)
NR
1 (2)
NR
NR
87%
24 (56)
23/14
60%
15.8 mo
(12.9-NE)
25%
NR
100%
NR
61%
4 (2-10)
61 (36-74)
3×106/kg
71%
At 320×106
42%
NR
5 (3-11)
64 (46-77)
40-480×106
Allogeneic
BCMA/4-1BB
Human
Ongoing
NR
BCMA/4-1BB
Human
13 mo (7-13)
1
ARI-002h (n=30)
Not fur­ther
devel­oped
NR
4%
4%
7/6
NR
NR
0%
17%
NR
NR
50%-75%
NR
60%
8 (2-18)
60 (42-74)
51-1178×106
Autologous
Mouse
4 mo
1/2
NA
ALLO-715
UNIVERSAL (n=43)
P-BCMA-101
PRIME (n=53)
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/180/2021675/180rodriguez-otero.pdf by guest on 09 December 2022
184 | Hematology 2022 | ASH Education Program
CLINICAL CASE (Con­t in­u ed)
Our patient under­
went apher­
e­
sis after the com­
ple­
tion of
screen­
ing tests and received bridg­
ing ther­
apy, with pro­
gres­sive dis­ease as the best response. The bridg­ing period
was com­pli­cated with acute kid­ney fail­ure and sev­eral infec­
tious com­pli­ca­tions. After recov­ery, he received stan­dard
fludarabine-cyclo­phos­pha­mide con­di­tion­ing before CAR
T-cell infu­sion, which was performed in Octo­ber 2020. He did
not develop CRS or ICANs but did have grade 4 cytopenia
requir­ing trans­fu­sion sup­port, granulocyte col­ony-stim­u­lat­ing
fac­tor, eryth­ro­poi­e­tin-stim­u­lat­ing agents, and thrombopoietin
ana­logues. After CAR T-cell treat­ment, the patient achieved a
strin­gent CR with MRD neg­a­tiv­ity (10−6) as well as a com­plete
met­a­bolic response (Deauville 3), using an F-fluorodeoxyglucose emis­sion tomog­ra­phy/com­puted tomog­ra­phy scan, that
con­tin­ues today. His cytopenia improved by month 12, and
the patient no lon­ger requires trans­fu­sions or med­i­ca­tion.
Improving the T-cell manufactur­ing pro­cess
Several stud­ies across dif­fer­ent CAR T-cell prod­ucts have shown
a pos­i­tive cor­re­la­tion between the depth and DoR and CAR
T-cell expan­sion or per­sis­tence.9,17 Therefore, sig­nif­i­cant effort
has been fos­tered to opti­mize the final CAR T-cell prod­uct. One
rel­e­vant aspect to con­sider is the immu­no­ge­nic­ity of approved
ani­mal-derived con­structs. Indeed, in the KarMMa trial, patients
with anti­drug antibodies (ADAs) increased over time, from 21%
(21 of 102) at month 3 to 65% (34 of 52) at month 12.9 Although
these antibodies did not seem to influ­ence CAR con­cen­tra­tion
or response, only patients with­out neg­a­tive ADAs responded to
CAR T-cell reinfusion, suggesting the poten­tial role of these ADAs
in CAR T-cell per­sis­tence or activ­ity. Humanized or fully human
con­structs may be a way to over­come the neg­a­tive impact of
immu­no­ge­nic­ity in CAR T-cell out­comes. Preliminary data from
sev­eral phase 1 and 2 tri­als have shown encour­ag­ing clin­i­cal
effi­cacy with high ORRs and deep and dura­ble responses.22,23
Moreover, in 1 aca­demic study eval­u­at­ing ARI-002h, a human­
ized BCMA CAR T, patients with an ongo­ing response and the
absence of sig­nif­i­cant tox­ic­ity to the first infu­sion were eli­gi­ble
for a reinfusion. Sixteen out of 30 patients received a sec­ond
infu­sion, and 5 patients showed an improved response, suggesting an ade­quate expan­sion of the CAR T-cell prod­uct.24,25
Alternatively, the use of non–scFv-based CARs, such as engineered bind­ing scaf­folds, recep­tors, or nat­u­ral ligands, might
addi­tion­ally serve to mit­i­gate immu­no­genic reac­tions.26
Higher lev­els of mem­ory T-cell phe­no­types have been asso­ci­
ated with bet­ter clin­i­cal out­comes, and sev­eral strat­e­gies, such
as CD4/CD8 in a 1:1 ratio, may yield a higher pro­por­tion of these
pop­u­la­tions in the final prod­uct. An inter­est­ing approach was
eval­u­ated in the phase 1 CRB-402 study. In this trial, ide-cel was
cul­tured with a PI3K inhib­i­tor (bb007), resulting in a dif­fer­ent CAR
T-cell prod­uct (bb21217) enriched with T cells hav­ing a mem­orylike phe­no­type and a supe­rior pro­lif­er­a­tive capac­ity upon adop­
tive trans­fer. Overall, 72 patients with a median of 6 prior lines
were infused. The ORR was 69%, with 36% of the patients achiev­
ing a CR or bet­ter. The median DoR was 23.8 months and was
not reached by patients in CR. Importantly, the per­sis­tence of
CAR T cells was prolonged, with detect­able CAR T cells in some
patients up to 24 months.27
One lim­i­ta­tion of targeted immunotherapies is the poten­tial
het­ero­ge­ne­ity of anti­gen expres­sion among tumor cells and the
loss or downregulation of anti­
gen after expo­
sure.18 Simultaneously targeting dif­fer­ent anti­gens could over­come this bar­rier,
and sev­eral tri­als are eval­u­at­ing this pos­si­bil­ity. GC012F is a dual
BCMA/CD19-targeted CAR T manufactured in 24 to 36 hours on
the FASTCAR® plat­form. Sixteen patients with RRMM and a median
of 5 prior lines of ther­apy were infused with GC012F (dose range:
100 000-300 000 CAR T cells per kilo­gram). The ORR was 94%,
with 56.3% of patients achiev­ing CR or bet­ter. The median DoR
was not reached at 7.3 months of median fol­low-up.28
Likewise, other tar­gets are being explored (Table 4). GPR5D
(G pro­tein–cou­pled recep­tor 5 fam­ily D)-directed CAR T cells are
Dr Prakash Singh Shekhawat
CAR T-cell ther­apy in mul­ti­ple mye­loma | 185
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/180/2021675/180rodriguez-otero.pdf by guest on 09 December 2022
been reported after anti-BCMA treat­ment, with the poten­tial
to limit BCMA CAR T-cell effi­cacy.15-17
Additionally, strat­e­gies to shorten vein-to-vein time are being
devel­oped, such as plat­forms for rapid CAR T-cell manufactur­
ing or uni­
ver­
sal off-the-shelf allo­
ge­
neic CAR. ALLO-715 is an
allo­ge­neic BCMA-targeting CAR T cell. It is manufactured with
a knock­out of the T-cell recep­tor alpha con­stant and CD52 to
min­i­mize the risk of graft-ver­sus-host dis­ease and to allow pro­
found T-cell deple­tion using an anti-CD52 anti­body to improve
engraft­ment of the CAR T cells, respec­tively.18 The UNIVERSAL
phase 1 study tested ALLO-715 in 43 RRMM patients using esca­
lat­ing doses and dif­fer­ent lympho-deplet­ing reg­i­mens. The best
clin­i­cal results were obtained with 320 × 106 cells in com­bi­na­
tion with fludarabine, cyclo­phos­pha­mide, and anti-CD52 anti­
body (ALLO-647). In this cohort, 24 patients were infused. The
ORR was 71%, and the median DoR was 8.3 months.19 Potential
immu­no­ge­nic­ity and short per­sis­tence are the main chal­lenges
when using allo­genic CAR T cells, and other cel­lu­lar plat­forms
are under inves­ti­ga­tion, such as nat­u­ral killer CAR or γδ T cells.18
Treating patients in ear­lier lines of ther­apy with less refrac­
tory dis­ease may help over­come some of the afore­men­tioned
prob­lems. Therefore, dif­fer­ent stud­ies are indeed eval­u­at­ing
the fea­si­bil­ity and effi­cacy of this ther­apy in the con­text of early
relapse or front­line dis­ease. In cohort A from the phase 2 CARTITUDE-2 study, patients with 1 to 3 prior lines and lenalidomiderefrac­tory dis­ease were included and treated with cilta-cel. The
median num­ber of prior lines was 2 (range, 1-3). Interestingly,
40% of the patients were already tri­ple-class refrac­tory. Nineteen out of 20 patients responded (95%), and 85% achieved at
least CR. The median PFS has not yet been reached, with a PFS at
12 months of 84% (95% CI, 59.1-94.7). Safety was com­pa­ra­ble to
that of the CARTITUDE-1 trial with­out new safety sig­nals.20 Likewise, in cohort B of the same study, patients with early relapse
(within the first 12 months) after front­line ther­apy with a proteasome inhibitor and immu­no­mod­u­la­tory drugs were included. A
total of 19 patients were infused. The ORR was 95%, and 79% of
patients achieved at least CR, with a 12-month PFS of 84% (95%
CI, 57.9-94.5) and a safety pro­file com­pa­ra­ble to that of the
CARTITUDE-1 trial.21 Altogether, these results sug­gest that CAR
T-cell ther­apy in ear­lier lines of ther­apy is safe and may yield high
and deep responses in dif­fer­ent unmet-need pop­u­la­tions. Still,
its effi­cacy needs to be con­firmed in the con­text of large phase
3 stud­ies both in early relapse (KarMMa-3, NCT03651128; CARTITUDE-4, NCT04181827) and the front­line set­ting (BMTCTN1902,
NCT05032820; CARTITUDE-5, NCT04923893; CARTITUDE-6,
NCT04181827) (Table 3).
Dr Prakash Singh Shekhawat
NDMM
high-risk
NDMM
NDMM not
intended for
ASCT
NDMM-TE
high-risk
NDMM TE
NDMM
high-risk
(R-ISS 3)
Multicohort
Multicohort
1-3 PL + Len R
≥3 PL RRMM
2-4 prior lines
Len-ref & antiCD38 exposed
1-3 prior lines
Len refrac­tory
NDMM postASCT
FUMANBA-2
(NCT05181501)
NCT04287660
CARTITUDE-5
NCT04923893
NCT04935580
CARTITUDE-6
NCT05257083
KarMMa-4
NCT04196491
CARTITUDE-2
NCT04133636
KarMMa-2
NCT03601078
KarMMa-7
NCT04855136
KarMMa-3
NCT03651128
CARTITUDE-4
NCT04181827
BMTCTN1902
NCT05032820
Ide-cel + Len main­te­nance
A: PVd or DPd
B: cilta-cel
A: DVd/DPd/
EloPd/Ixa-Rd/Kd
B: Ide-cel
(A): Ide-cel + iberdomide
main­te­nance
(B): Ide-cel + gamma
secretase inhib­i­tor
2a + b: R-ISS 3 + early relapse
2c: sub­op­ti­mal response to
ASCT
Cohort A: 1-3 prior lines
Len-ref
Cohort B: early relapse after
front­line treat­ment
Cohort C: relapse after BCMA
Cohort F: NDMM after front­
line
Cohort E: NDMM-Tx not
planned (high-risk)
Standard induc­tion×3 cycles
+ Ide-cel
A: DVRD×4 + ASCT +2×DVRD
+ Len
B: DVRD×6 + cilta-cel + Len
(2y)
2 cycles induc­tion + CAR T +
Len main­te­nance
A: VRD×8 + Rd
B: VRD (6+2) + cilta-cel
Ide-cel
Cilta-cel
Ide-cel
Ide-cel
Murine
Llama
Murine
Murine
Murine
Llama
Cilta-cel
Ide-cel
Murine
Ide-cel
Llama
Human
GC012F
Cilta-cel
Llama
Cilta-cel
NA
NA
BiRD + BCMA CAR T cell
Human
CT103A
VRD/PAD/PCD×3-ASCT
(if eli­gi­ble) >> CT103A
ScFv
ori­gin
CAR
name
Study design
2
3 ran­dom­ized
3 Randomized
2
2
2
1
3 ran­dom­ized
1/2
3 ran­dom­ized
3
1
Phase
40
419
381
181
181
157
13
750
20
650
20
20
n
BCMA
BCMA
BCMA
BCMA
BCMA
BCMA
BCMA
BCMA
BCMACD19
BCMA
BCMA
BCMA
Antigen
Recruiting
Recruiting
Recruiting
Active Not
recruiting
Recruiting
150-450×106
150-450×106
450×106
0.75×106/kg
450×106
ORR
ORR
PFS
PFS
Autologous
Autologous
Autologous
Autologous
Autologous
ORR
Recruiting
0.75×106/kg
Autologous
MRD neg­a­tiv­ity at
1 year
Not yet
enroll­ing
0.75×106/kg
Autologous
Recruiting
3×105/kg
Active Not
recruiting
Recruiting
Recruiting
0.75×106/kg
450×106
PFS
Safety
ORR, PFS, MRD
PFS
Not yet
enroll­ing
1×106/kg
2-3×107/kg
Status
Dose
Safety
Autologous
Autologous
Autologous
ORR at 4 weeks
MRD neg­a­tiv­ity
2-year PFS
Autologous
Autologous
End point
Cell source
ASCT, autol­o­gous stem cell trans­plan­ta­tion; BiRD, bortezomib, claritromicine, lenalidomide, dexa­meth­a­sone; DPd, daratumumab, pomalidomide, dexa­meth­a­sone; DVd, daratumumab,
bortezomib, dexa­meth­a­sone; EloPd, elotuzumab, pomalidomide, dexa­meth­a­sone; IxaRd, izaxomib, lenalidomide, dexa­meth­a­sone; Kd, carfilzomib, dexa­meth­a­sone; Len, lenalidomide; Len-ref,
lenalidomide refrac­tory; NA, not avail­­able; NDMM, newly diag­nosed mul­ti­ple mye­loma; NDMM-TE, newly diag­nosed trans­plant eli­gi­ble; PAD, bortezomib, adryamicine, dexa­meth­a­sone; PCD,
bortezomib, cyclo­phos­pha­mide, dexa­meth­a­sone; PVd, pomalidomide, bortezomib, dexa­meth­a­sone; Rd, lenalidomide-dexa­meth­a­sone; Tx, transplant.
Study
pop­u­la­tion
Trial
Table 3. Summary of ongo­ing stud­ies eval­u­at­ing CAR T cell ther­apy in ear­lier lines of ther­apy or in com­bi­na­tions with stan­dard of care treat­ment
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/180/2021675/180rodriguez-otero.pdf by guest on 09 December 2022
186 | Hematology 2022 | ASH Education Program
CAR T-cell ther­apy in mul­ti­ple mye­loma | 187
Peter MacCallum
Cancer Centre,
Australia
Zhejiang University
Tongji Hospital
Memorial Sloan
Kettering Cancer
Center
NCT01716364
NCT05016778
NCT05219721
NCT04555551
MCARH109
CAR-GPR5D
GPR5D-CAR T
LeY
NA
CAR2 antiCD38 A2
NKG2D-CAR
UCARTCS1
NA
κ.CARTs
CART-138
Human
Human
NA
NA
NA
NA
Human
NA
NA
1
1
1
1
1
1
17
18
15
6
NA
72
12
18
1
1
42
1
5
7
1/2
Murine
10
38
n
1
1
Murine
Murine
1
Phase
Murine
ScFv ori­gin
GPRC5d
GPRC5d
GPRC5d
Lewis Y
CD44v6 (+HSVTK sui­cide
gene)
CD38
NKG2D ligands
SLAMF7
(TALEN-targeted
gene
editing TCR
and SLAMF7)
SLAMF7
κ light chain
CD138
CD19
SLAMF7
Antigen
4-1BB
4-1BB
NA
CD28
CD28
NA
DAP10
4-1BB
CD28 or 4-1BB/
CD3z + induc­ible
caspase 9 (IC9)
cell sui­cide
CD28
4-1BB
4-1BB
NR
Costimulatory
domain
Autologous
Autologous
Autologous
Autologous
Autologous
Autologous
Autologous
Allogeneic
Autologous
Autologous
Autologous
Autologous
Autologous
Cell source
NA, not avail­­able; NR, not reached; PD, pro­gres­sive dis­ease; PR, par­tial response; SD, sta­ble dis­ease; VGPR, very good par­tial response.
EU Horizon 2020
Program
Cellectis S.A.
NCT04142619
EU-CART
National Cancer
Institute
NCT03958656
Sorrento
Therapeutics
Baylor College of
Medicine
NCT00881920
NCT03464916
Chi­nese PLA
General Hospital
NCT01886976
Celyad
University of
Pennsylvania
NCT02135406
NCT02203825
CARAMBA
Euro­pean Union &
CARAMBA
NCT04499339
CTL019
CAR name
Sponsor
Trial
Table 4. Clinical tri­als in mul­ti­ple mye­loma eval­u­at­ing non–BCMA-directed CAR T-cell con­structs
1.1-6×108
Lentiviral
NA
NA
NA
NA
1.0, 3.0, or
6.0×106/kg
0.5, 1.0, and
2.0×106/kg
25, 50, 150,
450×106
Retroviral
Retroviral
NA
NA
Lentiviral
ORR 69%
CR 25%
NA
NA
NA
NA
NA
NA
NA
NA
4 SD
1-3×107
NA
0.3-12.0×106
2.0×108
SD: 4; PD: 1
VGPR: 6; PR:
2; PD: 2
NR
Efficacy
Retroviral
NA
NA
Retroviral
0.44-1.51×107
NA
Sleeping
beauty
Retroviral
Dose
Transfer
method
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/180/2021675/180rodriguez-otero.pdf by guest on 09 December 2022
Dr Prakash Singh Shekhawat
also under inves­ti­ga­tion. In a first-in-human study of MCARH109,
17 patients were infused. Four doses were tested (25, 50, 150,
and 450×106 CAR+T cells). Fifty-nine per­cent of the patients had
prior BCMA ther­apy, includ­ing 47% with prior CAR T-cell treat­
ment. The ORR was 60% and 80% in patients with prior BCMA
ther­apy. Data are still pre­lim­i­nary since the median fol­low-up
was only 18 weeks.29
In addi­
tion, dif­
fer­
ent strat­
e­
gies aiming to improve CAR
potency and sen­si­tiv­ity are under eval­u­a­tion, and the pos­si­bil­
i­ties are end­less. Therefore, the fine-tun­ing of CAR den­sity in
the T cell,30 the co-infu­sion of CAR and chi­me­ric costimulatory
recep­tors,31 the use of HLA-inde­pen­dent T-cell recep­tors instead
of CAR,32 armored CAR T, or a com­bi­na­tion with gamma-secretase inhib­i­tors to increase BCMA den­sity in the tumor cell are
poten­tial ven­ues to improve.33 Clinical data, how­ever, are still
scanty (Figure 2).
Conclusions
In con­clu­sion, BCMA-directed CAR T-cell ther­a­pies have dem­
on­strated impres­sive clin­i­cal results in the con­text of advanced
tri­ple-class exposed RRMM, lead­ing to the approval of 2 CARs.
New modal­i­ties are in devel­op­ment to over­come some of the
cur­rent lim­i­ta­tions of CAR T-cell ther­apy, aiming to improve out­
comes, shorten vein-to-vein time, and reduce tox­ic­ity. The fur­
ther use of BCMA-directed CAR T-cell treat­ment in ear­lier lines of
ther­apy is sure to improve out­comes and, hope­fully, find the way
to cur­ing MM patients.
Conflict-of-inter­est dis­clo­sure
Paula Rodriguez-Otero: hon­o­raria: Abbvie, Amgen, Bristol Myers
Squibb/Celgene, GSK, Janssen, Kite Pharma, Oncopeptides,
Pfizer, Sanofi; con­sul­tancy: Celgene, GSK, Janssen, Pfizer.
Jesús F. San-Miguel: hon­o­raria: Abbvie, Amgen, Bristol Myers
Squibb, Celgene, Janssen, MSD, Novartis, Sanofi, Roche, Takeda.
188 | Hematology 2022 | ASH Education Program
Off-label drug use
Paula Rodriguez-Otero: nothing to disclose.
Jesús F. San-Miguel: nothing to disclose.
Correspondence
Paula Rodríguez-Otero, Clinica Universidad de Navarra, Av de PIO
XII, 36, 31008 Pamplona, Spain; e-mail: paurodriguez@unav​­.es.
References
1.
Fonseca R, Abouzaid S, Bonafede M, et al. Trends in over­all sur­vival and
costs of mul­ti­ple mye­loma, 2000-2014. Leukemia. 2017;31(9):1915-1921.
2. Walker BA, Mavrommatis K, Wardell CP, et al. A high-risk, dou­ble-hit, group
of newly diag­nosed mye­loma iden­ti­fied by geno­mic anal­y­sis. Leukemia.
2019;33(1):159-170.
3. Usmani SZ, Rodriguez-Otero P, Bhutani M, Mateos MV, Miguel JS. Defining
and treating high-risk mul­ti­ple mye­loma. Leukemia. 2015;29(11):2119-2125.
4. Garcés JJ, Bretones G, Burgos L, et al. Circulating tumor cells for com­pre­
hen­sive and mul­ti­re­gional non-inva­sive genetic char­ac­ter­iza­tion of mul­ti­ple
mye­loma. Leukemia. 2020;34(11):3007-3018.
5. Paiva B, Puig N, Cedena MT, et al. Measurable resid­ual dis­ease by next-gen­
er­a­tion flow cytom­e­try in mul­ti­ple mye­loma. J Clin Oncol Off J Am Soc Clin
Oncol. 2020;38(8):784-792.
6. Gandhi UH, Cornell RF, Lakshman A, et al. Outcomes of patients with mul­
ti­ple mye­loma refrac­tory to CD38-targeted mono­clo­nal anti­body ther­apy.
Leukemia. 2019;33(9):2266-2275.
7. Munshi NC, Anderson LD, Shah N, et al. Idecabtagene vicleucel (ide-cel;
bb2121), a BCMA-targeted CAR T-cell ther­apy, in patients with relapsed
and refrac­tory mul­ti­ple mye­loma (RRMM): ini­tial KarMMa results [ab­stract].
J Clin Oncol. 2020;38(suppl 15). Abstract 8503.
8. Berdeja JG, Madduri D, Usmani SZ, et al. Ciltacabtagene autoleucel, a B-cell
mat­u­ra­tion anti­gen-directed chi­me­ric anti­gen recep­tor T-cell ther­apy in
patients with relapsed or refrac­
tory mul­
ti­
ple mye­
loma (CARTITUDE-1): a
phase 1b/2 open-label study. Lancet (London, England). 2021;398(10297):314324.
9. Munshi NC, Anderson LDJ, Shah N, et al. Idecabtagene vicleucel in relapsed
and refrac­tory mul­ti­ple mye­loma. N Engl J Med. 2021;384(8):705-716.
10. Raje NS, Siegel DS, Jagannath S, et al. Idecabtagene vicleucel (ide-cel,
bb2121) in relapsed and refrac­tory mul­ti­ple mye­loma: ana­ly­ses of high-risk
sub­groups in the KarMMa study. Blood. 2020;136(suppl 1):37-38.
Dr Prakash Singh Shekhawat
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/180/2021675/180rodriguez-otero.pdf by guest on 09 December 2022
Figure 2. Possibilities for the future development of CAR T-cell therapy. Depicted are some of the novel new-generation CAR T
constructs, including third-generation CARs incorporating 2 different costimulatory domains; fourth-generation CARs (also named
“armored” CARs) incorporating transducible genes codifying for cytokines, antibodies, or receptors (among others); and novel target
(non-BCMA) and dual targeting strategies, including tandem CAR and dual CAR.
28. Jiang H, Dong B, Gao L, et al. Clinical results of a mul­ti­cen­ter study of the
first-in-human dual BCMA and CD19 targeted novel plat­form fast CAR-T cell
ther­apy for patients with relapsed/refrac­tory mul­ti­ple mye­loma. Blood.
2020;136(suppl 1):25-26.
29. Mailankody S, Diamonte C, Fitzgerald L, et al. Phase I first-in-class trial
of MCARH109, a G pro­tein cou­pled recep­tor class C group 5 mem­ber D
(GPRC5D) targeted CAR T cell ther­apy in patients with relapsed or refrac­
tory mul­ti­ple mye­loma. Blood. 2021;138(suppl 1):827.
30. Rodriguez-Marquez P, Calleja-Cervantes ME, Serrano G, et al. CAR den­sity
influ­ences anti­tu­moral effi­cacy of BCMA CAR-T cells and cor­re­lates with
clin­i­cal out­come. Blood. 2021;138(suppl 1):735.
31. Katsarou A, Sjöstrand M, Naik J, et al. Combining a CAR and a chi­me­ric
costimulatory recep­tor enhances T cell sen­si­tiv­ity to low anti­gen den­sity
and pro­motes per­sis­tence. Sci Transl Med. 2021;13(623):eabh1962.
32. Mansilla-Soto J, Eyquem J, Haubner S, et al. HLA-inde­
pen­
dent T cell
recep­
tors for targeting tumors with low anti­
gen den­
sity. Nat Med.
2022;28(2):345-352.
33. Cowan AJ, Pont M, Sather BD, et al. Safety and effi­cacy of fully human BCMA
CAR T cells in com­bi­na­tion with a gamma secretase inhib­i­tor to increase
BCMA sur­face expres­sion in patients with relapsed or refrac­tory mul­ti­ple
mye­loma. Blood. 2021;138(suppl 1):551.
34. Anderson LD, Munshi NC, Shah N, et al. Idecabtagene vicleucel (ide-cel,
bb2121), a BCMA-directed CAR T cell therapy, in relapsed and refractory
multiple myeloma: updated KarMMa results. J Clin Oncol. 2021;39(suppl
15):8016.
35. Oriol A, San-Miguel J, Kansagra A, et al. Idecabtagene vicleucel (IDE-CEL,
BB2121), a BCMA-directed CAR T cell therapy, in patients with relapsed
and refractory multiple myeloma: updated KarMMa results. EHA Library.
2021;324732.
36. Martin T, Usmani SZ, Berdeja JG, et al. Updated results from CARTITUDE-1:
phase 1b/2Study of ciltacabtagene autoleucel, a B-Cell maturation antigendirected chimeric antigen receptor T cell therapy, in patients with
relapsed/refractory multiple myeloma. Blood. 2021;138(suppl 1):549.
37. Mailankody S, Jakubowiak AJ, Htut M, et al. Orvacabtagene autoleucel
(orva-cel), a B-cell maturation antigen (BCMA)-directed CAR T cell therapy for patients (pts) with relapsed/refractory multiple myeloma (RRMM):
update of the phase 1/2 EVOLVE study (NCT03430011). J Clin Oncol.
2020;38(15):8504.
38. Chen W, Fu C, Cai Z, et al. Sustainable efficacy and safety results from
LUMMICAR study 1: a phase 1/2 study of fully human B-cell maturation
antigen-specific CAR T cells (CT053) in Chinese subjects with relapsed
and/or refractory multiple myeloma. Blood. 2021;138(suppl 1):2821.
39. Chunrui L, Wang D, Song Y, et al. A phase 1/2 study of a novel fully human
B-cell maturation antigen-specific CAR T cells (CT103A) in patients with
relapsed and/or refractory multiple myeloma. Blood. 2021;138(suppl 1):
547.
40. Gan An, et al. Blood. 2020;136 (suppl 1):29-30.
41. Costello CL, Cohen AD, Patel KK, et al. Phase 1/2 study of the safety and
response of P-BCMA-101 CAR-T cells in patients with relapsed/refractory
(r/r) multiple myeloma (MM) (PRIME) with novel therapeutic strategies.
Blood. 2020;136(suppl 1):29-30.
42. Mailankody S, Matous JV, Liedtke M, et al. Universal: an allogeneic firstin-human study of the amt-BCMA ALLO-715 and the anti-CD52 ALLO647 in relapsed/refractory multiple myeloma. Blood. 2020;136(suppl 1):
24-25.
© 2022 by The Amer­i­can Society of Hematology
DOI 10.1182/hema­tol­ogy.2022000396
Dr Prakash Singh Shekhawat
CAR T-cell ther­apy in mul­ti­ple mye­loma | 189
Downloaded from http://ashpublications.org/hematology/article-pdf/2022/1/180/2021675/180rodriguez-otero.pdf by guest on 09 December 2022
11. Jakubowiak A, Usmani SZ, Berdeja JG, et al. Efficacy and safety of ciltacabtagene autoleucel in patients with relapsed/refrac­tory mul­ti­ple mye­loma:
CARTITUDE-1 sub­group anal­y­sis. Blood. 2021;138(suppl 1):3938.
12. Shah N, Munshi NC, Berdeja JG, et al. Baseline cor­
re­
lates of com­
plete
response to idecabtagene vicleucel (ide-cel, bb2121), a BCMA-directed CAR
T cell ther­apy in patients with relapsed and refrac­tory mul­ti­ple mye­loma:
subanalysis of the KarMMa trial. Blood. 2021;138(suppl 1):1739.
13. Cohen AD, Mateos M-V, Cohen YC, et al. Efficacy and safety of cilta-cel
in patients with progressive MM after exposure to other BCMA-targeting
agents [published online ahead of print 12 September 2022]. Blood.
14. Hansen DK, Sidana S, Peres L, et al. Idecabtagene vicleucel (Ide-cel) chi­
me­ric anti­gen recep­tor (CAR) T-cell ther­apy for relapsed/refrac­tory mul­ti­
ple mye­loma (RRMM): real-world expe­ri­ence. J Clin Oncol. 2022;40(suppl
16):8042.
15. Da Vià MC, Dietrich O, Truger M, et al. Homozygous BCMA gene dele­tion in
response to anti-BCMA CAR T cells in a patient with mul­ti­ple mye­loma. Nat
Med. 2021;27(4):616-619.
16. Samur MK, Fulciniti M, Aktas Samur A, et al. Biallelic loss of BCMA as a resis­
tance mech­a­nism to CAR T cell ther­apy in a patient with mul­ti­ple mye­loma.
Nat Commun. 2021;12(1):868.
17. Cohen AD, Garfall AL, Stadtmauer EA, et al. B cell mat­u­ra­tion anti­genspe­cific CAR T cells are clin­i­cally active in mul­ti­ple mye­loma. J Clin Investig.
2019;129(6):2210-2221.
18. Manier S, Ingegnere T, Escure G, et al. Current state and next-gen­er­a­tion
CAR-T cells in mul­ti­ple mye­loma. Blood Rev. 2022;54(suppl 1):100929.
19. Mailankody S, Liedtke M, Sidana S, et al. Universal updated phase 1 data
val­i­dates the fea­si­bil­ity of allo­ge­neic anti-BCMA ALLO-715 ther­apy for
relapsed/refrac­tory mul­ti­ple mye­loma. Blood. 2021;138(suppl 1):651.
20. Cohen YC, Cohen AD, Delforge M, et al. Efficacy and safety of ciltacabtagene autoleucel (Cilta-cel), a B-cell mat­u­ra­tion anti­gen (BCMA)-directed
chi­me­ric anti­gen recep­tor (CAR) T-cell ther­apy, in
Download