Calcineurin: Form and Function

advertisement
PHYSIOLOGICAL REVIEWS
Vol. 80, No. 4, October 2000
Printed in U.S.A.
Calcineurin: Form and Function
FRANK RUSNAK AND PAMELA MERTZ
Section of Hematology Research and Department of Biochemistry and Molecular Biology, Mayo Clinic,
Rochester, Minnesota; and Department of Chemistry and Biochemistry, University of Massachusetts
Dartmouth, North Dartmouth, Massachusetts
I. Introduction
II. A Brief History and Overview of Calcineurin
A. Calcineurin: the early years
B. Calcineurin properties
III. Physiological Roles for Calcineurin
A. Lower eukaryotes
B. Higher eukaryotes
C. Inhibitors of calcineurin
IV. Calcineurin Structure
A. A dinuclear metal-binding phosphoesterase motif
B. Three-dimensional structure
C. Active site architecture
D. Metal ion requirements
V. Enzymatic Mechanism
A. Mechanism of phosphoryl group transfer: evidence for direct transfer to water
B. Catalytic role of the dinuclear metal center
C. Conserved active site residues
D. A model for the calcineurin catalytic mechanism
VI. Regulation
1483
1484
1484
1484
1490
1490
1494
1497
1500
1500
1500
1502
1503
1504
1504
1505
1505
1508
1510
Rusnak, Frank, and Pamela Mertz. Calcineurin: Form and Function. Physiol Rev 80: 1483–1521, 2000.—Calcineurin is a eukaryotic Ca2⫹- and calmodulin-dependent serine/threonine protein phosphatase. It is a heterodimeric
protein consisting of a catalytic subunit calcineurin A, which contains an active site dinuclear metal center, and a
tightly associated, myristoylated, Ca2⫹-binding subunit, calcineurin B. The primary sequence of both subunits and
heterodimeric quaternary structure is highly conserved from yeast to mammals. As a serine/threonine protein
phosphatase, calcineurin participates in a number of cellular processes and Ca2⫹-dependent signal transduction
pathways. Calcineurin is potently inhibited by immunosuppressant drugs, cyclosporin A and FK506, in the presence
of their respective cytoplasmic immunophilin proteins, cyclophilin and FK506-binding protein. Many studies have
used these immunosuppressant drugs and/or modern genetic techniques to disrupt calcineurin in model organisms
such as yeast, filamentous fungi, plants, vertebrates, and mammals to explore its biological function. Recent
advances regarding calcineurin structure include the determination of its three-dimensional structure. In addition,
biochemical and spectroscopic studies are beginning to unravel aspects of the mechanism of phosphate ester
hydrolysis including the importance of the dinuclear metal ion cofactor and metal ion redox chemistry, studies
which may lead to new calcineurin inhibitors. This review provides a comprehensive examination of the biological
roles of calcineurin and reviews aspects related to its structure and catalytic mechanism.
I. INTRODUCTION
The year 1999 marked the 20th anniversary of the
isolation of the Ca2⫹- and calmodulin-dependent protein
serine/threonine phosphatase calcineurin (206). During
the past 20 years, the biological roles of calcineurin have
progressed from a putative inhibitor of the calmodulinhttp://physrev.physiology.org
dependent phosphodiesterase (444) to the ground-breaking
discovery that it is the target of the immunosuppressant
drugs cyclosporin A (CsA) and FK506, pharmacological reagents that have been used to demonstrate it as a major
player in Ca2⫹-dependent eukaryotic signal transduction
pathways (238). In recent years, several milestones regarding calcineurin structure have been achieved including the
0031-9333/00 $15.00 Copyright © 2000 the American Physiological Society
1483
1484
FRANK RUSNAK AND PAMELA MERTZ
determination of the three-dimensional structure by X-ray
diffraction methods (124, 197) and biochemical, spectroscopic, and physical studies that are beginning to unravel its
catalytic mechanism (150, 259, 261, 262, 470, 471). Insight
into its physiological functions include mapping its subcellular localization (10, 106, 175, 219, 286, 306); the discovery
of its colocalization with other important signaling proteins
(365); and, aside from Ca2⫹/calmodulin, the finding of possible endogenous regulators of its activity including redox
and/or oxidative stress (45, 111, 345, 447, 470, 471) as well as
interacting proteins (224, 234, 277, 359, 400). The next generation of studies, which includes the use of transgenic
mouse technology, is beginning to reveal interesting yet
sometimes subtle roles for this enzyme in the whole organism (181, 254, 281, 320, 460, 477).
It will not be the attempt of this review to provide an
all-encompassing survey of calcineurin. In fact, several
excellent review articles on calcineurin and other protein
serine/threonine phosphatases are available, some quite
comprehensive (60, 130, 203, 205, 207, 208, 316, 371). In
addition, numerous specialized articles focusing on particular aspects of either calcineurin structure or function
have been published. A list of these appears in Table 1 for
the benefit of the reader who would prefer to be directed
to specific calcineurin-related topics. Rather, we focus on
a comprehensive treatise of some of the recent developments of calcineurin since the last major review was
published (371) (ca. 1990 to present).
II. A BRIEF HISTORY AND OVERVIEW
OF CALCINEURIN
A. Calcineurin: The Early Years
Calcineurin was first detected by Wang and Desai
(444) as a column fraction that inhibited the calmodulinTABLE
Volume 80
dependent cyclic nucleotide phosphodiesterase. Independently, Watterson and Vanaman (452) also obtained
highly purified fractions of calcineurin from bovine brain
extract by use of calmodulin-affinity chromatography but
erroneously referred to the 58- and 18-kDa subunits of
calcineurin as “affinity-purified phosphodiesterase.” Klee
and Krinks (206) are credited with the first purification of
calcineurin and hypothesized that it might be a regulatory
subunit of phosphodiesterase since it was demonstrated
to inhibit phosphodiesterase activity. Other groups subsequently showed that calcineurin inhibited the Ca2⫹/
calmodulin-dependent isozymes of cyclic nucleotide
phosphodiesterase and adenylate cyclase by competing
for calmodulin in a Ca2⫹-dependent fashion, and they
speculated that its function may be regulatory (435, 436,
445). Shortly thereafter, Klee et al. (204) coined the descriptive label “calcineurin” on the basis of its Ca2⫹-binding properties and localization to neuronal tissue (204), a
popularized name which is widely used to date and which
we will use throughout this review. At that time, the true
function of calcineurin had yet to be revealed. It was not
until pioneering work in the early 1980s in Philip Cohen’s
lab, investigating cellular extracts capable of dephosphorylating the ␣- and ␤-subunits of phosphorylase kinase,
that a fraction represented as protein phosphatase 2B
(PP2B) was demonstrated to be identical to Klee’s calcineurin (390, 391).
B. Calcineurin Properties
Biochemical studies during the 1980s continued and
determined many of the physical properties listed in Table
2 (61, 130, 208). Purified calcineurin is a heterodimer
consisting of a catalytic subunit, calcineurin A, and a
“regulatory” subunit, calcineurin B.
1. Calcineurin reviews listed according to subject
Review Topic
General aspects of calcineurin
Control of adenylyl cyclase
Neural roles for immunosuppressive drugs, immunophilins,
and calcineurin
Role of calcineurin in brain ischemia and injury
Calcineurin and NF-AT family of transcription factors
Calcineurin, Ca2⫹ signaling, and the cell cycle
Calcineurin in yeast and other lower eukaryotic organisms
Calcineurin, T lymphocyte activation, and the mechanism of
action of immunosuppressive agents
Structure of calcineurin and related phosphatases
Calcineurin in the mammalian nephron
Calcineurin localization via AKAP
Calcineurin and ion channel regulation
Calcineurin and apoptosis
Calcineurin regulation of dystrophin
Calcineurin and protein phosphatases in plants
Role of calcineurin in hypertension
Regulation of microtubules by calcineurin: tau phosphorylation
Reference No.
60, 61, 130, 145, 167, 202–205, 207, 208, 316, 371
13, 369
59, 119, 160, 253, 299, 333, 353, 384, 385, 434, 465, 474
285
173, 338–340
266, 276, 357, 358, 405
47, 68, 71, 141–143, 303, 368, 387
30, 46, 55, 133, 134, 155, 179, 189, 192, 222, 236, 237, 249, 255, 274,
311, 361, 363, 370, 378, 412, 413, 439, 455, 458, 472
22–24, 199, 215, 223, 244, 307, 351, 392, 411, 428, 430, 432, 433, 457
418
57, 103, 323, 366
404, 465
170
275
245, 368, 383
356
253
NF-AT, nuclear factor of activated T cells; AKAP, protein kinase A anchoring protein.
October 2000
TABLE
1485
CALCINEURIN: FORM AND FUNCTION
2. Physical properties of calcineurin
Physical Properties
Subunit structure and size
Calcineurin A
Calcineurin B
Domain structure
Calcineurin A
Calcineurin B
Activators
Cofactors
Co-/posttranslational modification
Reference No.
␣␤
57–59 kDa (mammals)
57–71 kDa (lower eukaryotes)
19–20 kDa
NH2 terminus, catalytic, calcineurin B binding,
calmodulin binding, autoinhibitory
four Ca2⫹-binding EF-hand motifs, myristoylation motif
Ca2⫹ and calmodulin
Fe3⫹, Zn2⫹
Phosphorylation of calcineurin A
Myristoylation of calcineurin B
Cloning efforts have provided evidence that all eukaryotic organisms possess one or more genes for each
subunit; Table 3 is a compilation of known calcineurin A
and calcineurin B gene sequences to date. Genes for
calcineurin A and B subunits have been identified in yeast,
filamentous fungi, protozoa, insects, and mammals. The
␣␤-quaternary structure of calcineurin observed in mammals is conserved in lower eukaryotic organisms. These
subunits are tightly associated and can only be dissociated by use of denaturants (271).
1. Calcineurin A
A) CLASSIFICATION. In addition to calcineurin, the serine/
threonine protein phosphatase family members include
protein phosphatases 1 (PP1), 2A (PP2A), and 2C (PP2C),
phosphatases essential for a number of signal transduction pathways in eukaryotic cells (61, 371). The original
classification of this family was proposed by Ingebritsen
and Cohen (167), separating almost all the serine/threonine phosphatase activity in mammalian tissue extracts
into two classes (60, 61, 167, 371). Type 1 protein phosphatases were found to dephosphorylate the ␤-subunit of
phosphorylase kinase, whereas type 2 protein phosphatases dephosphorylate the ␣-subunit of phosphorylase
kinase. Differences between the two types are also found
with inhibitors; type 1 is inhibited by phosphopeptide
inhibitors 1 and 2, whereas the type 2 class is not affected
by these inhibitors.
A difference in divalent metal ion dependence led to
the resolution of the type 2 enzymes into PP2A, PP2B
(calcineurin), and PP2C (60). PP2A was originally described as having no requirement for divalent metal ion,
calcineurin is regulated by Ca2⫹/calmodulin, and PP2C is
Mg2⫹ dependent.1 A more detailed discussion regarding
the importance of metal ion cofactors of these phospha-
1
Although PP2A was originally characterized as having no divalent
metal ion dependence (60), more recently it has also been found to be
stabilized or reactivated by divalent metal ions (43).
205, 371
72, 75, 153, 242, 310, 343, 467, 468
213, 221, 346
61, 130, 208
206, 390, 391
124, 195, 471
44, 136, 138, 257, 382, 420
1
tases is described in section IVD. Differences among type
2 members are also found with regard to sensitivity to
inhibition by macrolide inhibitors (see sect. IIIC1). PP2A
and PP1 are inhibited by okadaic acid, whereas calcineurin is specifically inhibited by the immunosuppressant drugs CsA and FK506, in the presence of cyclophilin
and FK506-binding protein (FKBP), respectively (238,
361).
Although different in metal ion dependence, sensitivity to inhibitors, and substrate specificity, PP1, PP2A, and
calcineurin have homologous amino acid sequences and
are evolutionarily related (26, 61, 371). In fact, the PP1/
PP2A/calcineurin superfamily ranks among one of the
most highly conserved enzyme families encountered (61).
Within the active site domain, PP1 shares 49% amino acid
identity with PP2A and 39% identity with calcineurin.
Recent work has found homologs of this family in cyanobacteria (373, 473) and the archea (229, 251, 386). With
the determination of the amino acid sequences of these
phosphatases, it was found that the original classification
of PP2C along side PP1, PP2A, and calcineurin does not
hold at the primary sequence level. Thus PP2C does not
share any homology with PP1/2A/calcineurin and is considered to be in a separate superfamily (33).
B) DOMAIN STRUCTURE. The active site of calcineurin is
located on the A subunit which, in mammals, is 57–59 kDa
depending on the isoform. The size of the catalytic subunit can be up to ⬃20% larger in lower eukaryotic species
[e.g., Saccharomyces cerevisiae, 63 and 69 kDa (72, 242,
467); Schizosaccharomyces pombe, 64 kDa (327, 468);
Drosophila melanogaster, 62 and 65 kDa (38, 132, 158);
Cryptococcus neoformans and Dictyostelium discoideum, 71 kDa (75, 310)]. Nevertheless, there is strict
conservation throughout all eukaryotic organisms such
that all calcineurin A genes encode for a polypeptide
consisting of a catalytic domain homologous to other
serine/threonine protein phosphatases and three regulatory domains at the COOH terminus that distinguish calcineurin from other family members (Fig. 1). These domains have been identified as the calcineurin B binding
1486
TABLE
FRANK RUSNAK AND PAMELA MERTZ
Volume 80
3. A comprehensive list of published calcineurin A and calcineurin B gene sequences
Organism
Gene
Alternative
Name
Tissue/Stage
Genbank Entry
Chromosomal
Location
Reference No.
Calcineurin A
Saccharomyces cerevisiae
Schizosaccharomyces
pombe
Dictyostelium discoideum
Aspergillus nidulans
Neurospora crassa
Caenorhabditis elegansd
Cryptococcus neoformanse
Drosophila melanogaster
CNA1a
CMP2a
ppb1⫹ b
YSCCALA1.Gb_pl1
YSCCALA2.Gb_pl2
Q12705.SWISSPROT
CMP1
CNA1
cna-1
CeCnA
Vegetative growth
Cell cycle G1/S
transition
Early mycelia
Neuronal, vulva
CnnA21EF
?
canA
cnaA⫹
PP2B 14D
Neuronal,
embryonic
Xenopus laevisg
Mouse
Ppp3cah,i
PP2BA␣,
PP2B␣1
Brain
Rat
Ppp3cbh
Ppp3cch
Ppp3cah,i
PP2BA␤
PP2BA␥
PP2BA␣,
Calna1
Brain
Testis
Brain
Ppp3cbh
PP2BA␤,
Calna2
327, 468
75
343
NEUCAM.Gb_pl
153
328, 456
310
132
DMU30493.Gb_in
Chromosome 2,
position 21EF
X chromosome,
position 14D
AF019569.GB_OV
MUSCALCAT.Gb_ro
38, 158
190, 191
P2BA_MOUSE.SW
P2BB_MOUSE.SW
RATCNRAA.Gb_ro
Chromosome 15
117
291
169, 354, 466
Brain
RNCALCA.Gb_ro (⌬ 10aa)
RATCNRAB.Gb_ro
Chromosome 15
354
220, 466
4q21 3 q24j
129, 190, 293
10q21 3 q22j
129
S46622.Gb_pr1
Chromosome 8l
265
292
YSCCNB.Gb_pl1
AF060553
AF076251.Gb_pl2
AF076252.Gb_pl2
AF076253.Gb_pl2
XI
Rabbit
Bovine
PP2Bw
PPP3CAh
Human
PPP3CAh,i
PP2BA␣,
PP2B␣1
Brain
RATCALAB.Gb_ro
RABPP2BWS.Gb_om
BTU33868.Gb_om
P2BA_BOVIN.SW
HUMCALCAT.Gb_pr1 (part.)
PPP3CBh,k
PP2BA␤,
PP2B␣2
Brain
E07798.Gb_pr1
HUMCNRA1.Gb_pr1
PP2B␤3k
PP2BA␥,
PP2B␣3
Teratocarcinoma
Testis
PPP3CCh
72, 242
DDCALNRNA.Gb_in
P2b_Emeni.SW
AF042082.Gb_pl1
DMCNA.Seq
CnnA14Df
XII
XIII
I or IIc
PP2BA␣
74
124
HUMCRNAB.Gb_pr1
Calcineurin B
Saccharomyces cerevisiae
Arabidopsis thaliana
a
AF034089.Gb_pl1
221
239
218
218
218
328, 456
213
NCCALCINB.Gb_pl1
NGU04380.Pep
346
Caenorhabditis elegansd
Neurospora crassa
CNB1
SOS3
CBL1
CBL2
CBL3
CeCnB
cnb-1
Naegleria gruberi
CNB1
Drosophila melanogaster
dCNB1
DROCALCB.Gb_in
dCNB2
CALB_DROME.Sw
U56245.Gb_In
Mouse
Rat
Asexual
development
Flagellar
differentiation
Ppp3r1h
Ppp3r2h
Ppp3r1h
Alt. Splice
Ppp3r2
h
PP2B␤1
PP2B␤2
CNB␣1,
Pp2b␤1
CNB␣2,
Pp2b␤2
CBLP
X chromosome,
position 4F
Chromosome 2,
position 43E
132
450
Brain
Testis
Brain
CALB_MOUSE.SW
RATCALCB.GB_RO
424
424
48
Testis
RATCALNB.GB_RO
48
Testis
Testis
RATCBLP.Gb_ro
289
TABLE
1487
CALCINEURIN: FORM AND FUNCTION
October 2000
3. (Continued)
Organism
Alternative
Name
Gene
Bovine
PPP3R1h
Human
h
PPP3R1
Tissue/Stage
Brain
Brain
Genbank Entry
BBCALCB.GB_OM
CALB_BOVIN.SW
HUMCNR.Pr
Chromosomal
Location
Reference No.
2, 304
2p16 3 p15
j
131
a
Tissue/stage refers to tissue or developmental stage of highest abundance.
Gene locus and chromosomal location according to
b
information posted at http://genome-www.stanford.edu/Saccharomyces/.
Sequence available at http://www.sanger.ac.uk/Projects/S_pombe/.
c
d
There is still an unresolved discrepancy whether the S. pombe ppb1⫹ gene resides on chromosome I (468) or chromosome II (327).
A complete
analysis of the C. elegans genome had not yet been completed at the time this review went to press. Evidence for both calcineurin A- and calcineurin B-like
genes has been found. Future information can likely be found at one of the C. elegans genome databases: http://stein.cshl.org/elegans/ or http://
e
f
g
wormsrv1.sanger.ac.uk/cgi-bin/ace/simple/worm.
Filobasidiella neoformans.
Two splice variants were identified (158).
GeneBank entry.
h
Gene nomenclature of human protein serine/threonine phosphatase genes according to Cohen (62). PP3CA corresponds to ␣-isoforms, PP3CB
corresponds to ␤-isoforms, and PP3CC corresponds to ␥-isoforms. Similar nomenclature has been adopted for mouse (http://www.informatics.jax.org/)
i
and rat (http://ratmap.gen.gu.se/) calcineurin genes.
Two major splice variants have been identified, both identical with the exception of a 10-amino
acid deletion between calmodulin-binding domain and autoinhibitory domain (see References 190, 354.) These have been referred to as PP2B␣1 and
j
k
PP2B␣2; note numerical subscript in Reference 265.
Chromosomal location according to References 117, 446.
Three major splice variants have
been identified. Two isozymes were designated CNA1 and CNA2 in the original report (129). CNA1 and CNA2 are referred to as PP2B␤1 and PP2B␤2,
respectively, while a third alternatively spliced isozyme was designated as PP2B␤3 in Reference 265 [note use of subscript to designate splice variant, in
contrast to the gene names used by Kincaid et al. to describe the ␣ (PP2B␣1), ␤ (PP2B␣2), and ␥ (PP2B␣3) isoforms of the catalytic subunit (118,
l
292)].
Chromosomal location according to Reference 292.
domain (56, 132, 164, 379, 451), the calmodulin-binding
domain (132, 191), and the “autoinhibitory” domain (137,
326), which binds in the active site cleft in the absence of
Ca2⫹/calmodulin (197) and inhibits the enzyme, acting in
concert with the calmodulin binding domain to confer
calmodulin regulation.
These domains have been identified through standard biochemical mapping procedures including primary
sequence comparisons, partial proteolysis experiments,
peptide interaction studies, site-directed mutagenesis
studies, and cross-linking studies. The X-ray structures of
calcineurin (see sect. IVB) confirm the identification of
residues involved in these regulatory domains and also
indicate that the autoinhibitory domain forms an ␣-helix
that binds to the substrate-binding cleft of the enzyme
(197). It has been shown that when Ca2⫹/calmodulin
binds to the enzyme, inhibition ceases and likely involves
a conformational change that exposes the active site.
Interestingly, Perrino (324) has provided recent evidence
for additional autoinhibitory elements within residues
that are situated between the calmodulin-binding and autoinhibitory domains noted in Figure 1.
The NH2 and COOH termini are highly variable between species as well as between calcineurin A genes
within the same organism (130, 188, 208). Particularly
striking are unusual amino acid compositions and sequences of particular isoforms. For example, the mammalian calcineurin A␤-isoform contains the sequence
MAAPEPARAAPPPPPPPPPPPGAD. . . at the NH2 terminus (117, 129, 220). The COOH terminus of the canA gene
product from Dictyostelium contains a fourfold repeat of
the sequence RXNSX(G/A)(E/D)LX as well as the repetitious sequence ..TTNNINPNSITTNENNSNEQLQQQQQQQQQQQPPTTTSTTT.. and, along with the NH2 terminus,
is enriched in glutamine and asparagine (75). The COOH
terminus of calcineurin A from C. neoformans is highly
enriched in proline, serine, threonine, and glycine (310).
The function of these variable domains is unknown, but
they may play a role in substrate recognition and/or localization.
C) CALCINEURIN PHOSPHORYLATION. Purified calcineurin from
bovine brain contains up to 0.6 equivalents of phosphate
(195), suggesting that it may be phosphorylated in vivo.
Calcineurin can be phosphorylated by protein kinase C
(138, 420), casein kinase I (382), and casein kinase II (136,
138, 257) in vitro. The site of phosphorylation by casein
kinase II has been determined to be the serine in the
sequence -RVFS(p)VLR- near the COOH terminus of the
calmodulin-binding domain (Fig. 1) (138, 257). Although
phosphorylation could be blocked by Ca2⫹/calmodulin,
the kinetic properties of the phosphorylated and dephosphorylated forms are similar (136, 138). Furthermore,
phosphorylated calcineurin still binds and is activated by
calmodulin. Whether phosphorylation represents a means
of regulating calcineurin activity in vivo remains to be
demonstrated.
2. Calcineurin B
A) SEQUENCE DIVERSITY AND ISOFORMS. The calcineurin B subunit is also highly conserved throughout evolution, with
mammalian calcineurin B showing 86% amino acid sequence identity with insect calcineurin B (i.e., Drosophila) and 54% identity with calcineurin B from S. cerevisiae
(Fig. 2). This high degree of conservation allows functional interchange of calcineurin B subunits between
mammalian and N. crassa catalytic subunits (423). The
gene for mammalian calcineurin B encodes a protein of
170 amino acids containing four Ca2⫹-binding EF-hand
motifs (Fig. 2) (2).
1488
FRANK RUSNAK AND PAMELA MERTZ
Volume 80
FIG. 1. Primary sequence and domain structure of calcineurin A. The amino acid sequence represents the rat
calcineurin A ␣-isoform reported by Saitoh et al. (354). CaM, calmodulin; CNB, calcineurin B; AI, autoinhibitory.
In mammals, there are two calcineurin B genes, one
which is ubiquitously expressed, while mRNA for the
second gene is found only in testes (48, 289, 424).
B) NH2-TERMINAL MYRISTOYLATION. The mature calcineurin B
protein is missing the initiator methionine, and the new
␣-amino group of glycine at position 2 is acylated with
myristic acid (1). This modification has been conserved
throughout evolution from yeast to mammals, suggesting
a crucial physiological role (71). To explore possible biological roles for calcineurin myristoylation, Heitman and
colleagues (483) generated a mutant of calcineurin B in
which glycine at position 2 was mutated to alanine,
thereby preventing myristoylation. Surprisingly, expression of the wild-type and mutant proteins in S. cerevisiae
demonstrated that myristoylation was not required for
membrane association nor for interaction with immunosuppressant drug complexes. Indeed, the nonmyristoylated protein exhibited full biological function. These results were subsequently confirmed in biochemical
experiments with purified myristoylated and nonmyristoylated calcineurin heterodimer which showed equivalent enzymatic activities, inhibition by the CsA/cyclophilin
immunosuppressant drug complex, and interactions with
a synthetic phospholipid monolayer (182). Interestingly,
the myristoylated protein exhibited substantial thermal
stability (⬃12°C) relative to the nonmyristolyated protein
(182). At present, it is unknown whether the biological
role of calcineurin B myristoylation is to impart increased
stability to the protein or whether there is another role yet
to be identified.
C) CALCIUM BINDING PROPERTIES. Klee et al. (204) were the
first to discover that calcineurin binds Ca2⫹. With the use
of flow dialysis, it was demonstrated that four Ca2⫹ bind
with high affinity [dissociation constant (Kd) ⱕ10⫺6 M]
and that the Ca2⫹-binding sites were localized to the
calcineurin B subunit. The complete primary sequence
determination of calcineurin B revealed homology with
calmodulin (35% identity) and troponin C (29% identity)
(2), most of which was confined to four Ca2⫹-binding
“EF-hand” motifs. More detailed thermodynamic aspects
of Ca2⫹ binding became possible when the recombinant
calcineurin B subunit was obtained via heterologous expression in Escherichia coli. Using the purified recombinant protein, Burroughs et al. (40) studied the metal binding properties using Eu3⫹ and Tb3⫹ luminescence
spectroscopy. Four Eu3⫹-binding sites were revealed, two
October 2000
CALCINEURIN: FORM AND FUNCTION
1489
FIG. 2. Multiple sequence alignment of calcineurin B sequences from diverse organisms. Calcineurin B sequences of
Saccharomyces cerevisiae (221), Neurospora crassa (213), rat brain (48), rat testes (289), human (131), bovine (2, 304),
Drosophila melanogaster (132, 450), and Naegleria gruberi (346) were aligned using the multiple sequence alignment
editor of the Wisconsin Package version 9.0, Genetics Computer Group (Madison, WI). The four Ca2⫹-binding EF-hand
motifs are indicated. The residues that participate in Ca2⫹ coordination are noted by an asterisk. The consensus
sequence is defined in which a residue is conserved in all 8 sequences.
with relatively low affinity (Kd values of 1 ⫾ 0.2 and
1.6 ⫾ 0.5 ␮M) and two with relatively high affinity (Kd
values of 0.14 ⫾ 0.020 and 0.020 ⫾ 0.010 ␮M). Tb3⫹ also
bound but with slightly weaker affinities (Kd values of
0.04 ⫾ 0.01 and 0.17 ⫾ 0.02 ␮M for the COOH-terminal
sites and 1–3 ␮M for the NH2-terminal sites). Direct Ca2⫹
binding to calcineurin B has also been studied by flow
dialysis, which found one high-affinity (Kd ⫽ 0.024 ␮M)
and three lower affinity sites (Kd ⫽ 15 ␮M) (176). The
NMR-active isotope 113Cd has been used as a Ca2⫹ surrogate to identify four similar but distinct metal binding
sites consisting of all-oxygen coordination of pentagonal
bipyramidal geometry as expected for an EF-hand Ca2⫹binding site (176), later confirmed in the X-ray structure
1490
FRANK RUSNAK AND PAMELA MERTZ
(124, 197). Ca2⫹ binding to individual sites of calcineurin
B has been studied using point mutants of this subunit
altered in each of the four EF-hands (104). This study
confirmed the higher Ca2⫹ affinity for COOH-terminal
EF-hand sites and also found that Ca2⫹ binding at these
sites is likely to be structural.
D) CALCINEURIN B HOMOLOGS. EF-hand proteins have been
classified into 39 distinct subfamilies containing anywhere from two to eight EF-hand domains (178). Calcineurin B proteins represent one subfamily of EF-hand
proteins based on sequence alignments and congruence
of domain and interdomain regions (302). In recent years,
a number of Ca2⫹-binding proteins containing EF-hand
domains have been identified from cloning studies and
shown to be homologous to calcineurin B. These include
NCS-1, a neuronal calcium sensor that inhibits rhodopsin
phosphorylation in a Ca2⫹-dependent fashion (81), modulates calmodulin targets (359), and may regulate exocytosis (264); a protein p22/CHP (for calcineurin homologous protein) that is required for constitutive membrane
traffic (25) and inhibits serum and GTPase stimulation of
the Na⫹/H⫹ exchanger NHE1 (234); CIB (for calcium- and
integrin-binding protein), a 22-kDa Ca2⫹-binding protein
that interacts with the cytoplasmic tail of the integrin ␣IIb
portion of the GPIIb/IIIa fibrinogen receptor (297); and a
protein product of the SOS3 gene of Arabidopsis thaliana involved in tolerance to the ionic component of salt
stress in plants (239). Homology to calcineurin B ranges
from 27 to 31% for NCS-1, CIB, and SOS3 and to up to 43%
for p22/CHP. Like calcineurin B, p22/CHP is myristoylated
while NCS-1, CIB, and SOS3 contain the requisite consensus sequences for myristoylation. The fact that a constitutively active form of yeast calcineurin in transgenic
tobacco plants resulted in increased salt tolerance provides evidence for a possible functional overlap between
SOS3 and calcineurin B (320). On the contrary, biochemical studies with NCS-1 suggested that it could replace
calmodulin rather than calcineurin B in activating calcineurin and other calmodulin-dependent enzymes (359).
Furthermore, although p22/CHP is completely congruent
with the calcineurin B family of proteins, arguments have
been forwarded that it is more appropriate to place it in a
separate subfamily (25). Undoubtedly, further studies are
TABLE
Volume 80
required to determine whether any or all of these proteins
can function in a fashion comparable to calcineurin B.
III. PHYSIOLOGICAL ROLES FOR CALCINEURIN
A. Lower Eukaryotes
Genetic methods for the selective deletion of one or
both calcineurin subunits to assess biological function by
noting the phenotype of the mutant strain are now possible in several eukaryotic organisms. In addition, the immunosuppressant drugs CsA and FK506, as specific calcineurin inhibitors, have provided complementary tools
for discerning the role of calcineurin in many eukaryotic
organisms (71, 142, 303, 387). Some of the most thorough
work investigating biological roles for calcineurin have
used the yeast S. cerevisiae as a model system. There are
two genes for the catalytic subunit of calcineurin in S.
cerevisiae (CNA1/CMP1 and CNA2/CMP2) and only one
gene for the B subunit (CNB1). Calcineurin is essential in
CsA- and FK506-sensitive yeast strains (36). Recent work
has begun to explore the role of calcineurin in slightly
more complex organisms such as N. crassa (153, 213, 335)
and D. discoideum (75, 159, 252). Furthermore, calcineurin has either been isolated or detected from the
human pathogens C. neoformans (310), Leishmania species (21, 342), the malarial parasite Plasmodium falciparum (87), helminth parasites (347), and schistosomes
(186). In some of these, growth can be inhibited by the
immunosuppressive agents FK506 and its analogs as well
as CsA (16, 87, 186, 309, 342), thus raising the possibility
that novel calcineurin inhibitors might be developed as
specific antifungal and antiparasitic agents. The following
sections detail what has been learned regarding physiological roles for calcineurin in lower eukaryotic organisms and is summarized in Table 4.
1. Saccharomyces cerevisiae
A) RECOVERY FROM PHEROMONE-INDUCED GROWTH ARREST. Haploid cells of S. cerevisiae produce one of two mating
4. Physiological roles of calcineurin in lower eukaryotic organisms
Organism
Saccharomyces cerevisiae
Schizosaccharomyces pombe
Dictyostelium discoideum
Neurospora crassa
Cryptococcus neoformans
Aspergillus nidulans
Paramecium tetraurelia
Function
Reference No.
Recovery from mating factor ␣-induced growth arrest, cation (e.g., Li⫹, Na⫹, Mn2⫹)
resistance, Ca2⫹ homeostasis, Ca2⫹-mediated G2 arrest, onset of mitosis
Cytokinesis, mating, nuclear and spindle pole body positioning, polarized growth, proper
septation, chloride homeostasis
Differentiation, stalk cell/spore formation
Hyphal growth/conidiation, maintenance of apical Ca2⫹ gradient, proper septation
Virulence, pH and CO2 homeostasis, temperature-sensitive growth, resistance to Li⫹
Cell cycle progression through G1/S, nuclear division, polarized growth, proper septation
Exocytosis
71, 72, 279, 461, 467
327, 399, 468
159
213, 335
310
303, 343
209, 282
October 2000
CALCINEURIN: FORM AND FUNCTION
pheromones, a-factor and ␣-factor. Exposure of haploid
strains to the opposite mating pheromone prepares cells
for mating by inducing cell cycle arrest in G1. This is
mediated by an elaborate signal transduction pathway
involving a rise in intracellular Ca2⫹ and activation of
calcineurin (47, 142). Growth arrest can be observed as a
zone of clearing surrounding a source of ␣-factor on a
lawn of cells and occurs within 24 h at 30°C. Escape from
␣-factor-induced cell cycle arrest involves three metabolic processes that have been referred to as recovery,
adaptation, and survival (287). Recovery is defined as the
ability of cells to resume growth after removal of the
pheromone, whereas adaptation is a process in which
cells eventually resume growth in the continuous presence of pheromone. Both recovery and adaptation may
involve common signaling components and can be observed by a shrinking of the zone of clearing and increasing turbidity within it, usually within ⬃24 h at 30°C.
Survival differs from recovery and adaptation in that it
describes whether a cell remains viable after exposure to
pheromone.
Coincident with the cloning of genes for the two
yeast calcineurin A subunits (CNA1/CMP1 and CNA2/
CMP2, Table 3), strains deficient in either subunit were
viable but failed to recover from ␣-factor-induced growth
arrest (72, 73, 242). Mata strains containing a single CNA1
or CNA2 mutation were twice as sensitive as wild type to
␣-factor-induced growth arrest, whereas the double mutant CNA1/CNA2 was four times as sensitive, as assessed
by the size of the halo after 24 h at 30°C (72, 73). Furthermore, once arrested, the double mutant failed to resume
growth. In contrast, the CNB1 mutant did not show an
increased sensitivity compared with wild type, but like
the CNA1/CNA2 double mutant, it failed to recover from
growth arrest. In wild-type cells, the immunosuppressant
drugs CsA and FK506 also inhibited recovery from ␣-factor-mediated growth arrest, and these required the presence of their respective immunophilins cyclophilin and
FKBP (107). In addition, expression of the CNA1/CMP1
gene increased in the presence of ␣-factor, the result of
5⬘-noncoding sequences in the CNA1/CMP1 gene matching closely the consensus sequence for the ␣-factor element (467).
As expected, the activator protein of calcineurin,
calmodulin, has also been shown to be required for escape from cell cycle arrest after exposure to pheromone
(287). Calmodulin mutants did not display increased sensitivity to ␣-factor, nor did these mutant strains appear to
be affected in either recovery or adaptation. Indeed, both
calcineurin and calmodulin mutants adapted as well as a
wild-type strain to low concentrations of pheromone, and
both mutants recovered after pheromone removal with
the same kinetics as the wild-type strain. The process that
appeared to be affected was survival, a result consistent
with previous work indicating that Ca2⫹ is also essential
1491
for survival after exposure to ␣-factor (166). Interestingly,
in addition to calcineurin, the Ca2⫹, calmodulin-dependent protein kinases (CMK1 and CMK2), yeast protein
kinase C (PKC1), and a mitogen-activated protein (MAP)
kinase (MPK1) are also required for recovery from
growth arrest, thus indicating that enzymes of opposing
function are required for surviving exposure to ␣-factor
(287, 301, 461).
One downstream signaling component in S. cerevisiae regulated by calcineurin is the yeast transcription
factor Crz1p/Tcn1p. Crz1p/Tcn1p is required for calcineurin-dependent induction of genes for the vacuolar
and secretory Ca2⫹ pumps, Pmc1p and Pmr1p, respectively; one of two genes encoding ␤-1,3 glucan synthase,
FKS2; and the gene for the plasma membrane Na⫹ pump,
PMR2 (Fig. 3) (263, 388). In addition, calcineurin has been
shown to regulate the high-/low-affinity state of the
plasma membrane K⫹ channel, Trk1p (269), and inhibit
the vacuolar H⫹/Ca2⫹ exchanger Vcx1p (69) by posttranlational mechanisms. Some of these are presented below
in more detail.
B) ADAPTATION TO SALT STRESS. The search for additional
phenotypes found that calcineurin-deficient yeast exhibited decreased tolerance to the monovalent cations Na⫹
and Li⫹, but not K⫹, Ca2⫹, and Mg2⫹ (269, 300). The role
of calcineurin in Na⫹/Li⫹ tolerance is thought to be mediated by transcriptional and posttranslational mechanisms. Adaptation to high salt stress requires the presence
of a plasma membrane Na⫹-ATPase involved in Na⫹ and
Li⫹ efflux, Pmr2p. Cells deficient in calcineurin accumulate Na⫹ and Li⫹ due to decreased expression of Pmr2p
(269). Although no changes in intracellular Ca2⫹ have
been observed after induction of the high-salt response,
evidence indicates that Ca2⫹ mediates this response.
Ca2⫹, via calmodulin activation of calcineurin, regulates
adaptation to high salt stress by induced expression of
Pmr2p (76, 154, 268), mediated by the transcription factor
Crz1p/Tcn1p (Fig. 3) (263, 388). The activity of Pmr2p is
also stimulated by Ca2⫹/calmodulin, thereby providing
both transcriptional and posttranslational regulation of
Na⫹ efflux mediated by Ca2⫹ (349, 454).
Cells deficient in CNB1 are unable to convert the K⫹
transport system (Trk1p, a K⫹ channel) to a high-affinity
state. In the high-affinity state, this pump has increased
affinity for K⫹, but the Michaelis constant (Km) for Na⫹ or
Li⫹ is unaffected, thereby resulting in increased Na⫹ uptake in calcineurin-deficient cells. The mechanism of this
regulation has been hypothesized to be direct or indirect
dephosphorylation of Trk1p by calcineurin (269).
Other proteins in addition to calcineurin are required
for salt tolerance such as the gene products of PDE1, a
low-affinity cAMP-dependent phosphodiesterase (154);
URE2, a regulator of nitrogen catabolite repression (462);
PMA1, the plasma membrane H⫹-ATPase (462); HAL3, a
protein involved in cell cycle control and ion homeostasis
1492
FRANK RUSNAK AND PAMELA MERTZ
Volume 80
FIG. 3. Physiological roles of calcineurin in Saccharomyces cerevisiae. Calcineurin has numerous roles in budding
yeast including recovery from ␣-factor-induced growth arrest, salt and temperature tolerance, Ca2⫹ homeostasis, and
Mn2⫹ tolerance (387). Many of these are mediated by Crz1p/Tcn1p, a calcineurin-dependent transcription factor
necessary for the transcriptional induction of Pmc1p, Pmr1p, Pmr2p, and Fks2 (heavy arrows) (263, 388). In addition,
calcineurin inhibits the activity of the vacuolar H⫹/Ca2⫹ exchanger (Vcx1p) and causes conversion of the K⫹ channel,
Trk1p, to the high-affinity state. The latter occur independently of Crz1p probably by posttranslational processes (narrow
arrows) (69). Cch1p, Ca2⫹ channel (313); Crz1p, calcineurin-responsive zinc finger transcription factor, product of the
CRZ1 gene; Pmcp1, high-affinity vacuolar Ca2⫹ pump, product of the PMC1 gene (69); Pmr1p: secretory Ca2⫹ pump,
product of the PMR1 gene; Tcn1p: alternative name for Crz1p protein; Vcx1p: low-affinity vacuolar H⫹/Ca2⫹ exchanger
(69). Mid1p is an alternative name for the plasma membrane Ca2⫹ channel.
(105); and STD1, a protein that interacts with the SNF1
protein kinase in two-hybrid and in vitro binding studies
(113). Thus multiple parallel pathways are necessary for
full induction of this response.
C) CALCIUM HOMEOSTASIS. Calcineurin is involved in the
regulation of Ca2⫹ pumps and exchangers responsible for
Ca2⫹ homeostasis in yeast (Fig. 3). These maintain cytoplasmic [Ca2⫹] in the range of 100 –300 nM (68, 78). In
addition, other ion transporters indirectly influence intracellular [Ca2⫹]. One of these is the vacuolar H⫹-ATPase,
which provides the driving force for Ca2⫹ sequestration
by the Ca2⫹/H⫹ exchanger encoded for by the VCX1 gene
(114, 144, 408). Two Ca2⫹-ATPases, Pmc1p and Pmr1p,
are responsible for depleting the cytosol of Ca2⫹. The
former is localized to the vacuole (70), while the latter is
important in the secretory pathway and localizes to the
Golgi (349). Mutants deleted in either Pmc1p or Pmr1p
cannot grow in media containing high Ca2⫹. Deletion of
the gene for either calcineurin subunit, or treatment of
cells with CsA or FK506, restores growth to either single
PMC1 or double PMC1/PMR1 mutants in high Ca2⫹ media
(70), indicating that calcineurin activation can have a
negative effect on growth. As noted above, activation of
calcineurin leads to transcriptional induction of the PMC1
and PMR1 genes via Crz1p/Tcn1p (263, 388).
Calcineurin mutants are also sensitive to extracellular Mn2⫹. Wild-type strains are able to prevent Mn2⫹
entry, whereas mutants exhibit an increased uptake phenotype (100), therefore indicating that the regulation of
Mn2⫹ homeostasis by calcineurin follows a different
October 2000
CALCINEURIN: FORM AND FUNCTION
mechanism than monovalent cation transport, in which
export is regulated by a P-type ATPase (269, 300). An
alternative hypothesis has been proposed in which
Pmr1p, the Golgi-localized Ca2⫹ pump, plays a role in
Mn2⫹ tolerance by sequestering Mn2⫹ to late compartments in the secretory pathway (263). Mn2⫹ may also be
transported into the vacuole via the Ca2⫹/H⫹ exchanger
Vcx1p (332).
D) ⟩-GLUCAN SYNTHASE AND CELL WALL SYNTHESIS. Calcineurin is
responsible for transcriptional regulation of FKS2, one of
two genes encoding ␤-1,3-glucan synthase (Fig. 3) (95,
480). Calcineurin-dependent regulation occurs through
Crz1p/Tcn1p (263, 388). The Fks1p protein is the predominant synthase expressed during optimum growth, but
expression of Fks2p is induced upon treatment of cells
with mating pheromone, high Ca2⫹, or growth on poor
carbon sources. Deletion of the FKS1 and CNB1 genes
results in lethality due to the inability to induce FKS2
(114). In fact, FKS1 mutants are hypersensitive to FK506
(95). These results suggest that calcineurin plays a role in
regulating cell wall structure.
2. Schizosaccharomyces pombe
Like the budding yeast, treatment of S. pombe with
FK506 or deletion of the ppb1⫹ gene, encoding for the
calcineurin A subunit (Table 3), is not lethal. However,
calcineurin in fission yeast appears to have distinct functions. Calcineurin-deficient S. pombe cells exhibit drastic
Cl⫺-sensitive growth (399) and are defective in cytokinesis, transport, nuclear and spindle pole body positioning,
cell shape (468), and sporulation (327). One function for
calcineurin in S. pombe that appears to overlap with S.
cerevisiae is the mating process, although the roles for
calcineurin in mating appear to be distinct in these two
organisms. In S. cerevisiae, calcineurin is required for the
cell to recover from or survive growth arrest after exposure to pheromone. It thus may function to assist cells to
reenter the cell cycle if they respond to ␣-factor but fail to
mate (see sect. IIIA1A). In S. pombe, calcineurin is required
for the mating response, and calcineurin mutants in this
organism are sterile (327, 468). Northern analysis indicates that the transcript for calcineurin varies during the
cell cycle and can be induced by nitrogen limitation, a
condition that favors mating in S. pombe (327). The latter
effect was dependent on the transcription factor ste1.
3. Neurospora crassa
N. crassa has been widely used as a model system for
studying eukaryotic gene expression. In this fungus, calcineurin is thought to play a major role in hyphal extension during mycelial growth and in determining apical
orientation. Thus calcineurin mRNA exhibited the highest
expression during early mycelial logarithmic growth but
was repressed before conidiation upon entry into station-
1493
ary phase (153). A similar pattern of protein expression
was observed but with about a 12-h lag behind message
expression. Expression of antisense mRNA to the catalytic subunit, treatment with CsA and FK506, or disruption of the calcineurin B gene caused growth arrest preceded by aberrant and increased hyphal branching and
entry into conidiation, consistent with a role in apical
growth (213, 335). Growth on two different carbon
sources, glutamate and sucrose, did not influence the
level of expression, indicating that calcineurin is not involved in mechanisms related to catabolite repression
(153).
4. Aspergillus nidulans
Similar to N. crassa, calcineurin A mRNA levels vary
during the cell cycle in A. nidulans, and disruption of the
cnaA⫹ gene resulted in growth arrest (303, 343). After
growth arrest in metaphase upon treatment with nocodazole followed by resuspension in fresh media to allow for
synchronous growth, it was shown that calcineurin A
message levels peak after the end of mitosis before DNA
replication, with the highest expression appearing at the
G1/S boundary. Inducible gene disruption by homologous
recombination revealed that the calcineurin A gene was
essential for normal growth, and disruption was lethal.
Further results indicated that the cnaA⫹ gene was required for early cell cycle events before DNA replication.
Taken together with other data, this study suggested that
calcineurin as well as other calmodulin targets may be
required during different periods of the cell cycle.
5. Cryptococcus neoformans
Fungal diseases are becoming an increasing health
problem, most notably in individuals infected with human
immunodeficiency virus (HIV); C. neoformans represents
one of these life-threatening infectious agents (6). Heitman and colleagues (5, 67, 309, 310) have begun to explore the role of calcineurin in this pathogen with the goal
of using novel CsA and FK506 analogs that have increased
specificity toward the fungal calcineurin compared with
the host (human) enzyme. Such compounds may eventually prove useful as antifungal agents with reduced toxicity and immunosuppressive effect toward the host. One
such candidate is the FK506 analog L-685,818 (18-hydroxy, 21-ethyl-FK506). L-685,818 is toxic to C. neoformans, mediated by binding and inhibiting the fungal calcineurin, but has reduced immunosuppressive activity in
humans (309). Growth studies in the presence of CsA and
FK506 indicate that these drugs are growth inhibitory at
37°C but not at 24°C and that they inhibit a common
target. Disruption of the calcineurin A gene resulted in
mutant strains that are viable at 24°C but do not survive
under conditions that mimic the host environment including elevated temperature, 5% CO2, or alkaline pH (310).
1494
FRANK RUSNAK AND PAMELA MERTZ
These mutant strains are no longer pathogenic, thus indicating that calcineurin is necessary for virulence in this
organism.
6. Dictyostelium discoideum
Calcineurin in the slime mold D. discoideum exhibited the familiar developmental pattern of expression as
noted above for the filamentous fungi, with the highest
level of expression during vegetative growth and decreasing expression during multicellular development (75).
CsA and FK506 had no effect on growth, a process that
can be separated from development in this organism
(159). However, these drugs do inhibit developmental
processes such as stalk cell spore formation and expression of prestalk and prespore developmental markers.
Volume 80
formed indicating that the parasitic calcineurin is functionally and structurally equivalent to mammalian calcineurin. A similar phenomenon was observed with calcineurin from the tapeworms Hymenolepsis microstoma
and Hymenolepsis diminuta such that calcineurin from
both organisms was inhibited by CsA complexed with
mammalian cyclophilin but not H. microstoma cyclophilin (347). This was not the case with calcineurin from
Schistoma mansoni (186) and Plasmodium falciparum
(87). One hypothesis to explain the lack of complex formation with calcineurin is that parasitic cyclophilins are
structurally different from mammalian cyclophilins, such
that cyclophilin residues surrounding the CsA binding site
that interact with calcineurin are not conserved in parasitic cyclophilins. Further studies are necessary to resolve
these interesting findings.
7. Other lower eukaryotic organisms
A gene for calcineurin B has been isolated in the
amoeboflagellate N. gruberi (346). mRNA levels are detectable in the amoebae and are cyclic, with peak abundance during flagellar formation, followed by a gradual
decline. In the unicellular organism Paramecium tetraurelia, calcineurin localization was investigated by use of
a specific antibody and immunocytochemical methods
(209). Calcineurin was largely localized to the cilia and
cell membrane, with only a diffuse staining pattern observed within the cell body. Further staining indicated
that there was no difference in either localization or abundance in cells prepared either in logarithmic or stationary
phase. Thus calcineurin abundance does not appear to
change during the cell cycle as it does in the simple fungi.
To further explore calcineurin’s role in P. tetraurelia,
anticalcineurin antibody or Ca2⫹/calmodulin-calcineurin
was microinjected into cells. Anticalcineurin antibody
blocked exocytosis after treatment with the exocytosis
trigger agent, aminoethyldextran, while microinjection of
a complex of Ca2⫹/calmodulin-calcineurin induced exocytosis. These results implicate calcineurin as the phosphatase previously shown to dephosphorylate a 63-kDa
protein hypothesized to be involved in trichocyst exocytosis (198).
Recently, calcineurin has been isolated from Leishmania major (342) and Leishmania donovani (21). Calcineurin was isolated by chromatographic separation of
cytosol from promastigotes where it was hypothesized to
be a key regulatory component in the life cycle of this
parasite. Interestingly, in L. major, extracellular growth is
not inhibited by CsA, and in fact, a high-affinity complex
of CsA with L. major cyclophilin forms [inhibitory constant (Ki) ⫽ 5.2 nM] but does not inhibit or form a tight
complex with calcineurin from that organism, suggesting
a possible mechanism for this organism’s resistance to
CsA (342). Interestingly, a complex between CsA, recombinant human cyclophilin, and L. major calcineurin was
B. Higher Eukaryotes
1. Calcineurin in plants
Evidence for a plant homolog of calcineurin was first
obtained by Luan et al. (246) who demonstrated, using
patch-clamp techniques, that CsA and FK506 blocked
Ca2⫹-dependent inactivation of K⫹ channels in Vicia
faba. A partially proteolyzed and constituitively active
form of calcineurin also inhibited K⫹ channel activity.
Furthermore, both CsA and FK506 inhibited a Ca2⫹-dependent phosphatase activity in cellular extract. Subsequent studies have provided additional evidence for calcineurin function in plants (reviewed in Ref. 245).
To date, however, calcineurin has not been successfully purified to homogeneity from plant tissue nor have
bona fide genes for either subunit been cloned. The closest contenders are two EF-hand Ca2⫹-binding proteins
encoded for by the SOS3 and AtCBL genes that are homologous to the calcineurin B subunit (218, 239). The
protein encoded by the SOS3 gene is 30% identical to
calcineurin B from various organisms, and mutations in
SOS3 render A. thaliana sensitive to Na⫹ (239). The SOS3
protein is also homologous to NCS-1 (30% identity), a
neuronal Ca2⫹ sensor in the recoverin family of EF-hand
proteins (see sect. IIB2D). The AtCBL proteins are most
homologous to calcineurin B (32% identity to rat calcineurin B) and can complement a yeast calcineurin B
mutation, indicating a calcineurin B-like physiological
function (218). Recent work, however, indicates that the
AtCBL proteins interact with a novel group of protein
kinases in a Ca2⫹-dependent fashion (372). A. thaliana
contains at least six AtCBL genes. The AtCBL and SOS3
proteins clearly play different roles since they are unable
to complement each other (218). It is intriguing that SOS3
and AtCBL encode for Ca2⫹-binding proteins, indicating
that salt stress in plants may be regulated by Ca2⫹-dependent signaling pathways (possibly via calcineurin) as has
October 2000
CALCINEURIN: FORM AND FUNCTION
been found in S. cerevisiae (see sect. IIIA1B). Further
evidence for this hypothesis was obtained in a study
demonstrating that overexpression of an activated form
of yeast calcineurin conferred salt tolerance in transgenic
tobacco plants (320). Similarly, genes for three of the
AtCBL isoforms appear to be stress regulated. Whether
SOS3 or the AtCBL proteins represent plant calcineurin B
homologs or just close relatives will hopefully be resolved
if a protein corresponding to plant calcineurin can be
isolated and/or cloned and shown to be a functional phosphatase.
2. Calcineurin in mammals
A) TISSUE DISTRIBUTION. Calcineurin is widely distributed
in mammalian tissues, with the highest levels found in
brain (168, 175, 216, 437). In addition, calcineurin A and B
subunits have been observed in adipose tissue, adrenal
cells (318, 319), bone osteoclasts (19), heart, hindbrain
and spinal cord (394), kidney (42, 418, 419), liver (135), B
and T lymphocytes (4, 50, 193), lung, medulla, olfactory
bulb, pancreas (112), placenta (314), platelets (406, 438),
retina (66), skeletal muscle (168), smooth muscle, spleen,
testis and sperm (278, 286, 396, 409), thymus (42, 193),
and thyroid (121).
Distinct tissue distribution is observed for the various isoforms of each subunit (42, 175, 219). An isoform of
the catalytic subunit encoded for by the PPP3CC gene
(␥-isoform, see Table 3) is testes specific (291, 292), as is
the product of the PPP3R2 gene encoding an isoform of
the regulatory subunit (48, 424). With the use of polyclonal antibodies that distinguish between the ␣- and
␤-isoforms of calcineurin A (encoded by the PPP3CA and
PPP3CB genes, respectively; Table 3), it was found that
calcineurin A␣ was more abundant than A␤ in the rat
brain and heart, but the relative abundance is reversed in
spleen, thymus, and lymphocytes (175, 219). These results
partly explain the recent finding that PPP3CA knock-out
mice produce T and B cells that mature normally, respond
to mitogenic stimulation, and remain sensitive to both
CsA and FK506, but are defective in in vivo antigenspecific T-cell responses (477). These PPP3CA-deficient
mice also accumulate hyperphosphorylated tau protein
and exhibit cytoskeletal changes in the hippocampus as a
result of reduced calcineurin A␣ activity (181). More recent studies indicate that synaptic depotentiation is completely abolished, indicating that calcineurin A␣ may play
a role in the learning and memory process (181, 485).
B) SUBCELLULAR DISTRIBUTION. Using a radioimmunoassay,
Cheung and colleagues (10) measured the subcellular
distribution of calcineurin in chick forebrain homogenate.
In that study, calcineurin was highly enriched in the cytoplasmic and microsomal fractions as well as synaptosomes. Subsequent studies have confirmed its predominance in the cytoplasm and synaptosomal cytosol (219,
1495
306). Politino and King (329) explored the physical association of calcineurin with synthetic phospholipid vesicles and showed that calcineurin binds small, acidic,
unilamellar vesicles in a Ca2⫹-dependent fashion. Furthermore, the phosphatase activity of calcineurin was profoundly affected by phosphatidylglycerol or phosphatidylserine, with up to a 23-fold increase in activity toward
phosphorylated histone, but inhibition using p-nitrophenylphosphate (p-NPP) or tyrosine phosphate. In a subsequent study it was hypothesized that the phospholipidbinding site was located on the calcineurin B subunit
(330). Using synthetic phospholipid monolayers, Kennedy
et al. (183) investigated the factors that contributed to
calcineurin-phospholipid interactions and found that calcineurin binding is myristoyl independent, mediated by
anionic phospholipids and/or diacylglycerol, and also affected by the presence of calmodulin.
There is overwhelming evidence for calcineurin in
the nucleus along with other calmodulin-binding proteins
such as casein kinase-2 and myosin light-chain kinase (34,
286, 336, 376, 426). In spermatids, calcineurin was localized to the nucleus, and its levels were most abundant
during the initial stage of nuclear elongation, with almost
no signal present in the cytoplasm (286). In the context of
signaling pathways that activate nuclear factor of activated T cells (NF-AT), Shibasaki et al. (376) have shown
that calcium induces an association between calcineurin
and NF-AT that results in colocalization of both molecules
to the nucleus.
Calcineurin has also been shown to be associated
with the cytoskeleton (106). The latter finding is of interest given that several substrates of calcineurin are colocalized to the cytoskeleton including tau (115, 122, 181),
microtubule-associated protein 2 (122, 284), tubulin (122),
dystrophin (275, 440), and dynamin (90, 240).
C) CALCINEURIN FUNCTION. Numerous functions have been
identified for calcineurin in higher eukaryotic organisms,
and it is beyond the scope of this review to cover all of
them comprehensively. Klee et al. (208) have provided a
recent update and cited a number of specific reviews
regarding calcineurin function (208). Table 5 is an attempt
to summarize a number of tissues, systems, and specific
substrates that are implicated to be regulated by calcineurin. In sections IIIB2D and IIIB2E, we provide a short
review of the role of calcineurin on two key systems of
importance in modern biology, apoptosis and cardiac hypertrophy.
D) CALCINEURIN AND APOPTOSIS. It has been recognized for
some time that calcineurin plays a role in programmed
cell death of T and B lymphocytes (32, 110, 116, 481).
Recently, it has also been shown that calcineurin plays a
role in apoptosis in neuronal cells via the cytochrome
c/caspase-3 pathway (17). In T-cell hybridomas, apoptosis
can be stimulated by ligation of the T-cell receptor/CD3
complex and has provided a useful in vitro model to
1496
FRANK RUSNAK AND PAMELA MERTZ
TABLE
Volume 80
5. Physiological roles of calcineurin in higher eukaryotic organisms
Organ or System
Bone
Brain
Gastric chief cells
Heart
Hematopoetic system
Kidney
Liver
Lymphocytes
Pancreas
Pituitary
Placenta
Skeletal muscle
Smooth muscle
Other functions
Function
Reference No.
Osteoclast bone resorption
Amyloid ␤-peptide formation
Apoptosis
Dephosphorylation of DARPP-32
Dephosphorylation of nitric oxide synthase
Endocytosis of synaptic vesicles via dynamin I
Gene regulation
Hormonal control
Ischemia
Long-term potentiation
Memory
Neuritogenesis
Regulation of Ion Channels
Glutamatergic receptors
Glutamate release
␥-Aminobutyric acid receptor
M channel modal gating
Nicotinic receptors
Voltage-gated Ca2⫹ channels
Voltage-gated Na⫹ channels
Voltage-gated K⫹ channels
Exocytosis and pepsinogen secretion
Cardiac hypertrophy
Progenitor mast cell IgE-dependent cytokine transcription
Thymic selection
Erythropoietin-dependent c-myb regulation
Immunosuppressant-induced hypertension
Simulation of Na⫹-K⫹-ATPase activity in renal tubule cells
Adrenergic, dopamanergic, and angiotensin receptor signaling
Glucocorticoid and mineralocorticoid signaling
Protein phosphorylation
Hepatoprotection
T-lymphocyte activation and cytokine signaling
Neutrophil chemokinesis
Lymphocyte degranulation
Apoptosis and FasL gene transcription
Angiotensin II regulation of immune responses
Macrophage effector function
Acinar cell amylase secretion
␤-Cell insulin secretion
Regulation of adenylyl cyclase
Hormone secretion
Epidermal growth factor urogastrone receptor dephosphorylation
Skeletal muscle hypertrophy
Inhibition of L-type Ca2⫹ channels
Spermatid motility
Integrin recycling, integrin/fibronectin interaction
Tumor cell autocrine growth
Dephosphorylation of Elk-1
DNA binding of p53 to HIV-1 long terminal repeat
19
83
17
84, 196
79, 102, 474
90, 225, 240
96, 123, 364, 465
59, 434
41, 285, 422
460, 485
123, 181, 254
49, 106
299, 465
7, 35, 79, 217, 231, 284, 415
308, 427
162
256
241
465, 484
51, 294
348
344
82, 86, 108, 171, 232, 233, 247, 267, 281, 290,
312, 337, 377, 397, 401, 402, 417, 441
180
442
250, 360
356
14
418
418
34, 135
101
46, 340
146
89
32, 110, 116, 157, 226, 243, 374, 375, 421,
429, 443, 463, 481
305
65
125
407
11, 13, 322
11, 12
314
88, 295, 367
362
409
228, 331
298
398, 414
126
HIV, human immunodeficiency virus.
investigate signaling pathways responsible for this biological phenomenon. Both CsA and FK506 inhibit this process, implicating calcineurin in the signaling pathway of
apoptosis which is known to involve a rise in intracellular
Ca2⫹ (110). Similarly, in the B-cell lymphoma cell lines
WEHI-231, B104, and BL60, apoptosis induced by crosslinking of surface immunoglobulin receptors was inhibited by these immunosuppressant drugs (32, 116).
In lymphocytes, calcineurin and NF-AT appear to
participate in apoptosis, in part by mediating the induc-
tion of Fas and Fas ligand which then interact and transduce the apopototic signal after T-cell receptor ligation
(157, 226, 243, 421). Using a constituitive (Ca2⫹- and calmodulin-independent) form of calcineurin, Shibasaki and
McKeon (375) demonstrated that calcineurin functions in
calcium-induced apoptosis in mammalian cells deprived
of growth factors and that this was a direct consequence
of calcineurin’s phosphatase activity. Interestingly, coexpression of Bcl-2 blocked calcineurin-induced apoptosis.
At least one mechanism for how this occurs was provided
October 2000
CALCINEURIN: FORM AND FUNCTION
subsequently by experiments which showed that Bcl-2
forms a complex with calcineurin that targets it to the
cytoplasmic membrane (374). Although still maintaining
phosphatase activity, calcineurin bound to Bcl-2 is unable
to promote nuclear translocation of NF-AT. Furthermore,
BAD, a proapoptotic member of the Bcl-2 family, is a
substrate of calcineurin. Dephosphorylation of BAD by
calcineurin enhances BAD heterodimerization with Bcl-xL
and apoptosis (443).
However, another intriguing hypothesis is that apoptosis is linked to cellular redox homeostasis. Wolvetang
et al. (463) showed that inhibitors of the plasma membrane NADH-oxidoreductase (PMOR) activity induce apoptosis through a signaling pathway involving calcineurin
(463). It was proposed that PMOR serves as a redox
sensor that can regulate the signals required for apoptosis
(227). The finding that calcineurin activity is sensitive to
redox state changes (45, 111, 345, 447, 470, 471) provides
support for this hypothesis and a means by which apoptosis could be regulated by the cellular redox potential.
E) IMPORTANCE OF CALCINEURIN IN CARDIOVASCULAR FUNCTION.
Recently, calcineurin and NF-AT have been implicated in
transducing signals responsible for cardiac morphogenesis and inducing cardiac hypertrophy (82, 232, 233, 281,
337, 377, 401, 402, 417). Thus disruption of the NF-ATc
gene in mice results in failure to develop normal cardiac
valves and septa, and the transgenic mice die from congestive heart failure in utero (82, 337). Overexpression of
calcineurin has also been shown to induce cardiac hypertrophy and heart failure in transgenic mice that could be
blocked by the immunosuppressant drug CsA (281). Furthermore, a transgenic mouse model for hypertrophy in
which tropomodulin-overexpressing transgenic mice develop progressive dilated cardiomyopathy has provided
evidence for increased calcineurin protein levels before
the onset of the hypertrophic phenotype, suggesting that
calcineurin may play an early regulatory role in this process (402). Similar results were found in skeletal muscle
from mice subject to overload (88) and confirmed later in
skeletal muscle cells virally transfected with insulin-like
growth factor I (295, 367). Some of these studies have
even proposed that immunosuppressant drugs such as
CsA and FK506 might be used to treat hypertrophy (312,
401). Indeed, in a subsequent study, Sussman et al. (401)
utilized an aortic banding model to induce hypertrophy
and showed that treatment with CsA, albeit an excessive
dose, resulted in significantly less hypertrophy. However,
although a few studies have confirmed this finding (267,
402), several other groups examining calcineurin’s role in
this process have failed to demonstrate any efficacy of
CsA (86, 247, 290, 476) and, in fact, Molkentin (280) has
responded by reporting that CsA protected against pressure-overload hypertrophy after 7 days but not after 21
days. Although the reason for some of these discrepancies
may be due to the dose of immunosuppressant drug used,
current hypotheses suggest that multiple signaling path-
1497
ways might be recruited to participate in the hypertrophic
response and that inhibition of one parallel pathway (i.e.,
calcineurin) might delay but not prevent hypertrophy
(108, 397, 441). Nevertheless, they implicate a possible
role for calcineurin and NF-AT in cardiac function.
Oxidative stress is also thought to play a role in
cardiomyopathy and heart failure (381). The possibility
that calcineurin may be regulated by oxidative stress
indicates that signaling pathways in which it is involved
may be important in mediating processes that lead to
cardiac dysfunction.
C. Inhibitors of Calcineurin
1. Natural product and synthetic inhibitors
A number of natural products have been isolated that
are potent inhibitors of calcineurin and other serine/threonine protein phosphatases. The most potent, specific,
and well-known inhibitors of calcineurin are the immunosuppressant drugs CsA and FK506 (Fig. 4), which inhibit
calcineurin when complexed with their respective cytoplasmic receptors cyclophilin and FKBP (see Table 1
entry for a number of reviews on these drugs). Interestingly, in vitro calcineurin inhibition by these immunosuppressant drug complexes only occurs when a physiological substrate is used to assay the enzyme such as
phosphocasein or phospho-RII peptide, a peptide whose
sequence represents the phosphorylation site of the regulatory subunit of cAMP-dependent protein kinase, a wellcharacterized and more physiological phosphopeptide
substrate (31). The use of p-NPP as substrate results in
activation of calcineurin by these immunosuppressant
drug complexes (238, 403).
A number of other compounds have demonstrated
inhibitory activity against calcineurin and other serine/
threonine protein phosphatases. Okadaic acid, often used
as a potent and specific inhibitor of PP2A, can also inhibit
PP1 and calcineurin at higher concentrations. The ID50 of
okadaic acid for PP2A has been measured to be ⬃1 nM,
while the ID50 values for PP1 and calcineurin are ⬃300 nM
and ⬃4 ␮M, respectively (29). The cyclic peptide microcycstin LR is a potent inhibitor of PP1 and PP2A, with a Ki
value ⬍0.1 nM. Although the inhibition of calcineurin by
microcystin LR occurs at over 1,000-fold higher concentrations, microcystin LR still is a relatively potent inhibitor of calcineurin (IC50 ⫽ 0.2 ␮M) (248). Dibefurin, a novel
fungal metabolite, also has modest inhibitory activity
against calcineurin (37).
Since the discovery of these natural product inhibitors, several new synthetic compounds have been found
to be reasonable inhibitors of calcineurin and other phosphatases. Tatlock et al. (410) utilized computational docking experiments and synthetic derivatives of the exo,exo7-oxabicyclo[2.2.1]heptane-2,3-dicarboxylic acid ring
1498
FRANK RUSNAK AND PAMELA MERTZ
Volume 80
FIG. 4. Natural product and synthetic inhibitors of calcineurin. For the metal-ligating phosphonate inhibitors, n
refers to the number of methylene units.
system of endothall (Fig. 4) to search for enhanced ligand
binding to calcineurin (410). Endothall is structurally related to the natural defensive toxin of blister beetles,
cantharidin (210), a potent inhibitor of PP1 and PP2A
(210) and a weak inhibitor of calcineurin (98). Substitution at the 5-endo position was hypothesized to provide
reasonable binding interactions that mimicked the interaction between the active site of calcineurin and its autoinhibitory domain. Incorporation of a trans-cyclopropylphenyl group at this position afforded the most potent
inhibitor, with an apparent Ki of 0.5 ␮M. Interestingly, the
tethered dicarboxylic acid moiety and bridgehead oxygen
atom of endothall and cantharidin derivatives were modeled to interact with the active site dinuclear metal center
(see sect. IVC ) (410).
A similar approach to inhibitor design incorporating pendant metal-coordinating groups that could anchor the inhibitor to the active site metal ions has been
introduced by Widlanski and colleagues (296). A variety
of alkylphosphonic acid derivatives containing an additional thiol or carboxylate group (Fig. 4) were explored
as inhibitors of alkaline phosphatase and purple acid
phosphatase. Assuringly, nearly all bound more tightly
than substrate p-nitrophenol and up to 55-fold tighter
than ethanylphosphonic acid, indicating that these additional function groups could improve binding affinity.
Whether binding occurs via direct metal ligation for
endothall and/or alkylphosphonic acid derivatives remains to be demonstrated by spectroscopic means. If
correct, these compounds could provide a route to the
design of more potent and selective metallophosphatase inhibitors.
Peptide inhibitors of calcineurin have been also been
introduced. One of these, a 25-residue peptide based on
the sequence of the autoinhibitory domain of the calcineurin A subunit from residues 457– 481 (Fig. 1), is a
October 2000
CALCINEURIN: FORM AND FUNCTION
relatively potent inhibitor of calcineurin phosphatase activity (137, 325). Recently, a high-affinity calcineurin-binding peptide was selected from a combinatorial peptide
library based on the calcineurin docking motif of NF-AT
(15). The peptide inhibited NF-AT activation and expression of NF-AT-dependent cytokine genes in T cells, but
did not inhibit calcineurin phosphatase activity toward
phospho-RII peptide, and thus did not affect the expression of other cytokines that require calcineurin but not
NF-AT. The latter point is significant because compounds
such as this peptide that selectively interfere with calcineurin-NF-AT interaction without disrupting calcineurin
phosphatase activity may prove to be less toxic immunosuppressants compared with CsA and FK506.
At least one other synthetic calcineurin inhibitor has
been reported, PD 144795, a benzothiophene derivative
shown to have anti-inflammatory and anti-HIV effects
(126). Transcriptional activity mediated by p53 and NF-␬B
were inhibited by both CsA and PD 144795. An in vitro
assay of calcineurin activity from Jurkat cell lysate also
indicates that PD 144795 led to dose-dependent inhibition
of calcineurin.
It was previously concluded by Enan and Matsumura
(93) that class II pyrethroid insecticides were potent inhibitors of bovine brain calcineurin, with IC50 values of
10⫺9 to 10⫺11 M. In that study, p-NPP and O-phospho-DLtyrosine were used as substrates in the inhibition assay.
However, in an independent study, none of the class II
pyrethroids was able to inhibit purified bovine calcineurin
using phospho-RII peptide (97). Calcineurin activity in rat
brain homogenate and in IMR-32 neuroblastoma cells in
culture was also not affected by pyrethroids, indicating
that these insecticides are not effective inhibitors of calcineurin (99).
The tyrphostins A8, A23, and A48, members of a
family of tyrosine kinase inhibitors, inhibited calcineurin
with IC50 values of ⬃10⫺5 M (258). However, the use of
p-NPP as substrate in these studies should be questioned
given the inhibition pattern of calcineurin noted above for
CsA, FK506, and the pyrethroid insecticides. A follow-up
study using phospho-RII peptide or other suitable phosphoprotein substrate may confirm yet another class of
calcineurin inhibitors.
2. Endogenous regulators
In addition to synthetic and natural product inhibitors of calcineurin, a number of endogenous cellular proteins have emerged as inhibitors of calcineurin protein
phosphatase activity and thus potential regulators of its in
vivo function. One of the first to be identified was a
79-kDa protein kinase A anchoring protein (AKAP79)
(58). AKAP79 associates with the regulatory subunit of
the cAMP-dependent protein kinase and localizes it to
postsynaptic densities. Using a yeast two-hybrid approach
1499
to search for proteins that interacted with AKAP79,
Scott and colleagues (58) identified a positive clone encoding the calcineurin A subunit. Immunofluorescence
studies demonstrated that calcineurin and the regulatory
subunit of protein kinase A were colocalized in rat hippocampal neurons via AKAP79. Interestingly, AKAP79
contained a domain homologous to FKBP, hypothesized to be the calcineurin binding domain. A synthetic
peptide based on this sequence was a noncompetitive
inhibitor of calcineurin activity (58). A subsequent study,
however, suggests that AKAP79 interacts with calcineurin
through a site distinct from the FKBP-homologous region
(177).
Another potential calcineurin regulatory protein is
cain/cabin 1, a 2,220-residue phosphoprotein isolated by
yeast two-hybrid screens of either rat hippocampal or
mouse T-cell cDNA libraries (224, 400). Cain/cabin 1 binds
to calcineurin and inhibits it in a noncompetitive fashion.
The interaction between cain/cabin 1 and calcineurin was
dependent on protein kinase C activation, and overexpression inhibited the transcriptional activation of
the interleukin-2 gene and prevented dephosphorylation of the transcription factor NF-AT. Recently, cain/
cabin 1 was found to regulate the transcription factor
MEF2, itself regulated via calcineurin-dependent pathways, by binding to MEF2 and sequestering it in an inactive state (469).
In an expression library screen searching for proteins
that interact with the ubiquitously expressed Na⫹-H⫹ exchanger NHE1, Lin and Barber (234) identified a novel
protein, CHP (see sect. IIB2D), that specifically bound
NHE1 and was critical for growth factor stimulation of
exchange activity. Overexpression of CHP in Jurkat and
HeLa cells resulted in inhibition of NF-AT nuclear translocation and transcriptional activity that was hypothesized to be the result of calcineurin inhibition (235). Indeed, the phosphatase activity of immunoprecipitated
calcineurin was inhibited 50% in cells overexpressing
CHP, whereas in a reconstitution assay, the activity of
purified calcineurin was inhibited nearly quantitatively in
a dose-dependent fashion. These results indicate that
CHP could represent yet another member of this emerging class of endogenous calcineurin regulators.
Recently, a protein of the African swine fever virus,
A238L, was found to display immunosuppressive activity
by inhibiting NF-AT-regulated gene transcription in vivo
(277). A238L coimmunoprecipitated with calcineurin after
viral infection of Vero cells, and calcineurin phosphatase
activity was inhibited in cellular extracts from viral-infected cells. It was hypothesized that A238L may enable
the virus to evade host defense systems by preventing
transcriptional activation of genes important for host immunity.
Although the classical mechanism for regulating calcineurin activity is via Ca2⫹/calmodulin, it is intriguing to
1500
FRANK RUSNAK AND PAMELA MERTZ
speculate that these and possibly other proteins can interact with calcineurin to regulate subcellular targeting
and/or activity toward specific substrates in novel yet
undefined ways.
IV. CALCINEURIN STRUCTURE
A. A Dinuclear Metal-Binding
Phosphoesterase Motif
Before the availability of any structural data, Averill
and colleagues (431, 432) predicted that the serine/threonine protein phosphatases were homologous to purple
acid phosphatases and therefore might contain an active
site dinuclear metal center. Their hypothesis was based
on a comparison of the primary sequences of serine/
threonine protein phosphatases with human, porcine, and
bovine purple acid phosphatases, enzymes which were
already well characterized and known to contain
dinuclear iron centers. Their prediction was correct, and
the authors were able to identify three of the metal ligands.
With increasing sequence data available, several
groups have completed comprehensive sequence alignments of serine/threonine protein phosphatases and have
identified a number of residues that are conserved in all
members of this family (26, 132, 212, 244, 371, 486). These
studies identified a “phosphoesterase motif” (Fig. 5) that
is conserved not only in PP1, PP2A, and calcineurin,
but in many other enzymes involved in the cleavage of
phosphoester bonds, including acid and alkaline phosphatases, bacterial exonucleases, diadenosine tetraphosphatase, 5⬘-nucleotidase, phosphodiesterase, sphingomyelin phosphodiesterase, an enzyme involved in RNA
debranching, and a phosphatase in the bacteriophage ␭
genome, ␭ protein phosphatase (244).
FIG. 5. The phosphoesterase motif as found in calcineurin and
purple acid phosphatase. The original motif was identified by Koonin
(212) and modified by Zhuo et al. (486). Residues that have been identified as metal ligands are shown in bold (for calcineurin A␣, these
represent residues Asp-90, His-92, Asp-118, and Asn-150; cf. Figs. 7A and
10). Conserved, nonligand residues also found in the active site (Asp121, Arg-122, and His-151) are underlined. The phosphoesterase motif of
plant (200) and mammalian purple acid phosphatases (92, 185, 380) is
also shown (bold residues are represented as Asp-135, Asp-164, Y-167,
and Asn-201, while the underlined residues are represented as Asp-169
and His-202 in the kidney bean enzyme; cf. Figs. 7B and 11). The motif
in the plant and mammalian purple acid phosphatases represents a
variation of the phosphoesterase motif. PP1, protein phosphatase 1.
Volume 80
Four of the residues in the phosphoesterase motif
(bold letters, Fig. 5) are ligands to a dinuclear metal
cofactor in PP1 and calcineurin. Thus it has been hypothesized that this motif provides a scaffold for an active site
dinuclear metal center in every member of the phosphoesterase family (244, 351). Support for this hypothesis was
provided in studies of ␭ protein phosphatase which demonstrated a spin-coupled dinuclear metal binding site by
spectroscopic means (272, 352), and in the recent determination of the three-dimensional structure of E. coli
5⬘-nucleotidase, a distant member of the metallophosphatase superfamily that has an active site containing two
Zn2⫹ separated by 3.3 Å (211). The conserved phosphoesterase motif suggests a common catalytic mechanism for
enzymes involved in phosphotransfer reactions, a hypothesis that seems to be the case for at least two of these,
calcineurin and ␭ protein phosphatase (see below).
The phosphoesterase motif is also found in purple
acid phosphatases, albeit in a slightly modified form (Fig.
5). Despite these differences, the phosphoesterase motif
in purple acid phosphatase has a similar ␤-␣-␤-␣-␤ fold
accommodating the dinuclear metal center (199, 201,
395).
B. Three-Dimensional Structure
The three-dimensional structures of several enzymes
in the metallophosphatase family have been solved. X-ray
structures (with highest resolution noted in parentheses)
of PP1 (2.1 Å) (91, 120), calcineurin (2.1 Å) (124, 197),
kidney bean purple acid phosphatase (2.65 Å) (199, 201,
395), mammalian purple acid phosphatase (1.55 Å) (128,
425), and the periplasmic 5⬘-nucleotidase from E. coli (1.7
Å) (211) have been solved. In these structures, the phosphoesterase motif described in the previous section is
represented as a ␤-␣-␤-␣-␤ scaffold for an active site
dinuclear metal center. The three ␤-strands of this motif
form a parallel pleated sheet that is capped by intervening
␣-helices. Two metal ions are positioned at the apex of
this fold forming a dinuclear metal center with 3.0 – 4.0 Å
between metal ions, with four of the metal ligands provided by residues in loops between ␤-sheets and ␣-helices.
A ribbon diagram representing the X-ray structure of
phosphate-inhibited calcineurin, complexed with the immunosuppressant drug complex FK506䡠FKBP, is shown in
Figure 6. The overall structure of the catalytic subunit
(shown in gray) is ellipsoidal and consists of a mixture of
␣-helices and ␤-sheets. The metal ions of the dinuclear
metal center are obscured in this diagram by the orange
van der Waals spheres of phosphate that form a bridge
between the two metal ions (see Fig. 10, below). The
calcineurin B-binding domain (cf. Fig. 1) is evident in this
structure as an ␣-helix that protrudes from the core of the
October 2000
CALCINEURIN: FORM AND FUNCTION
1501
FIG. 6. Ribbon diagram showing the three-dimensional structure of calcineurin complexed with FK506䡠FK506
binding protein (FKBP). The figure was created using the X-ray coordinates of Griffith et al. (protein data bank file 1TCO)
(124), using MOLSCRIPT (214) and Raster3D (20). The catalytic subunit calcineurin A is colored in gray, the calcineurin
B subunit is yellow, FKBP is green, and the immunosuppressant drug FK506 is shown in ball-and-stick fashion in gray.
The four Ca2⫹ of the B subunit are represented as blue spheres, and the myristoyl group of the B subunit is colored as
red van der Waals spheres. The phosphate molecule bound to the active site metal ions of the A subunit is represented
as orange van der Waals spheres.
molecule, forming the binding site for the calcineurin B
subunit (Fig. 6, yellow). Absent in this structure are the
calmodulin-binding and autoinhibitory domains, since a
truncated form of calcineurin missing these domains was
the source of protein for crystallization studies (124). In a
subsequent study that determined the structure of the
holoenzyme, it was found that the autoinhibitory domain
folds into an ␣-helix that binds to the substrate-binding
cleft of the catalytic domain, with one of the glutamate
residues forming hydrogen bonds to metal-coordinated
water molecules (197). Interestingly, the calmodulin domain was disordered in the X-ray structure, and therefore,
our knowledge of how this domain interacts with the
active site and autoinhibitory domains to confer calmodulin-regulation remains rudimentary.
The calcineurin B subunit in Figure 6 is colored in
yellow and is seen forming a complex with calcineurin A
via the calcineurin B-binding helix. Four Ca2⫹, shown as
blue spheres, are bound in the EF-hand domains of the B
subunit. The NH2-terminal myristoyl group of calcineurin
B (red van der Waals spheres) is situated in a hydrophobic
cleft between two amphipathic helices that are integral
components of two EF-hands (124). In contrast, the X-ray
structure of the nonmyristoylated protein shows the first
five residues of the NH2 terminus of calcineurin B as
disordered, suggesting increased mobility of the NH2 terminus relative to the acylated protein (197). Thus the
myristoyl group appears to be anchored via multiple hydrophobic contacts and may contribute to the overall
structural stability of the enzyme and may explain why
the acylated protein has increased thermal stability compared with the unmodified enzyme (182).
1502
FRANK RUSNAK AND PAMELA MERTZ
C. Active Site Architecture
The dinuclear metal cofactor of calcineurin has been
modeled as an Fe3⫹-Zn2⫹ cluster based on the presence of
near-stoichiometric quantities of Fe3⫹ and Zn2⫹ (195,
471), electron paramagnetic resonance (EPR) spectroscopic experiments (471), and X-ray diffraction studies
which show a dinuclear metal center separated by 3.14 Å
in a coordination environment shown in Figure 7A (197).
A similar coordination environment and metal-metal distance is also found in PP1 (120), although in that enzyme,
the metal ions could not be identified with certainty and
were therefore referred to as M1 and M2. In the X-ray
structure of calcineurin, the Fe3⫹ was modeled in the M1
site largely based on a comparison to the Fe3⫹-Zn2⫹ active
site metal cofactor of kidney bean purple acid phosphatase. In purple acid phosphatase, a tyrosine residue coordinates to the Fe3⫹ (M1 site) and gives rise to a tyrosineto-iron charge transfer band at 510 –550 nm, responsible
for the purple color of these enzymes (Fig. 7B) (223, 433).
As can be seen in Figure 7, this tyrosine ligand is missing
in calcineurin and is replaced by a histidine, thus explaining why calcineurin does not exhibit any appreciable
absorbance in the visible region of the optical spectrum
Volume 80
(471). The additional substitution of a histidine ligand
with a water molecule results in a net water-for-tyrosine
substitution at the Fe3⫹ site in calcineurin compared with
purple acid phosphatase. The coordination of the Zn2⫹
(M2) provided by amino acid side chains in these enzymes
is identical, with ligands provided by two histidines, an
aspartic acid that bridges to the Fe3⫹ and an asparagine
residue (compare Fig. 7, A and B). Figure 8 shows a
comparison of the active sites of kidney bean purple acid
phosphatase and calcineurin that illustrates the remarkable superposition of both metal ions and protein ligands
in these two enzymes.
In addition to coordination provided by protein side
chains, the metal ions in calcineurin and purple acid
phosphatase have one or more solvent molecules as additional ligands. All enzymes show bridging and terminal
solvent molecules, but at least in one X-ray structure of
calcineurin, an additional terminal solvent molecule was
modeled into the coordination sphere of the M1 site as
shown in Figure 7A (197). From the iron-oxygen bond
length (2.1 Å in purple acid phosphatase, Ref. 128), the
bridging solvent molecule is most likely a hydroxo group
based on distances obtained from model compounds
(223).
FIG. 7. A: schematic of the active site of human
calcineurin based on the 2.1-Å resolution structure described by Kissinger et al. (197). B: schematic of the
active site of kidney bean purple acid phosphatase
based on the 2.65-Å resolution structure described by
Klabunde et al. (201).
October 2000
CALCINEURIN: FORM AND FUNCTION
1503
FIG. 8. Comparison of the active sites of calcineurin
and kidney bean purple acid phosphatase. An alignment
of the active sites of calcineurin and purple acid phosphatase was performed based on the conserved His-Asp
(calcineurin His-151/Asp-121) pair using the program
Quanta (Quanta 97 Molecular Modeling Software Package from Molecular Simulations, 1997). The X-ray crystal
structures used were Protein Data Bank entries 1AUI
and 1KBP for human calcineurin (197) and kidney bean
purple acid phosphatase (201), respectively. The A subunit of calcineurin and a monomer form of purple acid
phosphatase were used to generate the alignments. Only
the metal ions and the active site residues shown in the
figure were used for the alignment. Calcineurin and purple acid phosphatase residues are indicated in red and
yellow, respectively. The Zn atom is green, and the Fe
atom is purple. The numbers denote the following residues: 1) His-92 in calcineurin, Tyr-167 in purple acid
phosphatase; 2) Asp-121 in calcineurin, Asp-169 in purple acid phosphatase; 3) His-151 in calcineurin, His-202
in purple acid phosphatase; 4) Asn-150 in calcineurin,
Asn-201 in purple acid phosphatase; 5) His-199 in calcineurin, His-286 in purple acid phosphatase; 6) His-281
in calcineurin, His-323 in purple acid phosphatase; 7)
His-325 in purple acid phosphatase; 8) Asp-90 in calcineurin, Asp-135 in purple acid phosphatase; 9) Asp-118
in calcineurin, Asp-164 in purple acid phosphatase.
D. Metal Ion Requirements
It has been known for some time that divalent cations
in assay buffers are necessary to achieve the high activities of purified calcineurin, with the best activators being
Mn2⫹ and Ni2⫹ (194, 230, 315). Mn2⫹ and Ni2⫹ have been
shown to increase the activity of calcineurin in the absence of Ca2⫹/calmodulin (315), to prevent inactivation,
or to restore activity following inactivation by exposure to
Ca2⫹/calmodulin (194). Similar divalent metal ion effects
have been observed with PP1 and PP2A (3, 28, 43, 53, 54,
94, 478) and with ␭-protein phosphatase (486). It is unclear why these divalent metal ions are potent activators.
One possibility is that Mn2⫹ or Ni2⫹ are native metal ions
that become lost during purification. Pallen and Wang
(317) incubated calcineurin with Ni2⫹ or Mn2⫹ and
showed that these metal ions are not dissociable by extensive dialysis or gel filtration but can be released after
prolonged exposure to Ca2⫹/calmodulin or by the use of
chelating reagents, both of which occur during a typical
purification protocol. To address this issue, Rao and Wang
(341) used an anticalcineurin immunoaffinity matrix to
rapidly purify calcineurin from crude bovine brain extract
in the absence of calmodulin. Analysis showed that the
immunoprecipitated calcineurin did not contain significant amounts of Ni2⫹ and Mn2⫹. Although no mention
was made of the Fe content, ⬃1 equivalent of Zn2⫹ was
found in all samples, suggesting that Zn2⫹ is an intrinsic
metal ion but not Ni2⫹ or Mn2⫹.
An alternative hypothesis to explain the mechanism
of divalent metal ion activation is that prolonged exposure of calcineurin to Ca2⫹/calmodulin promotes the release of the intrinsic Fe and Zn metal ions and subsequent
replacement by Mn2⫹ or Ni2⫹. The most efficient method
to purify calcineurin utilizes Ca2⫹/calmodulin affinity
chromatography. It is possible that calmodulin binding
exposes the active site and thereby promotes loss of Fe
and/or Zn while calcineurin is adsorbed to the matrix.
Alternatively, because elution of calcineurin from calmodulin-Sepharose requires the use of a metal chelator (e.g.,
EDTA), it is possible that the Fe3⫹ and/or Zn2⫹ may be
removed during elution. Thus far, neither hypothesis has
1504
FRANK RUSNAK AND PAMELA MERTZ
been rigorously tested. However, King and Huang (195)
demonstrated that there was no correlation between the
loss of enzymatic activity after calmodulin-dependent inactivation and the iron and zinc content. Both metal ions
remained tightly bound during prolonged exposure to
calmodulin. Nevertheless, additional studies have demonstrated that up to two equivalents each of Mn2⫹ and Ni2⫹
could bind to calcineurin (317, 482), a result consistent
with these metal ions occupying the sites of the dinuclear
metal cluster. In support of this was an EPR study following Mn2⫹ binding to ␭-protein phosphatase which found
that a dinuclear Mn2⫹-Mn2⫹ cluster was formed upon
addition of two equivalents of Mn2⫹ to the apoenzyme
(352). The situation with calcineurin is not as straightforward due to the presence of calmodulin and the calcineurin B subunit, each which can provide additional
divalent metal ion binding sites. Indeed, an EPR study
following Mn2⫹ binding to calcineurin in the presence of
calmodulin demonstrated 10 Mn2⫹ sites and attributed 4
each to calmodulin and calcineurin B and 2 to the catalytic subunit (453).
Clearly further work is needed to understand the mechanism whereby exogenous Ni2⫹ and Mn2⫹ activate calcineurin after prolonged exposure to Ca2⫹/calmodulin in
vitro and whether or not this activation also occurs in vivo.
V. ENZYMATIC MECHANISM
A. Mechanism of Phosphoryl Group Transfer:
Evidence for Direct Transfer to Water
Much of the initial work on the mechanism of calcineurin focused on determining whether a phosphoenzyme intermediate was formed during catalysis, since
other phosphatases are known to proceed by this mechanism (Fig. 9). For example, E. coli alkaline phosphatase
catalyzes phosphate ester hydrolysis by first transferring
the phosphoryl group to an active site serine residue to
Volume 80
form a transient phosphoenzyme intermediate (63, 64). In
the next step, the enzyme is regenerated for another
round of catalysis by hydrolysis of the phosphoenzyme
intermediate. Alkaline phosphatase is a metalloenzyme
containing a Mg2⫹ and a Zn-Zn dinuclear center reminiscent of the dinuclear metal sites of the serine/threonine
phosphatases and purple acid phosphatases (187). A
phosphoenzyme intermediate has also been demonstrated in the protein tyrosine phosphatases (52, 127,
479), enzymes that function without active site metal ions.
Steady-state experiments of calcineurin carried out
by Graves and colleagues (262) suggested that a phosphoenzyme intermediate was not formed during catalysis.
A linear relationship between log (V/K) and the pKa of the
leaving group was observed for a set of four substrates,
with a trend toward increasing velocity as the pKa of the
leaving group decreased (262). This result is consistent
with a direct transfer to water without formation of a
phosphoenzyme intermediate (Fig. 9). Additional experiments failed to demonstrate phosphotransferase activity
in the presence of alternate nucleophiles, which also is
consistent with a direct transfer mechanism (262). In a
subsequent study using p-NPP as substrate, product inhibition studies showed that both phosphate and the phenol
were competitive inhibitors of calcineurin (259). These
inhibition patterns are consistent with a random uni-bi
mechanism. In contrast, an ordered uni-bi mechanism is
expected for a phosphoenzyme intermediate, since the
product alcohol would be released before the phosphoenzyme intermediate is hydrolyzed. Although these results
are consistent with a direct transfer mechanism, they are
not conclusive, since they can also be interpreted to
indicate a mechanism involving a transient phosphoenzyme intermediate whose breakdown is not rate limiting.
The most definitive work that addressed whether a
phosphoenzyme intermediate was formed during catalysis for the metallophosphatases was performed with bo-
FIG. 9. Phosphatase catalysis: phosphoenzyme intermediate versus direct
transfer mechanisms. Shown are two
possibilities by which phosphatases catalyze phosphoryl group transfer to water.
Escherichia coli alkaline phosphatase
and protein tyrosine phosphatases represent two classes of phosphatases that utilize the phosphoenzyme intermediate
mechanism. The metallophosphatases
appear to proceed by direct transfer of
the phosphoryl group to a metal-coordinated water molecule, without the formation of a phosphoenzyme intermediate.
The “X” symbol represents an active nucleophile, e.g., a cysteine residue in the
protein tyrosine phosphatases and a
serine residue in alkaline phosphatase.
October 2000
CALCINEURIN: FORM AND FUNCTION
vine spleen purple acid phosphatase. Using a chiral
[18O,17O]phosphorothioate ester and carrying out the hydrolysis in [16O]H2O, Knowles and co-workers (288) demonstrated that purple acid phosphatase carries out hydrolysis with net inversion of configuration at the phosphorus
center, thus indicating that the phosphoryl group is transferred directly to water. Because of the similarity of the
active sites of calcineurin and PP1 to purple acid phosphatases, it is thought that the catalytic mechanism of the
serine/threonine phosphatases also proceeds by a similar
mechanism.
B. Catalytic Role of the Dinuclear Metal Center
Several pieces of data indicate that the dinuclear
metal center is a key component of the active site of
calcineurin. 1) As already mentioned, the dinuclear metal
center has a ligand environment similar to purple acid
phosphatases, enzymes which contain dinuclear Fe-Fe or
Fe-Zn centers previously demonstrated to be essential for
catalytic activity (223, 433). 2) Crystallographic (91, 124,
201) and spectroscopic (80, 334, 416, 448, 449, 470) studies indicate that the product of the reaction, phosphate,
and product analogs such as tungstate, molybdate, and
arsenate, coordinate the metal ions. 3) Redox titrations of
either the Fe3⫹-Zn2⫹ or Fe3⫹-Fe2⫹ forms of calcineurin
and purple acid phosphatase indicate a correlation between enzyme activity and the oxidation state of the
bound metal ions (8, 9, 18, 77, 470, 471).
The metal ions of the dinuclear center could function
in numerous ways to catalyze phosphate ester hydrolysis.
The Lewis acidity of the metal ion(s) could serve to
activate a solvent molecule, a well-known mechanism in
several metalloenzymes such as carbonic anhydrase
(132a) and adenosine deaminase (459). A metal-activated
water molecule has been proposed for purple acid phosphatases (85, 432). Alternatively, a metal-coordinated solvent molecule could serve as a general acid to donate a
proton to the leaving group, as has been proposed for
inorganic pyrophosphatase (140, 355). In addition to a
role in activation of the solvent nucleophile, the metal
ions in the serine/threonine phosphatases could be involved in other aspects of the catalytic mechanism. Metal
coordination of the phosphate ester could have several
positive effects acting to accelerate hydrolysis. Neutralization of the negative charge on the oxygen atoms of the
phosphate ester oxygen atoms would increase the electrophilicity of the phosphorus atom, making it more prone
to nucleophilic attack. During P-O bond scission, the
metal ions could stabilize the developing charge on the
leaving group. However, another possible role for the
metal ions could be to orient the substrate for in-line
attack. These are discussed in section VC in the context of
specific active site residues and the effect that mutagenesis of these residues has on catalytic efficiency.
1505
C. Conserved Active Site Residues
In addition to the metal ions and their cognate protein ligands, several conserved residues within 4 –9 Å of
the dinuclear metal cofactor are involved in catalysis. One
of these is a histidine residue that is not a metal ligand but
is within 5 Å of either metal ion. The conserved histidine
in calcineurin, His-151, is also conserved in PP1 (His-125),
purple acid phosphatases (His-202, kidney bean enzyme
numbering), and ␭-protein phosphatase (His-76). It is represented in the phosphoesterase motif as the underlined
residue of Figure 5 (244). This histidine is hydrogen
bonded to an aspartic acid that is also part of the phosphoesterase motif, Asp-121 in calcineurin, Asp-95 in PP1,
Asp-169 in kidney bean purple acid phosphatases, and
Asp-52 in ␭-protein phosphatase. Two arginines, Arg-122
and Arg-254, are also present in the active site of calcineurin. These arginines are also conserved in other
metallophosphatases and correspond to Arg-96 and Arg221 in PP1 (120) and Arg-53 and possibly Arg-162 in
␭-protein phosphatase. Figure 10 depicts the active site of
calcineurin in the phosphate-inhibited form showing the
conserved nonligand residues. The corresponding residues in kidney bean purple acid phosphatases are depicted in Figure 11. In purple acid phosphatase, two histidines (His-295 and His-296) are substituted for two
arginines in the serine/threonine phosphatase family.
The importance of these residues and their potential
role in catalysis has in most cases been addressed by
FIG. 10. Schematic of the active site of calcineurin complexed with
product phosphate. Shown is the dinuclear metal center of calcineurin,
along with the conserved His-Asp pair (His-151/Asp-121) and two arginine residues (Arg-122/Arg-254) that form a region of positive electrostatic potential in the active site.
1506
FRANK RUSNAK AND PAMELA MERTZ
FIG. 11. Schematic of the active site of purple acid phosphatase
complexed with product phosphate. Shown is the active site of purple
acid phosphatases with the conserved His-Asp pair (His-202/Asp-169)
and two histidines (His-296/His-295) thought to be important for the
enzyme’s mechanism (91, 124).
enzymatic and spectroscopic studies of enzymes in which
these residues have been altered by site-directed mutagenesis. The interpretation of these data assumes that
mutagenesis alters only the physical-chemical features of
the residue of interest and does not alter some other
requisite structural feature, (e.g., tertiary structure, metal
binding, etc). A review of the primary literature indicates
that this assumption has not always been defended by
rigorous structural studies. The reader is therefore cautioned that the roles ascribed to some of these active site
residues are still tenuous.
1. Role of the histidine/aspartate pair
in the metallophosphatases
Studies have shown that mutation of His-151 in calcineurin, or its analog in ␭-protein phosphatase, His-76,
leads to significant reductions in enzyme activity (272,
487). When its hydrogen-bonding partner (corresponding
to Asp-121 in calcineurin) is mutated, a 101- to 103-fold
decrease in the rate constant for catalysis (kcat) occurs
with little effect on Km (161, 475, 487). One role considered for this histidine has been as an active site nucleophile. As discussed above, the data argue against this
since a phosphoenzyme intermediate probably does not
form during the catalytic cycle. Alternatively, this histidine may have a role in binding substrate, orienting the
nucleophilic water molecule, and/or in acid/base catalysis. These possibilities are discussed in the following
sections.
Volume 80
A) ACID/BASE CATALYSIS. A series of kinetic studies were
undertaken to study the effect of mutating the conserved
histidine in calcineurin (His-151) and ␭-protein phosphatase (His-76) (272, 283). In both mutant enzymes there
were significant reductions in enzyme activity but only
small effects on Km (272, 283). This was the case using
p-NPP and the [P]-RII peptide as substrates for calcineurin, and p-NPP and phenyl phosphate for ␭-protein
phosphatase. These substrates were chosen to explore
the role of this histidine as a proton donor. A phosphate
ester such as p-NPP has a better leaving group (pKa of
p-nitrophenol ⫽ 7.2) compared with either phenyl phosphate or a phosphoseryl peptide (pKa of phenol ⫽ 10, pKa
serine side chain -OH ⬃14). Therefore, if protonation of
the leaving group occurs in the transition state, hydrolysis
of p-NPP will require less catalytic assistance than either
phenyl phosphate or a phosphoseryl peptide such as [P]RII peptide. As a result, one would expect larger effects on
relative kcat (wild-type kcat/mutant kcat) for substrates
with poorer leaving groups. The results showed less than
a threefold difference in relative kcat for calcineurin using
p-NPP versus the [P]-RII peptide, despite a difference in
pKa for the respective leaving groups of ⬎106 (272). With
␭-protein phosphatase, mutagenesis of His-76 to Asn
(H76N) resulted in the same 500- to 600-fold reduction in
kcat using either p-NPP or phenyl phosphate as substrates
(leaving group pKa difference ⬎103) (272). Ideally, more
than two substrates with varying pKa values should be
used in this type of Brønsted analysis.
A useful method to determine if a residue is acting as
an acid or base in catalysis is to follow the pH dependence
of the rate for wild-type and mutant enzymes. Often one
arm of the usual bell-shaped dependence curve is missing
for the mutant if the residue is acting as a general acid or
base. Using p-NPP as a substrate, pH dependence studies
for wild-type and H76N ␭-protein phosphatase were performed (156). The mutant enzyme had a lower catalytic
rate at every pH but still exhibited a bell-shaped pH curve.
However, kinetic data were difficult to obtain at low pH
due to substrate inhibition that necessitated higher metal
ion concentrations. In summary, the difficulties encountered were such that it was not possible to conclude
whether this histidine serves as a general acid in catalysis.
It was concluded that this residue may function more
decisively as a general acid in reactions of phosphoprotein substrates that have a much poorer leaving group. It
is noteworthy that the pH optimum for the mutant appeared to be shifted to pH 7.0, compared with 7.8 for the
wild-type enzyme. Also the Km for substrate increased at
high pH for the native enzyme to values that were similar
to those of the mutant enzyme at all pH values, indicating
that protonation of this residue may assist in substrate
binding.
B) KINETIC ISOTOPE EFFECTS. Kinetic isotope effect studies
have been performed with both calcineurin (150, 260, 261)
October 2000
CALCINEURIN: FORM AND FUNCTION
and ␭-protein phosphatase (156) to study the chemistry of
the transition state. In the case of calcineurin, the use of
D2O to measure a solvent isotope effect found no effect
on kcat but a modest isotope effect (1.35) on kcat/Km (260).
The lack of a solvent isotope effect on kcat may be due to
having some other step in the reaction mechanism other
than the proton transfer step be rate limiting, a situation
that was confirmed in a subsequent study that used heavy
atom isotopes of p-NPP (150). In the latter study it was
shown that P-O bond cleavage was partially rate limiting
at the pH optimum, and therefore, the isotope effects
were suppressed, a situation which could be improved
upon raising the pH from 7.0 to 8.5. For the ␭-protein
phosphatase reaction, P-O bond cleavage was fully rate
limiting, and the measured isotope effects represented the
intrinsic isotope effects on the bond-breaking step of the
catalytic mechanism. With both enzymes, the data indicate that the substrate of the reaction is the p-NPP dianion, the predominant form at neutral pH (pKa2 ⫽ 4.96).
The isotope effect studies also provide evidence for a
dissociative mechanism, in which the transition state has
substantial P-O bond cleavage before bond formation to
the nucleophilic water molecule. A dissociative mechanism has also been observed for the protein tyrosine
phosphatases (148, 151, 152) and for uncatalyzed reactions in solution (149). This result is important because it
demonstrates that the metal ions do not change the transition state of the phosphoryl transfer reaction to become
more associative, a hypothesis that was previously forwarded based on the idea that metal ions could stabilize
the extra negative charge in the transition state that results from bond formation to the nucleophile (139). Instead, the transition state of the phosphoryl group looks
very much the same as in the solution reaction.
The 15N kinetic isotope effects indicate that in calcineurin and wild-type ␭-protein phosphatase, there is
substantial charge neutralization of the leaving group in
the transition state (150, 156). In contrast, the magnitude
of this charge increases when His-76 of ␭-protein phosphatase is mutated to asparagine. This result suggests that
this histidine may be protonating the leaving group in the
transition state; its absence would lead to a greater charge
accumulation on the leaving group compared with the
wild-type enzyme. Interestingly, the magnitude of negative charge on the leaving group is smaller than expected
for the mutant enzyme compared with comparable studies with protein tyrosine phosphatases that have the active site general acid removed by mutagenesis. Indeed, the
magnitude of the isotope effect in the mutant is significantly less than would result from a full negative charge
on the departing p-nitrophenol product. One explanation
that was forwarded is that the metal ions may participate
in stabilizing the transition state of the reaction by neutralizing the developing negative charge on the leaving
group.
1507
C) POTENTIAL ROLE IN SUBSTRATE BINDING. Another possible
role of His-76/His-151 could be to assist in substrate binding. With ␭-protein phosphatase, the Km for p-NPP increased from 1 to 70 mM with increasing pH (156). In
contrast, the Km for p-NPP in the ␭-PP(H76N) mutant was
higher at low pH compared with wild type. As the pH was
increased to neutral pH and greater, the Km values for
both enzymes became similar. It is possible that at acidic
and neutral pH, where the histidine would be protonated,
this residue assists in substrate binding via electrostatic
interactions. For example, a proton on His-76 may form a
hydrogen bond with one of the substrate oxygen atoms of
the phosphorylated substrate. In X-ray structures of other
serine/threonine phosphatases with bound inhibitors,
phosphate or tungstate, this histidine is within hydrogen
bonding distance of the most solvent exposed oxygen
atom of the inhibitor (91, 124). A similar orientation is
observed with purple acid phosphatases with phosphate
or tungstate bound (128, 201, 425). This type of electrostatic interaction will not affect the isotope effect on
18
(V/K)nonbridge. However, if the substrate were actually
protonated in the transition state, this would have been
revealed in the 18(V/K)nonbridge isotope effect; this was not
observed (156). It is not clear if His-76 really does participate in substrate binding, since at basic pH it would be
deprotonated. The pH optimum for the wild-type enzyme
is ⬃8.0, where both the wild-type enzyme and mutant
have similar Km values for p-NPP.
D) GENERAL BASE CATALYSIS. Another possibility for His151/His-76 would be to act as a general base to deprotonate the metal-bound water molecule that is the nucleophile in the mechanism. In one X-ray structure of
calcineurin, the N⑀ of His-151 is hydrogen-bonded to one
of two terminal solvent molecules coordinated to the Fe
ion. Considering also Asp-121, which is hydrogen bonded
to His-151, the interaction of this histidine/aspartate pair
with the metal-coordinated solvent molecule is analogous
to the Asp-His-Ser catalytic triad of serine proteases,
where an Asp/His pair is important for interacting with
the nucleophilic serine residue. The “Asp-His-HO-metal”
motif in the serine/threonine phosphatases can be
thought of as a catalytic tetrad, with the metal ion serving
to lower the pKa of the nucleophilic water molecule while
the Asp-His functions as a catalytic base to assist in
hydroxide formation. Histidines in other metalloproteins
are thought to perform a similar role by acting as a base
to deprotonate a metal-bound water molecule or by assisting in stabilizing a metal-bound hydroxyl group. For
example, His-372 in E. coli alkaline phosphatase forms a
hydrogen bond with Asp-327 (a bidentate Zn ligand) and is
thought to lower the pKa of a zinc-bound water molecule
(464). Mutagenesis studies of Asp-327 to asparagine result
in increased enzyme activity, since the negative charge of
the hydroxyl group is more stable due to the loss of the
aspartate side chain and replacement with a neutral res-
1508
FRANK RUSNAK AND PAMELA MERTZ
idue. However, this comes at the expense of Zn binding
since a carboxylate is a better metal ligand than a carboxamide, and thus a higher concentration of Zn in the assay
is needed to achieve this increased activity. Examples of
other enzymes where a histidine is postulated to deprotonate a metal-coordinated water molecule include His320 in Klebsiella aerogenes urease (172, 321), His-231 in
thermolysin (27), and His-89 in Serratia nuclease (109).
E) ORIENTATION OF THE NUCLEOPHILIC HYDROXIDE SOLVENT MOLE-
Another potential role of His-76/His-151 could be to
position the metal-coordinated hydroxide for optimal inline attack of substrate, a mechanism analogous to the
concept of “orbital steering” proposed by Storm and Koshland (393) almost 30 years ago to provide an explanation
for the great rate accelerations seen in enzyme catalysis.
The theory of orbital steering postulates that a major
factor in catalytic enhancement is that an enzyme arranges the reaction trajectory to optimize the overlap of
attractive (bonding) orbitals and minimize the overlap of
repulsive (nonbonding) orbitals. This concept has been
debated and contested by numerous groups (39, 174, 270).
Nevertheless, in a recent study of isocitrate dehydrogenase by Koshland and co-workers (273), small structural
perturbations in isocitrate dehydrogenase were created to
evaluate the contribution of precise substrate alignment
to the catalytic rate of an enzyme. They found that small
changes in the orientation of substrates had large effects
on reaction velocity (103- to 105-fold decreases). Their
main conclusion was that orbital steering is an important
contribution to the catalytic power of enzymes.
A role in orientation of metal-bound water was proposed for His-238 of murine adenosine deaminase (459)
and is possible for the conserved histidine in the serine/
threonine phosphatases. Interestingly, small perturbations in the EPR spectra of the dinuclear metal center in
CN(H151Q) and ␭-PP(H76N) enzymes were observed
compared with the wild-type enzymes (272), indicating
subtle perturbations to the geometry of the dinuclear
metal center. X-ray structures of these mutants would be
valuable to obtain better information on whether substrate orientation may be affected.
CULE.
2. Role of arginines in the active site
As shown in Figure 10, there are two arginine residues in the active site of calcineurin, Arg-122 and Arg-254.
These arginine residues are conserved in other phosphatases. Mutagenesis of Arg-122 in calcineurin or its homologs in other phosphatases (Arg-96 in PP1 and Arg-53
in ␭-protein phosphatase) resulted in 102- to 103-fold decreases in kcat and only slight changes in Km (161, 283,
475, 487). An exception to this was a 20-fold increase in
Km when Arg-53 in ␭-protein phosphatase was mutated to
an alanine and assayed in the presence of Ni2⫹ (487).
However, when this mutant was assayed in the presence
Volume 80
of Mn2⫹, there were no significant changes in Km. Sitedirected mutagenesis studies of Arg-254 in calcineurin or
Arg-221 in PP1 resulted in an 200-fold reduction in kcat,
while values for Km increased 2- to 10-fold (161, 283). In
the X-ray structures of wild-type calcineurin and PP1, the
guanidinium groups of these arginine residues form salt
bridges with the oxygen atoms of bound phosphate or
tungstate and stabilize the anion inhibitor (91, 120, 124,
197). The purpose of these arginine residues may be to
provide electrostatic stabilization for binding the negatively charged phosphate ester. In addition, they may help
neutralize developing negative charge in the transition
state. Because the mutation of Arg-122 led to greater
decreases in kcat than mutation of Arg-254, it may be that
Arg-122 has a more important role in catalysis, perhaps
involving stabilization of the transition state. To note, in
the active site of kidney bean purple acid phosphatase
(Fig. 11), two histidine residues, His-295 and His-296,
appear in place of these arginine residues, and it has been
suggested that these substitutions might explain the lower
pH optimum of purple acid phosphatases (201). Krebs and
colleagues (201) have hypothesized that one of these
histidines may be the active site residue with an apparent
pKa of 6.9 that produces the basic pH portion of the
bell-shaped pH/kinetic profile. In an analogous manner, it
is possible that Arg-53 in ␭-protein phosphatase produces
the basic pH arm observed in the pH dependence studies
(156). In mammalian purple acid phosphatases, there is a
histidine residue (His-195) corresponding to His-296 of
kidney bean purple acid phosphatase, but Glu-194 substitutes for His-295 in the mammalian enzyme and is oriented away from the phosphate in the oxidized enzymephosphate complex.
D. A Model for the Calcineurin
Catalytic Mechanism
Taking into account the available biochemical and
chemical data, we propose a model for the catalytic mechanism of calcineurin and other members of this family
(Fig. 12). The first step of the mechanism involves association of the phosphate monoester, as the dianion, with
the enzyme (Fig. 12A). In this step, neutralization of negative charge by the metal ions may occur. X-ray structures
of calcineurin and PP1 show phosphate and other anionic
product inhibitors coordinated to both metal ions, inferring that the substrate phosphoryl group might also coordinate to one or both metal ions at some point during catalysis. In addition, the conserved arginines, Arg-122 and Arg254, may also play a role in binding substrate and
neutralizing charge by forming hydrogen bonds with the
oxygen atoms of the phosphoryl group. The intermediate in
Figure 12A shows the Zn atom and the two conserved arginines assisting in substrate binding/neutralization of charge.
October 2000
CALCINEURIN: FORM AND FUNCTION
1509
FIG. 12. Proposed mechanism for calcineurin. The mechanism shown in this figure poses His-151 acting as a general
base and a general acid. A: substrate binds to the active site, and His-151 deprotonates a Fe3⫹-bound water molecule to
form hydroxide. B: bond cleavage to the leaving group occurs to a greater extent than bond formation to the hydroxyl
group in the transition state since the transition states for both ␭-protein phosphatase and calcineurin were shown to be
dissociative. The Zn atom may aid in neutralization of the negative charge on the leaving group in the transition state.
The leaving group is protonated by His-151 and then leaves the active site. C: the product-inhibited state is shown with
phosphate bridging the two metal ions. D: a water molecule displaces bound phosphate, and the enzyme is ready for
another round of turnover. Another substrate then enters the active site and displaces the bridging water molecule,
which then becomes a terminal water molecule that must be deprotonated to a hydroxide molecule as is shown in A.
This neutralization of charge would make the substrate
more electrophilic and ready for attack by a nucleophile.
The interaction between the metal-bound water molecule and the conserved histidine His-151 in calcineurin is
also depicted in Figure 12. In the first step of this mechanism (Fig. 12A), His-151 is functioning as a general base
to remove the proton from the metal-bound water molecule. Alternatively, it may orient the solvent nucleophile
for optimal nucleophilic attack of the phosphate ester.
Kinetic isotope studies of calcineurin and ␭-protein
phosphatase show that the transition state of the reaction
is dissociative. A dissociative transition state is represented in Figure 12B, where bond cleavage to the leaving
group has occurred before bond formation to the nucleophile. The phosphoryl group in a dissociative mechanism
resembles the metaphosphate anion (147).
A metal-bound water hydroxide coordinated to Fe3⫹
is shown as the attacking nucleophile in Figure 12B, with
the Fe3⫹ functioning as a Lewis acid to lower the pKa of
the water molecule. Extensive redox studies by Yu and
co-workers (470, 471) have demonstrated a requirement
for Fe3⫹ by calcineurin and a loss of activity upon reduction to Fe2⫹, a result consistent with a decreased Lewis
acidity of Fe2⫹ versus Fe3⫹.
P-O bond scission in the transition state results in a
significant negative charge on the leaving group. Neutralization of the charge by protonation (general acid catalysis) or coordination to a metal ion would lower the energy
of the transition state and increase the rate of the reaction. Kinetic isotope studies indicate that considerable
charge is neutralized in the transition states of calcineurin
and ␭-protein phosphatase. His-151 may play a role in this
charge neutralization (Fig. 12B). It is also possible that
one of the metal ions (e.g., Zn2⫹ in Fig. 12B) neutralizes
1510
FRANK RUSNAK AND PAMELA MERTZ
the charge to the leaving group by coordination. Another
possibility is that a metal-bound solvent molecule acts as
a general acid as has been proposed for the mechanism of
inorganic pyrophosphatase (140, 355). In addition, the
two conserved arginines in the active site may also be
important for charge neutralization and transition state
stabilization. After bond cleavage and proton transfer to
the leaving group occur, the result is the product-inhibited
state that has a molecule of orthophosphate bridging the
two metal ions of the dinuclear center (Fig. 12C). Evidence for this intermediate is obtained in the X-ray structure of the product-inhibited enzyme which shows phosphate coordinated to both metal ions as depicted (124).
An identical coordination is observed in phosphate and
tungstate complexes of PP1 (91, 395) and the phosphate
complex of purple acid phosphatase (128, 201). Phosphate release, perhaps by exchange with a solvent molecule, regenerates the enzyme for another turnover.
VI. REGULATION
The classical mechanism by which calcineurin is regulated in vivo is via changes in intracellular Ca2⫹. Thus, in
a resting cell where [Ca2⫹] is low, calcineurin is unable to
bind calmodulin, and the enzyme exists in an inactive
form. In signaling pathways that lead to a rise in intracellular Ca2⫹, Ca2⫹ binding to calmodulin results in a conformational change, thereby allowing it to bind to calcineurin and activate its phosphatase activity (203). Ca2⫹
binding to calcineurin B also appears to play a role (389).
Calcineurin activity has also been shown to be affected by phospholipids, with either activation or inhibition resulting depending on the phospholipid and substrate investigated (163, 329, 330). Recently, recombinant
Dictyostelium calcineurin has been shown to be activated by arachidonic acid and unsaturated, long-chain
fatty acids (184). These effects may be physiologically
significant given the fact that calcineurin is found associated with membranes during fractionation.
Recently, an additional mechanism for regulating calcineurin involving redox reactions of active site metal
ions has been considered (350, 447). Previous studies
demonstrated that calcineurin is susceptible to redox regulation in vitro (470, 471), a process that may also occur
in vivo. Wang et al. (447) found that superoxide dismutase
protects calcineurin from inactivation and hypothesized
that this might occur by preventing oxidation of active
site metals ions. Recently, a number of groups have begun
testing this hypothesis and have provided evidence that
calcineurin activity can be affected by extracellular oxidants, in particular, H2O2 (45, 111, 345). Thus exposure of
cells to micromolar concentrations of H2O2 results in
inhibition of NF-AT (111, 345) or NF␬B-mediated processes (45) and appears to be mediated by calcineurin.
Volume 80
The mechanism for this regulation may reside in the
redox-active Fe3⫹ of the active site dinuclear metal center, which can toggle between reduced (Fe2⫹) and oxidized (Fe3⫹) states. Indeed, redox titrations by Yu and
co-workers (470, 471) previously demonstrated that the
mixed-valence oxidation state, either Fe3⫹-Zn2⫹ or Fe3⫹Fe2⫹, is required for enzyme activity; reduction to the
Fe2⫹-M2⫹ (M ⫽ Zn, Fe) state led to loss of activity (470,
471). At first glance, the in vivo results, which suggest that
oxidation (i.e., treatment with H2O2) results in inactivation, appear contrary to the in vitro results. One hypothesis that has been forwarded to reconcile this discrepancy
is that calcineurin may exist in different forms containing
either Fe-Zn or Fe-Fe dinuclear metal centers. Because
the Fe3⫹-Fe2⫹ form of calcineurin can lose activity by
oxidation to the Fe3⫹-Fe3⫹ state, it may be that calcineurin exists in oxidation-sensitive (Fe3⫹-Fe2⫹) and oxidation-inert (Fe3⫹-Zn2⫹) states in vivo.
Further work is obviously necessary to determine
whether calcineurin is a specific mediator for changes in
the redox state of the cytosol. If the dinuclear center of
calcineurin has a standard redox potential near the physiological state of the cytosol, a mechanism whereby small
alterations in the redox state could mediate changes in
enzyme activity would be highly tenable.
We apologize for any oversight that has resulted in a key
reference on calcineurin structure/function being excluded from
the bibliography. We gratefully acknowledge Drs. Robert Abraham, Ian Armitage, Howard Brockman, Alvan Hengge, and Ron
Victor for their collaborations. Also acknowledged are Timothy
Born, Alice Haddy, Michael Kennedy, Nicholas Reiter, Tiffany
Reiter, Robert Sikkink, Selene Swanson, Smilja Todorovic,
Daniel White, Janet Yao, and Lian Yu for their significant contributions during their tenure in F. Rusnak’s laboratory. We
thank Dr. Elizabeth Kurian for assistance with the program
Quanta. We are indebted to the editorial and review staff of
Physiological Reviews, in particular Dr. Susan Hamilton, for
support, helpful suggestions, and the opportunity to write this
review.
Financial support from National Institute of General Medical Sciences Grant GM-46865 and the Mayo Clinic Eagles Cancer Fund is noted.
Address for reprint requests and other correspondence: F.
Rusnak, Sect. of Hematology Research, Dept. of Biochemistry
and Molecular Biology, Mayo Clinic, 200 First St. SW, Rochester,
MN 55905 (E-mail: rusnak@mayo.edu).
REFERENCES
1. AITKEN A, COHEN P, SANTIKARN S, WILLIAMS DH, CALDER AG, SMITH A,
AND KLEE CB. Identification of the NH2-terminal blocking group of
calcineurin B as myristic acid. FEBS Lett 150: 314 –318, 1982.
2. AITKEN A, KLEE CB, AND COHEN P. The structure of the B subunit of
calcineurin. Eur J Biochem 139: 663– 671, 1984.
3. ALESSI DR, STREET AJ, COHEN P, AND COHEN PTW. Inhibitor-2 functions like a chaperone to fold three expressed isoforms of mammalian protein phosphatase-1 into a conformation with the specificity and regulatory properties of the native enzyme. Eur
J Biochem 213: 1055–1066, 1993.
October 2000
CALCINEURIN: FORM AND FUNCTION
4. ALEXANDER DR, HEXHAM JM, AND CRUMPTON MJ. The association of
type 1, type 2A, and type 2B phosphatases with the human T
lymphocyte plasma membrane. Biochem J 256: 885– 892, 1988.
5. ALSPAUGH JA, PERFECT JR, AND HEITMAN J. Signal transduction pathways regulating differentiation and pathogenicity of Cryptococcus
neoformans. Fungal Genet Biol 25: 1–14, 1998.
6. AMPLE NM. Emerging disease issues and fungal pathogens associated with HIV infection. Emerg Infect Dis 2: 109 –116, 1996.
7. ANKARCRONA M, DYPBUKT JM, ORRENIUS S, AND NICOTERA P. Calcineurin and mitochondrial function in glutamate-induced neuronal
cell death. FEBS Lett 394: 321–324, 1996.
8. ANTANAITIS BC, AISEN P, AND LILIENTHAL HR. Physical characterization of two-iron uteroferrin: evidence for a spin-coupled binuclear
iron cluster. J Biol Chem 258: 3166 –3172, 1983.
9. ANTANAITIS BC, AISEN P, LILIENTHAL HR, ROBERTS RM, AND BAZER FW.
The Novel “g ⫽ 1.74” EPR spectrum of pink and purple uteroferrin.
J Biol Chem 255: 11204 –11209, 1980.
10. ANTHONY FA, WINKLER MA, EDWARDS HH, AND CHEUNG WY. Quantitative subcellular localization of calmodulin-dependent phosphatase in chick forebrain. J Neurosci 8: 1245–1253, 1988.
11. ANTONI FA, BARNARD RJO, SHIPSTON MJ, SMITH SM, SIMPSON J, AND
PATERSON JM. Calcineurin feedback inhibition of agonist-evoked
cAMP formation. J Biol Chem 270: 28055–28061, 1995.
12. ANTONI FA, SHIPSTON MJ, AND SMITH SM. Inhibitory role for calcineurin in stimulus-secretion coupling revealed by FK506 and
cyclosporin A in pituitary corticotrope tumor cells. Biochem Biophys Res Commun 194: 226 –233, 1993.
13. ANTONI FA, SMITH SM, SIMPSON J, ROSIE R, FINK G, AND PATERSON JM.
Calcium control of adenylyl cyclase: the calcineurin connection.
Adv Second Messenger Phosphoprot Res 32: 153–172, 1998.
14. APERIA A, IBARRA R, SVENSSON L-B, KLEE C, AND GREENGARD P. Calcineurin mediates ␣-adrenergic stimulation of Na⫹,K⫹-ATPase activity in renal tubule cells. Proc Natl Acad Sci USA 89: 7394 –7397,
1992.
15. ARAMBURU J, YAFFE MB, LÓPEZ-RODRÍGUEZ C, CANTLEY LC, HOGAN PG,
AND RAO A. Affinity-driven peptide selection of an NFAT inhibitor
more selective than cyclosporin A. Science 285: 2129 –2133, 1999.
16. ARNDT C, CRUZ MC, CARDENAS ME, AND HEITMAN J. Secretion of
FK506/FK520 and rapamycin by Streptomyces inhibits the growth
of competing Saccharomyces cerevisiae and Cryptococcus neoformans. Microbiology 145: 1989 –2000, 1999.
17. ASAI A, QUI J-H, NARITA Y, CHI S, SAITO N, SHINOURA N, HAMADA H,
KUCHINO Y, AND KIRINO T. High level calcineurin activity predisposes
neuronal cells to apoptosis. J Biol Chem 274: 34450 –34458, 1999.
18. AVERILL BA, DAVIS JC, BURMAN S, ZIRINO T, SANDERS-LOEHR J, LOEHR
TM, SAGE JT, AND DEBRUNNER PG. Spectroscopic and magnetic
studies of the purple acid phosphatase from bovine spleen. J Am
Chem Soc 109: 3760 –3767, 1987.
19. AWUMEY EM, MOONGA BS, SODAM BR, KOVAL AP, ADEBANJO OA,
KUMEGAWA M, ZAIDI M, AND EPSTEIN S. Molecular and functional
evidence for calcineurin-A ␣ and ␤ isoforms in the osteoclast: novel
insights into cyclosporin A action on bone resorption. Biochem
Biophys Res Communun 254: 248 –252, 1999.
20. BACON DJ AND ANDERSON WF. A fast algorithm for rendering spacefilling molecule pictures. J Mol Graph 6: 219 –220, 1988.
21. BANERJEE C, SARKAR D, AND BHADURI A. Ca2⫹ and calmodulin-dependent protein phosphatase from Leishmania donovani. Parasitology 118: 567–573, 1999.
22. BARFORD D. Protein phosphatases. Curr Opin Struct Biol 5: 728 –
734, 1995.
23. BARFORD D. Molecular mechanisms of the protein serine/threonine
phosphatases. Trends Biochem Sci 21: 407– 412, 1996.
24. BARFORD D, DAS AK, AND EGLOFF M-P. Structure and mechanism of
protein phosphatases: insights into catalysis and regulation. Annu
Rev Biophys Biomol Struct 27: 133–164, 1998.
25. BARROSO MR, BERND KK, DEWITT ND, CHANG A, MILLS K, AND SZTUL
ES. A novel Ca2⫹-binding protein, p22, is required for constitutive
membrane traffic. J Biol Chem 271: 10183–10187, 1996.
26. BARTON GJ, COHEN PTW, AND BARFORD D. Conservation analysis and
structure prediction of the protein serine/threonine phosphatases.
Eur J Biochem 220: 225–237, 1994.
27. BEAUMONT A, O’DONOHUE MJ, PAREDES N, ROUSSELET N, ASSICOT M,
BOHUON C, FOURNIE-ZALUSKI M-C, AND ROQUES BP. The role of histi-
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
39.
40.
41.
42.
43.
44.
45.
46.
47.
48.
49.
1511
dine 231 in thermolysin-like enzymes. J Biol Chem 270: 16803–
16808, 1995.
BERNDT N AND COHEN PTW. Renaturation of protein phosphatase 1
expressed at high levels in insect cells using a baculovirus vector.
Eur J Biochem 190: 291–297, 1990.
BIALOJAN C AND TAKAI A. Inhibitory effect of a marine-sponge toxin,
okadaic acid, on protein phosphatases. Biochem J 256: 283–290,
1988.
BIERER BE, HOLLÄNDER G, FRUMAN D, AND BURAKOFF SJ. Cyclosporin
A and FK506: molecular mechanisms of immunosuppression and
probes for transplantation biology. Curr Opin Immunol 5: 763–
773, 1993.
BLUMENTHAL DK, TAKIO K, HANSEN RS, AND KREBS EG. Dephosphorylation of cAMP-dependent protein kinase regulatory subunit (type
II) by calmodulin-dependent protein phosphatase. J Biol Chem 261:
8140 – 8145, 1986.
BONNEFOY-BERARD N, GENESTIER L, FLACHER M, AND REVILLARD JP.
The phosphoprotein phosphatase calcineurin controls calcium-dependent apoptosis in B cell lines. Eur J Immunol 24: 325–329, 1994.
BORK P, BROWN NP, HEGYI H, AND SCHULTZ J. The protein phosphatase 2C (PP2C) superfamily: detection of bacterial homologues.
Protein Sci 5: 1421–1425, 1996.
BOSSER R, ALIGUÉ R, GUERINI D, AGELL N, CARAFOLI E, AND BACHS O.
Calmodulin can modulate protein phosphorylation in rat liver cells
nuclei. J Biol Chem 268: 15477–15483, 1993.
BOXER AL, MORENO H, RUDY B, AND ZIFF EB. FGF-2 potentiates
Ca2⫹-dependent inactivation of NMDA receptor currents in hippocampal neurons. J Neurophysiol 82: 3367–3377, 1999.
BREUDER T, HEMENWAY CS, MOVVA NR, CARDENAS ME, AND HEITMAN J.
Calcineurin is essential in cyclosporin A- and FK506-sensitive yeast
strains. Proc Natl Acad Sci USA 91: 5372–5376, 1994.
BRILL GM, PREMACHANDRAN U, KARWOWSKI JP, HENRY R, CWIK DK,
TRAPHAGEN LM, HUMPHREY PE, JACKSON M, CLEMENT JJ, BURRES NS,
KADAM S, CHEN RH, AND MCALPINE JB. Dibefurin, a novel fungal
metabolite inhibiting calcineurin phosphatase activity. J Antibiot
49: 124 –128, 1996.
BROWN L, CHEN MX, AND COHEN PTW. Identification of a cDNA
encoding a Drosophila calcium/calmodulin regulated protein phosphatase, which has its most abundant expression in the early
embryo. FEBS Lett 339: 124 –128, 1994.
BRUICE TC, BROWN A, AND HARRIS DO. On the concept of orbital
steering in catalytic reactions. Proc Natl Acad Sci USA 68: 658 – 661,
1971.
BURROUGHS SE, HORROCKS WD JR, REN H, AND KLEE CB. Characterization of the lanthanide ion-binding properties of calcineurin-B
using laser-induced luminescence spectroscopy. Biochemistry 33:
10428 –10436, 1994.
BUTCHER SP, HENSHALL DC, TERAMURA Y, IWASAKI K, AND SHARKEY J.
Neuroprotective actions of FK506 in experimental stroke: in vivo
evidence against an antiexcitotoxic mechanism. J Neurosci 17:
6939 – 6946, 1997.
BUTTINI M, LIMONTA S, LUYTEN M, AND BODDEKE H. Distribution of
calcineurin A isoenzyme mRNAs in rat thymus and kidney. Histochem J 27: 291–299, 1995.
CAI L, CHU Y, WILSON SE, AND SCHLENDER KK. A metal-dependent
form of protein phosphatase 2A. Biochem Biophys Res Commun 208: 274 –279, 1995.
CALALB MB, KINCAID RL, AND SODERLING TR. Phosphorylation of
calcineurin: effect on calmodulin binding. Biochem Biophys Res
Commun 172: 551–556, 1990.
CARBALLO M, MÁRQUEZ G, CONDE M, MARTÍN-NIETO J, MONTESEIRÍN J,
CONDE J, PINTADO E, AND SOBRINO F. Characterization of calcineurin
in human neutrophils. J Biol Chem 274: 93–100, 1999.
CARDENAS ME AND HEITMAN J. Role of calcium in T-lymphocyte
activation. Adv Second Messenger Phosphoprotein Res 30: 281–298,
1995.
CARDENAS ME, LORENZ M, HEMENWAY C, AND HEITMAN J. Yeast as
model T cells. Perspect Drug Disc Design 2: 103–126, 1994.
CHANG C-D, MUKAI H, KUNO T, AND TANAKA C. cDNA cloning of an
alternatively spliced isoform of the regulatory subunit of Ca2⫹/
calmodulin-dependent protein phosphatase (calcineurin B␣2). Biochim Biophys Acta 1217: 174 –180, 1994.
CHANG HY, TAKEI K, SYDOR AM, BORN T, RUSNAK F, AND JAY DG.
1512
50.
51.
52.
53.
54.
55.
56.
57.
58.
59.
60.
61.
62.
63.
64.
65.
66.
67.
68.
69.
70.
71.
72.
FRANK RUSNAK AND PAMELA MERTZ
Asymmetric retraction of growth cone filopodia following focal
inactivation of calcineurin. Nature 376: 686 – 690, 1995.
CHANTLER PD. Calcium-dependent association of a protein complex
with the lymphocyte plasma membrane: probable identity with
calmodulin-calcineurin. J Cell Biol 101: 207–216, 1985.
CHEN T-C, LAW B, KONDRATYUK T, AND ROSSIE S. Identification of
soluble protein phosphatases that dephosphorylate voltage-sensitive sodium channels in rat brain. J Biol Chem 270: 7750 –7756,
1995.
CHO H, KRISHNARAJ R, KITAS E, BANNWARTH W, WALSH CT, AND ANDERSON KS. Isolation and structural elucidation of a novel phosphocysteine intermediate in the LAR protein tyrosine phosphatase enzymatic pathway. J Am Chem Soc 114: 7296 –7298, 1992.
CHU Y, LEE EYC, AND SCHLENDER KK. Activation of protein phosphatase 1: formation of a metalloenzyme. J Biol Chem 271: 2574 –2577,
1996.
CHU Y, WILSON SE, AND SCHLENDER KK. A latent form of protein
phosphatase 1␣ associated with bovine heart myofibrils. Biochim
Biophys Acta 1208: 45–54, 1994.
CLIPSTONE NA AND CRABTREE GR. Calcineurin is a key signaling
enzyme in T lymphocyte activation and the target of the immunosuppressive drugs cyclosporin A and FK506. Ann NY Acad Sci 696:
20 –30, 1993.
CLIPSTONE NA, FIORENTINO DF, AND CRABTREE GR. Molecular analysis
of the interaction of calcineurin with drug-immunophilin complexes. J Biol Chem 269: 26431–26437, 1994.
COGHLAN VM, BERGESON SE, LANGEBERG L, NILAVER G, AND SCOTT JD.
A-kinase anchoring proteins: a key to selective activation of cAMPresponsive events? Mol Cell Biochem 127/128: 309 –319, 1993.
COGHLAN VM, PERRINO BA, HOWARD M, LANGEBERG LK, HICKS JB,
GALLATIN WM, AND SCOTT JD. Association of protein kinase A and
protein phosphatase 2B with a common anchoring protein. Science
267: 108 –111, 1995.
COHEN P. The role of protein phosphorylation in neural and hormonal control of cellular activity. Nature 296: 613– 620, 1982.
COHEN P. The structure and regulation of protein phosphatases.
Annu Rev Biochem 58: 453–508, 1989.
COHEN P AND COHEN PTW. Protein phosphatases come of age. J Biol
Chem 264: 21435–21438, 1989.
COHEN PTW. Nomenclature and chromosomal localization of human protein serine/threonine phosphatase Genes. Adv Prot Phosphatases 8: 371–376, 1994.
COLEMAN JE. Structure and mechanism of alkaline phosphatase.
Annu Rev Biophys Biomol Struct 21: 441– 483, 1992.
COLEMAN JE AND GETTINS P. Alkaline phosphatase p, solution structure, and mechanism. Adv Enzymol 55: 381– 452, 1983.
CONBOY IM, MANOLI D, MHAISKAR V, AND JONES PP. Calcineurin and
vacuolar-type H⫹-ATPase modulate macrophage effector functions. Proc Natl Acad Sci USA 96: 6324 – 6329, 1999.
COOPER NGF, MCLAUGHLIN BJ, TALLANT EA, AND CHEUNG WY. Calmodulin-dependent protein phosphatase: immunocytochemical localization in chick retina. J Cell Biol 101: 1212–1218, 1985.
CRUZ MC, DEL POETA M, WANG P, WENGER R, ZENKE G, QUESNIAUX
VFJ, MOVVA NR, PERFECT JR, CARDENAS ME, AND HEITMAN J. Immunosuppressive and nonimmunosuppressive cyclosporine analogs
are toxic to the opportunistic fungal pathogen Cryptoccus neoformans via cyclophilin-dependent inhibition of calcineurin. Antimicrobial Agents Chemother 44: 143–149, 2000.
CUNNINGHAM KW AND FINK GR. Ca2⫹ transport in Saccharomyces
cerevisiae. J Exp Biol 196: 157–166, 1994.
CUNNINGHAM KW AND FINK GR. Calcineurin inhibits VCX1-dependent
H⫹/Ca2⫹ exchange and induces Ca2⫹ ATPases in Saccharomyces
cerevisiae. Mol Cell Biol 16: 2226 –2237, 1996.
CUNNINGHAM KW AND KINK GR. Calcineurin-dependent growth control in Saccharomyces cerevisiae mutants lacking PMC1, a homolog of plasma membrane Ca2⫹ ATPases. J Cell Biol 124: 351–
363, 1994.
CYERT MS. The function of Ca2⫹/calmodulin-regulated phosphatase
in yeast. Adv Protein Phosphatases 7: 429 – 443, 1993.
CYERT MS, KUNISAWA R, KAIM D, AND THORNER J. Yeast has homologs
(CNA1 and CNA2 gene products) of mammalian calcineurin, a
calmodulin-regulated phosphoprotein phosphatase. Proc Natl Acad
Sci USA 88: 7376 –7380, 1991.
Volume 80
73. CYERT MS AND THORNER J. Regulatory subunit (CNB1 gene product)
of yeast Ca2⫹/calmodulin-dependent phosphoprotein phosphatases
is required for adaptation to pheromone. Mol Cell Biol 12: 3460 –
3469, 1992.
74. DA CRUZ E SILVA E AND COHEN PTW. Isolation of a cDNA likely to
encode a novel Ca2⫹-dependent/calmodulin-stimulated protein
phosphatase. Biochim Biophys Acta 1009: 293–296, 1989.
75. DAMMANN H, HELLSTERN S, HUSAIN Q, AND MUTZEL R. Primary structure, expression and developmental regulation of a Dictyostelium
calcineurin A homolog. Eur J Biochem 238: 391–399, 1996.
76. DANIELSSON A, LARSSON C, LARSSON K, GUSTAFSSON L, AND ADLER L. A
genetic analysis of the role of calcineurin and calmodulin in Ca2⫹dependent improvement of NaCl tolerance of Saccharomyces cerevisiae. Curr Genet 30: 476 – 484, 1996.
77. DAVIS JC AND AVERILL BA. Evidence for a spin-coupled binuclear
iron unit at the active site of the purple acid phosphatase from beef
spleen. Proc Natl Acad Sci USA 79: 4623– 4627, 1982.
78. DAVIS TN. Calcium in Saccharomyces cerevisiae. Adv Second Messenger Phosphoprotein Res 30: 339 –358, 1995.
79. DAWSON TM, STEINER JP, DAWSON VL, DINERMAN JL, UHL GR, AND
SNYDER SH. Immunosuppressant FK506 enhances phosphorylation
of nitric oxide synthase and protects against glutamate neurotoxicity. Proc Natl Acad Sci USA 90: 9808 –9812, 1993.
80. DAY EP, DAVID SS, PETERSON J, DUNHAM WR, BONVOISIN JJ, SANDS RH,
AND QUE L JR. Magnetization and electron paramagnetic resonance
studies of reduced uteroferrin and its “EPR-silent” phosphate complex. J Biol Chem 263: 15561–15567, 1988.
81. DE CASTRO E, NEF S, FIUMELLI H, LENZ SE, KAWAMURA S, AND NEF P.
Regulation of rhodopsin phosphorylation by a family of neuronal
calcium sensors. Biochem Biophys Res Commun 216: 133–140,
1995.
82. DE LA POMPA JL, TIMMERMAN LA, TAKIMOTO H, YOSHIDA H, ELIA AJ,
SAMPER E, POTTER J, WAKEHAM A, MARENGERE L, LANGILLE BL,
CRABTREE GR, AND MAK TW. Role of the NF-ATc transcription factor
in morphogenesis of cardiac valves and septum. Nature 392: 182–
186, 1998.
83. DESDOUITS F, BUXBAUM JD, DESDOUITS-MAGNEN J, NAIRN AC, AND
GREENGARD P. Amyloid ␤ peptide formation in cell-free preparations. J Biol Chem 271: 24670 –24674, 1996.
84. DESDOUITS F, SICILIANO JC, NAIRN AC, GREENGARD P, AND GIRAULT J-A.
Dephosphorylation of Ser-137 in DARRP-32 by protein phosphatases 2A and 2C: different roles in vitro and in striatonigral neurons.
Biochem J 330: 211–216, 1998.
85. DIETRICH M, MÜNSTERMANN D, SUERBAUM H, AND WITZEL H. Purple
acid phosphatase from bovine spleen: interactions at the active site
in relation to the reaction mechanism. Eur J Biochem 199: 105–113,
1991.
86. DING B, PRICE RL, BORG TK, WEINBERG EO, HALLORAN PF, AND LORELL
BH. Pressure overload induces severe hypertrophy in mice treated
with cyclosporine, an inhibitor of calcineurin. Circ Res 84: 729 –
734, 1999.
87. DOBSON S, MAY T, BERRIMAN M, VECCHIO CD, FAIRLAMB AH,
CHAKRABARTI D, AND BARIK S. Characterization of protein Ser/Thr
phosphatases of the malaria parasite, Plasmodium falciparum:
inhibition of the parasitic calcineurin by cyclophilin-cyclosporin
complex. Mol Biochem Parasitol 99: 167–181, 1999.
88. DUNN SE, BURNS JL, AND MICHEL RN. Calcineurin is required for
skeletal muscle hypertrophy. J Biol Chem 274: 21908 –21912, 1999.
89. DUTZ JP, FRUMAN DA, BURAKOFF SJ, AND BEIRER BE. A role of
calcineurin in degranulation of murine cytotoxic T lymphocytes.
J Immunol 150: 2591–2598, 1993.
90. EARNEST S, KHOKHLATCHEV A, ALBANESI JP, AND BARYLKO B. Phosphorylation of dynamin by ERK2 inhibits the dynamin-microtubule interaction. FEBS Lett 396: 62– 66, 1996.
91. EGLOFF M-P, COHEN PTW, REINEMER P, AND BARFORD D. Crystal
structure of the catalytic subunit of human protein phosphatase 1
and its complex with tungstate. J Mol Biol 254: 942–959, 1995.
92. EK-RYLANDER B, BILL P, NORGÅRD M, NILSSON S, AND ANDERSSON G.
Cloning, sequence, and developmental expression of a type 5,
tartarate-resistant, acid phosphatase from rat bone. J Biol Chem
266: 24684 –24689, 1991.
93. ENAN E AND MATSUMURA F. Specific inhibition of calcineurin by type
October 2000
94.
95.
96.
97.
98.
99.
100.
101.
102.
103.
104.
105.
106.
107.
108.
109.
110.
111.
112.
113.
114.
115.
CALCINEURIN: FORM AND FUNCTION
II synthetic pyrethroid insecticides. Biochem Pharmacol 43: 1777–
1784, 1992.
ENDO S, CONNOR JH, FORNEY B, ZHANG L, INGEBRITSEN TS, LEE EYC,
AND SHENOLIKAR S. Conversion of protein phosphatase 1 catalytic
subunit to a Mn2⫹-dependent enzyme impairs its regulation by
inhibitor 1. Biochemistry 36: 6986 – 6992, 1997.
ENG W-K, FAUCETTE L, MCLAUGHLIN MM, CAFFERKEY R, KOLTIN Y,
MORRIS RA, YOUNG PR, JOHNSON RK, AND LIVI GP. The yeast FKS1
gene encodes a novel membrane protein, mutations in which confer FK506 and cyclosporin A hypersensitivity and calcineurin-dependent growth. Gene 151: 61–71, 1994.
ENSLEN H AND SODERLING TR. Roles of calmodulin-dependent protein kinases and phosphatase in calcium-dependent transcription
of immediate early genes. J Biol Chem 269: 20872–20877, 1994.
ENZ A AND POMBO-VILLAR E. Class II pyrethroids: noninhibitors of
calcineurin. Biochem Pharmacol 54: 321–323, 1997.
ENZ A, ZENKE G, AND POMBO-VILLAR E. 7-Oxa[221]bicycloheptane-2,3dicarboxylic acid derivatives as phosphatase inhibitors. Bioorg
Med Chem Lett 7: 2513–2518, 1997.
FAKATA KL, SWANSON SA, VORCE RL, AND STEMMER PM. Pyrethroid
insecticides as phosphatase inhibitors. Biochem Pharmacol 55:
2017–2022, 1998.
FARCASANU IC, HIRATA D, TSUCHIYA E, NISHIYAMA F, AND MIYAKAWA T.
Protein phosphatase 2B of Saccharomyces cerevisiae is required
for tolerance to manganese, in blocking the entry of ions into the
cells. Eur J Biochem 232: 712–717, 1995.
FARGHALI H AND MASEK K. Immunopharmacologic agents in the
amelioration of hepatic injuries. Int J Immunopharm 20: 125–139,
1998.
FAST DJ, LYNCH RC, AND LEU RW. Cyclosporin A inhibits nitric oxide
production by L929 cells in response to tumor necrosis factor and
interferon-␥. J Interferon Res 13: 235–240, 1993.
FAUX MC AND SCOTT JD. More on target with protein phosphorylation: conferring specificity by location. Trends Biochem Sci 21:
312–315, 1996.
FENG B AND STEMMER PM. Interactions of calcineurin A, calcineurin
B, and Ca2⫹. Biochemistry 38: 12481–12489, 1999.
FERRANDO A, KRON SJ, RIOS G, FINK GR, AND SERRANO R. Regulation
of cation transport in Saccharomyces cerevisiae by salt tolerance
gene HAL3. Mol Cell Biol 15: 5470 –5481, 1995.
FERREIRA A, KINCAID R, AND KOSIK KS. Calcineurin is associated with
the cytoskeleton of cultured neurons and has a role in the acquisition of polarity. Mol Biol Cell 4: 1225–1238, 1993.
FOOR F, PARENT SA, MORIN N, DAHL AM, RAMADAN N, CHREBET G,
BOSTIAN KA, AND NIELSEN JB. Calcineurin mediates inhibition by
FK506 and cyclosporin of recovery from ␣-factor arrest in yeast.
Nature 360: 682– 684, 1992.
FORCE T, ROSENZWEIG A, CHOUKROUN G, AND HAJJAR R. Calcineurin
inhibitors and cardiac hypertrophy. Lancet 353: 1290 –1292, 1999.
FRIEDHOFF P, KOLMES B, GIMADUTDINOW O, WENDE W, KRAUSE KL, AND
PINGOUD A. Analysis of the mechanism of the Serratia nuclease
using site-directed mutagenesis. Nucleic Acid Res 24: 2632–2639,
1996.
FRUMAN DA, MATHER PE, BURAKOFF SJ, AND BIERER BE. Correlation
of calcineurin phosphatase activity and programmed cell death in
murine T cell hybridomas. Eur J Immunol 22: 2513–2517, 1992.
FURUKE K, SHIRAISHI M, MOSTOWSKI HS, AND BLOOM ET. Fas ligand
induction in human NK cells is regulated by redox through a
calcineurin-nuclear factors of activated T cell-dependent pathway.
J Immunol 162: 1988 –1993, 1999.
GAGLIARDINO JJ, KRINKS MH, AND GAGLIARDINO EE. Identification of
the calmodulin-regulated protein phosphatase, calcineurin, in rat
pancreatic islets. Biochim Biophys Acta 1091: 370 –373, 1991.
GANSTER RW, MCCARTNEY RR, AND SCHMIDT MC. Identification of a
calcineurin-independent pathway required for sodium ion stress
response in Saccharomyces cerevisiae. Genetics 150: 31– 42, 1998.
GARRETT-ENGELE P, MOILANEN B, AND CYERT MS. Calcineurin, the
Ca2⫹/calmodulin-dependent protein phosphatase, is essential in
yeast mutants with cell integrity defects and in mutants that lack a
functional vacuolar H⫹-ATPase. Mol Cell Biol 15: 4103– 4114, 1995.
GARVER TD, KINCAID RL, CONN RA, AND BILLINGSLEY ML. Reduction of
calcineurin activity in brain by antisense oligonucleotides leads to
1513
persistent phosphorylation of ␶ protein at Thr181 and Thr231. Mol
Pharmacol 55: 632– 641, 1999.
116. GENESTIER L, DEARDEN-BADET MT, BONNEFOY-BERARD N, LIZARD G,
AND REVILLARD JP. Cyclosporin A and FK506 inhibit activationinduced cell death in the murine WEHI-231 B cell line. Cell Immunol 155: 283–291, 1994.
117. GIRI PR, HIGUCHI S, AND KINCAID RL. Chromosomal mapping of the
human genes for the calmodulin-dependent protein phosphatase
(calcineurin) catalytic subunit. Biochem Biophys Res Commun
181: 252–258, 1991.
118. GIRI PR, MARIETTA CA, HIGUCHI S, AND KINCAID RL. Molecular and
phylogenetic analysis of calmodulin-dependent protein phosphatase (calcineurin) catalytic subunit genes. DNA Cell Biol 11: 415–
424, 1992.
119. GOLD BG. FK506 and the role of immunophilins in nerve regeneration. Mol Neurobiol 15: 285–306, 1997.
120. GOLDBERG J, HUANG H-B, KWON Y-G, GREENGARD P, NAIRN AC, AND
KURIYAN J. Three dimensional structure of the catalytic subunit of
protein serine/threonine phosphatase-1. Nature 376: 745–753, 1995.
121. GOTO S AND USHIO Y. Immunostaining for calcineurin, a Ca2⫹/
calmodulin-regulated protein phosphatase, in the diagnostic tumor
pathology. Brain Tumor Pathol 12: 23–30, 1995.
122. GOTO S, YAMAMOTO H, FUKUNAGA K, IWASA T, MATSUKADO Y, AND
MIYAMOTO E. Dephosphorylation of microtubule-associated protein
2, ␶ factor, and tubulin by calcineurin. J Neurochem 45: 276 –283,
1985.
123. GRAEF IA, MERMELSTEIN PG, STANKUNAS K, NEILSON JR, DEISSEROTH K,
TSEIN RW, AND CRABTREE GR. L-type calcium channels and GSK-3
regulate the activity of NF-ATc4 in hippocampal neurons. Nature
401: 703–708, 1999.
124. GRIFFITH JP, KIM JL, KIM EE, SINTCHAK MD, THOMSON JA, FITZGIBBON
MJ, FLEMING MA, CARON PR, HSIAO K, AND NAVIA MA. X-ray structure
of calcineurin inhibited by the immunophilin-immunosuppressant
FKBP12-FK506 complex. Cell 82: 507–522, 1995.
125. GROBLEWSKI GE, WAGNER ACC, AND WILLIAMS JA. Cyclosporin A
inhibits Ca2⫹/calmodulin-dependent protein phosphatase and secretion in pancreatic acinar cells. J Biol Chem 269: 15111–15117,
1994.
126. GUALBERTO A, MARQUEZ G, CARBALLO M, YOUNGBLOOD GL, HUNT SW
III, BALDWIN AS, AND SOBRINO F. p53 transactivation of the HIV-1
long terminal repeat is blocked by PD 144795, a calcineurin-inhibitor with anti-HIV properties. J Biol Chem 273: 7088 –7093, 1998.
127. GUAN K AND DIXON JE. Evidence for protein-tyrosine-phosphatase
catalysis proceeding via a cysteine-phosphate intermediate. J Biol
Chem 266: 17026 –17030, 1991.
128. GUDDAT LW, MCALPINE AS, HUME D, HAMILTON S, DE JERSEY J, AND
MARTIN JL. Crystal structure of mammalian purple acid phosphatase. Structure 7: 757–767, 1999.
129. GUERINI D AND KLEE CB. Cloning of human calcineurin A: evidence
for two isozymes and identification of a proline structural domain.
Proc Natl Acad Sci USA 86: 9183–9187, 1989.
130. GUERINI D AND KLEE CB. Structural diversity of calcineurin, a Ca2⫹
and calmodulin-stimulated protein phosphatase. Adv Protein Phosphatases 6: 391– 410, 1991.
131. GUERINI D, KRINKS MH, SIKELA JM, HAHN WE, AND KLEE CB. Isolation
and sequence of a cDNA clone for human calcineurin B, the Ca2⫹binding subunit of the Ca2⫹/calmodulin-stimulated protein phosphatase. DNA 8: 675– 682, 1989.
132. GUERINI D, MONTELL C, AND KLEE CB. Molecular cloning and characterization of the genes encoding the two subunits of Drosophila
melanogaster calcineurin. J Biol Chem 267: 22542–22549, 1992.
132a.HÅKANSSON K, CARLSSON M, SVENSSON LA, AND LILJAS A. Structure of
native and Apo carbonic anhydrase II and structure of some of its
anion-ligand complexes. J Mol Biol 227: 1192–1204, 1992.
133. HALLORAN PF. Molecular mechanisms of new immunosuppressants.
Clin Transplant 10: 118 –123, 1996.
134. HALLORAN PF, KUNG L, AND NOUJAIM J. Calcineurin and the biological
effect of cyclosporine and tacrolimus. Transplant Proc 30: 2167–
2170, 1998.
135. HANLEY RM, DEDMAN JR, AND SHENOLIKAR S. Identification of highaffinity calmodulin-binding proteins in rat liver. Am J Physiol Cell
Physiol 252: C272–C284, 1987.
136. HASHIMOTO Y, KING MM, AND SODERLING TR. Regulatory interactions
1514
137.
138.
139.
140.
141.
142.
143.
144.
145.
146.
147.
148.
149.
150.
151.
152.
153.
154.
155.
156.
FRANK RUSNAK AND PAMELA MERTZ
of calmodulin-binding proteins: phosphorylation of calcineurin by
autophosphorylated Ca2⫹/calmodulin-dependent protein kinase II.
Proc. Natl Acad Sci USA 85: 7001–7005, 1988.
HASHIMOTO Y, PERRINO BA, AND SODERLING TR. Identification of an
autoinhibitory domain in calcineurin. J Biol Chem 265: 1924 –1927,
1990.
HASHIMOTO Y AND SODERLING TR. Regulation of calcineurin by phosphorylation. Identification of the regulatory site phosphorylated by
Ca2⫹/calmodulin-dependent protein kinase II and protein kinase C.
J Biol Chem 264: 16524 –16529, 1989.
HASSETT A, BLÄTTLER W, AND KNOWLES JR. Pyruvate kinase: is the
mechanism of phospho transfer associative or dissociative? Biochemistry 21: 6335– 6340, 1982.
HEIKINHEIMO P, POHJANJOKI P, HELMINEN A, TASANEN M, COOPERMAN
BS, GOLDMAN A, BAYKOV A, AND LAHTI R. A site-directed mutagenesis
study of Saccharomyces cerevisiae pyrophosphatase. Functional
conservation of the active site of soluble inorganic pyrophosphatases. Eur J Biochem 239: 138 –143, 1996.
HEITMAN J, CARDENAS ME, BREUDER T, HEMENWAY C, MUIR RS, LIM E,
GOETZ L, ZHU D, LORENZ M, AND DOLINSKI K. Antifungal effects of
cyclosporine and FK 506 are mediated via immunophilin-dependent calcineurin inhibition. Transplant Proc 26: 2833–2834, 1994.
HEITMAN J, KOLLER A, CARDENAS ME, AND HALL MN. Identification of
immunosuppressive drug targets in yeast. Methods: A Companion
to Methods Enzymol 5: 176 –187, 1993.
HEMENWAY C AND HEITMAN J. Proline isomerases in microorganisms
and small eukaryotes. Ann NY Acad Sci 696: 38 – 46, 1993.
HEMENWAY CS, DOLINSKI K, CARDENAS ME, HILLER MA, JONES EW, AND
HEITMAN J. vph6 mutants of Saccharomyces cerevisiae require
calcineurin for growth and are defective in vacuolar H⫹-ATPase
assembly. Genetics 141: 833– 844, 1995.
HEMENWAY CS AND HEITMAN J. Calcineurin: structure, function, and
inhibition. Cell Biochem. Biophys 30: 115–151, 1999.
HENDEY B, KLEE CB, AND MAXFIELD FR. Inhibition of neutrophil
chemokinesis on vitronectin by inhibitors of calcineurin. Science
258: 296 –299, 1992.
HENGGE AC. Transfer of the PO2⫺
group. In: Comprehensive Bio3
logical Catalysis: A Mechanistic Reference. Reactions of Electrophilic Carbon, Phosphorus and Sulfur, edited by Sinnott M. San
Diego, CA: Academic, 1998, vol. 1, p. 517–542.
HENGGE AC, DENU JM, AND DIXON JE. Transition-state structures for
the native dual-specific phosphatase VHR and D92N and S131A
mutants. contributions to the driving force for catalysis. Biochemistry 35: 7084 –7092, 1996.
HENGGE AC, EDENS WA, AND ELSING H. Transition-state structures
for phosphoryl-transfer reactions of p-nitrophenyl phosphate. J Am
Chem Soc 116: 5045–5049, 1994.
HENGGE AC AND MARTIN BL. Isotope effect studies on the calcineurin phosphoryl-transfer reaction: transition state structure
and effect of calmodulin and Mn2⫹. Biochemistry 36: 10185–10191,
1997.
HENGGE AC, SOWA GA, WU L, AND ZHANG Z-Y. Nature of the transition
state of the protein-tyrosine phosphatase-catalyzed reaction. Biochemistry 34: 13982–13987, 1995.
HENGGE AC, ZHAO Y, WU L, AND ZHANG Z-Y. Examination of the
transition state of the low-molecular mass small tyrosine phosphatase. 1. Comparison with other protein phosphatases. Biochemistry 36: 7928 –7936, 1997.
HIGUCHI S, TAMURA J, GIRI PR, POLLI JW, AND KINCAID RL. Calmodulindependent protein phosphatase from Neurospora crassa. J Biol
Chem 266: 18104 –18112, 1991.
HIRATA D, HARADA S-I, NAMBA H, AND MIYAKAWA T. Adaptation to
high-salt stress in Saccharomyces cerevisiae is regulated by Ca2⫹/
calmodulin-dependent phosphoprotein phosphatase (calcineurin)
and cAMP-dependent protein kinase. Mol Gen Genet 249: 257–264,
1995.
HO S, CLIPSTONE N, TIMMERMANN L, NORTHROP J, GRAEF I, FIORENTINO
D, NOUSE J, AND CRABTREE G. The mechanism of action of cyclosporin A and FK506. Clin Immunol Immunopathol 80: S40 –S45,
1996.
HOFF RH, MERTZ P, RUSNAK F, AND HENGGE AC. The transition state
of the phosphoryl-transfer reaction catalyzed by the lambda Ser/
Thr protein phosphatase. J Am Chem Soc 121: 6382– 6390, 1999.
Volume 80
157. HOLTZ-HEPPELMANN CJ, ALGECIRAS A, BADLEY AD, AND PAYA CV. Transcriptional regulation of the human FasL promoter-enhancer region. J Biol Chem 273: 4416 – 4423, 1998.
158. HONG C-S AND GANETZKY B. Molecular characterization of neurally
expressing genes in the para sodium channel gene cluster of
Drosophila. Genetics 142: 879 – 892, 1996.
159. HORN F AND GROSS J. A role for calcineurin in Dictyostelium discoideum development. Differentiation 60: 269 –275, 1996.
160. HUANG EP. Synaptic plasticity: going through phases with LTP.
Curr Biol 8: R350 –R352, 1998.
161. HUANG H-B, HORIUCHI A, GOLDBERG J, GREENGARD P, AND NAIRN AC.
Site-directed mutagenesis of amino acid residues of protein phosphatase 1 involved in catalysis and inhibitor binding. Proc Natl
Acad Sci USA 94: 3530 –3535, 1997.
162. HUANG R-Q AND DILLON GH. Maintenance of recombinant type A
␥-aminobutyric acid receptor function: role of protein tyrosine
phosphorylation and calcineurin. J Pharmacol Exp Ther 286: 243–
255, 1998.
163. HUANG S, MERAT D, AND CHEUNG WY. Phosphatidylinositol modulates the response of calmodulin-dependent phosphatase to calmodulin. Arch Biochem Biophys 270: 42– 49, 1989.
164. HUSI H, LUYTEN MA, AND ZURINI MGM. Mapping of the immunophilinimmunosuppressant site of interaction on calcineurin. J Biol Chem
269: 14199 –14204, 1994.
166. IIDA H, YAGAWA Y, AND ANRAKU Y. Essential role for induced Ca2⫹
influx followed by [Ca2⫹]i rise in maintaining viability of yeast cells
late in the mating pheromone response pathway. J Biol Chem 265:
13391–13399, 1990.
167. INGEBRITSEN TS AND COHEN P. Protein phosphatases: properties and
role in cellular regulation. Science 221: 331–338, 1983.
168. INGEBRITSEN TS, STEWART AA, AND COHEN P. The protein phosphatases involved in cellular regulation: measurement of type-1 and
type-2 phosphatases in extracts of mammalian tissues; assessment
of their physiological roles. Eur J Biochem 132: 297–307, 1983.
169. ITO A, HASHIMOTO T, HIRAI M, TAKEDA T, SHUNTOH H, KUNO T, AND
TANAKA C. The complete primary structure of calcineurin A, a
calmodulin binding protein homologous with protein phosphatases
1 and 2A. Biochem Biophys Res Commun 163: 1492–1497, 1989.
170. IWATA M, OHOKA Y, KUWATA T, AND ASADA A. Regulation of T cell
apoptosis via T cell receptors and steroid receptors. Stem Cells 14:
632– 641, 1996.
171. IZUMO S AND AOKI H. Calcineurin—the missing link in cardiac hypertrophy. Nature Med 4: 661– 662, 1998.
172. JABRI E, CARR MB, HAUSINGER RP, AND KARPLUS PA. The crystal
structure of urease from Klebsiella aerogenes. Science 268: 998 –
1004, 1995.
173. JAIN J, LOH C, AND RAO A. Transcriptional regulation of the IL-2 gene.
Curr Opin Immunol 7: 333–342, 1995.
174. JENCKS WP AND PAGE MI. “Orbital steering,” entropy, and rate accelerations. Biochem Biophys Res Commun 57: 887– 892, 1974.
175. JIANG H, XIONG F, KONG S, OGAWA T, KOBAYASI M, AND LIU JO. Distinct
tissue and cellular distribution of two major isoforms of calcineurin. Mol Immunol 34: 663– 669, 1997.
176. KAKALIS LT, KENNEDY M, SIKKINK R, RUSNAK F, AND ARMITAGE IM.
Characterization of the calcium-binding sites of calcineurin B.
FEBS Lett 362: 55–58, 1995.
177. KASHISHIAN A, HOWARD M, LOH C, GALLATIN WM, HOEKSTRA MF, AND
LAI T. AKAP79 inhibits calcineurin through a site distinct from the
immunophilin-binding region. J Biol Chem 273: 27412–27419, 1998.
178. KAWASAKI H AND KRETSINGER RH. Calcium-binding proteins 1: EFhands. Protein Profile 2: 1– 490, 1995.
179. KAY JE, DOE SEA, AND BENZIE CR. The mechanism of action of the
immunosuppressive drug FK-506. Cell Immunol 124: 175–181, 1989.
180. KAYE RE, FRUMAN DA, BIERER BE, ALBERS MW, ZYDOWSKY LD, HO SI,
JIN Y-J, CASTELLS MC, SCHREIBER SL, WALSH CT, BURAKOFF SJ, AUSTEN
KF, AND KATZ HR. Effects of cyclosporin A and FK506 on Fc⑀
receptor type I-initiated increases in cytokine mRNA in mouse
bone marrow-derived progenitor mast cells: resistance to FK506 is
associated with a deficiency in FK506-binding protein FKBP12.
Proc Natl Acad Sci USA 89: 8542– 8546, 1992.
181. KAYYALI US, ZHANG W, YEE AG, SEIDMAN JG, AND POTTER H. Cytoskeletal changes in the brains of mice lacking calcineurin A␣. J Neurochemistry 68: 1668 –1678, 1997.
October 2000
CALCINEURIN: FORM AND FUNCTION
182. KENNEDY MT, BROCKMAN H, AND RUSNAK F. Contributions of myristoylation to calcineurin structure/function. J Biol Chem 271:
26517–26521, 1996.
183. KENNEDY MT, BROCKMAN H, AND RUSNAK F. Determinants of calcineurin binding to model membranes. Biochemistry 36: 13579 –
13585, 1997.
184. KESSEN U, SCHALOSKE R, AICHEM A, AND MUTZEL R. Ca2⫹/calmodulinindependent activation of calcineurin from Dictyostelium by unsaturated long chain fatty acids. J Biol Chem 274: 37821–37826,
1999.
185. KETCHAM CM, ROBERTS RM, SIMMEN RCM, AND NICK HS. Molecular
cloning of the type 5, iron-containing, tartarate-resistant acid phosphatase from human placenta. J Biol Chem 264: 557–563, 1989.
186. KHATTAB A, PICA-MATTOCCIA L, WENGER R, CIOLI D, AND KLINKERT M-Q.
Assay of Schistosoma mansoni calcineurin phosphatase activity
and assessment of its role in parasite survival. Mol Biochem Parasitol 99: 269 –273, 1999.
187. KIM EE AND WYCKOFF HW. Reaction mechanism of alkaline phosphatase based on crystal structures: two-metal ion catalysis. J Mol
Biol 218: 449 – 464, 1991.
188. KINCAID R. Calmodulin-dependent protein phosphatases from microorganisms to man: a study in structural conservatism and biological diversity. Adv Second Messenger Phosphoprotein Res 27:
1–23, 1993.
189. KINCAID RL. The role of calcineurin in immune system responses. J
Allergy Clin Immunol 95: 1170 –1177, 1995.
190. KINCAID RL, GIRI PR, HIGUCHI S, TAMURA J, DIXON SC, MARIETTA CA,
AMORESE DA, AND MARTIN BM. Cloning and characterization of
molecular isoforms of the catalytic subunit of calcineurin using
nonisotopic methods. J Biol Chem 265: 11312–11319, 1990.
191. KINCAID RL, NIGHTINGALE MS, AND MARTIN BM. Characterization of a
cDNA clone encoding the calmodulin-binding domain of mouse
brain calcineurin. Proc Natl Acad Sci USA 85: 8983– 8987, 1988.
192. KINCAID RL AND O’KEEFE SJ. Calcineurin and immunosuppression: a
calmodulin-dependent protein phosphatase acts as the “gatekeeper” to interleukin-2 gene transcription. Adv Protein Phosphatases 7: 543–583, 1993.
193. KINCAID RL, TAKAYAMA H, BILLINGSLEY ML, AND SITKOVSKY MV. Differential expression of calmodulin-binding proteins in B, T lymphocytes and thymocytes. Nature 330: 176 –178, 1987.
194. KING MM AND HUANG CY. Activation of calcineurin by nickel ions.
Biochem Biophys Res Commun 114: 955–961, 1983.
195. KING MM AND HUANG CY. The calmodulin-dependent activation and
deactivation of the phosphoprotein phosphatase, calcineurin, and
the effect of nucleotides, pyrophosphate, and divalent metal ions.
J Biol Chem 259: 8847– 8856, 1984.
196. KING MM, HUANG CY, CHOCK PB, NAIRN AC, HEMMINGS HC JR, CHAN
K-FJ, AND GREENGARD P. Mammalian brain phosphoproteins as substrates for calcineurin. J Biol Chem 259: 8080 – 8083, 1984.
197. KISSINGER CR, PARGE HE, KNIGHTON DR, LEWIS CT, PELLETIER LA,
TEMPCZYK A, KALISH VJ, TUCKER KD, SHOWALTER RE, MOOMAW EW,
GASTINEL LN, HABUKA N, CHEN X, MALDONADO F, BARKER JE, BACQUET
R, AND VILLAFRANCA JE. Crystal structure of human calcineurin and
the human FKBP12-FK506-calcineurin complex. Nature 378: 641–
644, 1995.
198. KISSMEHL R, TREPTAU T, HOFER HW, AND PLATTNER H. Protein phosphatase and kinase activities possibly involved in exocytosis regulation in Paramecium tetraurelia. Biochem J 317: 65–76, 1996.
199. KLABUNDE T AND KREBS B. The dimetal center in purple acid phosphatase. Struct Bonding 89: 177–198, 1997.
200. KLABUNDE T, STAHL B, SUERBAUM H, HAHNER S, KARAS M, HILLENKAMP
F, KREBS B, AND WITZEL H. The amino acid sequence of red kidney
bean Fe(III)-Zn(II) purple acid phosphatase. Eur J Biochem 226:
369 –375, 1994.
201. KLABUNDE T, STRÄTER N, FRÄLICH R, WITZEL H, AND KREBS B. Mechanism of Fe(III)-Zn(II) purple acid phosphatase based on crystal
structure. J Mol Biol 259: 737–748, 1996.
202. KLEE CB. Concerted regulation of protein phosphorylation and
dephosphorylation by calmodulin. Neurochem Res 16: 1059 –1065,
1991.
203. KLEE CB AND COHEN P. The calmodulin-regulated protein phosphatase. In: Molecular Aspects of Cellular Regulation: Calmodulin,
204.
205.
206.
207.
208.
209.
210.
211.
212.
213.
214.
215.
216.
217.
218.
219.
220.
221.
222.
223.
224.
225.
1515
edited by Cohen P and Klee CB. Amsterdam: Elsevier, 1988, vol. 5,
p. 225–248.
KLEE CB, CROUCH TH, AND KRINKS MH. Calcineurin: a calcium- and
calmodulin-binding protein of the nervous system. Proc Natl Acad
Sci USA 76: 6270 – 6273, 1979.
KLEE CB, DRAETTA GF, AND HUBBARD MJ. Calcineurin. Adv Enzymol
Relat Areas Mol Biol 61: 149 –200, 1988.
KLEE CB AND KRINKS MH. Purification of cyclic 3⬘,5⬘-nucleotide
phosphodiesterase inhibitory protein by affinity chromatography
on activator protein coupled to Sepharose. Biochemistry 17: 120 –
126, 1978.
KLEE CB, KRINKS MH, MANALAN AS, DRAETTA GF, AND NEWTON DL.
Control of calcineurin protein phosphatase activity. Adv Protein
Phosphatases 1: 135–146, 1985.
KLEE CB, REN H, AND WANG X. Regulation of the calmodulin-stimulated protein phosphatase calcineurin. J Biol Chem 273: 13367–
13370, 1998.
KLUMPP S, STEINER AL, AND SCHULTZ JE. Immunocytochemical localization of cyclic GMP, cGMP-dependent protein kinase, calmodulin
and calcineurin in Paramecium tetraurelia. Eur J Cell Biol 32:
164 –170, 1983.
KNAPP J, BOKNÍK P, HUKE S, BOMBOSOVÁ I, LINCK B, LÜSS H, MÜLLER
FU, MÜLLER T, NACKE P, SCHMITZ W, VAHLENSIECK U, AND NEUMANN J.
Contractility and inhibition of protein phosphatases by cantharidin.
Gen Pharmacol 31: 729 –733, 1998.
KNÖFEL T AND STRÖTER N. X-ray structure of the Escherichia coli
periplasmic 5⬘-nucleotidase containing a dimetal catalytic site. Nature Struct Biol 6: 448 – 453, 1999.
KOONIN EV. Conserved sequence pattern in a wide variety of phosphoesterases. Protein Sci 3: 356 –358, 1994.
KOTHE GO AND FREE SJ. Calcineurin subunit B is required for
normal vegetative growth in Neurospora crassa. Fungal Genet Biol
23: 248 –258, 1998.
KRAULIS P. MOLSCRIPT: a program to produce both detail and
schematic plots of protein structure. J Appl Crystallogr 24: 946 –
950, 1991.
KREBS B. Purple acid phosphatases and catechol oxidase and their
model compounds. Crystal structure studies of purple acid phosphatase from red kidney beans. In: Bioinorganic Chemistry: An
Inorganic Perspective of Life, edited by Kessissoglou DP. Dordrecht, Germany: Kluwer, 1995, p. 371–384.
KRINKS MH, MANALAN AS, AND KLEE CB. Calcineurin: a brain-specific
isozyme of protein phosphatase-2B (Abstract). Federation Proc 44:
707a, 1985.
KRUPP JJ, VISSEL B, HEINEMANN SF, AND WESTBROOK GL. Calciumdependent inactivation of recombinant N-methyl-D-aspartate receptors is NR2 subunit specific. Mol Pharmacol 50: 1680 –1688, 1996.
KUDLA J, XU Q, HARTER K, GRUISSEM W, AND LUAN S. Genes for
calcineurin B-like proteins in Arabidopsis are differentially regulated by stress signals. Proc Natl Acad Sci USA 96: 4718 – 4723,
1999.
KUNO T, MUKAI H, ITO A, CHANG C-D, KISHIMA K, SAITO N, AND TANAKA
C. Distinct cellular expression of calcineurin A␣ and A␤ in rat
brain. J Neurochemistry 58: 1643–1651, 1992.
KUNO T, TAKEDA T, HIRAI M, ITO A, MUKAI H, AND TANAKA C. Evidence
for a second isoform of the catalytic subunit of calmodulin-dependent protein phosphatase (calcineurin A). Biochem Biophys Res
Commun 165: 1352–1358, 1989.
KUNO T, TANAKA H, MUKAI H, CHANG C-D, HIRAGA K, MIYAKAWA T, AND
TANAKA C. cDNA cloning of a calcineurin B homolog in Saccharomyces cerevisiae. Biochem Biophys Res Commun 180: 1159 –1163,
1991.
KUNZ J AND HALL MN. Cyclosporin A, FK506 and rapamycin: more
than just immunosuppression. Trends Biochem Sci 18: 334 –338,
1993.
KURTZ DM JR. Oxo- and hydroxo-bridged diiron complexes: a chemical perspective on a biological unit. Chem Rev 90: 585– 606, 1990.
LAI MM, BURNETT PE, WOLOSKER H, BLACKSHAW S, AND SNYDER SH.
Cain, a novel physiological protein inhibitor of calcineurin. J Biol
Chem 273: 18325–18331, 1998.
LAI MM, HONG JJ, RUGGIERO AM, BURNETT PE, SLEPNEV VI, DE CAMILLI
P, AND SNYDER SH. The calcineurin-dynamin I complex as a calcium
1516
226.
227.
228.
229.
230.
231.
232.
233.
234.
235.
236.
237.
238.
239.
240.
241.
242.
243.
244.
245.
246.
247.
248.
249.
FRANK RUSNAK AND PAMELA MERTZ
sensor for synaptic vesicle endocytosis. J Biol Chem 274: 25963–
25966, 1999.
LATINIS KM, NORIAN LA, ELIASON SL, AND KORETZKY GA. Two NFAT
transcription factor binding sites participate in the regulation of
CD95 (Fas) ligand expression in activated human T cells. J Biol
Chem 272: 31427–31434, 1997.
LAWEN A, BAKER MA, AND MALIK S. Apoptosis and redox homeostasis: on a possible mechanism of action of Bcl-2. Protoplasma 205:
10 –20, 1998.
LAWSON MA AND MAXFIELD FR. Ca2⫹- and calcineurin-dependent
recycling of an integrin to the front of migrating neutrophils. Nature 377: 75–79, 1995.
LENG J, CAMERON AJM, BUCKEL S, AND KENNELLY PJ. Isolation and
cloning of a protein-serine/threonine phosphatase from an archaeon. J Bacteriol 177: 6510 – 6517, 1995.
LI H-C. Activation of brain calcineurin phosphatase towards nonprotein phosphoesters by Ca2⫹, calmodulin, and Mg2⫹. J Biol Chem
259: 8801– 8807, 1984.
LIEBERMAN DN AND MODY I. Regulation of NMDA channel function
by endogenous Ca2⫹-dependent phosphatase. Nature 369: 235–239,
1994.
LIM HW AND MOLKENTIN JD. Calcineurin and human heart failure.
Nature Med 5: 246 –247, 1999.
LIM HW AND MOLKENTIN JD. Reply to revisiting calcineurin in human
heart failure. Nature Med 6: 3, 2000.
LIN X AND BARBER DL. A calcineurin homologous protein inhibits
GTPase-stimulated Na-H exchange. Proc Natl Acad Sci USA 93:
12631–12636, 1996.
LIN X, SIKKINK RA, RUSNAK F, AND BARBER DL. Inhibition of calcineurin phosphatase activity by a calcineurin B homologous protein. J Biol Chem 274: 36125–36131, 1999.
LIU J. FK506 and cyclosporin, molecular probes for studying intracellular signal transduction. Immunol Today 14: 290 –295, 1993.
LIU J. FK506 and cyclosporin: molecular probes for studying intracellular signal transduction. Trends Pharmacol Sci 14: 182–188,
1993.
LIU J, FARMER JD JR, LANE WS, FRIEDMAN J, WEISSMAN I, AND SCHREIBER SL. Calcineurin is a common target of cyclophilin-cyclosporin
A and FKBP-FK506 complexes. Cell 66: 807– 815, 1991.
LIU J AND ZHU J-K. A calcium sensor homolog required for plant salt
tolerance. Science 280: 1943–1945, 1998.
LIU J-P, SIM ATR, AND ROBINSON PJ. Calcineurin inhibition of dynamin I GTPase activity coupled to nerve terminal depolarization.
Science 265: 970 –973, 1994.
LIU Q-S AND BERG DK. Actin filaments and the opposing actions of
CaM kinase II and calcineurin in regulating ␣7-containing nicotinic
receptors on chick ciliary ganglion neurons. J Neurosci 19: 10280 –
10288, 1999.
LIU Y, ISHII S, TOKAI M, TSUTSUMI H, OHKI O, AKADA R, TANAKA K,
TSUCHIYA E, FUKUI S, AND MIYAKAWA T. The Saccharomyces cerevisiae genes (CMP1 and CMP2) encoding calmodulin-binding proteins homologous to the catalytic subunit of mammalian protein
phosphatase 2B. Mol Gen Genet 227: 52–59, 1991.
LOBO FM, ZANJANI R, HO N, CHATILA TA, AND FULEIHAN RL. Calciumdependent activation of TNF family gene expression by Ca2⫹/
calmodulin kinase type IV/GR and calcineurin. J Immunol 162:
2057–2063, 1999.
LOHSE DL, DENU JM, AND DIXON JE. Insights derived from the
structures of the Ser/Thr phosphatases calcineurin and protein
phosphatase 1. Structure 3: 987–990, 1995.
LUAN S. Protein phosphatases and signaling cascades in higher
plants. Trends Plant Sci 3: 271–275, 1998.
LUAN S, LI W, RUSNAK F, ASSMANN SM, AND SCHREIBER SL. Immunosuppressants implicate protein phosphatase regulation of K⫹ channels in guard cells. Proc Natl Acad Sci USA 90: 2202–2206, 1993.
LUO Z, SHYU K-G, GUALBERTO A, AND WALSH K. Calcineurin inhibitors
and cardiac hypertrophy. Nature Med 4: 1092–1093, 1998.
MACKINTOSH C, BEATTIE KA, KLUMPP S, COHEN P, AND CODD GA.
Cyanobacterial microcystin-LR is a potent and specific inhibitor of
protein phosphatases 1 and 2A from both mammals and higher
plants. FEBS Lett 264: 187–192, 1990.
MACKINTOSH C AND MACKINTOSH RW. Inhibitors of protein kinases
and phosphatases. Trends Biochem Sci 19: 444 – 448, 1994.
Volume 80
250. MAGÓSCI M, APÁTI A, GÁTI R, AND KOLONICS A. Signalling mechanisms
and the role of calcineurin in erythropoiesis. Immunol Lett 68:
187–195, 1999.
251. MAI B, FREY G, SWANSON RV, MATHUR EJ, AND STETTER KO. Molecular
cloning and functional expression of a protein-serine/threonine
phosphatase from the hyperthermophilic archaeon Pyrodictum
abyssi TAG11. J Bacteriol 180: 4030 – 4035, 1998.
252. MALCHOW D, MUTZEL R, AND SCHLATTERER C. On the role of calcium
during chemotactic signalling and differentiation of the cellular
slime mold Dictyostelium discoideum. Int J Dev Biol 40: 135–139,
1996.
253. MANDELKOW E-M, BIERNAT J, DREWES G, GUSTKE N, TRINCZEK B, AND
MANDELKOW E. Tau domains, phosphorylation, and interactions
with microtubules. Neurobiol Aging 16: 355–363, 1995.
254. MANSUY IM, MAYFORD M, JACOB B, KANDEL ER, AND BACH ME. Restricted and regulated overexpression reveals calcineurin as a key
component in the transition from short-term to long-term memory.
Cell 92: 39 – 49, 1998.
255. MARKS AR. Cellular functions of immunophilins. Physiol Rev 76:
631– 649, 1996.
256. MARRION NV. Calcineurin regulates M channel modal gating in
sympathetic neurons. Neuron 16: 163–173, 1996.
257. MARTENSEN TM, MARTIN BM, AND KINCAID RL. Identification of the
site on calcineurin phosphorylated by Ca2⫹/CaM-dependent kinase
II: modification of the CaM-binding domain. Biochemistry 28:
9243–9247, 1989.
258. MARTIN BL. Inhibition of calcineurin by the tyrphostin class of
tyrosine kinase inhibitors. Biochem Pharmacol 56: 483– 488, 1998.
259. MARTIN BL AND GRAVES DJ. Mechanistic aspects of the low-molecular-weight phosphatase activity of the calmodulin-activated phosphatase, calcineurin. J Biol Chem 261: 14545–14550, 1986.
260. MARTIN BL AND GRAVES DJ. Isotope effects on the mechanism of
calcineurin catalysis: kinetic solvent isotope and isotope exchange
studies. Biochim Biophys Acta 1206: 136 –142, 1994.
261. MARTIN BL, JURADO LA, AND HENGGE AC. Comparison of the reaction
progress of calcineurin with Mn2⫹ and Mg2⫹. Biochemistry 38:
3386 –3392, 1999.
262. MARTIN BL, PALLEN CJ, WANG JH, AND GRAVES DJ. Use of fluorinated
tyrosine phosphates to probe the substrate specificity of the low
molecular weight phosphatase activity of calcineurin. J Biol Chem
260: 14932–14937, 1985.
263. MATHEOS DP, KINGSBURY TJ, AHSAN US, AND CUNNINGHAM KW. Tcn1p/
Crz1p, a calcineurin-dependent transcription factor that differentially regulates gene expression in Saccharomyces cerevisiae.
Genes Dev 11: 3445–3458, 1997.
264. MCFERRAN BW, GRAHAM ME, AND BURGOYNE RD. Neuronal Ca2⫹
sensor 1, the mammalian homologue of frequenin, is expressed in
chromaffin and PC12 cells and regulates neurosecretion from
dense-core granules. J Biol Chem 273: 22768 –22772, 1998.
265. MCPARTLIN AE, BARKER HM, AND COHEN PTW. Identification of a
third alternatively spliced cDNA encoding the catalytic subunit of
protein phosphatase 2B␤. Biochim Biophys Acta 1088: 308 –310,
1991.
266. MEANS TR. Calcium, calmodulin, and cell cycle regulation. FEBS
Lett 347: 1– 4, 1994.
267. MEGURO T, HONG C, ASAI K, TAKAGI G, MCKINSEY TA, OLSON EN, AND
VATNER SF. Cyclosporine attenuates pressure-overload hypertrophy
in mice while enhancing susceptibility to decompensation and
heart failure. Circ Res 84: 735–740, 1999.
268. MENDOZA I, QUINTERO FJ, BRESSAN RA, HASEGAWA PM, AND PARDO JM.
Activated calcineurin confers high tolerance to ion stress and alters
the budding pattern and cell morphology of yeast cells. J Biol
Chem 271: 23061–23067, 1996.
269. MENDOZA I, RUBIO F, RODRIGUEZ-NAVARRO A, AND PARDO JM. The
protein phosphatase calcineurin is essential for NaCl tolerance of
Saccharomyces cerevisiae. J Biol Chem 269: 8792– 8796, 1994.
270. MENGER FM AND GLASS LE. Contribution of orbital alignment to
organic and enzymatic reactivity. J Am Chem Soc 102: 5404 –5406,
1980.
271. MERAT DL AND CHEUNG WY. Calmodulin-dependent protein phosphatase: isolation of subunits and reconstitution to holoenzyme.
Methods Enzymol 139: 79 – 87, 1987.
272. MERTZ P, YU L, SIKKINK R, AND RUSNAK F. Kinetic and spectroscopic
October 2000
273.
274.
275.
276.
277.
278.
279.
280.
281.
282.
283.
284.
285.
286.
287.
288.
289.
290.
291.
292.
293.
CALCINEURIN: FORM AND FUNCTION
analyses of mutants of a conserved histidine in the metallophosphatases calcineurin and ␭ protein phosphatase. J Biol Chem 272:
21296 –21302, 1997.
MESECAR AD, STODDARD BL, AND KOSHLAND DE JR. Orbital steering in
the catalytic power of enzymes: small structural changes seen with
large catalytic consequences. Science 277: 202–206, 1997.
METCALFE SM AND RICHARDS FM. Cyclosporine, FK506, and rapamycin. Transplantation 49: 798 – 802, 1990.
MICHALAK M, FU SY, MILNER RE, BUSAAN JL, AND HANCE JE. Phosphorylation of the carboxyl-terminal region of dystrophin. Biochem
Cell Biol 74: 431– 437, 1996.
MIKOSHIBA K. The InsP3 receptor and intracellular Ca2⫹ signaling.
Curr Opin Neurobiol 7: 339 –345, 1997.
MISKIN JE, ABRAMS CC, GOATLEY LC, AND DIXON LK. A viral mechanism for inhibition of the cellular phosphatase calcineurin. Science
281: 562–565, 1998.
MIYAMOTO K, MATSUI H, TOMIZAWA K, KUWATA Y, ITANO T, TOKUDA M,
AND HATASE O. In situ localization of rat testis-specific calcineurin B
subunit isoform ␤1 in the developing rat testis. Biochem Biophys
Res Commun 203: 1275–1283, 1994.
MIZUNUMA M, HIRATA D, MIYAHARA K, TSUCHIYA E, AND MIYAKAWA T.
Role of calcineurin and Mpk1 in regulating the onset of mitosis in
budding yeast. Nature 392: 303–306, 1998.
MOLKENTIN JD. Calcineurin inhibition and cardiac hypertrophy. Science 282: 1007a, 1998.
MOLKENTIN JD, LU J-R, ANTOS CL, MARKHAM B, RICHARDSON J, ROBBINS
J, GRANT SR, AND OLSON EN. A calcineurin-dependent transcriptional pathway for cardiac hypertrophy. Cell 93: 215–228, 1998.
MOMAYEZI M, LUMPERT CJ, KERSKEN J, GRAS U, PLATTNER H, KRINKS
MH, AND KLEE CB. Exocytosis induction in Paramecium tetraurelia cells by exogenous phosphoprotein phosphatase in vivo and in
vitro: possible involvement of calcineurin in exocytotic membrane
fusion. J Cell Biol 105: 181–189, 1987.
MONDRAGON A, GRIFFITH EC, SUN L, XIONG F, ARMSTRONG C, AND LIU
JO. Overexpression and purification of human calcineurin ␣ from
Escherichia coli and assessment of catalytic functions of residues
surrounding the binuclear metal center. Biochemistry 36: 4934 –
4942, 1997.
MONTORO RJ, DÍAZ-NIDO J, AVILA J, AND LÓPEZ-BARNEO J. N-methyl-Daspartate stimulates the dephosphorylation of the microtubuleassociated protein 2 and potentiates excitatory synaptic pathways
in the rat hippocampus. Neuroscience 54: 859 – 871, 1993.
MORIOKA M, HAMADA J-I, USHIO Y, AND MIYAMOTO E. Potential role of
calcineurin for brain ischemia and traumatic injury. Prog Neurobiol
58: 1–30, 1999.
MORIYA M, FUJINAGA K, YAZAWA M, AND KATAGIRI C. Immunohistochemical localization of the calcium/calmodulin-dependent protein
phosphatase calcineurin, in the mouse testis: its unique accumulation in spermatid nuclei. Cell Tissue Res 281: 273–281, 1995.
MOSER MJ, GEISER JR, AND DAVIS TN. Ca2⫹-calmodulin promotes
survival of pheromone-induced growth arrest by activation of calcineurin and Ca2⫹-calmodulin-dependent protein kinase. Mol Cell
Biol 16: 4824 – 4831, 1996.
MUELLER EG, CROWDER MW, AVERILL BA, AND KNOWLES JR. Purple
acid phosphatase: a diiron enzyme that catalyzes a direct phospho
group transfer to water. J Am Chem Soc 115: 2974 –2975, 1993.
MUKAI H, CHANG C-D, TANAKA H, ITO A, KUNO T, AND TANAKA C. cDNA
cloning of a novel testis-specific calcineurin B-like protein. Biochem Biophys Res Commun 179: 1325–1330, 1991.
MÜLLER JG, NEMOTO S, LASER M, CARABELLO BA, AND MENICK DR.
Calcineurin inhibition and cardiac hypertrophy. Science 282: 1007,
1998.
MURAMATSU T, GIRI PR, HIGUCHI S, AND KINCAID RL. Molecular cloning of a calmodulin-dependent phosphatase from murine testis:
identification of a developmentally expressed nonneural isoenzyme. Proc Natl Acad Sci USA 89: 529 –533, 1992.
MURAMATSU T AND KINCAID RL. Molecular cloning and chromosomal
mapping of the human gene for the testis-specific catalytic subunit
of calmodulin-dependent protein phosphatase (calcineurin A). Biochem Biophys Res Commun 188: 265–271, 1992.
MURAMATSU T AND KINCAID RL. Molecular cloning of a full-length
cDNA encoding the catalytic subunit of human calmodulin-depen-
294.
295.
296.
297.
298.
299.
300.
301.
302.
303.
304.
305.
306.
307.
308.
309.
310.
311.
312.
313.
314.
1517
dent protein phosphatase (calcineurin A␣). Biochim Biophys Acta
1178: 117–120, 1993.
MURPHY BJ, ROSSIE S, DE JONGH KS, AND CATTERALL WA. Identification of the sites of selective phosphorylation and dephosphorylation of the rat brain Na⫹ channel ␣ subunit by cAMP-dependent
protein kinase and phosphoprotein phosphatases. J Biol Chem 268:
27355–27362, 1993.
MUSARÒ A, MCCULLAGH KJA, NAYA FJ, OLSON EN, AND ROSENTHAL N.
IGF-1 induces skeletal myocyte hypertrophy through calcineurin in
association with GATA-2 and NF-ATc1. Nature 400: 581–585, 1999.
MYERS JK, ANTONELLI SM, AND WIDLANSKI TS. Motifs for metallophosphatase inhibition. J Am Chem Soc 119: 3163–3164, 1997.
NAIK UP, PATEL PM, AND PARISE LV. Identification of ␣ novel calcium-binding protein that interacts with the integrin ␣iib cytoplasmic domain. J Biol Chem 272: 4651– 4654, 1997.
NAIR APK, HAHN S, BANHOLZER R, HIRSCH HH, AND MORONI C. Cyclosporin A inhibits growth of autocrine tumour cell lines by destabilizing interleukin-3 mRNA. Nature 369: 239 –242, 1994.
NAIRN AC AND SHENOLIKAR S. The role of protein phosphatases in
synaptic transmission, plasticity, and neuronal development. Curr
Opin Neurobiol 2: 296 –301, 1992.
NAKAMURA T, LIU Y, HIRATA D, NAMBA H, HARADA S-I, HIROKAWA T, AND
MIYAKAWA T. Protein phosphatase type 2B (calcineurin)-mediated,
FK506-sensitive regulation of intracellular ions in yeast is an important determinant for adaptation to high salt stress conditions.
EMBO J 12: 4063– 4071, 1993.
NAKAMURA T, OHMOTO T, HIRATA D, TSUCHIYA E, AND MIYAKAWA T.
Genetic evidence for the functional redundancy of the calcineurinand Mpk1-mediated pathways in the regulation of cellular events
important for growth in Saccharomyes cerevisiae. Mol Gen Genet
251: 211–219, 1996.
NAKAYAMA S AND KRETSINGER RH. Evolution of the EF-hand family of
proteins. Annu Rev Biophys Biomol Struct 23: 473–507, 1994.
NANTHAKUMAR NN, DAYTON JS, AND MEANS AR. Role of Ca2⫹/calmodulin binding proteins in Aspergillus nidulans cell cycle regulation.
Prog Cell Cycle Res 2: 217–228, 1996.
NARGANG CE, BOTTORFF DA, AND ADACHI Z. Isolation and characterization of a cDNA clone coding for the calcium-binding subunit of
calcineurin from bovine brain: an identical amino acid sequence to
the human protein. DNA Sequence 4: 313–318, 1994.
NATARAJ C, OLIVERIO MI, MANNON RB, MANNON PJ, AUDOLY LP, AMUCHASTEGUI CS, RUIZ P, SMITHIES O, AND COFFMAN TM. Angiotensin II
regulates cellular immune responses through a calcineurin-dependent pathway. J Clin Invest 104: 1693–1701, 1999.
NATARAJAN K, NESS J, WOOGE CH, JANOVICK JA, AND CONN PM. Specific identification and subcellular localization of three calmodulinbinding proteins in the rat gonadotrope: spectrin, caldesmon, and
calcineurin. Biol Reprod 44: 43–52, 1991.
NAVIA MA. Protein-drug complexes important for immunoregulation and organ transplantation. Curr Opin Struct Biol 6: 838 – 847,
1996.
NICHOLS RA, SUPLICK GR, AND BROWN JM. Calcineurin-mediated protein dephosphorylation in brain nerve terminals regulates the release of glutamate. J Biol Chem 269: 23817–23823, 1994.
ODOM A, DEL POETA M, PERFECT J, AND HEITMAN J. The immunosuppressant FK506 and its nonimmunosuppressive analog L-685,818
are toxic to Cryptococcus neoformans by inhibition of a common
target protein. Antimicrobial Agents Chemother 41: 156 –161, 1997.
ODOM A, MUIR S, LIU E, TOFFALETTI DL, PERFECT J, AND HEITMAN J.
Calcineurin is required for virulence of Cryptococcus neoformans.
EMBO J 16: 2576 –2589, 1997.
O’KEEFE SJ AND O’NEILL EA. Cyclosporin A and FK-506: immunosuppression, inhibition of transcription and the role of calcineurin.
Perspect Drug Disc Design 2: 85–102, 1994.
OLSON EN AND MOLKENTIN JD. Prevention of cardiac hypertrophy by
calcineurin inhibition. Hype or hope? Circ Res 84: 623– 632, 1999.
PAIDHUNGAT M AND GARRETT S. A homolog of mammalian, voltagegated calcium channels mediates yeast pheromone-stimulated
Ca2⫹ uptake and exacerbates the cdc1(ts) growth defect. Mol Cell
Biol 17: 6339 – 6347, 1997.
PALLEN CJ, VALENTINE KA, WANG JH, AND HOLLENBERG MD. Calcineurin-mediated dephosphorylation of the human placental mem-
1518
315.
316.
317.
318.
319.
320.
321.
322.
323.
324.
325.
326.
327.
328.
329.
330.
331.
332.
333.
334.
335.
336.
FRANK RUSNAK AND PAMELA MERTZ
brane receptor for epidermal growth factor urogastrone. Biochemistry 24: 4727– 4730, 1985.
PALLEN CJ AND WANG JH. Regulation of calcineurin by metal ions.
J Biol Chem 259: 6134 – 6141, 1984.
PALLEN CJ AND WANG JH. A multifunctional calmodulin-stimulated
phosphatase. Arch Biochem Biophys 237: 281–291, 1985.
PALLEN CJ AND WANG JH. Stoichiometry and dynamic interaction of
metal ion activators with calcineurin phosphatase. J Biol Chem
261: 16115–16120, 1986.
PAPADOPOULOS V, BROWN AS, AND HALL PF. Isolation and characterization of calcineurin from adrenal cell cytoskeleton: identification
of substrates of Ca2⫹-calmodulin-dependent phosphatase activity.
Mol Cell Endocrinol 63: 23–38, 1989.
PAPADOPOULOS V, BROWN AS, AND HALL PF. Calcium-calmodulindependent phosphorylation of cytoskeletal proteins from adrenal
cells. Mol Cell Endocrinol 74: 109 –123, 1990.
PARDO JM, REDDY MP, YANG S, MAGGIO A, HUH G-H, MATSUMOTO T,
COCA MA, PAINO-D’URZO M, KOIWA H, YUN D-J, WATAD AA, BRESSAN
RA, AND HASEGAWA PM. Stress signaling through Ca2⫹/calmodulindependent protein phosphatase calcineurin mediates salt adaption
in plants. Proc Natl Acad Sci USA 95: 9681–9686, 1998.
PARK I-S AND HAUSINGER RP. Site-directed mutagensis of Klebsiella
aerogenes urease: identification of histidine residues that appear to
function in nickel ligation, substrate binding, and catalysis. Protein
Sci 2: 1034 –1041, 1993.
PATERSON JM, SMITH SM, HARMAR AJ, AND ANTONI FA. Control of a
novel adenylyl cyclase by calcineurin. Biochem Biophys Res Commun 214: 1000 –1008, 1995.
PAWSON T AND SCOTT JD. Signaling through scaffold, anchoring, and
adaptor proteins. Science 278: 2075–2080, 1997.
PERRINO BA. Regulation of calcineurin phosphatase activity by its
autoinhibitory domain. Arch Biochem Biophys 372: 159 –165, 1999.
PERRINO BA, FONG Y-L, BRICKEY DA, SAITOH Y, USHIO Y, FUKUNAGA K,
MIYAMOTO E, AND SODERLING TR. Characterization of the phosphatase activity of a baculovirus-expressed calcineurin A isoform.
J Biol Chem 267: 15965–15969, 1992.
PERRINO BA, NG LY, AND SODERLING TR. Calcium regulation of calcineurin phosphatase activity by its B subunit and calmodulin.
J Biol Chem 270: 340 –346, 1995.
PLOCHOCKA-ZULINSKA D, RASMUSSEN G, AND RASMUSSEN C. Regulation
of calcineurin gene expression in Schizosaccharomyces pombe.
J Biol Chem 270: 24794 –24799, 1995.
PLOWMAN GD, SUDARSANAM S, BINGHAM J, WHYTE D, AND HUNTER T.
The protein kinases of Caenorhabditis elegans: a model for signal
transduction in multicellular organisms. Proc Natl Acad Sci USA
96: 13603–13610, 1999.
POLITINO M AND KING MM. Calcium- and calmodulin-sensitive interactions of calcineurin with phospholipids. J Biol Chem 262: 10109 –
10113, 1987.
POLITNO M AND KING MM. Calcineurin-phospholipid interactions:
identification of the phospholipid-binding subunit and analyses of a
two-stage binding process. J Biol Chem 265: 7619 –7622, 1990.
POMIÉS P, FRACHET P, AND BLOCK MR. Control of the ␣5␤1 integrin/
fibronectin interaction in vitro by the serine/threonine protein
phosphatase calcineurin. Biochemistry 34: 5104 –5112, 1995.
POZOS TC, SEKLER I, AND CYERT MS. The product of HUM1, a novel
yeast gene, is required for vacuolar Ca2⫹/H⫹ exchange and is
related to mammalian Na⫹/Ca2⫹ exchangers. Mol Cell Biol 16:
3730 –3741, 1996.
PRICE NE AND MUMBY MC. Brain protein serine/threonine phosphatases. Curr Opin Neurobiol 9: 336 –342, 1999.
PRIGGEMEYER S, EGGERS-BORKENSTEIN P, AHLERS F, HENKEL G, KÖRNER
M, WITZEL H, NOLTING H-F, AND KREBS B. XAS investigations on the
iron-zinc center of purple acid phosphatase from red kidney beans.
Inorg Chem 34: 1445–1454, 1995.
PROKISCH H, YARDEN O, DIEMINGER M, TROPSCHUG M, AND BARTHELMESS IB. Impairment of calcineurin function in Neurospora
crassa reveals its essential role in hyphal growth, morphology and
maintenance of the apical Ca2⫹ gradient. Mol Gen Genet 256:
104 –114, 1997.
PUJOL MJ, BOSSER R, VENDRELL M, SERRATOSA J, AND BACHS O. Nuclear
calmodulin-binding protiens in rat neurons. J Neurochem 60: 1422–
1428, 1993.
Volume 80
337. RANGER AM, GRUSBY MJ, HODGE MR, GRAVALLESE EM, DE LA BROUSSE
FC, HOEY T, MICKANIN C, BALDWIN HS, AND GLIMCHER LH. The transcription factor NF-ATc is essential for cardiac valve formation.
Nature 392: 186 –190, 1998.
338. RAO A. NF-ATp: a transcription factor required for the co-ordinate
induction of several cytokine genes. Immunol Today 15: 274 –280,
1994.
339. RAO A. NFATp, a cyclosporin-sensitive transcription factor implicated in cytokine gene induction. J Leukoc Biol 57: 536 –542, 1995.
340. RAO A, LUO C, AND HOGAN PG. Transcription factors of the NFAT
family: regulation and function. Annu Rev Immunol 15: 707–747,
1997.
341. RAO J AND WANG JH. Calcineurin immunoprecipitated from bovine
brain extract contains no detectable Ni2⫹ or Mn2⫹. J Biol Chem
264: 1058 –1061, 1989.
342. RASCHER C, PAHL A, PECHT A, BRUNE K, SOLBACH W, AND BANG H.
Leishmania major parasites express cyclophilin isoforms with an
unusual interaction with calcineurin. Biochem J 334: 659 – 667,
1998.
343. RASMUSSEN C, GAREN C, BRINING S, KINCAID RL, MEANS RL, AND MEANS
AR. The calmodulin-dependent protein phosphatase catalytic subunit (calcineurin A) is an essential gene in Aspergillus nidulans.
EMBO J 13: 2545–2552, 1994.
344. RAUFMAN J-P, MALHOTRA R, AND RAFANIELLO RD. Regulation of calcium-induced exocytosis from gastric chief cells by protein phosphatase-2B (calcineurin). Biochim Biophys Acta 1357: 73– 80, 1997.
345. REITER TA, ABRAHAM RT, CHOI M, AND RUSNAK F. Redox regulation of
calcineurin in T-lymphocytes. J Biol Inorg Chem 4: 632– 644, 1999.
346. REMILLARD SP, LAI EY, LEVY YY, AND FULTON C. A calcineurin-Bencoding gene expressed during differentiation of the amoeboflagellate Naegleria gruberi contains two introns. Gene 154: 39 – 45,
1995.
347. ROBERTS HC, STERNBERG JM, AND CHAPPELL LH. Characterization of
calcineurin from Hymenolepsis microstoma and H. diminuta and
its interactions with cyclosporin A. Parasitology 114: 279 –283,
1997.
348. ROEPER J, LORRA C, AND PONGS O. Frequency-dependent inactivation
of mammalian A-type K⫹ channel kv1.4 regulated by Ca2⫹/calmodulin-dependent protein kinase. J Neurosci 17: 3379 –3391, 1997.
349. RUDOLPH HK, ANTEBI A, FINK GR, BUCKLEY CM, DORMAN TE, LEVITRE
J, DAVIDOW LS, MAO J-I, AND MOIR DT. The yeast secretory pathway
is perturbed by mutations in Pmr1, a member of the Ca2⫹ ATPase
family. Cell 58: 133–145, 1989.
350. RUSNAK F, MERTZ P, REITER T, AND YU L. Regulation of the protein
phosphatase calcineurin by redox: implications for catalysis and
signal transduction. In: Iron Metabolism: Inorganic Biochemistry
and Regulatory Mechanisms, edited by Ferreira GC, Moura JJG,
and Franco R. New York: Wiley-VCH, 1999, p. 275–302.
351. RUSNAK F, YU L, AND MERTZ P. Metalloenzymes and signal transduction: the protein serine/threonine phosphatases, a novel class of
binuclear metal-containing enzymes. J Biol Inorg Chem 1: 388 –396,
1996.
352. RUSNAK F, YU L, TODOROVIC S, AND MERTZ P. The interaction of
bacteriophage ␭ protein phosphatase with Mn(II): evidence for the
formation of a [Mn(II)2] cluster. Biochemistry 38: 6943– 6952, 1999.
353. SABATINI DM, LAI MM, AND SNYDER SH. Neural roles of immunophilins and their ligands. Mol Neurobiol 15: 223–239, 1997.
354. SAITOH Y, MAEDA S, FUKUNAGA K, YASUGAWA S, SAKAMOTO Y, UYEMURA
K, SHIMADA K, USHIO Y, AND MIYAMOTO E. Nucleotide sequence of a
rat calcineurin A cDNA lacking a specific 30-base pair region.
Biomed Res 12: 215–218, 1991.
355. SALMINEN T, KÄPYLÄ J, HEIKINHEIMO P, KANKARE J, GOLDMAN A, HEINONEN J, BAYKOV AA, COOPERMAN BS, AND LAHTI R. Structure and
function analysis of Escherichia coli inorganic pyrophosphatase: is
a hydroxide ion the key to catalysis? Biochemistry 34: 782–791,
1995.
356. SANDER M, LYSON T, THOMAS GD, AND VICTOR RG. Sympathetic neural
mechanisms of cyclosporine-induced hypertension. Am J Hypertens 9: 121S–138S, 1996.
357. SANTELLA L. The role of calcium in the cell cycle: facts and hypotheses. Biochem Biophys Res Commun 244: 317–324, 1998.
358. SANTELLA L AND CARAFOLI E. Calcium signaling in the cell nucleus.
FASEB J 11: 1091–1109, 1997.
October 2000
CALCINEURIN: FORM AND FUNCTION
359. SCHAAD NC, DE CASTRO E, NEF S, HEGI S, HINRICHSEN R, MARTONE ME,
ELLISMAN MH, SIKKINK R, RUSNAK F, SYGUSH J, AND NEF P. Direct
modulation of calmodulin targets by the neuronal calcium sensor
NCS-1. Proc Natl Acad Sci USA 93: 9253–9258, 1996.
360. SCHAEFER A, MAGÓCSI M, STÖCKER U, FANDRICH A, AND MARQUARDT H.
Ca2⫹/calmodulin-dependent and -independent down-regulation of
c-myb mRNA levels in erythropoietin-responsive murine erythroleukemia cells. J Biol Chem 271: 13484 –13490, 1996.
361. SCHREIBER SL AND CRABTREE GR. The mechanism of action of cyclosporin A and FK506. Immunol Today 13: 136 –142, 1992.
362. SCHUHMANN K, ROMANIN C, BAUMGARTNER W, AND GROSCHNER K. Intracellular Ca2⫹ inhibits smooth muscle L-type Ca2⫹ channels by
activation of protein phosphatase type 2B and by direct interaction
with the channel. J Gen Physiol 110: 503–513, 1997.
363. SCHUMACHER A AND NORDHEIM A. Progress towards a molecular
understanding of cyclosporin A-mediated immunosuppression.
Clin Invest 70: 773–779, 1992.
364. SCHWANINGER M, BLUME R, OETJEN E, LUX G, AND KNEPEL W. Inhibition of cAMP-responsive element-mediated gene transcription by
cyclosporin A and FK506 after membrane depolarization. J Biol
Chem 268: 23111–23115, 1993.
365. SCOTT JD. Dissection of protein kinase and phosphatase targeting
interactions. Soc Gen Phys Ser 52: 227–239, 1997.
366. SCOTT JD AND MCCARTNEY S. Localization of A-kinase through anchoring proteins. Mol Endocrinol 8: 5–11, 1994.
367. SEMSARIAN C, WU M-J, JU Y-K, MARCINIEC T, YEOH T, ALLEN DG,
HARVEY RP, AND GRAHAM RM. Skeletal muscle hypertrophy is mediated by a Ca2⫹-dependent calcineurin signalling pathway. Nature
400: 576 –581, 1999.
368. SERRANO R. Salt tolerance in plants and microorganisms: toxicity
targets and defense responses. In: International Review of Cytology: A Survey of Cell Biology, edited by Jeon KW. San Diego, CA:
Academic, 1996, p. 1–52.
369. SHARMA RK AND KALRA J. Molecular interaction between cAMP and
calcium in calmodulin-dependent cyclic nucleotide phosphodiesterase system. Clin Invest Med 17: 374 –382, 1994.
370. SHENOLIKAR S. Protein phosphatase regulation by endogenous inhibitors. Semin Cancer Biol 6: 219 –227, 1995.
371. SHENOLIKAR S AND NAIRN AC. Protein phosphatases: recent progress.
Adv Second Messenger Phosphoprotein Res 23: 3–121, 1991.
372. SHI J, KIM KN, RITZ O, ALBRECHT V, GUPTA R, HARTER K, LUAN S, AND
KUDLA J. Novel protein kinases associated with calcineurin-B-like
calcium sensors in arabidopsis. Plant Cell 11: 2393–2406, 1999.
373. SHI L, CARMICHAEL WW, AND KENNELLY PJ. Cyanobacterial PPP family
protein phosphatases possess multifunctional capabilities and are
resistant to microcystin-LR. J Biol Chem 274: 10039 –10046, 1999.
374. SHIBASAKI F, KONDO E, AKAGI T, AND MCKEON F. Suppression of
signalling through transcription factor NF-AT by interactions between calcineurin and Bcl-2. Nature 386: 728 –731, 1997.
375. SHIBASAKI F AND MCKEON F. Calcineurin functions in Ca2⫹-activated
cell death in mammalian cells. J Cell Biol 131: 735–743, 1995.
376. SHIBASAKI F, PRICE ER, MILAN D, AND MCKEON F. Role of kinases and
the phosphatase calcineurin in the nuclear shuttling of transcription factor NF-AT4. Nature 382: 370 –373, 1996.
377. SHIMOYAMA M, HAYASHI D, TAKIMOTO E, ZOU Y, OKA T, UOZUMI H,
KUDOH S, SHIBASAKI F, YAZAKI Y, NAGAI R, AND KOMURO I. Calcineurin
plays a critical role in pressure overload-induced cardiac hypertrophy. Circulation 100: 2449 –2454, 1999.
378. SIEKIERKA JJ AND SIGAL NH. FK-506 and cyclosporin A: immunosuppressive mechanism of action and beyond. Curr Opin Immunol 4:
548 –552, 1992.
379. SIKKINK R, HADDY A, MACKELVIE S, MERTZ P, LITWILLER R, AND RUSNAK
F. Calcineurin subunit interactions: mapping the calcineurin B
binding domain on calcineurin A. Biochemistry 34: 8348 – 8356,
1995.
380. SIMMEN RCM, SRINIVAS V, AND ROBERTS RM. cDNA sequence, gene
organization, and progesterone induction of mRNA for uteroferrin,
a porcine uterine iron transport protein. DNA 8: 543–554, 1989.
381. SINGAL PK, KHAPER N, PALACE V, AND KUMAR D. The role of oxidative
stress in the genesis of heart disease. Cardiovasc Res 40: 426 – 432,
1998.
382. SINGH TJ AND WANG JH. Phosphorylation of calcineurin by glycogen
synthase (casein) kinase-1. Biochem Cell Biol 65: 917–921, 1987.
1519
383. SMITH RD AND WALKER JC. Plant protein phosphatases. Annu Rev
Plant Physiol Plant Mol Biol 47: 101–125, 1996.
384. SNYDER SH, LAI MM, AND BURNETT PE. Immunophilins in the nervous
system. Neuron 21: 283–294, 1998.
385. SNYDER SH, SABATINI DM, LAI MM, STEINER JP, HAMILTON GS, AND
SUZDAK PD. Neural actions of immunophilin ligands. Trends Pharmacol Sci 19: 21–26, 1998.
386. SOLOW B, YOUNG JC, AND KENNELLY PJ. Gene cloning and expression
and characterization of a toxin-sensitive protein phosphatase from
the methanogenic archeon Methanosarcina thermophila TM-1. J
Bacteriol 179: 5072–5075, 1997.
387. STARK MJR. Yeast protein serine/threonine phosphatases: multiple
roles and diverse regulation. Yeast 12: 1647–1675, 1996.
388. STATHOPOULOS AM AND CYERT MS. Calcineurin acts through the
CRZ1/TCN1-encoded transcription factor to regulate gene expression in yeast. Genes Dev 11: 3432–3444, 1997.
389. STEMMER PM AND KLEE CB. Dual calcium ion regulation of calcineurin by calmodulin and calcineurin B. Biochemistry 33: 6859 –
6866, 1994.
390. STEWART AA, INGEBRITSEN TS, AND COHEN P. The protein phosphatases involved in cellular regulation. 5. Purification and properties
of a Ca2⫹/calmodulin-dependent protein phosphatase (2B) from
rabbit skeletal muscle. Eur J Biochem 132: 289 –295, 1983.
391. STEWART AA, INGEBRITSEN TS, MANALAN A, KLEE CB, AND COHEN P.
Discovery of a Ca2⫹- and calmodulin-dependent protein phosphatase: probable identity with calcineurin (CaM-BP80). FEBS Lett 137:
80 – 84, 1982.
392. STODDARD BL AND FLICK KE. Calcineurin-immunosupressor complexes. Curr Opin Struct Biol 6: 770 –775, 1996.
393. STORM DR AND KOSHLAND DE JR. A source for the special catalytic
power of enzymes: orbital steering. Proc Natl Acad Sci USA 66:
445– 452, 1970.
394. STRACK S, WADZINSKI BE, AND EBNER FF. Localization of the calcium/
calmodulin-dependent protein phosphatase, calcineurin, in the
hindbrain and spinal cord of the rat. J Comp Neurol 375: 66 –76,
1996.
395. STRÄTER N, KLABUNDE T, TUCKER P, WITZEL H, AND KREBS B. Crystal
structure of a purple acid phosphatase containing a dinuclear
Fe(III)-Zn(II) active site. Science 268: 1489 –1492, 1995.
396. SU Q, ZHAO M, WEBER E, EUGSTER H-P, AND RYFFEL B. Distribution
and activity of calcineurin in rat tissues. Evidence for post-transcriptional regulation of testis-specific calcineurin B. Eur J Biochem 230: 469 – 474, 1995.
397. SUGDEN PH. Signaling in myocardial hypertrophy. Life after calcineurin? Circ Res 84: 633– 646, 1999.
398. SUGIMOTO T, STEWART S, AND GUAN K-L. The calcium/calmodulindependent protein phosphatase calcineurin is the major Elk-1
phosphatase. J Biol Chem 272: 29415–29418, 1997.
399. SUGIURA R, TODA T, SHUNTOH H, YANAGIDA M, AND KUNO T. Pmp1⫹, a
suppressor of calcineurin deficiency, encodes a novel MAP kinase
phosphatase in fission yeast. EMBO J 17: 140 –148, 1998.
400. SUN L, YOUN H-D, LOH C, STOLOW M, HE W, AND LIU JO. Cabin 1, a
negative regulator for calcineurin signalling in T lymphocytes. Immunity 8: 703–711, 1998.
401. SUSSMAN MA, LIM HW, GUDE N, TAIGEN T, OLSON EN, ROBBINS J,
COLBERT MC, GUALBERTO A, WIECZOREK DF, AND MOLKENTIN JD. Prevention of cardiac hypertrophy in mice by calcineurin inhibition.
Science 281: 1690 –1693, 1998.
402. SUSSMAN MA, WELCH S, GUDE N, KHOURY PR, DANIELS SR, KIRKPATRICK
D, WALSH RA, PRICE RL, LIM HW, AND MOLKENTIN JD. Pathogenesis of
dilated cardiomyopathy: molecular, structural, and population
analyses in tropomodulin-overexpressing transgenic mice. Am J
Pathol 155: 2101–2113, 1999.
403. SWANSON SK-H, BORN T, ZYDOWSKY LD, CHO H, CHANG HY, WALSH CT,
AND RUSNAK F. Cyclosporin-mediated inhibition of bovine calcineurin by cyclophilins A and B. Proc Natl Acad Sci USA 89:
3741–3745, 1992.
404. SWOPE SL, MOSS SJ, RAYMOND LA, AND HUGANIR RL. Regulation of
ligand-gated ion channels by protein phosphorylation. Adv Second
Messenger Phosphoprotein Res 33: 49 –78, 1999.
405. TAKUWA N, ZHOU W, AND TAKUWA Y. Calcium, calmodulin and cell
cycle progression. Cell Signal 7: 93–104, 1995.
406. TALLANT EA, BRUMLEY LM, AND WALLACE RW. Activation of a cal-
1520
407.
408.
409.
410.
411.
412.
413.
414.
415.
416.
417.
418.
419.
420.
421.
422.
423.
424.
425.
426.
FRANK RUSNAK AND PAMELA MERTZ
modulin-dependent phosphatase by a Ca2⫹-dependent protease.
Biochemistry 27: 2205–2211, 1988.
TAMURA K, FUJIMURA T, TSUTSUMI T, NAKAMURA K, OGAWA T, ATUMARU
C, HIRANO Y, OHARA K, OHTSUKA K, SHIMOMURA K, AND KOBAYASHI M.
Transcriptional inhibition of insulin by FK506 and possible involvement of FK506 binding protein-12 in pancreatic ␤ cells. Transplantation 59: 1606 –1613, 1995.
TANIDA I, HASEGAWA A, IIDA H, OHYA Y, AND ANRAKU Y. Cooperation
of calcineurin and vacuolar H⫹-ATPase in intracellular Ca2⫹ homeostasis of yeast cells. J Biol Chem 270: 10113–10119, 1995.
TASH JS, KRINKS M, PATEL J, MEANS RL, KLEE CB, AND MEANS AR.
Identification, characterization, and functional correlation of calmodulin-dependent protein phosphatase in sperm. J Cell Biol 106:
1625–1633, 1988.
TATLOCK JH, LINTON MA, HOU XJ, KISSINGER CR, PELLETIER LA, SHOWLATER RE, TEMPCZYK A, AND VILLAFRANCA JE. Structure-based design
of novel calcineurin (PP2B) inhibitors. Bioorg Med Chem Lett 7:
1007–1012, 1997.
TAYLOR WP AND WIDLANSKI TS. Charged with meaning: the structure
and mechanism of phosphoprotein phosphatases. Chem Biol 2:
713–718, 1995.
THOMSON AW, BONHAM CA, AND ZEEVI A. Mode of action of tacrolimus (FK506): molecular and cellular mechanisms. Ther Drug
Monit 17: 584 –591, 1995.
THOMSON AW AND STARZL TE. New immunosuppressive drugs: mechanistic insights and potential therapeutic advances. Immunol Rev
136: 71–98, 1993.
TIAN J AND KARIN M. Stimulation of Elk1 transcriptional activity by
mitogen-activated protein kinases is negatively regulated by protein phosphatase 2B (calcineurin). J Biol Chem 274: 15173–15180,
1999.
TRAYNELIS SF AND WAHL P. Control of rat GluR6 glutamate receptor
open probability by protein kinase A and calcineurin. J Physiol
(Lond) 503: 513–531, 1997.
TRUE AE, SCARROW RC, RANDALL CR, HOLZ RC, AND QUE L JR. EXAFS
studies of uteroferrin and its anion complexes. J Am Chem Soc 115:
4246 – 4255, 1993.
TSAO L, NEVILLE C, MURARO A, MCCULLAGH KJA, AND ROSENTHAL N.
Revisiting calcineurin and human heart failure. Nature Med 6: 2–3,
2000.
TUMLIN JA. Expression and function of calcineurin in mammalian
nephron: physiological roles, receptor signaling, and ion transport.
Am J Kidney Dis 30: 884 – 895, 1997.
TUMLIN JA, SOMEREN JT, SWANSON CE, AND LEA JP. Expression of
calcineurin activity and ␣-subunit isoforms in specific segments of
the rat nephron. Am J Physiol Renal Fluid Electrolyte Physiol 269:
F558 –F563, 1995.
TUNG HYL. Phosphorylation of the calmodulin-dependent protein
phosphatase by protein kinase C. Biochem Biophys Res Commun
138: 783–788, 1986.
TÓTH R, SZEGEZDI E, MOLNÁR G, LORD JM, FÉSÜS L, AND ZSUZSA S.
Regulation of cell surface expression of Fas (CD95) ligand and
susceptibility to Fas (CD95)-mediated apoptosis in activation-induced T cell death involves calcineurin and protein kinase C,
respectively. Eur J Immunol 29: 383–393, 1999.
UCHINO H, ELMÉR E, UCHINO K, LI P-A, HE Q-P, SMITH M-L, AND SEISJÖ
BK. Amelioration by cyclosporin A of brain damage in transient
forebrain ischemia in the rat. Brain Res 812: 216 –226, 1998.
UEKI K AND KINCAID RL. Interchangeable associations of calcineurin
regulatory subunit isoforms with mammalian and fungal catalytic
subunits. J Biol Chem 268: 6554 – 6559, 1993.
UEKI K, MURAMATSU T, AND KINCAID RL. Structure and expression of
two isoforms of the murine calmodulin-dependent protein phosphatase regulatory subunit (calcineurin B). Biochem Biophys Res
Commun 187: 537–543, 1992.
UPPENBERG J, LINDQVIST F, SVENSSON C, EK-RYLANDER B, AND ANDERSSON G. Crystal structure of mammalian purple acid phosphatase. J
Mol Biol 290: 201–211, 1999.
USUDA N, ARAI H, SASAKI H, HANAI T, NAGATA T, MURAMATSU T,
KINCAID RL, AND HIGUCHI S. Differential subcellular localization of
neural isoforms of the catalytic subunit of calmodulin-dependent
protein phosphatase (calcineurin) in central nervous system neurons: immunohistochemistry on Formalin-fixed paraffin sections
427.
428.
429.
430.
431.
432.
433.
434.
435.
436.
437.
438.
439.
440.
441.
442.
443.
444.
445.
446.
447.
Volume 80
employing antigen retrieval by microwave irradiation. J Histochem
Cytochem 44: 13–18, 1996.
VICTOR RG, THOMAS GD, MARBAN E, AND O’ROURKE B. Presynaptic
modulation of cortical synaptic activity by calcineurin. Proc Natl
Acad Sci USA 92: 6269 – 6273, 1995.
VILLAFRANCA JE, KISSINGER CR, AND PARGE HE. Protein serine/threonine phosphatases. Curr Opin Biotechnol 7: 397– 402, 1996.
VILLALBA M, KASIBHATLA S, GENESTIER L, MAHBOUBI A, GREEN DR, AND
ALTMAN A. Protein kinase C⌰ cooperates with calcineurin to induce
Fas ligand expression during activation-induced T cell death. J Immunol 163: 5813–5819, 1999.
VINCENT JB AND AVERILL BA. An enzyme with a double identity:
purple acid phosphatase and tartarate-resistant acid phosphatase.
FASEB J 4: 3009 –3014, 1990.
VINCENT JB AND AVERILL BA. Sequence homology between purple
acid phosphatases and phosphoprotein phosphatases: are phosphoprotein phosphatases metalloproteins containing oxide-bridged
dinuclear metal centers? FEBS Lett 263: 265–268, 1990.
VINCENT JB, CROWDER MW, AND AVERILL BA. Hydrolysis of phosphate
monoesters: a biological problem with multiple chemical solutions.
Trends Biochem Sci 17: 105–110, 1992.
VINCENT JB, OLIVIER-LILLEY GL, AND AVERILL BA. Proteins containing
oxo-bridged dinuclear iron centers: a bioinorganic perspective.
Chem Rev 90: 1447–1467, 1990.
WALAAS SI AND GREENGARD P. Protein phosphorylation and neuronal
function. Pharmacol Rev 43: 299 –349, 1991.
WALLACE RW, LYNCH TJ, TALLANT EA, AND CHEUNG WY. Purification
and characterization of an inhibitor protein of brain adenylate
cyclase and cyclic nucleotide phosphodiesterase. J Biol Chem 254:
377–382, 1978.
WALLACE RW, LYNCH TJ, TALLANT EA, AND CHEUNG WY. An endogenous inhibitor protein of brain adenylate cyclase and cyclic nucleotide phosphodiesterase. Arch Biochem Biophys 187: 328 –334,
1978.
WALLACE RW, TALLANT EA, AND CHEUNG WY. High levels of a heatlabile calmodulin-binding protein (CaM-BP80) in bovine neostriatum. Biochemistry 19: 1831–1837, 1980.
WALLACE RW, TALLANT EA, AND MCMANUS MC. Human platelet calmodulin-binding proteins: identification and Ca2⫹-dependent proteolysis upon platelet activation. Biochemistry 26: 2766 –2773,
1987.
WALSH CT, ZYDOWSKY LD, AND MCKEON FD. Cyclosporin A, the
cyclophilin class of peptidylprolyl isomerases, and blockade of T
cell signal transduction. J Biol Chem 267: 13115–13118, 1992.
WALSH MP, BUSAAN JL, FRASER ED, FU SY, PATO MD, AND MICHALAK
M. Characterization of the recombinant C-terminal domain of dystrophin: phosphorylation by calmodulin-dependent protein kinase
ii and dephosphorylation by type 2B protein phosphatase. Biochemistry 34: 5561–5568, 1995.
WALSH RA. Calcineurin inhibition as therapy for cardiac hypertrophy and heart failure. Circ Res 84: 741–743, 1999.
WANG C-R, HASHIMOTO K, KUBO S, YOKOCHI T, KUBO M, SUZUKI M,
SUZUKI K, TADA T, AND NAKAYAMA T. T cell receptor-mediated signaling events in CD4⫹ CD8⫹ thymocytes undergoing thymic selection:
requirement of calcineurin activation for thymic positive selection
but not negative selection. J Exp Med 181: 927–941, 1995.
WANG H-G, PATHAN N, ETHELL IM, KRAJEWSKI S, YAMAGUCHI Y, SHI2⫹
BASAKI F, MCKEON F, BOBO T, FRANKE TF, AND REED JC. Ca -induced
apoptosis through calcineurin dephosphorylation of BAD. Science
284: 339 –343, 1999.
WANG JH AND DESAI R. A brain protein and its effect on the Ca2⫹and protein modulator-activated cyclic nucleotide phosphodiesterase. Biochem Biophys Res Commun 72: 926 –932, 1976.
WANG JH AND DESAI R. Modulator binding protein: bovine brain
protein exhibiting the Ca2⫹-dependent association with the protein
modulator of cyclic nucleotide phosphodiesterase. J Biol Chem
252: 4175– 4184, 1977.
WANG MG, YI H, GUERINI D, KLEE CB, AND MCBRIDE OW. Calcineurin
A alpha (PPP3CA), calcineurin A beta (PPP3CB) and calcineurin B
(PPP3R1) are located on human chromosomes 4, 10q213q22 and
2p163p15, respectively. Cytogenet Cell Genet 72: 236 –241, 1996.
WANG X, CULOTTA VC, AND KLEE CB. Superoxide dismutase protects
calcineurin from inactivation. Nature 383: 434 – 437, 1996.
October 2000
CALCINEURIN: FORM AND FUNCTION
448. WANG X AND QUE L JR. Extended X-ray absorption fine structure
studies of the anion complexes of FeZn uteroferrin. Biochemistry
37: 7813–7821, 1998.
449. WANG X, RANDALL CR, TRUE AE, AND QUE L JR. X-ray absorption
spectroscopic studies of the FeZn derivative of uteroferrin. Biochemistry 35: 13946 –13954, 1996.
450. WARREN WD, PHILLIPS AM, AND HOWELLS AJ. Drosophila melanogaster contains both X-linked and autosomal homologues of the
gene encoding calcineurin B. Gene 177: 149 –153, 1996.
451. WATANABE Y, PERRINO BA, CHANG BH, AND SODERLING TR. Identification in the calcineurin A subunit of the domain that binds the
regulatory B subunit. J Biol Chem 270: 456 – 460, 1995.
452. WATTERSON DM AND VANAMAN TC. Affinity chromatrography purification of a cyclic nucleotide phosphodiesterase using immobilized
modulator protein, a troponin C-like protein from brain. Biochem
Biophys Res Commun 73: 40 – 46, 1976.
453. WEI Q, XIAO F, LU J, AND ZHOU J. ESR study on calcineurin. Sci
China (Ser B) 38: 1117–1122, 1995.
454. WEILAND J, NITSCHE AM, STRAYLE J, STEINER H, AND RUDOLPH HK. The
PMR2 gene cluster encodes functionally distinct isoforms of a
putative Na⫹ pump in the yeast plasma membrane. EMBO J 14:
3870 –3882, 1995.
455. WEISS A AND LITTMAN DR. Signal transduction by lymphocyte antigen receptors. Cell 76: 263–274, 1994.
456. WHEELAN SJ, BOGUSKI MS, DURET L, AND MAKALOWSKI W. Human and
nematode orthologs: lessons from the analysis of 1800 human
genes and the proteome of Caenorhabditis elegans. Gene 238:
163–170, 1999.
457. WIDLANSKI TS, MYERS JK, STEC B, HOLTZ KM, AND KANTROWITZ ER.
The road less travelled: taming phosphatases. Chem Biol 4: 489 –
492, 1997.
458. WIEDERRECHT G, LAM E, HUNG S, MARTIN M, AND SIGAL N. The
mechanism of action of FK-506 and cyclosporin A. Ann NY Acad
Sci 696: 9 –19, 1993.
459. WILSON DK, RUDOLPH FB, AND QUIOCHO FA. Atomic structure of
adenosine deaminase complexed with a transition-state analog:
understanding catalysis and immunodeficiency mutations. Science
252: 1278 –1284, 1991.
460. WINDER DG, MANSUY IM, OSMAN M, MOALLEM TM, AND KANDEL ER.
Genetic and pharmacological evidence for a novel, intermediate
phase of long-term potentiation suppressed by calcineurin. Cell 92:
25–37, 1998.
461. WITHEE JL, MULHOLLAND J, JENG R, AND CYERT MS. An essential role
of the yeast pheromone-induced Ca2⫹ signal is to activate calcineurin. Mol Biol Cell 8: 263–277, 1997.
462. WITHEE JL, SEN R, AND CYERT MS. Ion tolerance of Saccharomyces
cerevisiae lacking the Ca2⫹/CaM-dependent phosphatase (calcineurin) is improved by mutations in URE2 or PAM1. Genetics
149: 865– 878, 1998.
463. WOLVETANG EJ, LARM JA, MOUTSOULAS P, AND LAWEN A. Apoptosis
induced by inhibitors of the plasma membrane NADH-oxidase
involves Bcl-2 and calcineurin. Cell Growth Differ 7: 1315–1325,
1996.
464. XU X, QIN X-Q, AND KANTROWITZ ER. Probing the role of histidine-372
in zinc binding and the catalytic mechanism of Escherichia coli
alkaline phosphatase by site-specific mutagenesis. Biochemistry
33: 2279 –2284, 1994.
465. YAKEL JL. Calcineurin regulation of synaptic function: from ion
channels to transmitter release and gene expression. Trends Pharmacol Sci 18: 124 –134, 1997.
466. YAMADA T, KIM JK, MURAMATSU Y, SERIKAWA T, AND MATSUMOTO K.
Chromosomal assignments of the genes for the calcineurin a␣
(Calna1) and a␤ subunits (Calna2) in the rat. Cytogenet Cell Genet
67: 55–57, 1994.
467. YE RR AND BRETSCHER A. Identification and molecular characterization of the calmodulin-binding subunit gene (CMP1) of protein
phosphatase 2B from Saccharomyces cerevisiae. Eur J Biochem
204: 713–723, 1992.
1521
468. YOSHIDA T, TODA T, AND YANAGIDA M. A calcineurin-like gene ppb1⫹
in fission yeast: mutant defects in cytokinesis, cell polarity, mating
and spindle pole body positioning. J Cell Sci 107: 1725–1735, 1994.
469. YOUN H-D, SUN L, PRYWES R, AND LIU JO. Apoptosis of T cells
mediated by Ca2⫹-induced release of the transcription factor
MEF2. Science 286: 790 –793, 1999.
470. YU L, GOLBECK J, YAO J, AND RUSNAK F. Spectroscopic and enzymatic
characterization of the active site dinuclear metal center of calcineurin: implications for a mechanistic role. Biochemistry 36:
10727–10734, 1997.
471. YU L, HADDY A, AND RUSNAK F. Evidence that calcineurin accommodates an active site binuclear metal center. J Am Chem Soc 117:
10147–10148, 1995.
472. ZAV’YALOV VP, DENESYUK A, LUNDELL J, AND KORPELA T. Some new
aspects of molecular mechanisms of cyclosporin A effect on immune response. APMIS 103: 401– 415, 1995.
473. ZHANG C-C, FRIRY A, AND PENG L. Molecular and genetic analysis of
two closely linked genes that encode, respectively, a protein phosphatase 1/2A/2B homolog and a protein kinase homolog in the
Cyanobacterium Anabaena sp. strain PCC 7120. J Bacteriol 180:
2616 –2622, 1998.
474. ZHANG J AND STEINER JP. Nitric oxide synthase, immunophilins and
poly(ADP-ribose) synthetase: novel targets for the development of
neuroprotective drugs. Neurol Res 17: 285–288, 1995.
475. ZHANG J, ZHANG Z, BREW K, AND LEE EYC. Mutational analysis of the
catalytic subunit of muscle protein phosphatase-1. Biochemistry
35: 6276 – 6282, 1996.
476. ZHANG W, KOWAL RC, RUSNAK F, SIKKINK RA, OLSON EN, AND VICTOR
RG. Failure of calcineurin inhibitors to prevent pressure-overload
left ventricular hypertrophy in rats. Circ Res 84: 722–728, 1999.
477. ZHANG W, ZIMMER G, CHEN J, LADD D, LI E, ALT FW, WIEDERRECHT G,
CRYAN J, O’NEILL EA, SEIDMAN CE, ABBAS AK, AND SEIDMAN JG. T cell
responses in calcineurin A␣-deficient mice. J Exp Med 183: 413–
420, 1996.
478. ZHANG Z, BAI G, DEANS-ZIRATTU S, BROWNER MF, AND LEE EYC.
Expression of the catalytic subunit of phosphorylase phosphatase
(protein phosphatase-1) in Escherichia coli. J Biol Chem 267:
1484 –1490, 1992.
479. ZHANG Z-Y AND VANETTEN RL. Pre-steady-state and steady-state kinetic analysis of the low molecular weight phosphotyrosyl protein
phosphatase from bovine heart. J Biol Chem 266: 1516 –1525, 1991.
480. ZHAO C, JUNG US, GARRETT-ENGELE P, ROE T, CYERT MS, AND LEVIN
DE. Temperature-induced expression of yeast FKS2 is under the
dual control of protein kinase C and calcineurin. Mol Cell Biol 18:
1013–1022, 1998.
481. ZHAO Y, TOZAWA Y, ISEKI R, MUKAI M, AND IWATA M. Calcineurin
activation protects T cells from glucocorticoid-induced apoptosis.
J Immunol 154: 6346 – 6354, 1995.
482. ZHOU J, XIANG B, WEI Q, AND LU J. Interaction of metal ions, Mn2⫹
and Ni2⫹, with calcineurin. Chinese Sci Bull 41: 1029 –1032, 1996.
483. ZHU D, CARDENAS ME, AND HEITMAN J. Myristoylation of calcineurin
B is not required for function or interaction with immunophilinimmunosuppressant complexes in the yeast Saccharomyces cerevisiae. J Biol Chem 270: 24831–24838, 1995.
484. ZHU Y AND YAKEL JL. Calcineurin modulates G protein-mediated
inhibition of N-type calcium channels in rat sympathetic neurons.
J Neurophysiol 78: 1161–1165, 1997.
485. ZHUO M, ZHANG W, SON H, MANSUY I, SOBEL RA, SEIDMAN J, AND
KANDEL ER. A selective role of calcineurin A␣ in synaptic depotentation in hippocampus. Proc Natl Acad Sci USA 96: 4650 – 4655,
1999.
486. ZHUO S, CLEMENS JC, HAKES DJ, BARFORD D, AND DIXON JE. Expression, purification, crystallization, and biochemical characterization
of a recombinant protein phosphatase. J Biol Chem 268: 17754 –
17761, 1993.
487. ZHUO S, CLEMENS JC, STONE RL, AND DIXON JE. Mutational analysis of
a Ser/Thr phosphatase. J Biol Chem 269: 26234 –26238, 1994.
Download