RENAL AUTOREGULATION IN HEALTH AND DISEASE

advertisement
Physiol Rev 95: 405–511, 2015
doi:10.1152/physrev.00042.2012
RENAL AUTOREGULATION IN HEALTH AND DISEASE
Mattias Carlström, Christopher S. Wilcox, and William J. Arendshorst
Department of Medicine, Division of Nephrology and Hypertension and Hypertension, Kidney and Vascular Research
Center, Georgetown University, Washington, District of Columbia; Department of Physiology and Pharmacology,
Karolinska Institutet, Stockholm, Sweden; and Department of Cell Biology and Physiology, UNC Kidney Center,
and McAllister Heart Institute, University of North Carolina at Chapel Hill, Chapel Hill, North Carolina
Carlström M, Wilcox CS, Arendshorst WJ. Renal Autoregulation in Health and Disease. Physiol Rev 95: 405–511, 2015; doi:10.1152/physrev.00042.2012.—Intrarenal autoregulatory mechanisms maintain renal blood flow (RBF) and glomerular filtration
rate (GFR) independent of renal perfusion pressure (RPP) over a defined range (80 –180
mmHg). Such autoregulation is mediated largely by the myogenic and the macula densatubuloglomerular feedback (MD-TGF) responses that regulate preglomerular vasomotor tone primarily
of the afferent arteriole. Differences in response times allow separation of these mechanisms in the
time and frequency domains. Mechanotransduction initiating the myogenic response requires a sensing mechanism activated by stretch of vascular smooth muscle cells (VSMCs) and coupled to intracellular signaling pathways eliciting plasma membrane depolarization and a rise in cytosolic free calcium
concentration ([Ca2⫹]i). Proposed mechanosensors include epithelial sodium channels (ENaC), integrins, and/or transient receptor potential (TRP) channels. Increased [Ca2⫹]i occurs predominantly by
Ca2⫹ influx through L-type voltage-operated Ca2⫹ channels (VOCC). Increased [Ca2⫹]i activates inositol
trisphosphate receptors (IP3R) and ryanodine receptors (RyR) to mobilize Ca2⫹ from sarcoplasmic
reticular stores. Myogenic vasoconstriction is sustained by increased Ca2⫹ sensitivity, mediated by
protein kinase C and Rho/Rho-kinase that favors a positive balance between myosin light-chain kinase
and phosphatase. Increased RPP activates MD-TGF by transducing a signal of epithelial MD salt
reabsorption to adjust afferent arteriolar vasoconstriction. A combination of vascular and tubular
mechanisms, novel to the kidney, provides for high autoregulatory efficiency that maintains RBF and
GFR, stabilizes sodium excretion, and buffers transmission of RPP to sensitive glomerular capillaries,
thereby protecting against hypertensive barotrauma. A unique aspect of the myogenic response in the
renal vasculature is modulation of its strength and speed by the MD-TGF and by a connecting tubule
glomerular feedback (CT-GF) mechanism. Reactive oxygen species and nitric oxide are modulators of
myogenic and MD-TGF mechanisms. Attenuated renal autoregulation contributes to renal damage in
many, but not all, models of renal, diabetic, and hypertensive diseases. This review provides a summary
of our current knowledge regarding underlying mechanisms enabling renal autoregulation in health and
disease and methods used for its study.
L
I.
II.
III.
IV.
V.
VI.
VII.
VIII.
INTRODUCTION
MAJOR MECHANISMS AND METHODS ...
MECHANOSENSITIVE MECHANISM...
CONDUCTANCE CHANGES...
CALCIUM SIGNALING PATHWAYS...
MODULATING AGENTS
RENAL AUTOREGULATION IN DISEASE...
CONCLUSIONS
405
408
431
436
439
444
452
471
I. INTRODUCTION
A. Overview and Historical Perspective
Arteries from various vascular beds often share functional
characteristics. However, prominent differences exist in myogenic responses to changes in transmural pressure. These responses are greater in cerebral and renal than in mesenteric
and hindlimb arteries (748, 750, 859, 886). Vascular smooth
muscle cells (VSMCs) are derived from diverse embryological
and developmental origins, and such lineage may account for
heterogeneity of specialized function (1142, 1179). Renal vessels are formed by angiogenesis and vasculogenesis (479).
VSMCs express fast and slow contractile gene programs, accounting for phasic and tonic phenotypes (1213). The aorta
and efferent arterioles are examples of the tonic phenotype,
whereas small resistance arteries and arterioles including the
renal cortical radial (interlobular) artery and afferent arteriole
are examples of the phasic phenotype. Moreover, there are
significant differences in the magnitude of vasoconstrictor responses to KCl, norepinephrine (NE), and serotonin and to
endothelium-dependent and -independent vasodilation between mouse aorta and smaller arterial segments (804).
These variations likely reflect, in part, functional adaptations to meet the diverse roles of different arterial beds.
0031-9333/15 Copyright © 2015 the American Physiological Society
405
RENAL AUTOREGULATORY MECHANISMS
Cerebral artery tone is modulated primarily by local metabolic, paracrine, and mechanical factors such as the partial
pressure of carbon dioxide. The cerebral vasculature adjusts blood flow to the local metabolic demand independent
of systemic hemodynamics. The cerebral vascular myogenic
response mediates rapid and efficient autoregulation of cerebral blood flow that maintains a steady cerebral capillary
pressure (356, 739). In contrast, mesenteric arteries are
strongly influenced by transmitter release from perivascular
sympathetic nerves. They have a major role in the control of
peripheral vascular resistance and arterial blood pressure
(BP) (410). In the splanchnic circulation, the portal vein and
hepatic artery are arranged in parallel and supply blood to
the liver for detoxification and metabolism. The specialized
pulmonary circulation is characterized by its low vascular
pressure and resistance and inherent vasoconstrictor response to hypoxia. These organ-specific regulatory actions
interact with myogenic vasoconstriction.
Semple and DeWardener (1347) quantified the efficiency of
autoregulation by an “autoregulatory index” (AI) defined
as the steady-state change in RBF factored by the sustained
change in mean RPP (1347). An AI of zero indicates perfect
autoregulation, whereas an AI close to 1.0 or above signifies
very ineffective RBF autoregulation. Low AI values (⬍0.2),
reflecting effective steady-state RBF autoregulation, have
been reported in the kidneys of anesthetized dogs (90, 269,
745, 763, 799, 953, 954, 1295, 1344, 1531), mice (583,
751, 756, 1360), rabbits (80, 385, 902, 1141, 1270, 1623),
and rats (41, 45, 46, 269, 422, 709, 790, 833, 1105, 1131,
1188, 1270) as well as in conscious rats (270, 423, 668,
1195, 1376) and dogs (69, 97, 421, 510, 511, 753–755,
760, 799, 1172, 1172–1174, 1263, 1618) in both short and
prolonged settings (42, 1241). Autoregulation in the kidneys of conscious dogs and anesthetized rats is considerably
more complete than in the mesenteric or hindlimb vascular
beds (602, 750).
The kidney is richly perfused and renal blood flow (RBF)
normally accounts for ⬃25% of cardiac output. Autoregulation is a fundamental component of renal function. It
integrates intrinsic intrarenal mechanisms that stabilize
RBF and glomerular filtration rate (GFR) during changes in
renal perfusion pressure (RPP) over a defined range. This
requires that the renal vascular resistance (RVR) changes in
proportion to the RPP. The cortical radial arteries, and
especially the afferent arterioles, are the major preglomerular resistance vessels whose tone mediates most of pressureinduced autoregulation of RBF and GFR. The efferent arterioles usually do not participate in RBF autoregulation,
although they may contribute to autoregulation of GFR at
low RPP under certain conditions such as low-salt diet and
activation of the renin-angiotensin system (RAS) with high
angiotensin II (ANG II) levels. Renal autoregulation is
achieved primarily by a unique orchestrated action of two
major mechanisms: the myogenic response and the macula
densa tubuloglomerular feedback (MD-TGF) response. Together they adjust the tone of the preglomerular VSMCs to
buffer the challenge of a constantly changing RPP, whether
spontaneous or induced. The fast myogenic response relates
the tone of the afferent arterioles to their intraluminal pressure by intrinsic adjustments in the tension of VSMCs. The
more delayed MD-TGF response is activated by tubular
electrolyte delivery and reabsorption primarily by a luminal
Na⫹/K⫹/2Cl⫺ type 2 (NKCC2) cotransporter and a
Na⫹/H⫹ exchanger on the MD cells at the end of the thick
ascending limb (TAL) of Henle’s loop to elaborate positive
and negative modulators of vasoconstriction of the adjacent
afferent arteriole. The ensuing collective response entails
extensive, bidirectional, and time-dependent interactions
between these two mechanisms. Renal autoregulation is
intrinsic to the kidney and thus largely independent of extrinsic nerves or circulating hormones (428, 998, 1053,
1344, 1345, 1360, 1465, 1578).
The myogenic phenomenon was first described for large
conduit arteries by Bayliss et al. (76) more than a century
ago as part of his investigation of reactive hyperemia in the
kidney. Renal autoregulation to acute changes in RPP was
characterized originally in rabbits by Forster and Maes in
1947 (428) and in dogs by Selkurt et al. in 1949 (1345). The
renal myogenic response was initially described by Gilmore
et al. (473) in 1980 in hamster preglomerular arterioles. The
unique renal MD-TGF occurs within the juxtaglomerular
apparatus (JGA) that is composed of the MD cells at the
junction of the TAL of the loop of Henle and the distal
tubule, the granular renin-containing cells in the distal afferent arteriole and the extraglomerular mesangium that is
interspersed between, or surrounds, these cells. The anatomy of different cell types in the JGA is shown in FIGURE 1.
The concept of the JGA coordinating tubular and glomerular functions was introduced originally by Goormaghtigh
in 1937 (487), with subsequent hypotheses by Guyton,
Harsing, and Thurau in the early 1960s (533, 581, 1465,
1466). The specific contribution of MD-TGF to renal autoregulation of GFR was demonstrated initially by the laboratories of Navar and Schnermann in the 1970s (1022,
1064, 1070, 1190, 1191, 1317).
406
A myogenic response is observed in the renal, cerebral, coronary, pulmonary, mesenteric, and skeletal muscle vasculature (300 –302, 703, 739, 1249). This response limits vascular wall strain by reciprocal adjustments in vessel radius
during changes in the transmural pressure gradient. This
assumes that the wall thickness remains the same, as described by the LaPlace relationship. Mechanotransduction
in VSMCs to increased PP is intrinsic to VSMC and entails
stretch or distortion of the plasma membrane coupled to
mechanisms that decrease membrane conductance, thereby
depolarizing the resting membrane potential (EM) from approximately ⫺40 mV to activate voltage-operated Ca2⫹
channels (VOCCs). The ensuing increase in Ca2⫹ entry el-
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
PSF
Grease
block
Efferent
arteriole
Glomerulus
PGC
Granular renincontaining cells
Afferent
arteriole
Blood
Connecting
tubule
Extraglomerular
mesangial cells
Macula
densa
cells
Distal
tubule
Proximal
convoluted
tubule
Thick
ascending
limb
FIGURE 1. Schematic diagram illustrating the anatomy of the juxtaglomerular apparatus (JGA) showing cell
types involved in renal autoregulation, i.e., myogenic response and macula densa (MD)-mediated tubuloglomerular feedback (MD-TGF). MD cells are located at the junction of the thick ascending limb of Henle’s loop and
the distal convoluted tubule. Their basolateral membrane contacts some extraglomerular mesangial cells,
which in turn are contiguous with vascular smooth muscle cells and renin-containing juxtaglomerular granular
cells at the end of the afferent arteriole. Endothelial nitric oxide synthase (eNOS) is expressed primarily in the
endothelium of the afferent arteriole and neuronal NOS (nNOS) in the MD cells. MD-TGF can be assessed
conveniently by measuring proximal stop-flow pressure (PSF) upstream of a grease block in the nephron as an
index of glomerular capillary pressure (PGC) during variations in fluid delivery to MD cells by perfusion from
another micropipette (not shown) placed downstream from the block. Recent in vitro and in vivo PSF measurements have demonstrated functional implications of a connecting tubule glomerular feedback (CT-GF) in
regulating afferent arteriolar vasomotor tone.
evates cytosolic Ca2⫹ concentration ([Ca2⫹]i) that subsequently mobilizes Ca2⫹ stored in the sarcoplasmic reticulum (a process termed Ca2⫹-induced Ca2⫹ release, CICR),
combined with signaling pathways that enhance the sensitivity of myosin to a given [Ca2⫹]i (i.e., Ca2⫹ sensitivity).
These coordinated actions activate myosin light chain kinase (MLCK) and/or inhibit myosin light chain phosphatase (MLCP) to promote contraction of VSMCs (1328,
1392). This sequence of events is summarized in FIGURE 8.
The relative importance of Ca2⫹ influx, release, and sensitization varies considerably among published studies.
The MD-TGF is sensitive to renal metabolism. It relates the
GFR to the NaCl delivery to and reabsorption by MD cells.
MD NaCl delivery, in turn, is related to the filtered load of
NaCl and reabsorption in nephron segments upstream of
the MD, all of which therefore can affect MD-TGF. Accordingly, MD-TGF integrates tubular and vascular function
and is thus a unique renal mechanism of autoregulation.
Activated MD cells generate signaling molecule(s) such as
ATP and/or its breakdown product adenosine that traverse
the interstitial space and extraglomerular mesangial cells,
likely involving connexins (Cxs), to stimulate Ca2⫹ signaling and contract the afferent arteriole. The ATP and aden-
osine responses are primarily mediated by purinergic P2X
receptors and adenosine A1 receptors, respectively, on
VSMC and possibly mesangial cells.
Mesangial cells may contribute to MD-TGF by providing
an extraglomerular conduit pathway, but these cells lack
the expression of contractile ␣-smooth muscle actin in vivo
that is present in vitro in cell culture, and therefore probably
are not contractile in life (341, 378, 796). Because of difficulty in isolating fresh mesangial cells from glomeruli, they
are usually studied in culture to obtain a homogeneous
population (3, 607, 1615). However, it is recognized that
phenotypic changes take place as a function of passage
number, for example, in the relative amounts of contractile
proteins. VSMC in culture can undergo phenotypic switching from a nondividing contractile phenotype to a proliferative dedifferentiated phenotype characterized by reduced
mRNA and protein levels of VSMC markers (1142). Thus
results for contraction mechanisms using cultured mesangial cells should be interpreted with caution.
Although the general operation of MD-TGF in renal autoregulation is appreciated, the understanding of its fine details is still evolving. It is generally accepted that an increase
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
407
RENAL AUTOREGULATORY MECHANISMS
in RPP initially increases single-nephron glomerular filtration rate (SNGFR) and might reduce proximal tubule reabsorption, thereby increasing NaCl delivery from Henle’s
loop to the MD (FIGURE 1). Increased NaCl reabsorption by
MD cells generates a signal that mediates contraction of the
afferent arteriole. Thus MD-TGF functions as a negativefeedback circuit that increases preglomerular vascular resistance to restrain single-nephron glomerular blood flow
(SNGBF) and glomerular capillary hydraulic pressure
(PGC), thereby stabilizing GFR and RBF during changes in
RPP. The MD-TGF system is sensitive to factors that regulate tubular reabsorption. Therefore, MD-TGF relates renal
autoregulation to extracellular fluid volume (ECV) homeostasis. MD-TGF is sensitive to tubular metabolism because
it couples epithelial cell salt transport to preglomerular vascular tone utilizing metabolic intermediates such as ATP
and adenosine as well as other tubular and vascular metabolites or signaling molecules such as nitric oxide (NO), superoxide anion (O2⫺), prostaglandin endoperoxide (PGH2),
thromboxane A2 (TxA2), carbon monoxide (CO), and
prostaglandin E2 (PGE2) (FIGURE 6). Ultimately, both myogenic and MD-TGF mechanisms impact regulation of RBF
and GFR, primarily by varying the degree of afferent arteriolar vasoconstriction (51, 703, 1436). Their individual
contributions, as well as their interactions, determine the
overall autoregulatory efficacy.
This review focuses on mechanisms responsible for renal
autoregulation in health and its role and regulation in several models of hypertension, chronic kidney disease (CKD)
and diabetes mellitus (DM). GFR normally is autoregulated
with RBF since both respond in parallel to induced changes
in preglomerular resistance. However, GFR is determined
not only by PGC but also by SNGBF and by the capillary
ultrafiltration coefficient (KUF). Selective changes in efferent
arteriolar resistance, for example, during inhibition of the
RAS at low RPP, can dissociate autoregulation of GFR from
RBF (550, 1261). This review concentrates on the changes
in preglomerular vascular resistance responsible for regulation of RBF and glomerular perfusion. The reader is referred to recent reviews of renal autoregulation (286, 748,
933, 934) and of mathematical modeling of renal autoregulatory mechanisms (179, 228, 395, 541, 628, 805, 877,
973, 1023, 1350, 1614).
B. Functional Significance
Renal autoregulation subserves four important functions.
First, it sets a basal level of vasomotor tone upon which
neurohumoral agents, paracrine/autocrine substances, and
metabolic factors adjust RVR (48, 1065, 1071). Second, it
buffers the transmission of RPP to the sensitive glomerular
capillaries and parenchyma, thereby preventing intrarenal
barotrauma during systemic arterial hypertension (108).
Impaired renal autoregulation contributes to progressive
pressure-sensitive glomerular and interstitial injury in many
408
models of hypertension, CKD, and DM as described in
detail in section VII. Third, it isolates the subtle tubular
transport processes and sodium (Na⫹) excretion from
abrupt changes in GFR that would induce large fluctuations
in tubular fluid flow and peritubular physical forces (532).
Fourth, a somewhat less complete regulation of medullary
blood flow during changes in RPP has been proposed as a
mechanism mediating pressure-natriuresis and may be important in maintaining salt and water balance and controlling BP, as discussed in section IIA2.
Unfortunately, there are only limited studies of renal autoregulation in normal human subjects or those with hypertension. A study of African-American hypertensive patients
reported higher values for GFR and RBF than matched
Caucasians (FIGURE 2, A–C) (826). A pressor infusion of NE
increased the GFR only in the African-Americans, which led
the authors to propose that African-American hypertensives had impaired renal autoregulation, although the RBF
was maintained independent of RPP in both groups. In
another study of hypertensive patients, the systolic BP was
reduced incrementally by infusion of the vasodilator sodium nitroprusside (FIGURE 2, D AND E) that releases NO
(20). The GFR and effective renal plasma flow were wellmaintained across the range of systolic BPs of 170 to 130
mmHg in patients with moderate hypertension, but both
fell progressively in the range of 200 –135 mmHg in those
with severe hypertension. This demonstrated preservation
of renal autoregulation in patients with moderate hypertension, but a breakdown in autoregulation in those with severe hypertension. Thus renal autoregulation can provide a
remarkably stable GFR and RBF during changes in RPP in
human subjects with hypertension. The observation of impaired autoregulation in hypertensive African-Americans
and those with severe hypertension is consistent with the
hypothesis that renal autoregulation normally functions to
protect the kidney from barotrauma. Its failure in these two
populations increases risk for the development of nephrosclerosis and hypertensive CKD.
II. MAJOR MECHANISMS AND METHODS
USED TO STUDY RENAL
AUTOREGULATION
Autoregulation has been characterized in different preparations of the renal circulation ranging from in vivo
measurements of RPP-induced changes in whole kidney
RBF, GFR, and single-nephron hemodynamics during induced changes in RPP to direct assessment of pressureinduced contraction of isolated, perfused afferent arterioles, or flow-induced MD-TGF responses in isolated
JGAs. Examples of autoregulatory myogenic responses
from six commonly used preparations are described below and are highlighted in FIGURE 3, A–F.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
160
150
*
140
130
White Black
ERBF
1,500
**
1,400
1,300
1,200
1,100
1,000
900
C
GFR
200
P<0.001
180
160
140
*
120
100
White Black
White Black
D
GFR
150
130
110
90
70
**
50
130 140 150 160 170 180 190 200
E
CHipurran (ml · min-1 ·1.73 m-2)
P<0.001
B
CCr (ml · min-1 ·1.73 m-2)
SBP (mmHg)
170
P<0.001
GFR (ml · min-1 ·1.73 m-2)
BP
180
ERBF (ml · min-1 ·1.73 m-2)
A
ERPF
450
400
350
300
250
**
200
150
130 140 150 160 170 180 190 200
FIGURE 2. Studies of renal autoregulation in hypertensive human subjects. Blood pressure (BP) was
determined at regular intervals, and the glomerular filtration rate (GFR) was measured from the clearance of
inulin or creatinine and the effective renal plasma flow (ERPF) from the clearance of hippuran and used to
calculate the effective renal blood flow (ERBF). A–C: data were compared in 33 Caucasian (white) and 29
African-Americans (black) matched patients with essential hypertension in the basal state (open boxes) and
during infusion of norepinephrine (0.05 ␮g·kg⫺1·min⫺1 for 30 min) to increase BP (closed boxes). The ERBF
was higher in African-American than in Caucasian subjects but did not change during a short-term elevation of
BP by norepinephrine in either patient group. However, the GFR also was higher in the African-Americans and
increased further with the rise in BP, implying a selective defect in GFR autoregulation. Compared with
Caucasians in the basal state: *P ⬍ 0.05; **P ⬍ 0.01. [Data from Kotchen et al. (826).] D and E: data were
compared in patients with moderate hypertension (average: 170/108 mmHg; n ⫽ 10, open circles) and
severe hypertension (average: 198/127 mmHg; n ⫽ 10; closed circles), in the basal state and during graded
infusion of sodium nitroprusside (0.2–12 ␮g·kg⫺1·min⫺1) to provide incremental reductions in systolic BP
(SBP). After acute BP normalization, the GFR and the ERPF were reduced significantly in patients with severe
hypertension (⫺44.7 ⫾ 6.8 and ⫺41.6 ⫾ 8.3%, respectively), but did not change in those with moderate
hypertension, implying that renal autoregulation was preserved in subjects with essential hypertension, but was
impaired or absent in those with severe hypertension. **P ⬍ 0.01 vs. recordings in the basal state with high
BP. [Data from Almeida et al. (20).]
A. In Vivo Studies Using Whole Kidney
Preparations
1. Steady-state autoregulation
A) STEADY-STATE AUTOREGULATION OF WHOLE KIDNEY BLOOD
FLOW.
The BP in animal studies can be increased by carotid
arterial occlusion to elicit baroreceptor activation of the
sympathetic nervous system. Graded reductions of the elevated RPP by constriction of the suprarenal aorta for 1–2
min can provide a range of RPPs above, at, and below the
ambient RPP (FIGURE 3A) (41, 46). Ideally, this should be
achieved by coupling an aortic or renal artery constrictor to
a servo-controlled pressure sensor to provide more stable
levels of RPP (46, 376, 549, 610). The kidneys should be
denervated to minimize the influence of changes in renal
sympathetic nerve activity. Alternatively, the RPP can be
increased by constriction of the abdominal aorta below the
renal vessels. This can be combined with occlusion of the
superior mesenteric and celiac arteries to increase the RPP,
with a further rise in PP achieved by infusion of a cocktail of
vasoconstrictors such as NE and arginine vasopressin
(AVP) (1247). A limitation of this method is that the infused
hormones themselves affect renal vascular tone and may
modulate autoregulation. Another approach has been to
study an isolated, pump-perfused kidney of dogs (1106,
1108, 1110, 1112, 1120) or rats (1285, 1323). The oxygenated perfusate usually contains albumin or a colloid, but
has a low hematocrit to minimize erythrocyte damage by
the perfusion pump. The reduced red cell mass limits the
vessel wall shear stress and NO release, but this is offset by
some degree by less oxyhemoglobin-mediated oxidation of
NO, whose influence may thereby become more pronounced. Perfusion is usually at a fixed mean RPP rather
than a pulsatile pressure. Limitations are that the isolated
kidneys become edematous, fail to reabsorb salt normally,
and have limited autoregulatory efficiency (870, 871, 1332,
1468).
B) STEADY-STATE AUTOREGULATION IN DIFFERENT REGIONS OF THE
KIDNEY. Most current studies of renal cortical blood flow use
a fiberoptic probe connected to a laser Doppler flowmeter
to estimate the erythrocyte (RBC) velocity in a small tissue
area (⬃1 mm3). Highly efficient autoregulation was reported in the outer-, mid- and deep-cortical regions of rat
kidneys during hydropenia (990, 1195, 1248, 1252) and
during acute expansion of ECV (990, 1248). Autoregulation of cortical blood flow generally parallels that of the
whole kidney.
Measurement of medullary blood flow is considerably more
complex (283, 1256, 1423). For example, measurements of
average RBC velocity by Doppler methods applied to the
medulla do not differentiate between descending and ascending flow in vasa recta capillaries. Volume flux depends
on the number of vessels being perfused and the fluid reabsorbed by the tubules. This may explain in part why the
efficiency of medullary autoregulation has been more vari-
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
409
RENAL AUTOREGULATORY MECHANISMS
able among studies and species and in some depends on the
state of hydration.
Both cortical and medullary blood flow in nondiuretic rats
are autoregulated efficiently between RPP of 90 –160
mmHg (267, 288, 990, 1248, 1417).
An early study that measured the RBC velocity in individual
vessels on the surface of the renal papilla reported excellent
autoregulation during hydropenia and volume expansion
(267). However, other studies using similar methodology
110
RBF (% basal)
100
Rat dynamic autoregulation: time domain
RVR
(mmHg)
B
Rat steady state autoregulation
120
RVR
(% of perfect AR)
A
reported weak autoregulation of RBC velocity in descending and ascending vasa recta of euvolemic rats, with AI
values of 0.5– 0.8 (391). Many subsequent studies in rats
that have used the laser Doppler technique to measure RBC
velocity in multiple capillaries in a fixed volume of tissue
reported highly efficient autoregulation of blood flow in the
outer and inner portions of the renal medulla in hydropenia
(990), but this became substantially less efficient during
ECV expansion (277, 411, 717, 990, 1243, 1251, 1252,
1254, 1257, 1476). However, another study reported only
moderate medullary autoregulation (AI ⫽ 0.4) in hydro-
100
105
AI < 0.1
90
80
70
AI < 0.05
60
50
40
45
55
65
75
85
95
50
TGF: ~35% of total
0
MR: ~50% of total
-50
0
105 115 125 135 145
60
80
100
120
Time after release (sec)
Rat dynamic autoregulation: frequency domain
D
Rat juxtamedullary preparation
140
2.0
AI = 0.32
MR - 65%
CCB
120
Diameter
(% of control)
Admittance gain
40
20
Renal PP (mmHg)
C
3rd
90
1.0
TGF - 35%
0.5
MR - 65%
100
TGF block
TGF - 35%
80
Control
60
0.25
0.01
0.1
1
60
E
100
80
120
140
Renal PP (mmHg)
Frequency (Hz)
Rat hydronephrotic kidney
F
Mouse isolated perfused afferent arteriole
0
-5
Active wall tensions
(dynes/cm)
Afferent arteriolar diameter
(% change)
400
-10
-15
-20
300
200
100
-25
0
80
100
120
140
160
Renal PP (mmHg)
410
180
40
60
80
100
Arteriolar PP (mmHg)
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
120
140
CARLSTRÖM ET AL.
penic rats with little change (AI ⫽ 0.5) during ECV expansion (1248). An important study concluded that ECV expansion increases renal medullary blood flow by increasing
the number of capillaries perfused with moving RBCs without a change in their velocity in either ascending or descending vasa recta (415). Further evidence of incomplete medullary blood flow autoregulation came from the observation that the hydrostatic pressure in the vasa recta
capillaries was autoregulated less completely than in the
cortical peritubular capillaries (1248).
The renal medulla is perfused from cortical efferent arterioles. Therefore, the finding that medullary perfusion in the
rat can be poorly autoregulated, especially during ECV expansion, while deep cortical blood flow is well autoregulated located the postglomerular vasculature as the site of
impaired control of medullary blood flow (990).
Highly efficient autoregulation of both cortical and medullary blood flow that is independent of the ECV is observed
in whole kidney studies in the dog (950, 951, 954, 960, 961)
and in the rabbit (381). On the other hand, one study reports that blood flow to the exposed papilla of the dog was
poorly autoregulated (AI ⫽ 0.82) compared with whole
kidney RBF (AI ⫽ 0.33) (1423), but this may have been a
consequence of the impaired function of the exposed papilla.
In summary, in contrast to cortical blood flow, the degree of
autoregulation of medullary blood flow in the rat is controversial. Although most studies have reported efficient autoregulation of medullary blood flow during hydropenia, this
appears to become less efficient during ECV expansion, at
least in the rat (277, 990, 1243, 1251, 1252, 1254, 1257).
However, most studies in dogs and rabbits report efficient
medullary autoregulation during both hydropenia and ECV
expansion. Autoregulatory responses of the preglomerular
vasculature can be assessed directly in the in vitro rat juxtamedullary nephrovascular (JMN) preparation. As discussed in section IIB3, these studies generally have reported
efficient autoregulation, involving both myogenic and MDTGF components.
2. Pressure-natriuresis
The reports of highly efficient overall renal autoregulation,
and often also of intrarenal distribution of perfusion in the
dog, rabbit, and nondiuretic rat, pose a conundrum for the
understanding of how the kidney regulates NaCl and fluid
excretion to maintain long-term BP homeostasis. Renal
Na⫹ excretion changes acutely with RPP, even in studies in
which the intrarenal mechanisms maintain a stable blood
flow in the renal cortex and medulla. Further uncertainty
derives from the recognition that steady-state measurements of proximal reabsorption and peritubular capillary
fluid uptake do not permit firm conclusions regarding causality (1490). Thus the rate of reabsorption must correlate
with the effective reabsorptive pressure determined by Starling physical forces and the peritubular capillary reabsorptive coefficient. Moreover, the events that initiate the response may be transient yet the GFR and the filtered Na⫹
load remain constant during changes in PP. This indicates
that the changes in Na⫹ excretion must be due to variations
in tubular Na⫹ reabsorption. The key question is: What
intrarenal mechanism(s) can mediate pressure-natriuresis
when the pressures in the tubules and vasculature downstream of the afferent arteriole remain stable because of
highly efficient autoregulation?
A step change in RPP usually yields almost no change in
steady-state whole kidney RBF, GFR, efferent arteriolar
blood flow (956, 1636), or intrarenal hydrostatic pressures
FIGURE 3. Data redrawn from 6 published studies to illustrate different methods for assessing renal autoregulation or the contributions of
myogenic and macula densa-tubuloglomerular feedback (MD-TGF) mechanisms. A: measurements of renal blood flow (RBF) by electromagnetic
flow meter recordings in anesthetized rats as a function of induced changes in renal perfusion pressure (PP). The autoregulatory index (AI)
⬍0.05 implies almost perfect independence of RBF from renal PP across the range 105–145 mmHg (41, 46). B: study of dynamic renal
autoregulation in the time domain in anesthetized rats (749). The renal PP had been reduced previously by a suprarenal aortic clamp to ⬃95
mmHg. Abrupt release of the aortic clamp led to a rapid increase in renal PP (RPP) to ⬃110 mmHg, which was maintained throughout the 2
min. The renal vascular resistance (RVR) fell initially, but this was followed rapidly by a 3-component rise. Over the first 5 s, the increase restored
⬃50% of the final RVR response to the level of perfect autoregulation. This was attributed to the myogenic response (MR) and was followed over
10 –20 s by a further rise in RVR restoring ⬃35% of the RVR. This was attributed to macula densa-tubuloglomerular feedback (MD-TGF)
response and was followed by a third, slower component contributing ⬃15% of the final RVR response. C: study of dynamic renal autoregulation
in the frequency domain by transfer function analysis in rats of measurements of spontaneous changes of renal PP, measured by a pressure
transducer in the aorta, and RBF measured by ultrasonic blood flow meter (297). The tracing shows normalized admittance gain (a relative
autoregulatory index) as a function of frequency of spontaneous fluctuations in renal PP. Values at or above unity imply very weak to no active
autoregulation. Between 0.3 and 0.1 Hz oscillations in renal PP, the admittance gain fell steeply, corresponding to the myogenic response (MR).
There was a further sharp fall in admittance gain between 0.03 and 0.01 Hz, corresponding to the delayed MD-TGF mechanism. D:
measurements of the steady-state diameter of the afferent arteriole of the rat juxtamedullary nephron (JMN) preparation during induced
changes in renal PP (1020). An increase in renal PP, from 60 to 140 mmHg, reduced arteriolar diameter by ⬃25–30% (control). A progressive
autoregulatory reduction in diameter during increased renal PP was attenuated by blockade of MD-TGF (by furosemide) and was totally abolished
by the CCB nimodipine that eliminated all active tone. E: directly visualized changes in diameter of the rat afferent arteriole in the hydronephrotic
kidney (HNK) preparation lacking MD-TGF (586). An increase in renal PP from 80 to 180 mmHg reduced the arteriolar diameter (i.e., myogenic
response). F: development of active wall tension as a function of arteriolar PP in a mouse isolated perfused afferent arteriole (861). The slope
of this line defined the myogenic response over the PP range tested, and its intercept on the x-axis defined the threshold.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
411
RENAL AUTOREGULATORY MECHANISMS
(42, 276, 1167, 1188, 1241). Spontaneously hypertensive
rats (SHR) had excellent RBF and GFR autoregulation (41,
43, 465, 672, 706, 1247) and stable mean PGC and peritubular capillary pressures in their superficial cortical
nephrons, which were attributed to proportionate adjustments in preglomerular vascular resistance with RPP (42,
58, 394, 1064, 1247, 1564, 1567). Nevertheless, these SHR
displayed a pressure-natriuresis, albeit dampened compared with normotensive Wistar-Kyoto (WKY) rats that
also autoregulated efficiently (42, 276, 412, 808). Early
micropuncture studies reported that steady-state proximal
tubular reabsorption and fluid delivery to the distal tubule
of superficial cortical nephrons were maintained during an
acute change in RPP, implying minimal change in input
signal to modulate MD-TGF (42, 43, 342, 535, 901, 1067)
and minimal change in fractional Na⫹ delivery to the late
distal tubule in the superficial cortex (844). One in vivo
microperfusion study in rats reported that ⬃25 mmHg increase in RPP inhibited proximal Na⫹ reabsorption by almost 40%, but a similar decrease in RPP did not affect Na⫹
reabsorption (798). However, previous free-flow micropuncture studies from the same laboratory but in a different strain of rat reported unchanged proximal tubular
Na⫹ reabsorption of superficial nephrons, but decreased
reabsorption of deep nephrons, in response to increased
RPP (535).
Nevertheless, other investigators have reported that increased RPP inhibited Na⫹ reabsorption by the proximal
tubule of surface nephrons (248, 1245). Marsh, Roman,
and colleagues reported that increased RPP reduced proximal tubular reabsorption rapidly and increased fluid delivery to Henle’s loop despite highly efficient autoregulation of
RBF and GFR (249, 1248). Reduced absolute and fractional proximal tubular reabsorption after increased RPP
has been confirmed in other studies in rats, but autoregulatory efficiency was not verified (814, 995, 996).
A more consistent finding has been that increased RPP inhibits Na⫹ reabsorption by proximal tubules of deep
nephrons (to a greater extent than superficial nephrons)
(535, 1245). This likely reflects pressure-induced reductions in Na⫹ and fluid reabsorption in straight descending
portions of the late proximal tubule.
In studies in which an increased PP inhibited proximal tubular reabsorption, the anticipated increase in Cl⫺ reabsorption by Henle’s loop in response to the increased tubular fluid delivery was lacking (248, 249, 1245). Studies of
fractional Li⫹ excretion as an index of fractional proximal
Na⫹ reabsorption concluded that increased RPP reduces
fractional Na⫹ reabsorption in the proximal (536, 736) and
distal tubules (736). Increased RPP also reduced Na⫹ reabsorption by the nephron segments distal to Henle’s loop,
since pharmacological blockade of distal tubular Na⫹ reabsorption markedly attenuated the pressure-natriuresis re-
412
sponse (956). It is uncertain whether the medullary collecting duct (CD) also participates in these changes in reabsorption with PP (1245, 1394). Thus changes in Na⫹
reabsorption by Henle’s loop and the distal nephron can
participate in pressure-natriuresis.
The preglomerular vasculature of the juxtamedullary cortex maintained an appropriate myogenic and MD-TGF response that stabilized SNGFR in deep nephrons during increases in RPP (535, 1248). Similarly, the afferent arterioles
of the JMN preparation responded appropriately to
changes in PP (see below).
Nevertheless, the hydrostatic pressure in vasa recta capillaries and the cortical renal interstitial hydrostatic pressure
(RIHP), which in turn is thought to reduce tubular reabsorption, both increase with PP. An increase in RPP increased RIHP by only a few mmHg in nondiuretic rats, but
to a larger extent after volume expansion (459, 486, 524,
790, 792). Indeed, both the cortical and medullary RIHP
increased after volume expansion. Decapsulation of the
kidney reduced the changes in RIHP with PP and attenuated
pressure-natriuresis (459, 537, 791, 792). Changes in RIHP
independent of a change in RPP are commonly observed,
but may be more associative than causal since changes in
hydrostatic pressure are less influential on transepithelial
transport than changes in osmotic pressure (493, 539).
Since the efferent arterioles from the deep cortical nephrons
supply the medullary vasa recta capillaries, any attenuation
of autoregulation by the preglomerular vasculature of JMN
would likely transmit some of the changed RPP into the
vasa recta capillaries and the medullary interstitium. Consistent with this prediction, the hydrostatic pressure in vasa
recta capillaries was autoregulated less completely than in
renal cortical capillaries (1248).
Three factors have been identified that could increase the
medullary RIHP during increased PP in diuretic states. First,
the descending vasa recta provide a fourfold greater flow
resistance than the ascending vasa recta perhaps because of
the fewer descending capillaries that results in a markedly
lesser vascular cross-sectional area (71, 1149, 1643). Second, the hydrostatic pressure in the medulla is determined
primarily by the outflow resistance of the CDs (972). Third,
the descending vasa recta are surrounded by contractile
pericytes that can generate a myogenic response (908).
Large-diameter human outer medullary descending vasa
recta contracted to increased luminal pressure, whereas
smaller vasa recta did not (1348).
The medullary RIHP should determine the cortical RIHP
since the kidney is an encapsulated organ (277, 790, 1140).
Thus this sequence of events provides a mechanism
whereby cortical RIHP and proximal fluid reabsorption
could vary with RPP despite no change in PGC. Indeed, Na⫹
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
reabsorption was closely related to even very small changes
in cortical RIHP (1607).
Paracrine agents including NO, prostaglandins (PGs), 20HETE, and ATP have been proposed to modulate pressurenatriuresis. The slope of the natriuretic response to increased RPP is attenuated by inhibition of NOS (385, 411,
411, 429, 524, 666, 736, 946, 947, 955, 957, 959, 962,
963, 1277, 1476), or COX (190, 486, 536, 538, 797, 1074,
1258), or cytochrome P-450 4A to reduce 20-HETE generation (340, 1612). Moreover, renal interstitial ATP concentration paralleled RPP and regulated both MD-TGF (1090,
1091) and Na⫹ transport via epithelial sodium channels
(ENaC) in the CDs (1475).
Activation of AT1 receptors can inhibit pressure-natriuresis
(317, 321, 513, 515, 547, 569, 946, 995–997, 1074, 1258,
1516, 1549), whereas activation of AT2 receptors enhanced
pressure-natriuresis in some (514, 516) but not all studies
(911, 918). Inhibition of ACE-dependent ANG II generation enhanced pressure natriuresis (1074), whereas chronic
ANG II infusion increased BP and suppressed pressure natriuresis (1549). Moreover, clamping systemic ANG II levels during acute hypertension blunted the magnitude of the
pressure-natriuresis response (884). Increased RAS activity,
AVP (433), renal sympathetic nerve activity (RSNA) (328,
337, 376, 775, 806, 1130, 1382), and O2⫺ or hydrogen
peroxide (H2O2), especially in the renal medulla (273, 734,
1026, 1099), may also inhibit pressure natriuresis.
Despite this evidence of the importance of weak autoregulation of the medullary blood flow during volume expansion for initiating pressure-natriuresis, there remain two
divergent schools of thought concerning the major cause of
pressure-natriuresis in an autoregulating kidney with an
unchanged overall filtered load of Na⫹ and increased RIHP.
The major difference between them concerns the mediating
factor(s) and the relative importance of a primary change in
renal cortical NO generation or in medullary blood flow
(FIGURE 4, A AND B).
Majid and Navar demonstrated strong correlations between intrarenal levels of NO, RIHP, and pressure-natriuresis in dog kidneys exhibiting highly efficient autoregulation of cortical and medullary blood flow (FIGURE 4A) (955,
957–959, 961, 963). During changes in RPP, the Na⫹ excretion, renal cortical and medullary NO production, and
RIHP changed in parallel. Inhibition of renal NOS suppressed pressure-natriuresis in autoregulating kidneys without changing the filtered Na⫹ load (537). However, although administration of a NO donor during NOS inhibition increased basal Na⫹ excretion and RIHP, it failed to
normalize the slope of the pressure-natriuresis or RIHP relationships. This suggests that a combination of pressuredependent changes in renal NO production and RIHP are
required to mediate pressure-natriuresis.
A
Renal PP
Excellent corticol and medullary autoregulation
Vessel wall
shear stress
Arteriolar
vasoconstriction
Vascular
NO production
Interstitial
hydrostatic pressure
No change in
corticol or
medullary blood
flow or GFR
Tubular Na+ reabsorption
Renal Na+ excretion
B
Renal PP
Incomplete
medullary autoregulation
Excellent corticol
autoregulation
Medullary and
papillary blood flow
No change in corticol
blood flow or GFR
Medullary interstitial
hydrostatic pressure
Corticol interstitial
hydrostatic pressure
Na+ reabsorption
in the proximal
tubule
Renal Na+
excretion
FIGURE 4. Flow diagrams summarizing current hypotheses for
interactions between renal autoregulation and pressure natriuresis
in response to increased renal perfusion pressure (PP) with excellent autoregulation of cortical blood flow with and without efficient
autoregulation of medullary blood flow. A: proposed intrarenal
events causing pressure-natriuresis when both renal cortical and
medullary blood flows are excellently autoregulated. B: proposed
intrarenal events causing pressure-natriuresis when renal cortical
blood flow is excellently autoregulated but autoregulation of medullary blood flow is less complete. GFR, glomerular filtration rate; RBF,
renal blood flow; NO, nitric oxide; ROS, reactive oxygen species. For
details, see text (sect. IIA2).
Endothelial NO is released in response to increased vessel
wall shear stress. The constriction of the renal resistance
vessels during increased PP would increase wall stress without changing blood flow (171, 747, 1039, 1184, 1416,
1425) and could thereby provide a signal for NO release in
fully autoregulating kidneys. NO inhibits Na⫹ reabsorption in the TAL and the CDs (379, 460, 608, 1185, 1419),
whereas the actions in the proximal tubule are more controversial, with reports of NO inhibiting (1419) or stimu-
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
413
RENAL AUTOREGULATORY MECHANISMS
lating (1560) Na⫹ reabsorption. However, studies in mice
with deletion of specific NOS isoforms concluded that NO
derived from nNOS or eNOS stimulated Na⫹ reabsorption and H⫹ secretion by the proximal tubule (1561).
Thus it is unlikely that the proximal tubule is the primary
site of NO-mediated inhibition of tubular reabsorption
during increased PP. As mentioned previously, blockade
of distal nephron Na⫹ transport markedly attenuated
pressure-natriuresis (956), demonstrating the importance of active transport inhibition in mediating this response. The TAL and the CD are rich sources of NO and
sites at which NO reduced Na⫹ reabsorption (1129,
1148, 1304, 1627). Indeed, medullary NO preferentially
inhibited Na⫹ reabsorption in the CDs (1304), which
participated in inhibition of fractional NaCl reabsorption during increased PP (1394).
Carey and colleagues have proposed a modified view for the
participation of NO in pressure natriuresis via effects on the
generation of cGMP and activation of PKG. Increased NOcGMP-PKG signaling modulated pressure-natriuresis during elevated PP by increasing RIHP and inhibiting proximal
tubular Na⫹ reabsorption independent of changes in GFR
or renal cortical or medullary blood flow (736, 737, 900).
Renal cortical interstitial cGMP accumulated in proximal
tubule cells where it activated PKG, inhibited tubular Na⫹
reabsorption, thereby increasing Na⫹ excretion independent of measurable hemodynamic changes (736, 737,
1153). This was regional specific since infusion of cGMP
into the renal cortex produced a natriuresis, whereas infusion into the renal medulla was ineffective (737).
Knox and colleagues presented evidence that increased RPP
increases prostaglandin E2 (PGE2) production by COX-1.
The PGE2 was believed to derive from endothelial cells of
the preglomerular vasculature and, during increased RIHP,
to inhibit proximal tubular Na⫹ reabsorption (486, 509,
536, 538, 790, 797, 823, 1207). The mechanism may involve PGE2-induced enhancement of paracellular Na⫹
backflux from the interstitium into the tubular lumen during increased RIHP that would decrease net Na⫹ reabsorption (1207). PGE2 also inhibited Na⫹ reabsorption by the
TAL and the CDs (37, 150, 151, 1061).
An alternative view of pressure natriuresis is based largely
on studies in the rat where there is reportedly less efficient
autoregulation of medullary and papillary blood flow during states of relatively high Na⫹ excretion (FIGURE 4B)
(277, 1248). Cowley, Mattson, Roman, and colleagues
have implicated medullary blood flow as a major determinant of Na⫹ excretion and BP (274, 277). They reported
that medullary blood flow and NO were reduced, and ROS
production was increased, during states of Na⫹ retention
and hypertension (275, 989, 991, 1670), whereas increased
medullary blood flow was associated with states of a low BP
(273, 274, 1014). NO could have a causal role also in the
414
medulla since its blockade attenuated the increases in renal
medullary blood flow and Na⫹ excretion during increased
RPP (411). An increased RPP in volume-expanded rats did
not change RBF, GFR, or SNGFR in either superficial or
deep cortical nephrons (1248). Nevertheless, proximal tubular reabsorption was decreased by 10%, and NaCl delivery to the bend of Henle’s loop of JMN was nearly doubled
(1245), thereby likely providing a signal to activate MDTGF. As is summarized in FIGURE 4B, these investigators
have demonstrated that RPP increases papillary blood flow
and vasa recta capillary hydrostatic pressure that is translated into an increased medullary RIHP which, after transmission to the cortex, reduces proximal tubular Na⫹ reabsorption. Increased Na⫹ delivery to the TAL of Henle’s
loop may stimulate production of O2⫺ which can be countered by flow-induced NO production by the vasa recta
capillaries. The balance of vasoconstrictor O2⫺ and vasodilator NO in the renal medulla is held to regulate the medullary blood flow by determining the degree of contraction
of pericytes in the outer medullary vasa recta, which, in
turn, can determine the strength of the pressure-natriuresis
(1099).
Moreover, the oxygen tension (PO2) of the renal parenchyma could also account for a relatively impaired medullary and papillary autoregulation. The kidneys have a high
O2 supply that is necessary to match the demand for tubular
O2 consumption related to active tubular Na⫹ reabsorption. The PO2 in the proximal and distal tubules, the efferent
arterioles, and the interstitium of the outer cortex of rat
kidneys were all similar and average 40 – 45 mmHg, indicating diffusional O2 equilibrium (1582). The PO2 in the
glomerulus, measured by a different technique, also averaged 45 mmHg (1331). These values were all lower than
PO2 measured in the renal vein (i.e., 55– 65 mmHg), indicating a preglomerular diffusional shunt for O2 between the
arterial and venous systems, likely via the arcuate vessels
(1605). This shunt stabilized renal cortical PO2 during
changes in arterial PO2 (883, 1605) but maintained a low
renal parenchymal PO2. The PO2 in the deep cortex and the
outer medulla of rat (FIGURE 5) (1260, 1582, 1583) and
mouse kidneys (860) was reduced to 30 – 40 mmHg. A
study in the rat hydronephrotic kidney (HNK) preparation
reported substantial reductions in the myogenic response as
the perfusate PO2 was reduced from 80 to 20 mmHg (936).
Thus a lower parenchymal PO2 in the medulla might contribute to a lower efficiency of the myogenic response (FIGURE 5) and thereby to the relatively impaired medullary
autoregulation. However, studies in dogs reported that
hypoxia (PO2 between 40 – 45 mmHg) did not affect autoregulation of RBF or GFR (26), but the diffusional O2 shunt
rendered renal PO2 largely independent of arterial PO2, at
least in the rat (1605). It remains to be demonstrated that
medullary PO2 varies with PP as would be required to modulate pressure natriuresis.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
A
B
Myogenic responses and PO2
PO2 in
rat kidney
0
20
-10
30
40
-20
60
-30
PO2 (mmHg)
0
PO2 (mmHg)
Change in arteriolar diameter (%)
10
OC
IC
10
20
30
40
50
80
-40
80
120
140
160
180
PO2 in
mouse kidney
OC
OM
FIGURE 5. Oxygen tension (PO2) in different renal
compartments and its influence on the myogenic response of afferent arterioles. A depicts changes in
the diameter of afferent arterioles during increases in
renal PP (from 80 to 180 mmHg) in the atubular rat
HNK preparation (reflecting myogenic responses)
during perfusion with fluids of different PO2 ranging
from 20 to 80 mmHg. Note the graded impairment
of myogenic responses across the range of PO2 from
80 to 20 mmHg (936). B and C present values for
PO2 measured with microelectrodes in the kidney. In
B, PO2 values are shown for the outer cortex (OC) and
inner cortex (IC) in anesthetized rats (1582). In C,
PO2 values are shown for the outer cortex and outer
medulla (OM) of anesthetized mice (860).
Renal PP (mmHg)
In summary, the renal pressure-natriuresis response is important for control of ECV and BP. Although a rise in RIHP
and reduction in tubular reabsorption following an increase
in RPP are crucial events, the underlying mechanisms are
still debated. The contrasting two major hypotheses for
how the kidney senses a rise in RPP and generates a pressure-natriuresis, despite excellent autoregulation of the cortical nephrons, are summarized in FIGURE 4, A AND B. Functional abnormalities in disease models will be discussed
later in the sections on hypertension, CKD, and DM (see
sect. VII).
3. Conventional macula densa tubuloglomerular
feedback
The anatomical arrangement of different cell types in the
JGA that are involved in macula densa (MD)-mediated tubuloglomerular feedback (TGF) is shown schematically in
FIGURE 1. MD-TGF relates afferent arteriolar tone to MD
cell tubular electrolyte delivery and reabsorption. It can be
studied in superficial nephrons of anesthetized animals by
microperfusion of Henle’s loop with artificial tubular fluid
to vary NaCl reabsorption by the MD segment (48, 1066,
1075, 1305, 1316, 1513). Details about methodologies and
approaches to investigate MD-TGF by renal micropuncture
have been reviewed (923, 1311, 1513). Afferent arteriolar
responses have been assessed from changes in the SNGFR,
PGC, or proximal tubular stop-flow pressure (PSF) upstream
from an oil or wax block in the proximal tubule. MD-TGF
in longer deep cortical nephrons has been assessed in the
exposed papilla by perfusing fluid through the TAL of Henle’s loop and measuring glomerular responses at an upstream site of the same nephron (1042, 1043). The afferent
arteriole is the effector site, and its terminal segment is the
most responsive site for MD-TGF activation (200, 695,
703, 1020). Ambient MD-TGF activity has been assessed
from the difference in SNGFR measured from the proximal
tubule (which prevents activation of the MD) and the distal
tubule with MD function intact. This approach characterizes MD-TGF at zero and ambient, but not at supranormal
rates of tubular fluid flow.
The role of MD-TGF in autoregulation has been evaluated by measuring vascular function during changes in
RPP before and during inhibition of MD-TGF by blocking the flow of tubular fluid to the MD of the test nephron
or by blocking NKCC2 activation of MD cells with a
loop diuretic. Inhibition of MD function abolished the
effects of MD-TGF signaling on afferent arteriolar diameter, PGC, single-nephron blood flow, and GFR of superficial and deep cortical nephrons and attenuated autoregulation of SNGFR during changes in RPP (200, 1019,
1022, 1064, 1075, 1249, 1636). A contemporary approach has been to assess afferent arteriolar or glomerular responses using genetic mouse models that lack MDTGF such as the adenosine A1 receptor knockout (155,
205, 686, 1123, 1312, 1318, 1508) and other transgenic
models with impaired MD-TGF signaling (152, 496,
806, 942, 1093). Autoregulation was less complete after
inhibition of MD-TGF by genetic deletion of adenosine
A1 receptors (583, 751) or P2X1 receptors (686) or Cx40
(756, 1395), all of which impair MD-TGF responses.
Glomerular autoregulation has been reported to be less
efficient in the absence of MD-TGF in several species,
including rats (1019, 1022, 1313, 1444), dogs (351, 760,
1064, 1090, 1091, 1622), and mice (583). For example,
the MD-TGF in the rat contributed ⬃50% to autoregulation of SNGFR when the RPP was reduced from 115 to
95 mmHg and 30% when it was reduced from 95 to 80
mmHg (1313). Studies made in dogs indicated that a
reduction in RPP from 124 to 94 mmHg in kidneys in
which RBF and GFR were efficiently autoregulated led to
a 30% reduction in proximally measured SNGFR (when
flow to the MD is zero), whereas distally determined
SNGFR (when delivery to the MD is intact) was unchanged (1613). Steady-state RBF autoregulation was reduced ⬃50% in nonfiltering dog kidneys (1284). In mice,
a 15 mmHg reduction in RPP reduced RBF and GFR by 6
and 12%, respectively, but the adjustments were less
complete (with 15 and 40% respective reductions) in
mice lacking functional MD-TGF (583). Thus MD-TGF
clearly contributes to autoregulation of RBF and GFR.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
415
RENAL AUTOREGULATORY MECHANISMS
delay was ⬃30 s in the dog kidney (86). This response was
completed during the following 15–20 s in rodent kidneys
(FIGURE 3B) (296, 296, 629, 748, 749) or 20 – 40 s in dog
kidneys (748, 753). This delay sets up MD-TGF-based oscillations in proximal tubular pressure (PPT) at ⬃0.035 Hz
(626, 893, 1512) which are in phase with fluctuations in
early distal tubule fluid flow and [Cl⫺] that are signals for
activation of MD-TGF. All of these cycled at 25–35 s (35
mHz) (627). Measurement of these oscillations has been
used to assess endogenous MD-TGF activity during
changes in RPP.
Afferent arteriolar autoregulatory responses in the JMN
preparation were reduced 25–50% by elimination of MDTGF by papillectomy or furosemide in the rat (686, 1028,
1364, 1437) or mouse (686, 687). Papillectomy or tubular
oil block inhibited afferent arteriolar responses similar to
that of furosemide, indicating that the primary effect of
furosemide in this preparation was on MD transport and
not directly on the afferent arteriole. However, furosemide
also attenuated the autoregulation of GFR when measured
in proximal tubules in the JMN preparation, which prevents delivery of tubular fluid to the MD and negates the
MD-TGF (200, 1022). This might represent nephronnephron interactions whereby inhibition of MD-TGF in
neighboring nephrons attenuates the myogenic response in
the test nephron. Other studies reported that furosemide
abolished pressure-dependent autoregulation of afferent arterioles in the JMN preparation (485). Furosemide also inhibited MD-TGF-mediated constriction of the afferent arteriole in the doubly perfused, isolated JGA (695). These
studies reinforce the conclusion that MD-TGF contributes
importantly to renal autoregulation.
A number of vasoactive factors may link MD transport of
electrolytes by the NKCC2 cotransporter, ROMK channel,
Na⫹/H⫹ exchanger, and Na⫹-K⫹-ATPase to the release of
vasoconstrictors including ATP, adenosine, PGH2, TxA2, and
O2⫺ and vasodilators such as NO, PGE2, and carbon monoxide (CO) to modulate afferent arteriolar tone (FIGURE 6). The
following sections discuss the role of purinergic signaling
molecules (e.g., ATP and adenosine) in regulation of MDTGF and renal autoregulation, and the role of Cxs and
extraglomerular mesangial cells.
MD-TGF in rodent kidneys had a delay of 10 –15 s before
initiation of a PSF response following a rapid step change in
RPP and increased tubular fluid flow (87, 296, 1547); the
A) ATP RECEPTOR SIGNALING.
ATP can be released into the
interstitium by activated MD cells (83, 203, 1064, 1510);
Renal PP
Connecting tubule NaCl delivery
Macula densa NaCl delivery
ENaC transport
NO
NKCC2 transport
Na+/H+ exchange
EETS
PGE2
NO
O2·-
PGH2
TxA2
Adenosine
ATP
PGE2
Afferent arteriolar tone
FIGURE 6. Flow diagram for the modulation of renal afferent arteriolar tone by conventional macula densa
tubuloglomerular feedback (MD-TGF) and connecting tubule glomerular feedback (CT-GF). Changes in renal
perfusion pressure (PP) are transduced by alterations in tubular fluid NaCl delivery to the MD and the CT.
Subsequent changes in tubular transport lead to release of specific mediators and modulators that impact on
afferent arteriolar vasomotor tone. Vasoconstrictor agents increase (⫹) arteriolar tone. Vasodilator agents
reduce (⫺) arteriolar tone. ENaC, epithelial Na⫹ channel; NKCC2, Na⫹-K⫹-Cl⫺ cotransporter; EETs, epoxyeicosatrienoic acids; PGE2, prostaglandin E2; NO, nitric oxide; O2·⫺, superoxide; PGH2, prostaglandin H2; TxA2,
thromboxane A2; ATP, adenosine triphosphate; CO, carbon monoxide. For details, see text (sects. IID and VIE ).
416
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CO
CARLSTRÖM ET AL.
however, uncertainty exists regarding the mechanism(s) of
the associated vasoconstriction of the parent afferent arteriole (FIGURE 6). ATP is thought to directly activate P2 receptors on the afferent arteriole or to activate different purinergic receptors on extraglomerular mesangial cells to initiate a Ca2⫹ wave that is transmitted to the afferent arteriole
through gap junctional coupling (83, 85, 682, 1071, 1176).
Alternatively, ATP and AMP may activate MD-TGF indirectly following metabolism by nucleotidases to adenosine
(1510) (discussed in sect. IIA3B). ATP, adenosine, and other
endogenous agents can activate vasoconstrictor purinergic
homomeric P2X1, heteromeric P2X1/4, or adenosine A1 receptors, respectively, on preglomerular VSMCs (521, 565,
1510, 1540). Activation of P2X receptors by ATP induced
ligand-gated cation influx, membrane depolarization, and
Ca2⫹ entry via VOCCs that were reinforced by inositol
trisphosphate (IP3) receptor-mediated Ca2⫹ mobilization
(565, 683, 685, 1197). Studies examining the potential role
of P2X receptors in renal autoregulation, in particular MDTGF responses in different nephron populations, have
yielded conflicting results. ATP and P2X1 receptors are required for MD-TGF responses in JMN of both rats and
mice (683, 686, 690). Knockout of P2X1 receptors attenuated MD-TGF responses and the contribution of MD-TGF
to autoregulation of total or juxtamedullary blood flow
(686). In one rat study, peritubular capillary infusion of
ATP caused transient reduction in PSF and attenuated subsequent maximal MD-TGF response of superficial
nephrons by more than 80% (1013). Continuous administration of ATP caused P2X receptor desensitization and attenuated MD-TGF, suggesting that P2 receptor signaling
participated in MD-TGF (1013). However, emerging evidence from assessment of MD-TGF in superficial nephrons
elicited by elevated loop flows in vivo does not support a
role for P2 receptors in superficial nephrons. For example,
PSF measurements revealed normal MD-TGF in mice lacking P2X1 receptors (1312), and administration of the P2
receptor inhibitors PPADS or suramin did not alter the PSF
response (1306).
Stepwise reductions in RPP in dogs elicited parallel reductions in whole kidney RVR and renal cortical interstitial
ATP concentration (1090). Moreover, stimulation of MDTGF by increasing distal fluid delivery using acetazolamide
increased renal interstitial ATP, whereas blockade of MD
reabsorption with furosemide reduced ATP and renal autoregulation (1091). Thus a positive correlation between interstitial ATP and RVR adjustments has been demonstrated, and changes in ATP concentration could provide a
renal interstitial signal of changed RPP and an effector
agent for vasoconstriction by activation of purinergic signaling. Desensitization, saturation, or blockade of the P2purinoceptors attenuated whole kidney (952, 1131, 1438)
or nephron autoregulation via blockade of the MD-TGF
(683, 686, 690). Available evidence indicates that P2X1 receptors participate in renal autoregulation primarily by reg-
ulation of the MD-TGF mechanism in the JMN (203, 683,
1065, 1312, 1510).
In contrast to the renal microcirculation, myogenic autoregulation in the cerebral vasculature depended strongly on
P2Y4 and P2Y6 pyrimidine receptors (142). Pharmacological blockade or molecular suppression of either of these P2
receptors using antisense oligodeoxynucleotides reduced
myogenic constrictor tone by ⬃45%. Thus the roles of purinergic agonists and receptors in autoregulation vary according to the vascular bed.
B) ADENOSINE RECEPTOR SIGNALING. Other studies strongly implicate adenosine in the MD-TGF regulation of RBF, GFR,
and PGC (203, 1316, 1459, 1510, 1511). Vasoconstriction
occurs primarily via activation of adenosine A1 receptors on
the afferent arteriole (155, 778, 1216, 1316, 1320, 1426,
1540), but A1 receptors on mesangial cells also may modulate Ca2⫹ signaling (1122). Activation of high-affinity A1
receptors caused vasoconstriction by stimulation of G␣i
proteins that inhibited cAMP production and by activation
of phospholipase C (PLC) that liberated IP3 and mobilized
intracellular Ca2⫹ (555, 1510). Activation of the lower affinity adenosine A2 receptors elicited vasodilatation by
stimulation of G␣s proteins that activated the cAMP/protein kinase A (PKA) pathway and stimulated NOS (188,
555, 557). A2 receptor activation attenuated PP-induced
arteriolar vasoconstriction by opening KATP channels to
cause hyperpolarization of the plasma membrane (1450).
A2 receptor-mediated vasodilation during MD-TGF can
oppose A1 receptor-induced vasoconstriction (187, 188).
Maximum stimulation of A2 receptors impaired renal autoregulation of afferent arteriolar tone in one study (409), but
not of whole kidney RBF in another (1198).
Adenosine formed by nucleotidase metabolism of ATP and
AMP mediates the MD-TGF response in superficial
nephrons of mice (205, 647, 1123, 1459). A1 receptor antagonists inhibit MD-TGF markedly (1320). Mice genetically deficient in A1 receptors lacked MD-TGF in superficial
nephrons (155, 1426), whereas MD-TGF was enhanced in
mice overexpressing A1 receptors (1125). Steady-state autoregulation of RBF was reduced by 40 –50% in A1⫺/⫺ mice
(583, 751). This was ascribed to a weaker contribution of
MD-TGF while the myogenic component was well preserved. Selective deletion of A1 receptors in VSMCs reduced
MD-TGF responsiveness by 65%, thereby demonstrating
that the main effect of A1 receptors is to mediate MD-TGFinduced afferent arteriolar vasoconstriction rather than to
activate MD cells or enhance signal transmission through
extraglomerular mesangial cells (896, 1122). Thus activation of A1 receptors mediates MD-TGF in superficial
nephrons and contributes significantly to autoregulation in
rodents, but these receptors are probably not required for
the myogenic response. Pharmacological blockade of vascular A1 receptors in the isolated rabbit JGA preparation
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
417
RENAL AUTOREGULATORY MECHANISMS
abolished MD-TGF (1216), similar to inhibition of 5’-nucleotidase conversion of AMP to adenosine (1229).
However, some pharmacological studies in dogs, rats, and
mice failed to detect a major role of adenosine or A1 receptors in MD-TGF signaling or afferent arteriolar autoregulatory responses (438, 686). Systemic administration of an
A1 receptor antagonist did not impair RBF autoregulation
in the dog kidney (654) and did not blunt PP-induced
changes in afferent arteriolar diameter in the mouse JMN
preparation (686). Propagation of a MD-TGF-induced
Ca2⫹ wave along the afferent arteriole in rats also was
unaffected by A1 receptor blockade (1176).
In summary, activated MD cells release ATP that can be
metabolized to adenosine. Uncertainties about the roles of
adenosine and A1 receptors and ATP and P2X1 receptors in
preglomerular autoregulatory responses to MD activation
are based on controversies concerning their respective contributions to MD-TGF. There are studies supporting an
involvement of both adenosine and ATP in the MD-TGF
response, and contrasting conclusions may be explained by
different experimental designs or perhaps by regional differences in ATP and adenosine signaling within the kidney.
Available evidence demonstrates that P2X1 receptors participate in renal autoregulation primarily by regulation of
the MD-TGF mechanism in JMN, whereas A1 receptors are
essential for MD-TGF responses in superficial nephrons.
Nevertheless, it seems clear that neither of these purinergic
agents nor their specific receptors mediates the renal myogenic response.
C) CXS AND EXTRAGLOMERULAR MESANGIAL CELLS. Cxs form gap
junctions to connect cells. They provide intercellular channels that can facilitate the exchange of ions, second messengers, electrical signals, and metabolites. Cxs coupled multiple cell types in the JGA in the kidney, including endothelial, VSMCs, and juxtaglomerular granular cells (JGC)
(553, 1545). The preglomerular vasculature and the JGA
expressed primarily Cx37, Cx40, Cx43, and Cx45 (50,
553, 1439, 1546). Cx40 predominated in endothelial cells
and Cx45 in VSMCs (553), although the latter has been
questioned (1296). The endothelium of the proximal afferent arteriole possessed all of these Cxs, whereas the VSMCs
in the terminal segment of the arteriole expressed only
Cx40. Moreover, Cx40 was the major isoform in the JGA
where it was expressed on endothelial cells, granular reninproducing cells, and extraglomerular mesangial cells. However, MD cells lacked Cxs, and therefore any effects of
Cx40 on autoregulation likely involved transmission of
MD-TGF signals to the afferent arteriole.
Destruction of glomerular mesangial cells in rats by mesangiolysis or pharmacological disruption of gap junctions prevented MD-TGF responses (1218) and impaired GFR autoregulation (706, 1438), but left RBF autoregulation rela-
418
tively intact (706). Administration of 5’-nucleotidase
improved GFR autoregulation following mesangiolysis, implicating attenuated adenosine signaling in the impaired
GFR autoregulation (1444).
Putative inhibitory peptides directed against specific Cxs
have implicated them in autoregulation, but have not yet
distinguished between myogenic and MD-TGF mechanisms. Blockade of Cx37 and Cx40, but not Cx43, in normal rats increased the BP and reduced RBF autoregulation
(1439), whereas blockade of Cx37, Cx40, or Cx43 all impaired the RBF autoregulation in Zucker lean rats (1441).
Cx37 and Cx40 in Zucker lean rats were implicated in
transducing purinergic signals involved in autoregulatory
responses, likely involving MD-TGF (1438). RBF autoregulation following putative Cx inhibition was improved by
increasing adenosine production from AMP with 5’-nucleotidase.
There is some concern about the specificity of inhibitors of
gap junction function. Nonspecific effects of the gap junction inhibitors heptanol and 18␤-glycyrrhetinic acid have
been noted even at concentrations that have no or little
effect on gap junctions (979). Cx-mimetic peptides appear
to be more selective in inhibiting gap junctions. However,
the most definitive results have derived from Cx gene-deleted mice.
Genetic deletion of Cx40 and Cx43 increased renin synthesis and secretion and disrupted the regulation of renin
release by intravascular pressure and ANG II (545,
1546). Restoration of Cx40 in renin-producing JGC but
not endothelial cells reduced hypertension and renin secretion of Cx40 null mice (879). Cx40 knockout mice
had impaired whole kidney RBF autoregulation with a
markedly reduced MD-TGF component (756). However,
they retained an unchanged myogenic and third mechanism and a preserved dampening of the myogenic response by NO. Thus Cx40 contributes to RBF autoregulation via MD-TGF-mediated signal transduction from
the MD. This is independent of NO but likely includes
traffic across extraglomerular mesangial cells that are
heavily endowed with Cx40 (756). These findings and
conclusions are supported by the disruption of autoregulatory adjustments of the afferent arteriole in the JMN
preparation in Cx40-deficient mice (1395). These mice
lack the MD-TGF mechanism, whereas the myogenic response is preserved. Micropuncture studies of superficial
nephrons in kidneys lacking Cx40 demonstrated 50%
attenuation in MD-TGF (1122). In contrast, Cx37-deficient mice retain a highly efficient RBF autoregulation
with normal contributions of myogenic and MD-TGF
components (A. Just, personal communication). Thus
Cx40 appears to be a key element for transmission of
MD-TGF signal(s) that contribute to renal autoregulation.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
Conducted vasomotor responses in preglomerular arterioles are mediated largely by passive electronic spread of the
EM transmitted through gap junctions in the arteriolar wall
connecting endothelial cells and/or VSMCs, but the specific
Cxs involved remain controversial. Electrically induced
Ca2⫹ responses conducted along rat isolated cortical radial
arteries were preserved after inhibitory peptides directed
against Cx37, 40, 43, and 45 (1399), whereas propagation
in mice was abolished by genetic deletion of Cx40 (1395).
Interestingly, the Ca2⫹ response to NE along this arterial
segment was unaffected by deletion of Cx40. In vivo studies
demonstrated that vasoconstriction produced by NE, or
vasodilation elicited by acetylcholine (ACh), was preserved
in the renal vasculature of mice lacking Cx40 (756). Moreover, mechanical stimulation of cultured glomerular endothelial cells triggered a Ca2⫹ wave propagated from cell to
cell, which was dependent on Cx40, extracellular ATP, and
P2 receptors, but was independent of L-type VOCCs
(1473). On the other hand, the propagation of mechanically
induced Ca2⫹ waves was slower in cultured VSMCs deficient in Cx45 and in control VSMCs treated with a peptide
that inhibited Cx45 gap junctional communication (554).
Direct gap junction coupling was suggested since Ca2⫹
wave propagation was unaffected by the ATP P2 receptor
blocker suramin. In contrast, VSMC-specific deletion of
Cx45 did not affect conduction of vasomotor responses to
high KCl, ACh, or adenosine along the VSMC layer of
mouse cremaster arterioles tested in vivo (1296). A Ca2⫹
wave originating by activation of MD cells and passing
through extraglomerular mesangial cells was dependent on
ATP and P2 receptors and was blocked by pharmacological
uncoupling of gap junctions (1176). ATP traversed Cx43
hemichannels in astrocytes (766). As discussed in more detail in section VIG, major differences have been noted between positive in vitro and negative in vivo findings in these
studies.
The Cx-like channel pannexin 1 was expressed in VSMCs
of mouse afferent and efferent arterioles and may play a role
in purinergic signaling (552). Pannexin 1 also was localized
to the apical membrane of several tubular segments where it
mediated ATP release into tubular fluid.
Gap junctional communication has been implicated in myogenic responses of nonrenal arteries. Nonselective inhibition of gap junctions attenuated myogenic contractions of
cerebral arteries (856), whereas inhibition of Cx37 and
Cx43 attenuated myogenic vasoconstriction of rat mesenteric resistance arteries (369). VSMCs were hyperpolarized
and had diminished changes in [Ca2⫹]i. Although gap junctions mediated pressure-induced VSMC cell depolarization,
changes in [Ca2⫹]i and myogenic vasoconstriction in vitro,
they were not required for vasoconstriction to PE. Genetic
alteration of Cx40 in endothelial cells of mouse mesenteric
arteries demonstrated its role in transmission of electrical
(but not chemical) signals that increased the sensitivity of
the myogenic response (215).
In summary, in vivo and in vitro observations suggest a role
for gap junctions and Cxs, particularly Cx40, in MD-TGF
signaling to the afferent arteriolar cells via mesangial cells.
Further studies are required to elucidate how Cxs, either as
monomers or multimers, coordinate MD-TGF signal transmission and renal autoregulation. Cxs also are implicated in
Ca2⫹ signaling and electrotonic and purinergic signaling
along the preglomerular vasculature which may participate
in nephron synchronization and renal autoregulation. The
initial signaling between MD and JGC is diffusive while
transmission from JGC to the AA is mainly electrotonic.
The role of Cxs in modulating autoregulation is also discussed in section VIG.
4. Dynamics of autoregulatory responses
Steady-state assessments of RBF autoregulation in response
to graded changes in RPP have been used widely to study
the effectiveness of autoregulation and the range of RPP
over which autoregulation takes place (FIGURE 3A). Dynamic analyses allow a more detailed evaluation of the contributions of individual underlying mechanisms to the overall adjustments in vascular resistance or admittance.
A) STUDIES IN THE TIME DOMAIN.
Measurements of dynamic
responses of RVR in the time domain have employed a
sudden, single-step change in RPP within the autoregulatory range. Such studies revealed a fingerprint of distinct
intrarenal mechanisms as a function of time after the change
in RPP based on clearly separable time constants with
minimal nonlinear interactions. A rapid 20 mmHg step increase in RPP led to a biphasic or triphasic increase in RVR in
dogs (753, 755, 1465, 1578), rats (423, 748 –750, 756,
1313, 1342, 1640), and mice (299, 462, 505, 751, 756).
This approach separated the most rapid myogenic component (0 – 8 s in rodents) from the slower MD-TGF component
(10 –30 s in rodents) (FIGURE 3B) (748, 749, 753). The early
myogenic response was initiated by an immediate change in
transmural pressure since a similar rapid response followed a
step reduction of extravascular tissue pressure (57, 265). The
onset of the MD-TGF-dependent changes in RVR was delayed by 10 –15 s and completed within 20 –30 s in rodent
kidneys (FIGURE 3B) (296, 749, 1547) and within 30 – 45 s
in dog kidneys (753).
Autoregulatory kinetics have been characterized in normal
rats using the isolated, blood-perfused JMN preparation. A
50 mmHg increase in RPP elicited a biphasic contractile
response with a transition point at ⬃24 s. In the absence of
MD-TGF (papillectomy), the response of myogenic origin
retained the same rate of contraction but terminated 11 s
earlier at 13 versus 24 s with intact MD-TGF (1547), suggesting that MD-TGF enhances the myogenic mechanism
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
419
RENAL AUTOREGULATORY MECHANISMS
by prolonging its response. This important study by Walker
and Navar is discussed in more detail in section IIC.
Additional autoregulatory mechanisms have been identified
recently. A third mechanism has been proposed to increase
RVR progressively from 30 to 120 s after the abrupt increase in RPP. It contributed 10 –20% to overall autoregulation in mice and rats (FIGURE 3B) and 30% in dogs (299,
749, 751, 753, 1626). Its participation may increase at low
RPP in the rat (749). It was preserved in knockout mice
lacking neuronal NOS, inducible NOS, and adenosine A1
receptors (299, 751) and was independent of MD-TGF in
some, but not all, studies (749 –751, 753). The mediators of
this response are unclear, but may reflect an interaction
of known mechanisms, a delayed engagement of MD-TGF
of long JMN, a contribution of renal autocrine/paracrine
factors, or a component of pressure-natriuresis. ATP and
P2X1 receptors may also mediate the third mechanism, or
perhaps ANG II acting on AT1 receptors (284, 491, 1397).
Inhibition of MD-TGF by drugs or by genetic deletion of
adenosine A1 receptors unmasked an apparent fourth
mechanism that operated within the 5- to 25-s time window
after an abrupt increase in RPP and contributed ⬃25% to
overall autoregulation (751). This time-frame overlapped
that of the conventional MD-TGF and was unmasked only
when MD-TGF was prevented by furosemide. However,
some MD-TGF might operate independent of furosemidesensitive NKCC2 cotransport, perhaps modulated by a
Na/H exchanger in MD cells (84, 431, 800, 1177, 1178,
1210).
The influence of the connecting tubule glomerular feedback
(CT-GF) vasodilator system, involving Na⫹ reabsorption
via ENaC in the CT, is presently unclear. However, a vasodilatory phase observed 60 –200 s after a step increase in
RPP in the rat (1342) may represent the contribution of the
CT-GF (see sect. IIB2). Thus a third, and possibly a fourth,
mechanism may contribute to renal autoregulation in the
dog, rat, and mouse kidneys. Their importance and underlying signaling pathways require further study.
Several rat studies have characterized dynamic adjustments
of the renal vasculature following a brief period of complete
occlusion of the aorta (1342, 1397, 1398, 1626), but this
evoked metabolic and vasoactive factors, in addition to basic pressure-dependent autoregulatory mechanisms. Studies
in the rat of complete renal ischemia for 30 s followed by a
rapid single step increase in RPP reported time-dependent
oscillations in the RBF responses at ⬃10, 35, and 115 s,
reflecting the operation of three distinct autoregulatory
mechanisms (1626), likely the myogenic, MD-TGF, and
third mechanisms.
A strength of dynamic analysis of vascular adjustments in
the time domain is that in addition to providing useful in-
420
sight into individual mechanisms and their relative conributions, it provides reliable quantiative data for total autoregulation that agree with steady-state AI values for RBF autoregulation.
B) STUDIES IN THE FREQUENCY DOMAIN. Kidneys normally exhibit oscillations in microvascular tone. Transfer function
analyses of the gain of renal vascular admittance (the reciprocal of impedance and similar to steady-state conductance
that is the reciprocal of resistance) to spontaneous or imposed changes in RPP can distinguish between autoregulatory components in the frequency domain (19, 230, 297,
629, 631, 722, 971, 1274, 1357, 1358, 1563, 1569, 1570).
This has been investigated in conscious dogs (753, 755,
760, 1618), mice (668), rabbits (722), and rats (927, 1182,
1183, 1376). Renal autoregulation in conscious animals is
efficient and similar to that recorded using the same methodology under anesthesia. One discordant study reported
no blood flow autoregulation in the kidneys of conscious
rats during spontaneous fluctuations in BP between 0.03
and 2 Hz (723).
An admittance gain of less than 0 dB or a relative normalized admittance gain of 1.0 (analogous to an AI of 1.0)
indicates active autoregulatory adjustments within the vasculature to minimize RBF changes during fluctuations in
RPP. An admittance gain above 0 dB or 1.0 relative unit
implies that passive vascular elasticity contributes to the
amplification of the effect of RPP fluctuations on RBF. A
lower relative value between 1.0 and zero quantitates the
relative strength of overall autoregulation with the fractional variation in RBF being smaller than that of the RPP.
As with AI, a normalized admittance gain of zero indicates
perfect autoregulation. A representative recording of normalized admittance gain as a function of frequency for normal rats is shown in FIGURE 3C (297). Interpretation of
admittance gain commonly is best achieved by reading from
right to left, from high to low frequencies. A decline in the
gain values indicates the onset of an active contribution to
autoregulation. With the use of this methodology, the myogenic mechanism operating in the 0.20 – 0.10 Hz frequency
band accounts for ⬃65% of total autoregulation of RBF,
whereas the MD-TGF mechanism in the 0.04 – 0.011 Hz
frequency band appeared to contribute ⬃35% of total autoregulation. The slower apparent third mechanism operates at frequencies of ⬃0.01 Hz and below which is below
the frequency range normally examined with confidence.
Most studies of dynamic autoregulation utilizing frequency
analysis have reported that the myogenic response in normal
rat kidneys reduces the admittance gain from approximately
⫹7.5 to ⫺5 dB (or fraction gain from ⬃2.3 to 0.56) over the
frequency range of 0.20 to 0.07 Hz, whereas the MD-TGF
response reduces the admittance gain by a lesser degree from
⫺2.5 to ⫺12.5 dB (fractional gain from 0.75 to 0.24) over the
frequency range 0.05 to 0.020 Hz (FIGURE 7B). Summarized
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
B
RVR after a step increases in renal PP
150
100
50
Control
L-NAME
0
Frequency domain analysis
Admittance gain (dB)
RVR (percent of
perfect autoregulation)
A
15
10
5
0
-5
Control
L-NAME
-10
-15
-50
0
5 10 15 20 25 30 35 40 45
50 55 60
0.01
Time after renal PP increase (sec)
Afferent arteriolar
diameter (µm)
24
Afferent arteriolar myogenic
response in HNK
20
16
12
Control
L-NAME
8
4
80
100
120
140
160
180
Perfusion Pressure (mmHg)
D
Change in afferent
arteriolar diameter (µm)
C
0.1
1
Frequency (Hz)
Juxtamedullary
nephron preparation
0
-10
-20
-30
Control
L-SMTC
-40
100
130
160
Perfusion Pressure (mmHg)
FIGURE 7. Effect of nitric oxide synthase (NOS) inhibition of renal autoregulation mediated by myogenic and
macula densa tubuloglomerular feedback (MD-TGF) mechanisms. Illustrations of the effects of inhibition of NOS
with L-nitro-arginine methyl ester (L-NAME) on whole kidney autoregulation or renal blood flow (RBF), renal
vascular resistance (RVR), and myogenic responses of individual vessel segments. A depicts time-dependent
changes in RVR in rat kidneys following an abrupt step increase in renal perfusion pressure (PP) by release of
a suprarenal aortic clamp at time zero. After generalized inhibition of NOS with L-NAME, the initial increase in
RVR (corresponding to the myogenic response over the first 5 s) was greatly exaggerated and accounted for
almost all of the autoregulatory increase in RVR without major contributions of other autoregulatory components, such as the MD-TGF component seen in control kidneys from 5 to 25 s or the more delayed and gradual
third component thereafter (750). B shows the effect of general NOS inhibition with L-NAME on the transfer
function analysis of the admittance gain as a function of frequency for spontaneous changes in renal PP and
RBF in rat kidneys. After inhibition of NOS with L-NAME, the reduction in admittance between 0.07 and 0.2 Hz
(corresponding to the myogenic response) was much more abrupt and the regression of admittance on
frequency was steeper. Whereas control rat kidneys had a second reduction in admittance between 0.03 and
0.01 Hz (corresponding to the MD-TGF), this was hard to distinguish from the myogenic change in admittance
after L-NAME (1358). C depicts changes in the diameter of the afferent arterioles of the rat hydronephrotic
kidney (HNK) preparation as a function of renal PP. The decline in diameter with increasing renal PP (myogenic
response) was similar in control rats and those given L-NAME, indicating no change in the myogenic response
in this preparation that lacks renal tubules and MD cells (1358). D shows the change in afferent arteriolar
diameter in the rat JMN preparation to increased renal PP before and after inhibition of nNOS with SMTC
(S-methyl-L-thiocitruline) (662). The arteriolar autoregulatory response was increased after inhibition of MD
nNOS.
data for gain expressed in dB are from References 19, 286,
297, 629, 631, 748, 749, 1358, 1570. Data for fractional
gain are from References 109, 499, 1193. The myogenic
response in the kidneys of conscious mice reduces the admittance gain from ⫹7 to ⫺4 dB over the frequency range
of 0.3 to 0.1 Hz, whereas MD-TGF reduces the gain from
⫺1 to ⫺5 dB over the range of 0.04 to 0.01Hz (668). Thus
the myogenic mechanism provides the majority of the buffering of spontaneous fluctuations in RPP in rats and mice.
An early comparison of methodologies in conscious resting
dogs revealed that stepwise changes in mean RPP produced
stronger RBF autoregulation than overall autoregulation
analyzed by the transfer function admittance gain during
spontaneous fluctuations in RPP (760). Dynamic autoregu-
lation measured by the transfer function was only one-third
as strong as steady-state adjustments produced by the step
changes (⫺6.3 vs. ⫺19.5 dB). Just et al. (760) concluded
that MD-TGF was not required for RBF autoregulation
under normal physiological conditions, but contributed
during more extreme changes in PP.
Simultaneous measurements of steady-state RBF autoregulation and frequency domain analysis of whole kidney autoregulation in normal rats confirmed these methodological
differences and yielded disparate AIs of 0.2 and 0.4, respectively (109). Moreover, the methodological differences
were more pronounced in rats with reduced renal mass
(RRM). Whereas steady-state RBF autoregulation in 3/4
nephrectomized (Nx) rats was reduced compared with uni-
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
421
RENAL AUTOREGULATORY MECHANISMS
nephrectomized (UNx) or control rats (AIs: 0.7 vs. 0.2 or
0.1, respectively), frequency analysis revealed no defect in
autoregulatory efficiency, with fractional admittance gain
of 0.4 – 0.5 at very low frequencies in all three groups (109).
Such a methodological difference may represent insensitivity of admittance analysis at low frequencies (⬍0.05 Hz)
compounded by low coherence of RBF and PP determinations at these frequencies and the inclusion of time points
when RPP is below the lower limit of renal autoregulation.
Moreover, caution should be taken when interpreting dynamics at low PP, since this may interfere with natural
operating frequencies of the myogenic and MD-TGF components resulting in added complexity due to nonlinear dynamics, interactions, and time-dependent variability, as discussed in section IIC. Another considerations is that the
steady-state method assesses predominantly the response of
RVR to mean RPP, whereas the dynamic method assesses
primarily the admittance response to slow fluctuations in
RPP (⬍0.25 Hz), which constitute a substantially smaller
component of the total BP power. Bidani, Griffin, and associates (108, 109, 496, 499, 933) concluded the steadystate results were more accurate depictions of the overall
impaired autoregulation that leads to hypertensive glomerular damage after RRM. More systematic comparisons of
the characteristics and contributions of intrarenal mechanisms and their interactions need to be made using these
different approaches. Such results will advance the field by
providing a better understanding of the strengths and weaknesses of the static and dynamic methodologies.
Most frequency domain studies have found that normalized
admittance gain is close to unity (or 0 dB) or greater for
frequencies ⬎0.3 Hz, indicating little autoregulation and a
passive, compliant vasculature (FIGURE 3C) (286, 297, 629,
748). Nevertheless, one study reported that forcing of RPP
between 1 and 6 Hz produced strong, sustained adjustments in afferent arteriolar diameter in the HNK preparation (927). Renal autoregulation clearly acts normally as a
high-pass filter in adjusting afferent arteriolar diameter to
limit changes in PGC caused by fluctuations in RPP. PGC has
a pulse pressure of 7–10 mmHg which is roughly one-third
of arterial BP pulsations (47, 148, 347). This demonstrates
damping of the very rapid pulsations at heart rate frequencies prior to pressure transmission into the glomerulus, as is
expected from the known preglomerular resistance. Nevertheless, high systolic BP may periodically exceed the autoregulatory limit and thereby contribute to glomerular injury
and eventual fibrosis (108, 282).
Dynamic analysis reported an active 0.01- to 0.02-Hz lowfrequency component of RBF autoregulation in conscious
and anesthetized rats, which may represent the third mechanism (753, 755, 1375). An early study with limited ability
to quantify fast events (⬎0.08 Hz) reported an apparent
fourth mechanism operating in the 0.04 – 0.08 Hz frequency band in the dog (1618).
422
An advantage of the admittance analysis is that it can reveal
the contributions of myogenic and MD-TGF mechanisms
without necessitating blockade of MD-TGF. However, a
limitation is that it does not provide absolute values for the
relative strengths, especially of MD-TGF and slower frequency components (109, 289, 297, 631, 1313).
Oscillations of arteriolar blood flow and PPT in the renal
cortex were synchronous with oscillations of total RBF
(1636), indicating that they originate from the vasculature.
Recordings of SNBF in the efferent arteriole by laser Doppler flowmetry and of PPT in nephron pairs from the surface
of rat kidneys identified two oscillatory modes: a faster (0.1
to 0.2 Hz) myogenic mode and a slower (0.02– 0.05 Hz)
MD-TGF mode. A third oscillation at 0.0125– 0.0130 Hz
was thought to reflect an interaction between MD-TGF and
myogenic systems or possibly the third mechanism. Multiple studies indicate that the myogenic mechanism of normal
rats and SHR generates spontaneous oscillations of 0.1– 0.3
Hz, whereas MD-TGF produces oscillations of ⬃0.035 Hz
(289, 297, 628, 629, 631). However, the oscillations were
more irregular in rats with spontaneous or renovascular
hypertension (230, 625). Step changes in RPP elicited vascular oscillations at two frequencies corresponding to the
myogenic and MD-TGF responses (631).
5. Relative contributions of the different mechanisms
Most in vivo studies reported a ⬃40 –55% contribution of
myogenic and 25–35% contribution of MD-TGF responses
to dynamic autoregulation in dog, mouse, and rat kidneys
to a sudden, step increase in RPP (286, 748 –751, 753, 756).
Aside from frequency analysis of admittance gain, the individual components have been studied only after selective
blockade of MD-TGF, since prevention of the myogenic
response, for example, by blockade of VOCCs, abolished
vascular responses to all autoregulatory mechanisms (749,
1064). Blockade of MD reabsorption with loop diuretics
partially impairs overall renal autoregulation in kidneys of
dogs (749, 760, 1622), rats (749 –751, 1626), and mice
(668).
Clausen and Aukland, using a previously described box
technique (1601), placed a rat kidney in an airtight chamber
and acutely reduced the atmospheric pressure of the chamber by 35 mmHg. This reduced RIHP and increased vascular transmural pressure in this case by stretching the vessel
from outside. These authors observed a rapid, robust renal
vasoconstriction that was unaffected by blockade of MDTGF by furosemide (265). They concluded that the myogenic vasoconstriction is much stronger than that induced
by MD-TGF.
The MD-TGF component of renal autoregulation can be
abolished by ureteral obstruction (297), furosemide (19,
631, 749, 754, 1358), or genetic deletion of adenosine A1
receptors (751). As expected, all TGF components were
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
absent from the admittance pattern in the HNK preparation
lacking tubules, but the myogenic mechanism at 0.1– 0.3 Hz
persisted in all of these conditions and generally became
stronger and faster in the absence of MD-TGF (287, 1358),
as might be expected in the absence of MD-derived NO (see
sect. VIC). Overall, the myogenic mechanism appears to be
dominant in that it is essential and sufficient for efficient
RBF autoregulation. MD-TGF tends to provide a secondary
or back-up system to enhance total efficiency when the
myogenic mechanism is attenuated.
One study of frequency analysis concluded that an increase
in RPP increased the strength of both the myogenic and
MD-TGF mechanisms (1636). The efficacy of the myogenic
mechanism varies over time after a change in PP (247, 875).
There are additional actions of furosemide besides inhibition of NKCC2 transport in MD cells that may contribute
to its effects on renal autoregulation. Furosemide has a
direct vasorelaxant action in isolated, preconstricted afferent arterioles of mice by blockade of the NKCC1 transporter (1124). Indeed, inhibition of this transporter in the
afferent arteriole of the HNK preparation lacking MD-TGF
attenuated ⬃70% of myogenic response to changes in RPP
(1568). The natriuresis following blockade of tubular NaCl
reabsorption with furosemide increased the PPT (170, 749,
1124), which might attenuate the afferent arteriolar myogenic response if the transmural pressure was reduced, although RIHP was reportedly unchanged (789). On the
other hand, swollen tubules could compress peritubular
capillaries to increase vascular resistance and reduce RBF
(1124). Nevertheless, in one study a furosemide diuresis did
not change SNGBF, although RIHP was increased (1489).
Thus reduced autoregulation after furosemide may not necessarily indicate an exclusive contribution of MD-TGF. In
other studies, furosemide reduced RBF or GFR in intact
kidneys (1124), but the mechanism was not clear. Thus
furosemide may increase or decrease RBF and GFR by
mechanisms that have little to do with inhibition of MDTGF.
In summary, the myogenic response normally contributes
⬃50%, MD-TGF ⬃35%, and the third mechanism ⬃10 –
20% to dynamic renal autoregulation following a step increase in RPP (FIGURE 3B). Frequency analysis admittance
gain suggests less effective overall autoregulation with a
rather stronger contribution of the myogenic (⬃65%) than
the MD-TGF mechanism (⬃35%) (FIGURE 3C). The myogenic response contributes ⬃65% and MD-TGF ⬃35% at
high RPP in the JMN preparation (FIGURE 3D). There are
variable interactions between myogenic and MD-TGF
mechanisms, as discussed in section IIC.
B. In Vitro Studies in Isolated Preparations
The myogenic response has been studied from videometric
measurements of the diameter of afferent arterioles in re-
sponse to changes in RPP in the JMN preparation after
inactivation of MD-TGF (FIGURE 3D) or in the HNK preparation lacking tubules and thus devoid of any MD-TGF or
CT-GF (FIGURE 3E). Moreover, use of single isolated, perfused afferent arterioles allows calculation of the slope of
the regression of active wall tension on PP, which provides
a quantitative measure of the myogenic response throughout the PP range studied (FIGURE 3F). The intercept of this
regression on the x-axis indicates the threshold for initiating
a myogenic response. Myogenic and MD-TGF mechanisms
also have been studied using the isolated double-perfused
JGA preparation (702, 703).
1. The isolated JGA
Bell and associates (801) developed a rabbit nephrovascular
unit in which a glomerulus was attached to its MD segment
that was perfused via the TAL to characterize epithelial
transport properties. Early studies of renin release from the
JGC of the microdissected afferent arteriole were performed with the glomerulus attached or absent (698, 1377).
Mechanisms regulating renin release for JGC were also investigated in the HNK preparation (162, 163). Subsequent
studies demonstrated pressure-dependent renin secretion
from a perfused rabbit afferent arteriole with parent glomerulus present (122, 1181). In 1987, Skott and Briggs
(1378) used the isolated JGA to show that luminal NaCl
concentrations influenced MD signaling that regulated
renin secretion from JGC. Shortly thereafter, Ito and Carretero (702, 703) characterized the myogenic and MD-TGF
responses in a rabbit JGA preparation that was doubleperfused via the MD and the afferent arteriole. Activation
of MD reabsorption by increasing tubular [NaCl] constricted the afferent arteriole adjacent to the glomerulus,
whereas a pressure-induced myogenic constriction occurred
in more proximal segments of the afferent arteriole. This
preparation has been used commonly to characterize MD
metabolites that modulate the strength of afferent arteriolar
responses to MD-TGF signaling. Examples include potentiation by ATP, adenosine, O2⫺ (913, 1217), and attenuation by NO and CO (702, 1225, 1226). Recent studies in
the mouse isolated JGA indicated that MD cells can detect
variations in tubular fluid flow, independent of the salt
composition, by activation of primary cilia on MD cells,
which subsequently increases [Ca2⫹]i in afferent arteriolar
VSMCs and causes vasoconstriction (1374).
2. The isolated connecting tubule and glomerular/
afferent arteriolar unit
A variation of the isolated JGA is a tubulovascular unit
consisting of an individual CT and attached afferent arteriole and glomerulus (822, 1035, 1230). Many CTs, especially in the outer cortex, return to their parent glomerulus
and attached afferent arteriole (339, 1538). Three-dimensional tracing of 168 nephrons indicated that the terminal
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
423
RENAL AUTOREGULATORY MECHANISMS
portion of the distal convoluted tubule contacted the afferent arteriole of the parent glomerulus in 90% of shortlooped and 75% of long-looped nephrons of the mouse
kidney (1215). Such findings reaffirmed previous observations of 100% contacts in superficial nephrons and 60% of
midcortical nephrons of the rat kidney (339). Earlier serial
sections of 60 nephrons revealed that the late distal tubule
came within 3 ␮m of the afferent arteriole in 90% of superficial glomeruli, 87% from midcortical glomeruli, and 73%
of juxtamedullary glomeruli (67).
Carretero and associates have conducted a series of in vitro
and in vivo studies establishing the functional implications
of CT-GF regulation of afferent arteriolar vasomotor tone.
Tubular flow-dependent reabsorption of Na⫹ by the rabbit
or rat CT was mediated primarily by the amiloride/benzamil-sensitive ENaC (1230). Increased Na⫹ reabsorption
promoted epithelial cell production of PGE2 by COX-1 and
an EET by a cytochrome P-450 epoxygenase (1219, 1221),
both of which dilate the afferent arteriole. PGE2 induced
endothelium-independent dilatation via activation of the
EP4 receptor (1223). Earlier immunolabeling studies of
principal cells that express ENaC suggested colocalization
of membrane-associated PGE synthase with COX-1 but not
COX-2 (176, 1539). In contrast, COX-2 but not COX-1
was expressed in MD cells in both rats and mice. COX-2
was also expressed in the intra- and extraglomerular mesangium.
Tubular ANG II enhanced CT-GF, probably by simulating
Na⫹ reabsorption in the CT (1219). The effect of ANG II on
CT-GF was mediated via a protein kinase C (PKC)/NOX2/O2⫺ pathway (1227). In contrast to the conventional MDTGF, this suggests a vasodilator component to ANG II action due to CT-GF. Moreover, inhibition of tubular NOS
paradoxically increased the vasodilation induced by CTGF, suggesting that NO is not a mediator of the vasodilation but that NO rather inhibits Na⫹ reabsorption in CT,
thereby preventing the generation of the signal for the vasodilator response (1230). In vivo micropuncture studies
showed that increased tubular NaCl delivery activated the
CT-GF circuit to reduce the estimated PGC and offset the
MD-TGF-mediated constriction of the preglomerular vasculature (1558). CT-GF contributed also to acute resetting
of MD-TGF during changes in ECV and Na⫹ excretion
(1554). Such vasodilation and resetting were more pronounced in Dahl SS than in SHR rats consuming a high-salt
diet, thereby leading to increased PGC in the Dahl SS rats
(1555). In the absence of the MD-TGF circuit, activation of
CT-GF in vitro blunts the magnitude of the PP-induced
myogenic contraction of the afferent arteriole (Carretero
and associates, personal communication).
3. The juxtamedullary nephrovascular unit
Casellas and Navar (202) developed an in vitro JMN preparation perfused with low hematocrit blood or an albumin
424
solution to study the microvascular reactivity and the MDTGF of deep cortical nephrons. A rat kidney was sectioned
longitudinally and the papilla reflected to expose the overlying pelvic mucosa on the inner surface of the cortex where
the glomerular arterioles were visualized directly. Increased
RPP reduces the luminal diameters of the cortical radial
artery and afferent arteriole, but the diameter of the efferent
arteriole is unchanged or even increased (193, 199 –201,
580, 611, 1281, 1395, 1437). This preparation allows separation of the myogenic response from the MD-TGF without the potential complication attendant on the use of loop
diuretics by preventing MD perfusion by a tubular oil block
or papillectomy (200, 202, 686, 1020, 1281, 1395, 1437).
A doubling of RPP reduced afferent arteriolar diameter by
15–30% (1020, 1437), which was sufficient to yield nearperfect autoregulation of blood flow since resistance to flow
is a fourth power relation to the radius of the vessel. The
reduction in luminal diameter was limited to 7–10% after
elimination of MD-TGF by papillectomy, oil blockade to
halt tubular flow, or furosemide (FIGURE 3D) (1020, 1437).
Abolition of an active response with a Ca2⫹-free solution
increased arteriolar diameter by 20% when RPP was doubled (1020, 1437). Thus the modest reduction in arteriolar
diameter during a doubling of RPP after blockade of MDTGF still represents a considerable increase in vascular
tone. Afferent arteriolar autoregulatory constriction following increased RPP from 60 to 140 mmHg was attenuated by ⬃35% by blockade of MD-TGF, indicating the
myogenic mechanism accounts for ⬃65% (FIGURE 3D).
Both mechanisms were abolished by blockade of VOCCs
with nimodipine.
Measurements of hydrostatic pressures along the preglomerular vasculature during increased RPP indicated that
20% of the increase in resistance occured before the end of
the cortical radial artery, 65% by the end of the afferent
arteriole, and 80% along the composite preglomerular vascular tree (201). The terminal glomerular segment of the
afferent arteriole was twice as responsive as the upstream
segments (200, 1020, 1437). Afferent arteriolar blood flow
remained constant during changes in RPP above 95 mmHg
(580, 1437). Increasing RPP from 80 to 120 to 180 mmHg
increased PGC only modestly from 45 to 50 to 53 mmHg,
respectively (1281). Similar studies reported a 0.10- to
0.19-mmHg change in PGC per mmHg change in RPP
within the autoregulatory range (193, 201). Thus this preparation retains excellent autoregulation of nephron flow
and PGC within the physiological range.
4. The isolated perfused afferent arteriole
Edwards (375) developed a rabbit isolated, perfused renal
arteriolar preparation in 1983 to study the myogenic response and reactivity to vasoactive agents. Initial studies
reported maintenance of luminal diameter of cortical radial
arteries and afferent arterioles or constriction up to 11% to
increased PP compared with dilation of efferent arterioles
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
(375). Thereafter, the myogenic response was characterized
in isolated preglomerular vessels from dogs (562, 777,
1059), rabbits (697, 747), rats (700, 1224, 1637), and more
recently from mice (462, 728, 729, 861, 864, 866). A doubling of PP led to an 11% reduction in diameter of mouse
afferent arterioles that was complete within 5 s due to a
linear increase in active wall tension as a function of PP
(FIGURE 3F) (861). While this preparation excludes nonvascular paracrine factors, it requires an artificial perfusate free
of erythrocytes. Moreover, it is technically challenging to
measure the precise PP at the tip of the perfusion pipette.
Myogenic contraction, assessed as active wall tension, is
determined as the difference between the tension in physiological versus Ca2⫹-free EGTA-containing solutions.
Myogenic contraction was associated with increased
VSMC [Ca2⫹]i and increased ROS measured by fluorometric markers (864, 1224). The slope of the calculated regression of active wall tension on PP quantifies the myogenic
response with the intercept on the x-axis defining the
threshold PP for initiating a response (FIGURE 3F). The
threshold pressure was 35– 40 mmHg for the mouse isolated afferent arteriole (861, 864), 60 – 80 mmHg for the
preglomerular vasculature of the rat (700, 1637) and rabbit
(747), and 80 mmHg for the dog (562, 1059). The estimated or directly measured PGC in the euvolemic rat averaged 55 mmHg (42, 47, 58, 74, 87, 118, 199, 320, 334,
416, 1008, 1488) and was calculated to be similar in the
mouse (861). PGC was estimated to be 60 –70 mmHg in the
dog kidney (86, 1068, 1070, 1073, 1138, 1139) and 31
mmHg in the anesthetized rabbit (323). Thus these data
suggest that myogenic responses generally start at PPs below ambient levels of PGC, except in the rabbit, and may
thereby contribute to both dilation of the afferent arteriole
during reductions in RPP and constriction during increases
in PP.
5. The HNK
Steinhausen and associates (1413, 1415) developed the atubular, postischemic HNK preparation to facilitate direct
visualization of the renal cortical microcirculation and
glomeruli. This preparation allows intravital microscopic
measurements of vessel diameter and erythrocyte velocity in
an exclusive vascular preparation perfused with blood or
artificial solutions. During the development of hydronephrosis, the kidney parenchyma becomes progressively
thinner due to tubular atrophy, resulting in a tissue sheet of
cortical vessels of ⬃200 ␮m thick. Approximately 6 – 8 wk
after unilateral ureteral obstruction, there was complete
tubular atrophy (162, 1415). This preparation has been
used to investigate mechanisms of renin release and electrophysiological characteristics of JGC of the afferent arteriole
(162, 164). In addition to being readily amenable to study
of electrophysiology of the renal microcirculation, the
HNK provides an opportunity to investigate the myogenic
response of various segments of the preglomerular vasculature in the absence of input from either MD-TGF or CT-GF.
Videometric measurements were made of the diameter of
renal arterioles during perfusion in situ with blood or in
vitro with an artificial saline solution (1413, 1415). Loutzenhiser and associates (586) have used the HNK extensively to characterize the myogenic response in the absence
of any TGF. Increased RPP contracted cortical radial arteries and afferent arterioles but not efferent arterioles (FIGURE
3E) (586). Furthermore, frequency domain analysis reported a relatively well-maintained myogenic mechanism
with a threshold for activation of 60 mmHg (287) and
preglomerular vasoconstriction of 7–23% with a doubling
of RPP (586) that was complete within 5–10 s (802, 927,
931). The EM of cortical radial arterial and afferent arteriolar VSMCs averaged ⫺40 mV at 80 mmHg, which was
just below the threshold required to initiate a myogenic
response (928). Overall, RBF autoregulation in this preparation, which lacks MD-TGF, was weak when perfused
with a salt solution devoid of colloid or erythrocytes. This
was apparent from a steady-state AI of ⬃0.85 over the RPP
range of 60 –140 mmHg and a dynamic autoregulatory
fractional admittance gain of only ⬃0.8 in the PP range of
100 –140 mmHg (287). These values are considerably
higher than the same determinations in efficiently autoregulating intact kidneys during inhibition of MD-TGF (287,
297, 749, 753, 760). The HNK preparation suffers from
some other limitations including the low viscosity of a cellfree perfusion solution which influences normal NO production and precludes its uptake into oxyhemoglobin. Nevertheless, the preparation responded well to ACh-induced
NO release (1486).
C. Interactions Between MD-TGF and the
Myogenic Mechanisms
Interactions between the myogenic and MD-TGF mechanisms provide renal autoregulation with unique opportunities to regulate preglomerular vascular tone (200, 231, 245,
285, 625). The most distal portion of the afferent arteriole
was the major effector site for both the myogenic and the
MD-TGF responses (200). Studies have demonstrated positive, negative, or absent interactions between these two
mechanisms. The participation of a glomerular feedback
includes both the conventional MD-TGF and the more recently described CT-GF. It is firmly established that the
MD-TGF influences the myogenic response, which in turn
impacts the MD-TGF (244 –246, 630, 1163, 1205, 1341,
1401, 1402, 1547).
Interactions between the oscillating frequencies of the myogenic and the MD-TGF mechanisms have been observed in
whole kidney blood flow and single-nephron studies. Early
studies of myogenic fluctuations in SNGBF in the superficial
cortex revealed large amplitude oscillations at 0.1– 0.2 Hz
in response to inhibition of MD-TGF with furosemide,
whereas the myogenic oscillations were strongly damped
when MD-TGF was stimulated maximally by tubular per-
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
425
RENAL AUTOREGULATORY MECHANISMS
fusion (1636). This suggested that MD-TGF tonically inhibits the amplitude of intrinsic vascular tone driven by the
myogenic mechanism. Also inhibited were MD-TGF-dependent communications with adjacent afferent arterioles
and glomeruli (626). Recent wavelet analysis of blood flow
and tubular pressure confirmed that MD-TGF modulated
the frequency and amplitude of oscillations initiated by
myogenic mechanisms (1400). These dynamic interactions
occur within and between nephrons and are nonlinear and
time-varying (405, 630, 973, 975). Conversely, delivery of
the tubular fluid to the MD and the CT is altered by the
myogenic mechanism and thereby modulates TGF responses. Such coupling has been modeled extensively with
descriptions of nonlinear modulation of the amplitude and
frequency of myogenic oscillations by MD-TGF (244, 245,
395, 875, 973, 973, 975, 976, 1205, 1258, 1400). These
interactions appeared to vary between cortical and medullary nephrons (974) and may contribute to fluctuations in
the RPP set-point for RBF autoregulation (1172). The interactions were more variable in SHR than in SD rats (247,
626). Based on simulation studies (395), Holstein-Rathlou
and associates concluded that the myogenic response needs
to be active to have a MD-TGF mechanism participate in
RBF autoregulation. Indeed, there are examples when MDTGF is present and yet RBF autoregulation is very weak to
absent.
However, studies using frequency domain analysis have
concluded that autoregulatory responses in one afferent arteriole can generate both positive and negative interactions
with others (231, 245, 246, 628). MD-TGF can modulate
both the frequency and amplitude of oscillations arising
from the myogenic mechanism. Most nephrons in normotensive rats originating from a common renal cortical radial
artery had highly regular periodic synchronization of MDTGF and myogenic modes. However, the interactions in
SHR were more complex with transient, irregular coupling
and synchronization (231, 875, 1205, 1401). Also, the lowfrequency (0.01 Hz) oscillatory component of renal autoregulation, perhaps representing the third component,
modulateed both the myogenic response and MD-TGF activity (1163, 1341, 1375), with weaker interactions in SHR
than in SD rats (1375).
In vivo micropuncture studies by Schnermann and Briggs
(1308) reported that the maximum MD-TGF-mediated reduction in the estimated PGC was 13 mmHg when RPP was
120 mmHg, but was reduced to 3 mmHg when RPP was
lowered to 80 mmHg. Thus the magnitude MD-TGF increased in parallel with RPP, and effective autoregulation of
the estimated PGC was strongly dependent on tubular flow
past the MD. These results demonstrate two-way positive
interactions between myogenic and MD-TGF mechanisms
in vivo. The MD-TGF response is required for a full myogenic response, and conversely, the degree of myogenic contraction regulates the reactivity of MD-TGF.
426
Recent studies have provided further evidence of a positive
reinforcement between the MD-TGF and myogenic mechanisms. Dynamic studies using frequency analysis in the rat
found that MD-TGF inhibition with furosemide attenuated
characteristics associated with myogenic reactivity (reducing the slope of gain reduction and the associated phase
peak) (1358), suggesting removal of positive MD-TGF
modulation of the myogenic mechanism. In the conscious
mouse, furosemide eliminated the frequency domain signature of MD-TGF and reduced the shoulder operating frequency of the myogenic mechanism (from 0.3 to 0.2 Hz),
indicative of a slowing of the myogenic response (668).
Dynamic studies by Walker and Navar (1547) on the rat
JMN preparation in the time domain revealed a positive
interaction between MD-TGF and the myogenic response
of afferent arterioles. A rapid step increase in RPP produced
a prompt but relatively short-lived (⬃13 s) monotonic myogenic vasoconstriction of the afferent arteriole when MDTGF was inactivated by papillectomy, whereas there was a
biphasic contractile response with the initial myogenic response lasting ⬃24 s when MD-TGF was intact, indicating
positive interaction between these two mechanisms (1547).
With MD-TGF operational, a biphasic response included a
decline in afferent arteriolar diameter after 24 s, albeit at a
slower rate, to reach a 26% reduction at 93 s. The initial
myogenic response had a time constant of 6 –11 s (0.16 –
0.09 Hz), whereas that of MD-TGF was 37 s (0.027 Hz).
The delay in the initiation of the myogenic response was 6 s,
and the rate of myogenic contraction was the same whether
MD-TGF was engaged or inoperative. There was no evidence for a very slow third mechanism in this study. Autoregulation of the in vitro JMN preparation was stronger in
the initial but not the sustained phase after MD-TGF was
activated by increasing NaCl delivery to the MD (662). It is
noteworthy that inhibition of MD-TGF with furosemide
did not affect the myogenic response in vivo after NOS was
inhibited (749, 750), a finding consistent with an action of
furosemide to inhibit MD-TGF and NO generation and
modulation of myogenic tone.
However, several studies suggested weak to no interaction
between MD-TGF and the myogenic mechanism. Transfer
function analysis revealed that furosemide inhibited the
MD-TGF signature at 0.03– 0.05 Hz with a weak effect of
the myogenic response between 0.1– 0.2 Hz in the rat, reducing the gain at frequencies below 0.15 without changing
the peak of the phase angle (19). Similar inhibition of MDTGF had little effect on the myogenic response in the dog
(760) or mouse (299), but furosemide paradoxically augmented the myogenic response in adenosine A1 receptor
deficient mice lacking MD-TGF (751). However, inhibition
of MD-TGF by ureteral obstruction in the rat abolished the
MD-TGF signature of admittance gain without affecting
dynamic features of the myogenic response (297). Consistent with a lack of effect of MD-TGF on myogenic respon-
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
siveness, activation of MD-TGF by inhibition of proximal
NaCl reabsorption by acetazolamide to increase MD delivery of NaCl had little influence on the myogenic mechanism
in vitro (662) or in vivo (749). An increased myogenic tone
elicited by reducing perirenal pressure in the isolated perfused rat kidney was unaffected by MD-TGF inhibition
with furosemide (265). Thus myogenic vasoconstriction
can function independently from MD-TGF.
RVR responses to a sudden, step increase in RPP gave the
impression that the initial myogenic response was enhanced
during inhibition of MD-TGF with furosemide, increasing
its participation in the first 10 s from 40 to 60% of the total
response in the dog (751, 753) and from 40 to 55% of total
efficiency with an enhanced response rate in the rat (749,
750). However, the appearance of coupling between MDTGF and myogenic mechanisms depends crucially on the
protocol used and the operational status of MD-TGF while
RBF is normal and PP is varied within the autoregulatory
range. During frequency analysis to generate a transfer
function of admittance gain, spontaneous or forced oscillations of RPP continually change tubular flow past MD cells
to modulate MD-TGF activity in positive and negative directions. In contrast, analysis based on a sudden step change
in RPP usually starts with a reduced PP and diminished
tubular flow and MD-TGF activity. The resultant vasodilation tends to offset or counter the stimulus for rapid myogenic vasoconstriction when RPP increases. This may explain the increased strength and speed of the myogenic response observed following a rapid increase in RPP when
MD-TGF is inhibited since the opposing preexisting vasodilator input to arterial tone is removed. This spurious result is less evident during inhibition of the more symmetrical
activity of MD-TGF during frequency analysis of admittance gain. Inhibition of MD-TGF by acute volume expansion tends to increase the contribution of the myogenic
mechanism (from ⬃33 to 50%) of total autoregulatory adjustments in RVR to a sudden step increase in RPP in the
wild-type mouse (462, 505). Again, this could be explained
by removing the restraining effect of preexisting MD-TGFinduced vasodilation at low tubular flow on the initial myogenic vasoconstriction following the increase in RPP. Total
steady-state autoregulatory efficiency was unaffected by
such inhibition of MD-TGF. Collectively, these results suggest that the myogenic mechanism is more designed to protect against increases in BP and that the MD-TGF negativefeedback system is more supportive of RBF and GFR during
reductions in BP.
Furosemide might cause some renal vasodilation by increasing NaCl delivery to the CT and stimulating CT-GF. Afferent arteriolar vasodilation mediated by CT-GF attenuated
the strength of vasoconstriction by MD-TGF and myogenic
mechanisms (1554, 1555, 1558) (O. A. Carretero, personal
communication). However, furosemide had little effect on
the contribution of the putative third mechanism in either
the dog or rat (749, 753), although it was abolished by
furosemide in a mouse study (751). Furosemide may also
impact autoregulation by inhibiting Na⫹-K⫹-2Cl⫺ transport in VSMC to affect preglomerular vascular tone (522,
835, 1124, 1568). Effects of loop diuretics on NKCC1
transport by VSMC and their impact on myogenic tone of
the afferent arteriole in vitro are discussed in section IVA.
Four mechanisms have been proposed to account for observed positive interactions between the MD-TGF and the
myogenic responses. First, a MD-TGF-induced constriction
of the terminal afferent arteriole should increase the upstream intravascular pressure and thereby enhance the signal for a myogenic contraction in more proximal afferent
arteriolar segments (200, 1021). Second, MD-TGF triggered a Ca2⫹ wave that spread from the MD throughout the
JGA to the parent afferent arteriole that could enhance
myogenic responsiveness (1176). Third, membrane depolarization of the afferent arteriole by an active MD-TGF
should enhance Ca2⫹ entry through VOCCs and thereby
myogenic contractions. Fourth, MD-TGF-induced depolarization of the afferent arteriolar VSMCs can be transmitted
upstream (976, 1176) to constrict neighboring afferent arterioles stemming from the same cortical radial artery (231,
762). Electrical stimulation of the distal afferent arteriole
initiated electrical and Ca2⫹ entry responses that were conducted upstream along the afferent arteriole (527) and cortical radial artery (1279), presumably mediated by gap
junctions (527). Depolarizing current caused vasoconstriction and hyperpolarizing current caused vasodilation
(1414). High KCl-induced depolarization also elicited a
constrictor response of upstream arterioles and neighboring
afferent arterioles, with a long mechanical length constant
of ⬃325 ␮m (1279, 1544). The strength of internephron
coupling was greater in SHR than in SD rats (1544). Such
coupling could account for synchronized MD-TGF-mediated contractions in pairs or triplets of nephrons (630).
MD-TGF initiated stronger interactions with afferent arterioles originating from a common cortical radial artery in
SHR than in SD rats (231).
The molecular mechanisms by which MD-TGF interacts
with the myogenic mechanism are not certain. Positive interactions may reflect vasoconstrictor signaling Ca2⫹ pathways or perhaps involving ATP and P2X1 receptors or adenosine and A1 receptors, as discussed in section IIA3, A and B.
Alternatively, as is discussed in section VIB, NO generated
by neuronal NOS in MD cells during activation of MDTGF could account for a negative interaction since endogenous NO markedly inhibits both the strength and the
speed of the myogenic response.
In summary, it is clear that there are communications and
interactions among vascular and tubular mechanisms mediating autoregulation in the kidney, but their physiological
significance remains elusive. Both reinforcing and opposing
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
427
RENAL AUTOREGULATORY MECHANISMS
interactions between myogenic and MD-TGF mechanisms
have been reported. This may relate to coupling based on
the balance of positive and negative modulators of vasomotor tone of afferent arterioles, which can be generated by
activated MD or CT cells, as summarized in FIGURE 6 and
discussed in more detail in section VI. The complexity is
increased further by the fact that the degree of information
transfer can vary in a nonlinear manner over time and MDTGF may contribute intermittently to autoregulation. The
cellular events and consequences of interactions between
MD-TGF as well as CT-GF mechanisms and the myogenic
response are poorly understood, but modulation of NO and
Ca2⫹ signaling appear to be involved. As is discussed in
section VIC, NO modulates the renal myogenic response
and interactions between tubular and vascular mechanisms,
but the primary source is still controversial.
D. Kinetics of the Myogenic Response
Both the magnitude and the kinetics of the myogenic response vary among different circulatory beds. There is a
tendency for the speed of the myogenic response to be greatest in the smallest diameter arterioles. The renal vasculature
has the fastest and most complete response, followed by the
cerebral and then mesenteric arterioles.
The myogenic response of renal afferent arterioles of mice,
rats, and dogs (internal ID: 8 –20 ␮m) is often complete
within 5–10 s, whereas in other vascular beds it ranges from
⬃10 –15 s for small cremaster muscle arterioles (15–25 ␮m)
(613, 613, 615) to 60 –120 s for larger mesenteric and cremaster arterioles (⬃100 ␮m) (303, 740, 1671) and up to 5
min for mesenteric arteries and skeletal muscle arterioles
(⬎180 ␮m) (243, 667, 1002). This diversity is probably
related to a specific set of VSMC mechanosensation-contraction signaling systems. Whereas the depolarization and
Ca2⫹ entry through VOCCs are ubiquitous, distinct upstream signaling systems appear to provide tissue-specific
mechanisms of adaptation to local metabolic needs involving the supply of oxygen and fuels that are triggered by
changes in BP.
The myogenic response in the renal microcirculation is engaged within 1 s and is complete in ⬍5–10 s in intact rodent
kidneys (265, 462, 505, 749 –751, 756, 1640), mouse isolated afferent arterioles (861) or rabbit (747), or rat afferent
arteriole of the JMN (1547) or the HNK preparation (927).
This time frame corresponds closely to the 0.1– 0.3 Hz frequency band for the myogenic admittance transfer function
(FIGURE 3C) (109, 289, 297, 629, 749). Slower responses of
40 s were reported in one study of rat pressurized JMN
arterioles (1637) and of 60 –120 s in hamster preglomerular
arterioles transplanted to the cheek pouch (473).
As already discussed in the previous section, Walker and
Navar (1547) detected two components in the autoregula-
428
tory contractile response to a single step increase in PP of
the JMN preparation. The prolonged delay (⬃6 s) in the
initiation of the myogenic response reported in those studies
of the JMN preparation may reflect methodological problems, since much more rapid responses have been reported
in other studies among isolated afferent arterioles or in the
HNK preparation.
Loutzenhiser et al. (927) used the rat HNK preparation to
characterize the kinetics of the myogenic response in the
absence of any tubular influence. The vasoconstrictor response to an 80-mmHg rise in RPP was a monotonic contraction after a delay of ⬃0.3 s, with a time constant of 4 s.
The onset of the myogenic response of the HNK, based on
changes in arteriolar wall diameter, was very rapid and
averaged ⬃0.30 – 0.35 Hz, which is considerably faster than
values reported in vivo. A possible contributor to this unusually rapid response is the use of a low viscosity, colloid,
and RBC-free saline perfusion solution as well as the lack of
tubular modulatory agents. The vasodilator response to an
80 mmHg reduction in RPP was slower and more complex.
After a delay of 1 s, the response was biexponential with
time constants of 1 and 14 s, requiring almost 25 s for
complete relaxation. Based on those results, the authors
concluded that rapid pulsations should evoke sustained vasoconstriction due to the faster contraction when RPP rose
compared with the more sluggish vasodilatation when PP
fell. Consistent with this prediction, the magnitude of afferent arteriolar contraction in the HNK preparation was related to the peak systolic pressure for PP oscillations between 1 and 6 Hz. These investigators proposed that the
primary function of the myogenic response at the normal
heart rate frequency is to protect glomerular capillaries
from the damaging effects of the systolic BP rather than to
maintain a constant GFR to regulate body fluid homeostasis
(933, 934).
In contrast to these results for the HNK preparation, most
whole kidney studies of frequency domain analysis of renal
vascular admittance have reported a frequency for the myogenic response of 0.20 – 0.25 Hz (FIGURE 3C), as discussed
earlier in section IIA5. Moreover, Just and Arendshorst
(749, 751) reported that intact kidneys had a similar rate of
myogenic contraction and relaxation to a 20-mmHg step
change in RPP. Both responses were complete within 8 s
(749, 751). The myogenic contraction to an increase in RPP
in intact kidneys occurred slightly earlier, with a shorter
delay time (0.39 vs. 0.53 s) and contributed a smaller fraction of total autoregulation of ⬍37% compared with the
myogenic relaxation to a decreased RPP which was ⬍52%
(750). Moreover, the time for complete contraction of 5.1 s
was longer than for complete relaxation of 2.6 s. Accordingly, the in vivo renal myogenic response buffered increased and decreased RPP similarly and thus should respond primarily to the mean RPP, rather than preferentially
to the systolic BP. Support for this conclusion comes from
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
earlier studies that failed to detect an influence of arterial
pulse pressure on RBF autoregulation (1238, 1344). The
observed dynamic values for autoregulation in the studies
of Just and Arendshorst (750) were likely underestimates
since the in vivo changes in RPP, although abrupt, were not
instantaneous.
strictors, e.g., ANG II or AVP, to increase RPP had no
major effect on the renal myogenic response (750, 1193,
1563). The rate of change in RPP is another factor that may
modulate renal autoregulatory responses. During stepwise
increases in RPP in conscious rats, larger changes in RVR
were associated with more rapid changes in PP (423).
These striking differences reported for the myogenic contraction between the in vitro atubular HNK and in vivo
intact kidney preparations may relate to the presence of
active tubular cells only in the latter, since MD metabolites
impact on VSMC kinetics. Specifically, NO production affects the myogenic mechanism, as discussed in section VIC.
Moreover, activation of MD cells can accelerate the rate of
relaxation associated with the generation of PGE2 (218,
1036, 1334) and NO (298, 1036). Indeed, Loutzenhiser et
al. demonstrated later that adding exogenous NO or cGMP
to the HNK preparation to more closely simulate the in vivo
paracrine modulation, preferentially reduced the delay in
the vasodilation response to a reduced RPP (R. Loutzenhiser, personal communication), thereby reducing the time
difference between myogenic relaxation and constriction
which more closely approximated to those observed in vivo.
The myogenic response appears to be somewhat faster in
conscious than in anesthetized animals. Studies by Cupples,
Bidani, Griffin and colleagues reported the resonance frequency of the myogenic mechanism to be 0.23– 0.25 Hz in
conscious rats (886, 1193), compared with the general finding of 0.18 – 0.21 Hz in rats anesthesized with isofluorane
or barbiturates (109, 297, 499, 597, 749, 1274, 1358,
1569, 1570, 1572). There were similar responses during
anesthesia with thiobutabarbital (597, 1274) and pentobarbital (297, 749), but halothane slowed the operating frequency of the myogenic mechanism in SD, SHR, and WKY
rats to 0.10 – 0.15 Hz (230, 289, 631, 886, 1636).
The resonance peak in admittance gain of dynamic autoregulation is an index of the natural frequency of the system
(628). The dynamic autoregulation of arterial blood flow in
euvolemic Wistar and SHR rats was excellent in the renal
and in the superior mesenteric circulation, although the
myogenic response was faster (0.22 Hz) in the denervated
kidney than in the denervated gut (0.13 Hz) (7). This was
confirmed in two other studies (886, 1572). In contrast,
blood flow to the hindquarters lacks any dynamic autoregulatory signature.
Cupples and associates (1572) reported that the range of
RPP tested in the HNK preparation perfused with artificial
medium determined the speed of the myogenic response
assessed as transfer function admittance gain. Thus an increase in mean RPP from 80 to 140 mmHg increased the
myogenic operating frequency from ⬃0.265 to 0.37 Hz. In
a similar in vivo study in Wistar rats, inhibition of NOS
with L-NAME increased the RPP from 115 to 140 mmHg
and increased the myogenic resonance frequency from 0.21
to 0.25 Hz. This change was attributed primarily to PP since
it was prevented by maintaining PP by aortic clamping during NOS inhibition. This pressure sensitivity was unique to
the kidney as the myogenic resonance frequency in the mesenteric circulation (0.15– 0.16 Hz) was unaffected by LNAME hypertension. In another study by the same laboratory, however, hypertension (115–145 mmHg) induced by
L-NAME was reported to increase the strength of the myogenic response (both the slope and the amplitude of admittance gain in the 0.1– 0.2 Hz window) without affecting its
speed; the myogenic resonance frequency was unchanged at
0.18 – 0.20 Hz (1569). Other studies employing vasocon-
Neurohormonal and paracrine agent also modulate the
speed and strength of the myogenic response, as exemplified
by the effect of NO, discussed earlier and first reported by P.
Persson and associates (1626). ANG II is reported to augment the myogenic resonance peak from 0.23 to 0.27 Hz in
rats and increases the fractional gain by 25% and the slope
of the reduction in gain, whereas phenylephrine (PE) decreases fractional gain by 30% and the slope of the gain
reduction without affecting the resonance peak (0.24 – 0.25
Hz). PE had little effect on MD-TGF (1193). The myogenic
resonance peak also was increased by ANG II in conscious
dogs (755). As is discussed in section VIC, NOS inhibition
in the intact kidney markedly accelerates and magnifies the
myogenic mechanism.
Expression of different isoforms of smooth muscle myosin
heavy chain (MHC) has been proposed to explain the more
rapid contraction of the afferent arteriole compared with
the efferent arteriole. The B-isoform of MHC inserts into
the ATP binding pocket to activate VSMC contraction. The
afferent arteriole expressed predominantly the rapidly recycling MHC-B isoform, whereas the efferent arteriole expressed only the slower-cycling, alternatively 5’-spliced
MHC-A isoform (1161, 1361). However, the rate of afferent arteriolar contraction was unaltered in mice with a mutated form of MHC-B (1161). Thus variable expression of
VSMC MHC isoforms could not explain the differential
speed of contraction of afferent and efferent arterioles.
In summary, myogenic tone changes rapidly with PP and
adjusts the RVR to the systolic or mean BP according to the
preparation used. The more rapid constriction than relaxation reported in preparations lacking tubules likely relates
to the absence of MD-derived NO and PGE2 that are implicated in accelerating relaxation in vivo. The times of contraction and relaxation in vivo, when measured in response
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
429
RENAL AUTOREGULATORY MECHANISMS
to a step increase or decrease in RPP, are more symmetrical.
Moreover, the myogenic response appears to be somewhat
faster in conscious animals.
E. Vasomotion
The myogenic mechanism of most microvessels generates
oscillations in vascular tone and organ blood flow termed
“vasomotion” (1, 1169) which oscillate in vivo with frequencies of 0.05– 0.2 Hz, likely representing summation of
vasomotion in many individual small arteries (2, 854, 1087,
1408). The underlying mechanisms are still not fully understood and may even differ between vascular beds. Although
vasomotion can be induced by PP and involve similar Ca2⫹
signaling pathways as autoregulation, it has some unique
features.
Vasomotion entails the coordinated activation of VSMCs
whose oscillations in EM and [Ca2⫹]i evoke rhythmic contractions. Phasic cycles in Ca2⫹ signaling were mediated
both by periodic changes in Ca2⫹ influx and in Ca2⫹ mobilization from intracellular stores (1, 979). One proposed
sequence of events is that activation of either IP3 or RYRs/
channels on the sarcoplasmic reticulum mobilizes Ca2⫹ that
activates the plasma membrane ClCa channels to allow Cl⫺
efflux that depolarizes the membrane. The resultant activation of L-type VOCC allows Ca2⫹ entry that increases
[Ca2⫹]i further to promote MLC phosphorylation and
cross-bridge formation leading to contraction (1, 123, 529,
531). Ca2⫹ sparks (i.e., highly localized intracellular Ca2⫹
transients generated by RyRs acting on the sarcoplasmic
reticulum adjacent to the plasma membrane) also can elicit
global Ca2⫹ waves and oscillations (845). An alternative
view is that Ca2⫹ mobilization activates BKCa to cause hyperpolarization and reduce Ca2⫹ entry through L-type
VOCC and thereby cause vasodilation. Subsequently, Ca2⫹
is taken up into the sarcoplasmic reticulum via Ca2⫹-ATPase to restore the basal tone. A third scenario derives from
studies of isolated irideal arterioles whose vasomotion was
independent of EM and the endothelium and is mediated by
IP3 induced changes in [Ca2⫹]i (544).
A periodic depolarization of VSMCs generates electrical
signals that propagate along a vessel via cell-to-cell-coupling involving gap junction connections to cause synchronized oscillations in global Ca2⫹ signaling in the vasculature network (979). Oscillations are evoked by NE, KCl, or
other agents that depolarize the EM (1333, 1349). They are
dependent on gap junctions and Cxs, notably Cx37 (981).
Rhythmic Ca2⫹ signaling of VSMCs may oscillate rapidly
(⬍10 s) but can be dampened by neurotransmitters, vasoactive hormones, or local autocrine and paracrine factors
such as NO. Release of cGMP in some vascular beds may
facilitate Ca2⫹ activation of ClCa channels, perhaps via bestrophin 3, to reduce EM, thereby locking the gap junctions
between VSMCs into electrical phase (979).
430
The renal afferent arteriole studied in vitro is viewed normally as a damped oscillator that does not exhibit spontaneous vasomotion. However, in vivo oscillations of superficial cortical blood flow are readily detected, as mentioned
earlier in section IIA4B. Indeed, measurements of SNGBF
on the kidney surface in vivo reveal two oscillations driven
by the more rapid myogenic mechanism (0.15– 0.25 Hz)
and the slower MD-TGF (0.02– 0.05 Hz) (289, 625, 629,
1340, 1400, 1636). MD-TGF-mediated oscillations in PPT
have a period of 25–35 s. This slow process may oscillate
because of the relatively long delay in filtered tubular fluid
reaching the MD segment at the end of Henle’s loop to elicit
a change in arteriolar tone (626, 627, 631, 892, 893). The
faster myogenic oscillations observed in vivo with a period
of 5–10 s represent a rhythmic oscillation of vascular tone
and diameter of preglomerular vessels mediated by a coordinated and periodic fluctuation of intrinsic VSMC activity.
Afferent arteriolar oscillations of ⬃0.2 Hz with amplitudes
of ⬃1 ␮m have been detected in these renal vessels (289).
The cellular mechanisms for vasomotion of the preglomerular vasculature in vivo are thought to resemble those of the
myogenic mechanism, but have been studied more extensively in extra-renal vessels, as reviewed below.
Vasomotion was diminished in hamster skin by anesthetics
(268, 869, 1087), primarily by alternating the amplitude of
vasomotion in large arteries (653). The operating frequency
of the myogenic mechanism in the renal vasculature of SD
rats, based on transfer function gain, was 0.18 – 0.20 Hz
under barbiturate or isoflorane anesthesia, but was reduced
to ⬃0.14 Hz under halothane anesthesia (289).
Stimulation of Ca2⫹ entry in VSMCs of the HNK preparation through L-type VOCC by BAY K 8644 induced
large, 10-␮m oscillations of afferent arteriolar diameter
with a period of ⬃20 s which were dependent on RyRmediated Ca2⫹ mobilization from the sarcoplasmic reticulum (1411, 1443). An increased Ca2⫹ release was associated with an increased and more rapid constriction
which depended on Rho-kinase (1340). NE can stimulate
afferent arteriolar vasomotion and cycling of [Ca2⫹]i in
VSMCs (930, 1637). Pacemaker areas enriched in L-type
VOCCs occur at branch points along the cortical radial
arteries (477).
The requirement of the endothelium for vasomotion and
the influence of NO varied among vascular beds (1, 817).
NO acted as a stimulant in some, whereas in others, endothelium-derived factors dampened or even abolished vasomotion by desynchronizing the Ca2⫹ signals in VSMCs. For
example, vasomotion in mesenteric basilar or hamster
cheek pouch arteries was stimulated by increased NO or by
reduced KATP channel activity (1, 312) and abolished by
endothelial removal or NOS inhibition (312, 530). NO and
cGMP can modulate vasomotion at gap junctions to synchronize Ca2⫹ oscillations by inhibiting Ca2⫹ mobilization,
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
Ca2⫹ entry, or Ca2⫹ sensitivity (1, 530). NOS inhibition
abolished vasomotion in hamster cheek pouch arterioles by
activating K⫹ channels. cGMP facilitated vasomotion in
some (1, 477), but not other studies (312). Myogenic tone
and vasomotion were diminished in small mesenteric arteries in mice lacking caveolae with deficient NO generation
(1430).
In contrast, NO suppressed vasomotion of the rat middle
cerebral artery (336, 1645) independent of sGC but mediated by RyRs and VOCCs resulting in less synchronous
Ca2⫹ oscillations and propagating Ca2⫹ waves (1645).
TRPC1 channels in rat cerebral arteries generate oscillations in EM generated by Na⫹ entry following activation of
PLC and Ca2⫹ entry through L-type VOCCs (1620). Endothelial dysfunction in the cerebral vasculature led to lowfrequency (4 –12 cpm) spontaneous oscillations of blood
flow (635). Inhibition of oxidative stress caused oscillations
in cerebral blood flow mediated by TxA2 and TP receptors
(24, 635).
The endothelium of the small cerebral basilar arteries modulated vasomotion by direct myoendothelial coupling by
Cx37 and Cx40 that synchronized [Ca2⫹]i waves in adjacent VSMCs (542). Endothelial signaling by EDHF or TP
receptors, but not by AT1 or adrenergic receptors, also elicited vasomotion in the hamster cheek pouch microcirculation (1532). Pial arteriolar vasomotion was modulated by
somatosensory cortical neuronal activation (1536) and was
blocked by NE or ACh (1239). NOS inhibition or dibutyryl
cGMP failed to resynchronize [Ca2⫹]i waves, but K⫹ channel blockade depolarized constricted vessels and abolished
vasomotion (542).
Communication between endothelial cells was enhanced by
gap junctions consisting of Cx37, Cx40, and Cx43,
whereas VSMCs were more weakly coupled by Cx37. The
two cell types were linked by myoendothelial gap junctions
using Cx37 and Cx40. Electrotonic coupling of EM spread
a hyperpolarization wave along endothelial and VSMCs
accompanied by a propagated Ca2⫹ wave. The endothelium was essential for vasomotion in some (530, 651, 992,
1117, 1169), but not all (544, 868), studies of mesenteric
arteries and may even be promoted by removal of the endothelium or inhibition of NOS (98, 970), which can desynchronize Ca2⫹ signaling in VSMCs (1346),
PP regulated vasomotion and autoregulation. An increase
in PP of mesenteric or cerebral arteries (528) reduced the
amplitude but increased the frequency of rhythmic contractions, independent of the endothelium (529, 1135, 1525).
This was dependent on RyR-mediated Ca2⫹ mobilization
and Ca2⫹ entry through VOCCs (529, 1525). Moreover,
cyclic variations in BP in vivo modulated the arterial diameter and VSMC [Ca2⫹]i (818, 819).
Hamster cheek pouch arterioles had spontaneous vasomotion with oscillations of 20 mV in EM and 14 ␮m in diameter, which were most frequent in the smallest arterioles
(143, 268, 294). Isolated hamster cheek pouch and rat cerebral and mesenteric arterioles had both rate- and pressure-sensitive vasomotion (303, 1135, 1525). Small-amplitude, positive-pressure steps elicited monophasic constriction, whereas higher pressure steps elicited biphasic
constrictions that were only partially dependent on a myogenic mechanism.
Measurements of oscillations of blood flow have been used
to study myogenic mechanisms in humans. Spectral analysis
of microcirculatory peripheral blood flow in the skin of the
hand and sternum by laser Doppler flowmetry showed two
principal spectral rhythms of 0.04 – 0.15 Hz and ⬃0.17 Hz
(907), which reflected myogenic activity (1262). Whereas
the skin of the leg had more myogenic (0.05– 0.14 Hz) activity than the forearm at 33°C (622), this was reversed by
warming to 42°C. Insulin increased the amplitude of the
myogenic vasomotion in human skin, (1262), as in rat skeletal muscle (1085). A large-amplitude ⬃0.1 Hz oscillation
in cerebral cortical perfusion was observed in conscious
humans in association with wave-like propagation and oscillations in the diameter of pial arterioles (1211). The lowfrequency oscillations (⬃0.1 Hz) of cerebral blood flow
were reduced in the aged (1533).
In summary, a key component of vasomotion is spontaneous oscillation of EM that is generally mediated by either
ClCa or BKCa channels to modulate Ca2⫹ entry through
L-type VOCCs. This is followed by a prominent role for
RyRs in triggering Ca2⫹ mobilization. The contribution of
voltage-dependent channels and the endothelium varies
among vessels (543). Cell coupling by gap junctions enhances synchronization of oscillations in Ca2⫹ release
among adjacent cells, at least in isolated vessels in vitro. The
speed of the myogenic mechanism varies by vascular bed. It
is generally most rapid in the renal microcirculation (0.2
Hz), followed by the cerebral and mesenteric vascular beds
(⬃0.15 Hz), with slower vasomotion in the cutaneous network (0.05– 0.15 Hz). Smaller arteries and arterioles have
higher frequency oscillation than large vessels. Recent studies of oscillation in blood flow have given promise of a new
technique to study myogenic mechanisms in the human circulation.
III. MECHANOSENSITIVE MECHANISM
INITIATING THE MYOGENIC RESPONSE
Mechanotransduction implies an integrated interaction of
the extracellular environment with the three-dimensional
structure of the vessel wall that links changes in membrane
stretch to intracellular signaling mechanisms (678, 899).
Pressure-induced deformation of extracellular matrix proteins and their cell surface receptors such as integrins are
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
431
RENAL AUTOREGULATORY MECHANISMS
thought to initiate contraction and cytoskeletal remodeling
through modulation of ion channels, membrane depolarization, increased [Ca2⫹]i, and actomyosin crossbridge cycling (304, 617). Thus integrin interaction with the extracellular matrix transduced extracellular mechanical forces
such as shear and tension into intracellular biochemical
signals to affect VSMC contractility (64, 213, 304). Myogenic constriction of VSMCs resulted from activation of
VOCCs secondary to reduced EM (301, 562, 810). Activation of mechanosensitive ion channels initiated pressureinduced depolarization and subsequent Ca2⫹ signaling and
myogenic constriction.
was initiated by localized RyR-dependent Ca2⫹ sparks that
amplified global [Ca2⫹]i waves (214). Blockade of RyRs
inhibited integrin-mediated myogenic constriction of afferent arterioles (FIGURE 8) (64). RyR-dependent Ca2⫹ signaling was independent of Ca2⫹ entry via L-type VOCCs in the
renal (214), but not in nonrenal vessels (525). Ca2⫹-induced Ca2⫹ release in the afferent arteriole may be more
critical in the initiation of the myogenic response than in the
sustained phase when Ca2⫹ entry through VOCCs is essential. Mechanical force that perturbed the interactions between fibronectin and ␣5␤1 integrins prolonged the oscillatory Ca2⫹ signals and the tractional force developed by
renal VSMCs (65).
A. Integrins
How integrins detect mechanical distortion is poorly understood. The mechanosensitive myogenic response of cerebral
arteries involves cytoskeleton remodeling and actin polymerization (262). Changes in the intraluminal pressure of
arterioles may cause conformational changes of the extracellular matrix and fibronectin that expose binding sites for
integrins, which trigger an intracellular signaling cascade.
Alternatively, integrins may sense the distending force directly and signal across the extracellular-integrin-cytoskeleton axis. Disruption of the cytoskeleton can impair the
myogenic response. Deletion of the serum response factor in
endothelial cells or VSMCs of tail arteries markedly attenuated myogenic constriction and reduced actin polymerization, myosin light chain, and myosin light chain kinase
(1234) while maintaining vasoconstriction produced by PE
or U-46619. At present, the specific actions of particular
integrins in the renal microcirculation are unclear, but they
likely constitute an important component of arteriolar
mechanosensation.
Integrins are a diverse family of transmembrane heterodimeric proteins composed of one ␣- and one ␤-subunit that
link the extracellular matrix to the cytoskeleton of VSMCs.
Deformation of integrins during changes in cell shape or
stress conducts pressure-induced signals across the plasma
membrane (304, 525, 614) to activate protein tyrosinephosphorylation cascades and mitogen-activated protein
kinases (MAPKs) including extracellular receptor kinases
(ERKs) (614). Although inhibition of ERK in rat cerebral
arteries inhibited myogenic constriction and KCl-induced
tone (857, 1600), the dynamics of PP-induced tyrosine
phosphorylation in rat cremaster arterioles was considerably slower than the myogenic response, which argues
against a causal relationship (1045, 1046). Both L-type
VOCCs and BK channels in isolated VSMCs and cremaster
arterioles were regulated by ␣5␤1 integrins that interacted
with fibronectin in the extracellular matrix to cause c-Srcmediated phosphorylation (305, 525). Thus both VOCCs
that excite the myogenic response and BK channels that
generally inhibit the response are under constitutive control
by an interaction between ␣5␤1 integrins and fibronectins.
The balance of their activities could determine the degree of
pressure-induced myogenic tone and perhaps vascular remodeling (525).
Rat afferent arteriolar VSMCs expressed the ␣3-, ␣5-, ␣Vand the ␤1-, ␤3-integrin subunits (1638). A synthetic peptide with an arginine-glycine-aspartic acid (RGD) motif
that interacts with integrin binding sites mobilized Ca2⫹
and contracted VSMCs (1638). In contrast, a different peptide with a RGD sequence reduced [Ca2⫹]i and dilated rat
cremaster arterioles (291). Thus there is an extensive diversity and regional specificity among this group of proteins
with other residues possibly playing a role. As discussed in
the introduction, the embryological origins of VSMCs differ according to their vascular bed and thus may contribute
to this diversity.
Shear stress in preglomerular VSMCs was transduced by
cell surface ␣5␤1 integrins into cytoskeleton-dependent
Ca2⫹ release from sarcoplasmic reticular stores (64). This
432
B. ENaCs
The ENaC in the CT and CDs is a trimeric protein complex
composed of ␣-, ␤-, and ␥-subunits, which facilitates the
Na⫹ reabsorption that initiates CT-GF. A vascular ␤-ENaC
has been proposed to function as a pressure- or stretchsensitive mechanosensor by forming an ion channel with
degenerin (DEG) that is a closely related neuronal acidsensing ion channel (ASIC) that is tethered to the extracellular matrix and cytoskeletal proteins (343, 344, 346,
1431). The non-voltage-gated ␤-ENaC-DEG complex may
permit Na⫹ entry in response to increased intraluminal
pressure and thereby cause depolarization of VSMC to initiate the myogenic response (FIGURE 8) (728). Stretch activation of ␤-ENaC in mouse freshly isolated renal VSMCs
increased Na⫹ entry (258), whereas in the CD, shear stress
gated and opened ENaCs (1356).
Rat cerebral arterial VSMCs expressed ␤- and ␥-ENaC subunits (344). Most, but not all, of the myogenic constrictor
tone was inhibited by blockade of ENaC with amiloride (1
␮M) or benzamil (30 nM) (344). Likewise, amiloride, ben-
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
PERFUSION PRESSURE
VOCC
ENaC
+Deg
GPCR
PLC
Membrane
events
TRPC
Integrins
DAG
ADP
ribosyl
cyclase
Em
p47phox
NAPDH
oxidase
cADPR
[Ca2+]i
GPCR
RyR
ROS
PLC
Cytosolic
signaling
Ca2+ induced
IP3 R
Ca2+ release
IP3
DAG
Rho / rho
kinase
PKC
Actin / myosin
crosslinking
Myosin light
chain kinase
Myosin light
chain phosphotase
Contractile
response
MYOGENIC CONTRACTION
FIGURE 8. An overview of the major pathways proposed to mediate or modulate the renal afferent arteriolar
myogenic contractile response to an increase in perfusion pressure (PP). The mechanisms are grouped as
membrane events, cytosolic signaling events, and contractile responses. VOCC, voltage-operated calcium
channel; ENaC, epithelial Na⫹ channel; DEG, degenerin; GPCR, G protein-coupled receptor; TRPC, transient
receptor potential channel; PLC phospholipase C; DAG, diacylglycerol; Em, membrane potential; cADPR, cyclic
ADP-ribose; NADPH, nicotinamide adenine dinucleotide phosphate; [Ca2⫹]i, cytosolic concentration of ionized
calcium; RyR, ryanodine receptor; ROS, reactive oxygen species; IP3R, inositol trisphosphate receptor; PKC,
protein kinase C. For discussion, see text.
zamil (1 ␮M) or siRNA knockdown of ␤- or ␥-ENaC subunits attenuated some of the myogenic constriction in rat
posterior cerebral arteries (793). Myogenic constriction of
middle cerebral arteries of ASIC2 knockout mice was very
weak, whereas constrictor responses to high KCl and PE
were normal, suggesting specificity to mechanical stimulation (454). DEG-like currents in cerebral arterial VSMCs
were inhibited by ROS (257). ␤-ENaC in rat posterior cerebral arteries colocalized with TRPM4 but not TRPC6
(793, 794), although all three channels participated in myogenic constriction.
ENaC ␣- and ␥ -subunit expression in mesenteric arteries
was reduced by high-salt diet, whereas ␤-ENaC movement
to the plasma membrane of VSMCs was increased (730).
Benzamil (1 ␮M) inhibited the myogenic response only in
arteries from rats fed a high-salt diet (730).
There is controversy about the presence of mRNA and protein for specific ENaC subunits in VSMCs of the cortical
radial artery and afferent arteriole. Early studies of single
VSMCs in the renal microcirculation, either freshly isolated
or cultured from relatively large cortical radial arteries of
rats or mice, reported expression of ␤- and ␥-, but not ␣-,
ENaC protein subunits (729). However, subsequent immunofluorescent studies in VSMC freshly isolated from cortical radial arteries and afferent arterioles revealed expression of ␣- as well as ␤- and ␥-subunits of ENaC (522). The
finding of the ␣-subunit poses a potential problem with
ENaC serving as a mechanosensitive channel since it renders the ion channel constitutively active and thus not selectively responsive to a change in transmural pressure
(522). On the other hand, another study failed to detect
mRNAs for any ENaC subunits in VSMCs disassociated
from these arterioles (1574), but all three subunits were
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
433
RENAL AUTOREGULATORY MECHANISMS
evident in intact cortical radial arteries. This implicates expression of ENaC on endothelial or adventitial cells rather
than on VSMCs, which would not be compatible with a
mechanosensing function.
Varying results have been reported for pharmacological
blockade of ENaC channels depending on the concentrations of antagonists. Apparent blockade of ENaC activity
with 5 ␮M amiloride or benzamil or genetic knockdown of
either ␤- or ␥-ENaC subunits attenuated myogenic constriction of cortical radial arteries (344, 728, 729). Moreover,
ANG II infusion reduced the immunoreactive expression of
␤- and ␥-ENaC in rat interlobar arteries and reduced myogenic contraction (731). However, the majority of the myogenic response and the regulation of RVR in vivo occur in
smaller, more distal cortical radial arteries and afferent arterioles rather than in the relatively large interlobar arteries
(201, 611, 1071). Nevertheless, recent studies have reported an impaired myogenic constriction of afferent arterioles isolated from mice with genetically reduced ␤-ENaC
expression (462).
A mouse model of genetically reduced (⬃50%) ␤-ENaC
expression had a 50% reduction in the myogenic component of RBF autoregulation in the 0- to 5-s time window, yet
a well-maintained overall steady-state autoregulation
(505). This likely implies that the MD-TGF contribution
was upregulated since inhibition of MD-TGF by acute volume expansion revealed a 75% reduction in the whole kidney myogenic response in mice with genetic reductions in
␤-ENaC expression and almost complete loss of autoregulation over 30 s after a step change in PP (462).
Recent studies by Drummond and associates using VSMCs
freshly isolated from the mouse afferent arteriole have
shown less frequent and smaller stretch-activated Na⫹ currents in the VSMCs from mice with reduced ␤-ENaC expression, but preserved VOCC and BK channel activity
(258, 344). Thus ␤-ENaC appears to be required for normal
mechanically gated currents in renal VSMCs. Their disruption reduces myogenic constriction of afferent arterioles.
Cortical radial arteries and afferent arterioles isolated from
␤-ENaC knockdown mice have a weakened myogenic tone,
but constricted normally to PE and KCl (462), which implies that ENaC is upstream from myogenic contractions.
Mice with partial gene knockdown of ␤-ENaC became hypertensive and developed renal inflammation and damage
that may be related to reduced renal autoregulatory capacity (345).
The myogenic response of the rat perfused JMN preparation after papillectomy to eliminate MD-TGF was blunted
by blockade of ENaC with 10 ␮M amiloride or benzamil
added to the bath or 5 ␮M amiloride added to the luminal
perfusate (522). Similarly, relatively high concentrations of
inhibitors of ENaC (amiloride and benzamil, 10 ␮M)
434
blocked more than two-thirds of the autoregulatory response of the afferent arteriole, whereas vasoconstriction
produced by ATP or 20-HETE was unaffected (1051). The
authors concluded that ENaC is an essential component of
myogenic responses.
Different conclusions were reported in a recent study with
the HNK preparation where the myogenic response or the
EM was unaltered by relatively low concentrations (ⱕ1–3
␮M) of amiloride or benzamil which inhibited ⬎80%
ENaC activity in renal tubules (926, 1574). However,
higher concentrations (ⱖ5 ␮M), similar to those used in the
earlier studies, were confirmed to inhibit myogenic constriction, but this was ascribed to nonspecific inhibition of Na⫹
transporters and ion channels other than ENaC (926).
Moreover, myogenic constriction in the rat HNK preparation was independent of extracellular [Na⫹] (1574). Benzamil (ⱖ1 ␮M), which also inhibits the NCX exchanger,
depolarized and potentiated, rather than inhibited, the
myogenic constriction of the afferent arteriole in this study
(1574). However, another study of afferent arterioles of the
HNK reported that lowering extracellular [Na⫹] led to a
slowly developing attenuation of myogenic constriction,
with abolition of the myogenic response when [Na⫹] was
below 50 mM (1446).
Aldosterone regulates ENaC expression and activity in the
distal nephron, but less is known about its role in the regulation of ENaC in VSMCs and endothelium. Elevated aldosterone levels are causally linked to vascular fibrosis, inflammation, and remodeling (152, 816). Stretch of rat cultured aortic VSMCs increased nicotinamide adenine
dinucleotide phosphate (NADPH) oxidase activity that was
dependent on aldosterone and mineralocorticoids (1114).
Aged mice lacking mineralocorticoid receptors in their
VSMCs were hypotensive and had a 40% reduced myogenic tone in their mesenteric arteries, weaker TP receptorinduced contraction, and an 80% reduction in the expression and activity of L-type VOCC, whereas BKCa function
was normal (994). Aldosterone rapidly (⬍5 min) contracted rabbit isolated afferent and efferent arterioles via
nongenomic mechanisms mediated by L- and/or T-type
VOCC (53, 54).
ENaC is also located in the vascular endothelium of mesenteric arteries where it inhibited NO production (851),
which might be another pathway whereby ENaC modulates the myogenic mechanism.
In summary, genetic studies with knockdown of ␤-ENaC in
the preglomerular vasculature have reported an attenuated
myogenic response, implicating ␤-ENaC as a critical
mechanosensor. However, the pharmacological evidence
for a role of ENaC is more controversial because of lack of
specificity of the antagonists at the concentrations tested.
Integrins could function as force transducers linked to
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
ENaC as part of a mechanosensory complex that couples
the extracellular matrix to the cytoskeleton. In turn, ENaC
may be linked directly or indirectly to TRP channels. Further investigation is required to clarify whether and how
ENaC subunits function as mechanosensors in the renal
microcirculation.
C. Transient Receptor Potential Channels
and Stretch-Activated G Protein-Coupled
Receptors
Transient receptor potential (TRP) channels are a family of
voltage-insensitive, homo- or heterotetrameric proteins that
function either as nonselective cation channels allowing entry of Ca2⫹ and Na⫹ to depolarize the EM or as highly
selective Ca2⫹ channels to directly increase [Ca2⫹]i (FIGURE
8) (330, 366, 368, 681, 1352). Either mechanism could
cause vasoconstriction. TRPC1, TRPC3, and TRPC6 are
abundant in VSMCs. TRP channels resided in signaling
complexes in caveolar lipid raft domains in the plasma
membrane (23). Mechano-gating of TRPC1 and TRPC6
can be transduced through scaffolding proteins and actin
filaments of the cytoskeleton.
The mRNAs and proteins for multiple canonical (C) TRPs
have been identified in rat renal resistance vessels and glomeruli. The mRNAs for TRPC1, 3, 4, 5, and 6 were expressed
in renal cortical radial arteries and afferent arterioles (387),
and the proteins for TRPC1, 3, 5, and 6 have been detected
in glomeruli and preglomerular vessels (387, 1278). TRPC1
was expressed both in the afferent and efferent arterioles
(1448). Immunoreactive TRPC1, 3, and 6 were expressed
on rat afferent arteriolar VSMCs and TRPC3 on cortical
radial arteries (1278).
Mechanical stretch of the plasma membrane directly activated TRPC6, presumably via a conformational change to
open the channel, independent of GPCR ligand binding or
PLC activation, in HEK293 and CHO cells overexpressing
the ion channel (1403). DAG generation after GPCR activation can engage TRPC3 and TRPC6 channels in VSMC
throughout the vasculature (623, 679). In cerebral vessels,
TRPC6 and TRP melastatin (M) 4 and vanilloid (V) channels transduce myogenic responses (141, 366, 371, 681,
1352, 1403, 1594). Cerebral arterial VSMCs also express
TRPC3 that is activated by GPCR agonists such as UTP but
not by pressure, suggesting coupling of TRPC3 and TRPC6
to distinct excitatory stimuli (1212). TRPC1 was present in
VSMC of cerebral arteries, but it is not clear if it has a
mechanosensitive role in the vasculature (331).
These findings led to the concept that membrane stretch
activates a TRPC6-dependent constriction of cerebral arteries. A proposed pathway is that membrane stretch activates
PLC to produce DAG, which directly activates TRPC6 to
mediate depolarizing cation current to trigger a myogenic
response (679). Thus pressure-induced depolarization of
cerebral artery myocytes was blocked by PLC inhibition
(1383), and mechanosensitive channels in VSMCs resembling TRPC channels were inhibited by PLC blockade and
activated by a DAG analog (1154). However, mechanosensitive channel activity and myogenic responses of cerebral
arteries were not affected by deletion of TRPC1 (331) or
TRPC6 (332, 490). 20-HETE potentiated the activity of
TRPC6 channels of the mesenteric artery, thereby enhancing myogenic constriction (680).
TRPC6 on the glomerular podocyte slit diaphragm functioned as a mechanosensor coupled to Ca2⫹-activated potassium (KCa) channels (348). Stretch activation of cation
entry in the podocyte was independent of PLC or PLA2
activity (25). Gain-of-function mutations of TRPC6 caused
podocyte dysfunction and focal and segmental glomerulosclerosis (FSGS) (372, 837), whereas deletion of TRPC5
protected mice from albuminuria (1290).
IP3 constricted cerebral arteries via TRPC3 channel activation leading to Na⫹ influx and membrane depolarization,
independent of sarcoplasmic reticular Ca2⫹ release (1629).
A similar mechanism operated in rat mesenteric arteries
where IP3 coupling of TRPC3 channels to caveolae resulted
in Na⫹ entry (10).
TRPM4 in cerebral artery myocytes was a highly selective
monovalent cation channel activated by stretch or Ca2⫹
entry through TRPC6 channels and/or Ca2⫹ mobilization
mediated by RyRs (370, 1032). Other studies in cerebral
arteries reported that transmural pressure activated a signaling cascade involving activation of Src tyrosine kinase
and PLC with subsequent IP3 generation that mobilized
Ca2⫹ from the sarcoplasmic reticulum (480 – 482). The resultant increase in [Ca2⫹]i in turn activated plasma membrane TRPM4 that increased Na⫹ entry and depolarized
the plasma membrane to allow Ca2⫹ entry and contraction.
The Ca2⫹ activation of TPRM4 may be enhanced by Ca2⫹
entry through TRPC6 channels. In contrast, TRPV4 in mesenteric arteries acted as a nonselective cation channel activated by stretch or EETs to cause activation of K⫹ channels
that hyperpolarized the plasma membrane and relaxed the
vessel (367).
A provocative observation was that hyposmotic cell swelling and stretch-activated G protein-coupled receptors (GPCRs) that caused a conformational change independent of
ligand binding which conferred a mechanosensitive action
on TRPC6 in heterologous systems overexpressing GPRC
and TRPC6 (1403). GPCR activation in VSMCs linked to
Gq/11-dependent DAG generation can engage TRPC6 channels in VSMCs independent of PKC (FIGURE 8) (1143).
Subsequent studies demonstrated that stretch activation of
GPCRs stimulated cation flux through TRPC3 and TRPC6
via PLC-dependent DAG formation (1000). Diverse GP-
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
435
RENAL AUTOREGULATORY MECHANISMS
CRs may activate this pathway including ANG II type 1
(AT1), endothelin type A (ETA), vasopressin type 1A (V1A),
and muscarinic type 5 (M5) receptors (1000, 1673). In the
absence of ANG II, certain antagonists of AT1 receptors,
such as losartan or candesartan, prevented stretch-induced
myogenic tone of cerebral and renal cortical radial arterial
VSMCs via reductions in ERK and ROS (999, 1000). The
myogenic tone of cerebral arteries was enhanced by activation of GPCRs linked to PLC-sensitive G proteins (1137),
whereas TRPC3 in cerebral arteries was activated by IP3,
independent of Ca2⫹ release from the sarcoplasmic reticulum (1629). Collectively, these data suggest that membrane
stretch causes an agonist-independent conformational
change in the GPCR, which activates PLC and downstream
G␣q/ll signaling to cause DAG activation of TRPC6, resulting in Na⫹ and/or Ca2⫹ entry, membrane depolarization,
and a myogenic response (1000, 1421).
A recent study extends these findings to individual mesenteric arteries and isolated kidneys. This demonstrated a requirement for AT1A receptors for myogenic responses in the
absence of its native ligand (1294). However, the downstream signaling relies on an ion channel distinct from
TRPC6 or a Kv channel distinct from KCNQ3, 4, or 5 to
activate VSMCs and elevate vascular resistance.
Renal function studies raise some questions about the role
of GPCR in the myogenic response. Although ANG II may
potentiate the renal myogenic response, as discussed in section VIA, activation of other GPCR by natural agonists
such as adrenoceptors by catecholamines (760, 1193,
1570), ETA receptors by endothelin (1357), or V1A receptors by AVP (1563) do not normally impact the renal myogenic response. Frequency analysis in the kidney generally
has found that inhibition of ANG II formation (597) or AT1
receptor antagonism (329) has little effect on the myogenic
mechanism, although a recent study reported that prolonged ANG II administration potentiated the renal myogenic response (1193). In other studies, blockade of AT1
receptors with losartan or candesartan did not inhibit myogenic tone in rat mesenteric arteries (519, 985) and only
blunted myogenic constriction of cerebral arteries weakly
(1000).
G␣q/ll-coupled receptor activation in cerebral arteries by
uridine triphosphate or thromboxane prostanoid (TP) receptor activation by U-46619 did not enhance mechanosensitivity of TRPC-like currents or amplify myogenic responsiveness (36). Increasing the PP of rat isolated mesenteric
arteries activated ERK1/2 that was inhibited by an antagonist of AT1 but not AT2 receptors (383, 984). The myogenic
response was also blocked during ACE inhibition, suggesting dependence on ANG II formation.
At present, limited data have implicated an agonist-independent, stretch-activated GPCR pathway in TRP channel
436
activation of the myogenic contractile response in renal microvessels. Thus myogenic tone of renal arteries was diminished by the AT1 receptor antagonist losartan independent
of ANG II (1000). Cortical radial arteries from regulator G
protein signaling 2 knockout mice had enhanced myogenic
tone in addition to the expected exaggerated vasoconstriction to ANG II, ET-1, and PE (604).
Activation of TRPC channels in the afferent arteriole enhanced Ca2⫹ signaling by ANG II (402) and NE where they
functioned as store-operated cation channels (SOC) independent of L-type VOCCs (387, 1278). The myogenic response of cortical radial arteries and afferent arterioles of
the HNK depended on ill-defined mechanosensitive cation
channels (1446, 1447). GPCRs can activate redox signaling
in afferent arteriolar VSMCs (399, 400, 1606) where TRP
channels themselves are redox sensitive because of oxidative modifications of specific cysteine residues (507, 1005,
1433). VOCCs also were redox regulated (126). NO activated TRPC channels via cysteine S-nitrosylation (1639).
The polycystins represent a special class of TRP cation
channels. VSMC TRP polycystin-1 and -2 (TRPP1 and -2)
cation channel proteins modulated the myogenic response
in mesenteric arteries (1057). Deletion of TRPP1 in VSMCs
reduced stretch-activated channel activity and myogenic
tone (1353), but depletion of TRPP2 rescued both stretchactivated channel opening and the myogenic response, indicating extensive interaction. TRPP2 interacted with the
cytoskeleton and an actin crosslinking protein that was central for stretch-activated channel regulation.
In summary, cation entry through TRP channels subserves
two functions in VSMC: to relate changes in transmural
pressure or stretch to depolarization of the plasma membrane and to respond to GPCR stimulation and PLC activation to modulate SOC function. TRP channels are attractive candidates as mechanosensors, but their linkage to GPCRs remains an intriguing, but uncertain, possibility. The
specific roles of particular TRP channels acting as mechanosensors initiating membrane depolarization and increases in [Ca2⫹]i to trigger the myogenic response of afferent arterioles requires further clarification. Integrins, signaling through the cytoskeleton and possible TRP channels are
attractive candidates, but are incompletely studied in the
renal microcirculation. The role of stretch-activated ENaC
channels is controversial and still evolving.
IV. CONDUCTANCE CHANGES MEDIATING
MYOGENIC CONTRACTION
A. Chloride Channels
The intracellular [Cl⫺] in VSMCs is high due to accumulation by a Cl⫺/HCO3⫺ exchanger and the NKCC1 cotransporter. Because the [Cl⫺] in VSMCs is above its electro-
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
chemical equilibrium, activation of plasma membrane Cl⫺
channels permits Cl⫺ efflux, leading to depolarization of the
plasma membrane, with subsequent opening of VOCCs to
promote Ca2⫹ entry (489, 1324, 1442). However, the resting Cl⫺ permeability of VSMCs (621), arcuate arteries
(1200), and renal mesangial cells (1118) is sufficiently low
that the EM is unaffected by removal of extracellular Cl⫺.
Likewise, the basal tone of rabbit perfused afferent arterioles is unchanged in Cl⫺-free media (726). These results
indicate that there is little basal Cl⫺ conductance in afferent
arterioles.
On the other hand, VSMCs isolated from arcuate or cortical
radial arteries display a Ca2⫹-activated Cl⫺ conductance
(ClCa) that depolarizes the plasma membrane after GPCR
activation and leads to Ca2⫹ mobilization (489, 664). ClCa
channels in afferent arterioles contribute to membrane depolarization (559), Ca2⫹ entry (1324), GPCR activation
(726, 1442), and increased [Ca2⫹]i produced by ANG II or
ET-1 (451, 727, 1324, 1435). GPCR activates Cl⫺ channels
in mesangial cells to depolarize the plasma membrane and
enhance [Ca2⫹]i signaling (834, 987).
Activated Cl⫺ channels also can enhance myogenic tone by
providing a Cl⫺ influx pathway as a charge compensation
for influx of Ca2⫹ during activation of VOCCs (559, 726).
Indeed, the addition of extracellular Cl⫺ enhanced high
K⫹-induced depolarization and VOCC-dependent contractions of rabbit afferent arterioles with an EC50 of 82 mM
(559), that is close to [Cl⫺] in extracellular fluid, while
removal of extracellular Cl⫺ attenuated Ca2⫹ influx via
VOCCs in both mesangial cells (834) and rat aortic VSMCs
(1650).
Although Cl⫺ channels are important in vasoconstriction of
afferent arterioles produced by ANG II and NE (559, 726,
1407), neither blockade of Cl⫺ channels nor reducing extracellular [Cl⫺] inhibits myogenic responses of the HNK
preparation (1435, 1442). Thus ClCa regulated VOCCs
during stimulation with GPCR agonists, but these channels
do not appear to be required for the myogenic response of
afferent arterioles (726, 1442).
The electroneutral NKCC1 is present in VSMCs throughout the vascular system, in contrast to the NKCC2 isoform
that is expressed exclusively in the kidney in the TAL of
Henle’s loop and MD cells. NKCC1 may regulate myogenic
tone of arteries and arterioles. Ion entry via NKCC1 increased intracellular [Cl⫺] that caused membrane depolarization and thereby Ca2⫹ entry through VOCCs and vasoconstriction (1126). Indeed, afferent arteriolar myocytes
expressed NKCC1 but not NKCC2 (1568), and inhibition
of this vascular NKCC1 in the mouse isolated and perfused
afferent arteriole by furosemide or bumetanide caused vasodilation (1124) and inhibited most of the myogenic response of the afferent arteriole in the rat HNK preparation
lacking tubules (1568). Inhibition of NKCC1 by bumetanide decreased myogenic tone in mesenteric arteries as well
as contraction by high KCl, PE, and UTP (820).
However, as discussed in section IVC, such pronounced
inhibition of myogenic tone has not been observed during
administration of furosemide or bumetanide in vivo. RBF in
vivo is usually unchanged or slightly reduced during furosemide-induced natriuresis (749 –751, 753, 760, 1124,
1183, 1489), although increased RBF has been observed in
some cases (170, 351, 760). Tubular hydrostatic pressure
and interstitial hydrostatic pressure was increased during
furosemide-induced natriuresis (170, 1489), which may
contribute to a fall in GFR.
Molecular candidates for Cl⫺ channels are transmembrane
protein 16A (TMEM16A) or anoctamin 1 (ANO1), bestrophins, as well as the classic ClC-3 (1367). Both “classic”
and cGMP-dependent ClCa channels are expressed in
VSMCs. Bestrophin-3 is reported as a cGMP-dependent
Ca2⫹-activated Cl⫺ conductance and ClC-3, a voltagegated channel (980, 982). Membrane stretch in cerebral
arteries activated TMEM16A channels in VSMCs and
caused membrane depolarization and the myogenic response via cation entry through mechanosensitive nonselective cation channels, with Ca2⫹ activation of TMEM16A
facilitating Cl⫺ efflux (165). Disruption of TMEM16A
(ANO1) and thereby preventions of Ca2⫹-activated Cl⫺
currents relaxed VSMC of aorta and carotid arteries (599).
Blockade of Cl⫺ channels in pressurized rat cerebral arteries
caused hyperpolarization (⫺10 to ⫺15 mV) that inhibited
myogenic tone (1080).
Bestrophin-3 was important for Ca2⫹ oscillations and synchronization of VSMC during vasomotion in the rat mesenteric small arteries in vivo (153). siRNA knockdown of
bestrophins in rat mesenteric arteries did not affect arterial
contraction but inhibited the rhythmic contractions, vasomotion (295). Downregulation of TMEM16A also reduced
expression of bestrophins and suppressed vasomotion and
contractions produced by GPCR agonists (NE, AVP) and
KCl, suggesting this ClCa channel participates in depolarization of VSMC.
Thus Cl⫺ channels are involved in the regulation of the EM
of VSMC and vasoconstriction of mesenteric arteries and
myogenic tone of cerebral arteries. Cl⫺ channels in the renal
vasculature participate in vasoconstriction by GPCR agonists, but apparently not the myogenic response.
B. Potassium Channels
The resting EM of VSMCs of pressurized afferent arterioles
is approximately ⫺40 mV. This is ascribed largely to a K⫹
diffusion potential. The EM was near the threshold for activation of L-type VOCCs (847, 928). Opening K⫹ channels
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
437
RENAL AUTOREGULATORY MECHANISMS
and potentiating K⫹ efflux from VSMCs elicits vasodilation
primarily by hyperpolarization that inactivates L-type
VOCCs and reduces Ca2⫹ entry. The large-conductance,
Ca2⫹-activated K⫹ channel (BKCa) in the cerebral circulation was a major determinant of the VSMC EM (1081).
The BKCa channel has been implicated in several steps of the
myogenic response, including channel closure causing
membrane depolarization, and channel opening causing hyperpolarization to oppose excessive pressure-induced vasoconstriction. A special case in cerebral arteries is spontaneous or pressure-induced Ca2⫹ sparks representing Ca2⫹ release from RyR channels in the sarcoplasmic reticulum,
which can activate BKCa to hyperpolarize the plasma membrane, limit Ca2⫹ entry via VOCCs, and diminish myogenic
tone. Accordingly, inhibition of BKCa in many nonrenal
vascular beds has a variable effect, but often depolarizes the
EM and enhances stretch-induced depolarization, Ca2⫹ entry, and contraction (720, 1595).
Caveolae in VSMC of cerebral arteries can modulate the
myogenic response. VSMC of mice lacking caveolin-1 responded to an increased pressure with a diminished depolarization, an increase in [Ca2⫹]i, and myogenic response
(11) via activation of BKCa channels and hyperpolarization.
NOS inhibition did not restore myogenic tone in arteries
deficient in caveolin-1, arguing against excess NO production as the cause for inhibition of the myogenic response.
Genetic ablation of caveolin-1 in murine cerebral artery
VSMCs increased Ca2⫹ sparks and the frequency of transient BKCa currents without altering Ca2⫹ entry via VOCC
(233).
Stretch of VSMCs in coronary arteries/arterioles and mesenteric arteries stimulated an outward K⫹ current through
BKCa, whereas blockade of BKCa channels enhanced
stretch-induced depolarization (338, 1628). Inhibition of
BKCa channels increased myogenic contraction of pressurized mouse mesenteric arteries, presumably by favoring
membrane depolarization and Ca2⫹ entry through VOCCs
(836).
VSMCs exhibit heterogeneity in BKCa activity or their sensitivity that may contribute to tissue-specific differences in
the regulation of myogenic vasoconstriction (525, 618,
1632). For example, the BKCa is more sensitive to increased
[Ca2⫹]i resulting from Ca2⫹ sparks generated by Ca2⫹ release from sarcoplasmic reticulum in cerebral than skeletal
muscle cremaster arteries. The resultant vasodilation attenuated myogenic vasoconstriction in cerebral arteries
(1632). The variation in sensitivity might be explained by
the number of ␤1 subunits of the BKCa channel (1632).
The extent to which BKCa are open during basal conditions
and modulate basal EM in the renal vasculature is not clear.
BKCa channels are present on VSMCs of preglomerular ar-
438
teries and arterioles (390, 467, 949, 1200), but usually are
closed under basal conditions. Thus their blockade did not
modify resting RBF or renal vasoconstriction elicited by
ANG II or NE (949). Moreover, BKCa channels in pressurized afferent arterioles of JMN also were quiescent and did
not buffer vasoconstriction by ANG II (390, 1200). ACh
caused dose-dependent renal vasodilatation in isolated perfused rat kidneys, which was blunted by NOS inhibition but
not by inhibition of sGC. The vasodilation was mediated by
KCa3.1 channels sensitive to charbdotoxin which is characteristic of BKCa channels (1370).
On the other hand, an increase in PP of renal arterioles was
reported to increase 20-HETE generation from arachidonate, which closed BKCa channels and thereby caused membrane depolarization and vasoconstriction (672, 943,
1667). Blockade of 20-HETE production inhibited myogenic responses of arcuate arteries and impaired RBF autoregulation (464, 777, 1668, 1669).
Voltage-gated and inward-rectifier K⫹ channels (KV and
KIR) in afferent arterioles can regulate basal vascular tone
(936, 1484). KIR 2.1 in the afferent arteriole contributed to
K⫹ currents and EM of VSMCs (238). Enhanced rectification of KIR in the distal cortical radial artery and afferent
arteriole hyperpolarized the plasma membrane and attenuated myogenic tone (239, 240). Stimulation of PKC by
ANG II or ET-1 potentiated the myogenic response of afferent arterioles of the HNK preparation by inhibition of
delayed rectifier K⫹ channels which depolarizes the plasma
membrane (802).
Activation of ATP-dependent K⫹ channels (KATP) in JMN
had no effect on basal afferent arteriolar tone (665), but
blunted PE-induced constriction of rabbit afferent arterioles (924). Activation of these channels by calcitonin generelated peptide dilated the afferent arteriole in the HNK
preparation (1232) by causing hyperpolarization of the
plasma membrane that inhibited Ca2⫹ entry through
VOCCs (1231). Moreover, activation of KATP by pharmacological agents or hypoxia hyperpolarized the plasma
membrane and inhibited myogenic responses of afferent
arterioles (936, 1232). However, KATP channels in the dog
did not affect the MD-TGF response (782) or the autoregulation of coronary blood flow (1060).
A recent study of the renal vasculature in rats in vivo demonstrated that pharmacological blockade of individual K⫹
channels (e.g., BKCa, Kir, KATP, Kv) had little effect on large
whole kidney RVR responses to GPCR agonists such as
ANG II or NE and their activation of VOCCs (1396). However, concurrent closure of multiple types of K⫹ channels
effectively attenuated agonist-induced renal vasoconstriction, demonstrating considerable redundancy among the
K⫹ channels in the renal microcirculation. Although voltage-dependent K⫹ channels were tonically active in coro-
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
nary arteries, they were not required in vivo for autoregulation (99). KCNQ (Kv7) K⫹ channels in cerebral arteries
were inhibited by membrane stretch and thereby could
modulate myogenic vasoconstriction (1663), although genetic deletion of KCNQ3, 4, or 5 did produce constriction
of mouse mesenteric arteries or isolated perfused kidneys
(1294).
The myogenic tone of skeletal muscle arterioles results from
an interaction between Ca2⫹ entry through L-type VOCC
secondary to pressure-induced membrane depolarization
and CICR and Ca2⫹ mobilization mediated by RyR. BKCa
modulation of EM played only a small role in limiting myogenic vasoconstriction in this vascular bed (827). However,
inhibition of RyRs in mouse pressurized mesenteric arteries
increased myogenic constriction, presumably by reducing
BKCa activity to favor membrane depolarization and Ca2⫹
entry through VOCCs (836).
In summary, most evidence suggests that BKCa channels are
normally quiescent in the renal microcirculation and do not
respond to changes in RPP or limit the myogenic response,
in contrast to their central role in the cerebral circulation.
However, 20-HETE is reported to enhance myogenic constriction by inhibiting BKCa activity in some situations. Activation of other K⫹ channels that hyperpolarize the EM
also can inhibit myogenic responsiveness under specific experimental conditions such as hypoxia, whereas inhibition
of BKCa channels in some settings can depolarize VSMCs
and favor vasoconstriction. MD-TGF responses depend on
ROMK channels in MD that mediate K⫹ recycling after
reabsorption by NKCC-2 (1312).
C. Sodium Channels
Evidence relating to the proposal that ENaC subunits are
expressed on VSMCs of preglomerular vasculature and
function as mechanosensors in the myogenic response was
discussed in section IIIB. The VSMC plasma membrane
Na⫹/Ca2⫹ exchanger (NCX1) normally promotes Ca2⫹ entry and Na⫹ extrusion (1655), in contrast to the heart
where NCX1 primarily mediates Ca2⫹ extrusion (327).
Normally, three Na⫹ are exchanged for one Ca2⫹ (119).
NCX contributes to the myogenic response by favoring
Ca2⫹ entry both in nonrenal vessels and also in the renal
circulation where it enhances vasoconstrictor responses.
Ca2⫹ entry via NCX1 is implicated in pressure-induced
myogenic constriction in mouse mesenteric and hindlimb
arteries (713, 1655), rat cremaster first-order arterioles
(150 ␮m) (1206), and rat posterior cerebral arteries (771).
Reducing extracellular Na⫹ around isolated cremaster
muscle arterioles promoted NCX1 activity and Ca2⫹ entry
to enhance myogenic vasoconstriction (1206), whereas decreasing the activity of the VSMC NCX impaired arteriolar
myogenic reactivity (1206). Knockout of NCX1 in VSMCs
reduced vasoconstriction elicited by PE, high KCl, PP, or
L-type Ca2⫹ channel currents in mouse mesenteric arteries
and lowered BP (1655). Myogenic and ANG II responses of
mesenteric arteries were blunted in mutant mice lacking the
NCX1 in VSMCs (1655, 1660) accompanied by attenuated
Ca2⫹ entry and myogenic responses (1010, 1655).
A PKC regulated NCX1 in the afferent arteriole usually
operates in the “reverse” mode with Ca2⫹ entering VSMCs
down its concentration gradient to drive Na⫹ efflux and
promotes vasoconstriction (120, 430, 1079). Reducing extracellular [Na⫹] in the isolated perfused kidney promoted
Ca2⫹ entry into afferent arteriolar VSMCs via NCX to
increase [Ca2⫹]i (1078, 1079) and RVR (1335). Inhibition
or genetic deletion of NCX1 reduced the afferent arteriolar
[Ca2⫹]i responses and the vasoconstriction to ANG II (402,
1660).
The expressions of NCX1, sarcoplasmic reticular Ca2⫹ATPase-2 (SERCA2), and TRPC6 in VSMCs are inversely
related to the activity of the ␣2-subunit of Na⫹-K⫹-ATPase
(120), whose reductions in mesenteric arterial VSMCs have
been linked to increased Ca2⫹ signaling and augmented
myogenic- and PE-induced vasoconstriction mediated by
IP3 release of Ca2⫹ from the sarcoplasmic reticulum and
Ca2⫹ entry through SOC channels (119). Mesenteric arteries expressing 50% of the normal amount of the ␣2-subunit
of the Na⫹ pump had augmented myogenic contraction
(1654). VSMC NCX1, SERCA2, and TRPC6 were overexpressed in fourth-order mesenteric arteries of Milan genetically hypertensive rats which had exaggerated myogenic
responses (905).
However, NCX can operate in the “forward” mode in
VSMCs, thereby mediating Na⫹ influx in exchange for
Ca2⫹ efflux, for example, in the vasculature of the isolatedperfused kidney (1335) or skeletal muscle arteries where
blockade of NCX enhanced myogenic tone, presumably by
reducing Ca2⫹ extrusion (634). NCX activity in VSMC can
be inhibited by oxidative stress that also increased [Ca2⫹]i
(1505).
Thus Na⫹/Ca2⫹ exchange via NCX1 has variable effects
depending on whether it functions in the forward or reverse
mode. It typically functions in the reverse mode in renal
afferent arteriolar VSMCs to facilitate Ca2⫹ entry down its
concentration gradient, thereby expelling Na⫹ and promoting myogenic contractions. Its activity is linked to Na⫹-K⫹ATPase via change in the ␣2 subunit which can account for
the enhanced myogenic tone, and perhaps some of the hypertension, in the Milan genetically hypertensive rat.
V. CALCIUM SIGNALING PATHWAYS
INVOLVED IN AUTOREGULATORY
RESPONSES
An abrupt increase in RPP in mouse microperfused outer
cortical afferent arterioles increases [Ca2⫹]i within seconds,
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
439
RENAL AUTOREGULATORY MECHANISMS
followed by a sharp decline to a half-maximal level (865),
whereas the juxtamedullary afferent arteriole has a rather
more sustained rise in [Ca2⫹]i (1637). Both preparations
demonstrate a maintained myogenic contraction despite declining [Ca2⫹]i, indicating increased sensitivity of the contractile machinery to [Ca2⫹]i. There are comprehensive reviews on VSMC Ca2⫹ signaling during generalized myogenic vasoconstriction (169, 301, 302, 478, 614, 1001,
1249, 1329) and its role in modulating renal vascular microcirculatory responses (49, 1249, 1280). Also recommended are excellent mathematical models of renal autoregulation and the Ca2⫹ dynamics of the myogenic response
of the afferent arteriole (228, 374, 395, 628, 805, 1350,
1351, 1614). The models incorporate ionic transport and
cell membrane potential with contraction dynamics of the
afferent arteriolar VMSC.
A. Calcium Entry and Mobilization
Vasoconstriction results from a combination of increased
[Ca2⫹]i and augmented Ca2⫹ sensitivity (FIGURE 8). Intracellular Ca2⫹ signaling involves both Ca2⫹ entry across the
plasma membrane and Ca2⫹ release from intracellular sarcoplasmic reticular stores (1329, 1625, 1630). Ca2⫹ entry
into preglomerular VSMCs can be mediated by L-type and
perhaps T-type VOCCs, voltage-independent receptor-operated channels (e.g., TRPCs, TRPMs, and GPCR-coupled
channels), and SOC. GPCRs normally recruit more than
one type of cation channel. Both the sustained myogenic
tone and the MD-TGF-mediated vasoconstriction are
highly dependent on membrane depolarization and Ca2⫹
entry through L-type VOCC. As discussed in later sections,
Ca2⫹ sensitivity of the contractile proteins is primarily regulated by PKC and Rho/Rho kinase.
The mechanisms underlying mechanosensation and initiation of the myogenic response vary among vascular beds,
but usually involve an initial depolarization of the plasma
membrane that activates Ca2⫹ entry through L-type
VOCCs that leads to activation of myosin light-chain phosphorylation and vasoconstriction. A decreased intraluminal
pressure has the opposite effects, thereby resulting in vasodilation. Removal of extracellular Ca2⫹, or inhibition of
L-type VOCC activity, abolishes the myogenic response.
Mechanotransduction is a key triggering process, but the
details are poorly understood. Mechanosensitive initiating
elements include ion channels (e.g., TRPC, ENaC) or extracellular proteins (e.g., integrins). Pressure activation of
ENaC or nonselective cation TRPC channels may cause
depolarization by favoring Na⫹ entry. There is considerable
regional diversity in subsequent Ca2⫹ signaling and Ca2⫹
sensitivity of the actin/myosin contractile apparatus responsible for regulation of local blood flow and capillary hydrostatic pressure.
The myogenic response of the afferent arteriole is initiated
by depolarization of its VSMCs which, in turn, activate
440
L-type VOCCs to allow Ca2⫹ entry. Although the precise
mechanisms responsible for the initial depolarization of the
plasma membrane are unclear, they may entail the opening
of a cation channel to allow Na⫹ or Ca2⫹ entry, the opening
of a Cl⫺ channel to permit Cl⫺ exit, or the closing of K⫹
channels to prevent K⫹ efflux. All of these possibilities are
under current investigation.
Little is known about the role of Ca2⫹ mobilization resulting from Ca2⫹-induced Ca2⫹ release (CICR) from the sarcoplasmic reticulum following stimulation by IP3 and RyRs
and less still about Ca2⫹ release from lysosome-like acidic
vesicles (49, 1456). Ca2⫹ entry through VOCCs can amplify CICR from the sarcoplasmic reticulum therby enhancing arteriolar contraction (401, 1457). Depletion of sarcoplasmic reticular Ca2⫹ stores in nonrenal vessels can promote Ca2⫹ entry through TRPC or Orai channels. Orai
channels on VSMCs mediate Ca2⫹ entry by Ca2⫹ releaseactivated Ca2⫹ channels after activation a stromal interaction molecule (STIM) protein linking the sarcoplasmic reticulum to the plasma membrane (1390, 1482, 1658). Although renal resistance arterioles have store-operated Ca2⫹
entry channels (397, 398, 925, 1448, 1616), their molecular
identity is not yet known and their role in regulating myogenic or MD-TGF mediated autoregulatory tone has not
been studied.
1. Calcium entry through VOCCs
The renal preglomerular arteries and arterioles expressed
mRNA for three types of VOCCs of the L (Cav 1.2), T (Cav
3.1 and 3.2), and P/Q (Cav 2.1) subtypes (35, 558, 560,
595, 633). Depolarization of the plasma membrane led to
Ca2⫹ entry through L-type VOCCs in the afferent arteriole,
which was blocked by the dihydropyridine Ca2⫹ channel
blockers (CCB) (191, 558, 935). The depolarization mechanism adapts over time. Thus 2 days of hypertension enhanced expression of the ␣1c-subunit of the Cav1.2 Ca2⫹
channel in the rat main renal artery (1175), which may
contribute to sustaining the vasoconstriction. Increasing the
transmural pressure of a dog cortical radial artery from 20
to 120 mmHg reduced the EM from ⫺57 to ⫺38 mV and
increased the contractile tone of its VSMCs (562), whereas
increasing the luminal pressure from 60 to 100 and then to
145 mmHg reduced the EM from ⫺45 to ⫺38 and to ⫺33
mV, respectively (920). Pressurized afferent arterioles in the
HNK preparation had a resting EM of ⫺40 to ⫺45 mV that
was close to the activation threshold of ⫺30 mV of L-type
VOCCs (928, 1388).
Increasing the luminal pressure from 40 to 140 mmHg in
cerebral arteries of WKY rats and cats reduced the EM of
VSMCs from ⫺57 to ⫺41 mV, followed by an action potential spike and a contraction (561, 564). Harder and associates (466, 563) presented evidence that increased cerebral transmural pressure increased 20-HETE that activated
PKC which inhibited BKCa channels. This depolarized the
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
VSMCs and activated L-type VOCCs to increase [Ca2⫹]i.
Such a chain of events mediated autoregulation of cerebral
blood flow (466, 563).
Ca2⫹ entry through L-type VOCCs is required for renal
myogenic and MD-TGF autoregulatory adjustments of the
preglomerular vasculature (FIGURE 3D). Autoregulation is
abolished following blockade of VOCCs, or removal of
extracellular Ca2⫹ from the isolated perfused kidney of the
dog (1106, 1108, 1110, 1112, 1120) or rat (1285, 1323) or
the kidney of the dog (760, 1069, 1089, 1120) or the rat in
vivo (102, 499, 646, 749, 852). CCBs prevent preglomerular myogenic responses of isolated cortical radial arteries or
afferent arterioles of the dog (562) or rat or mouse (861,
890) or rabbit (52). Ca2⫹ entry via VOCCs in the renal
vasculature is essential also for autoregulatory adjustments
of the JMN preparation (192, 194, 200, 201, 688, 1020,
1281, 1437) and the chronic HNK model (586, 587, 593,
929, 931, 1446). CCBs eliminated the MD-TGF response in
the JMN (194, 201) and the isolated JGA preparations (82).
Pharmacological studies implicated T-type VOCCs in the
myogenic response of afferent arterioles (407). However,
recent patch-clamp studies of voltage-activated Ca2⫹ currents indicated that freshly isolated VSMCs of the rat afferent arteriole have a high density of L-type channels, but do
not express functionally active T-type VOCCs (1388).
Myogenic tone in the perfused hindlimb was attributed exclusively to Ca2⫹ entry through Cav l.2 L-type channels
(1024), whereas both L- and T-type VOCCs participate in
the myogenic response in rat cerebral arteries (4, 576).
The regulation of CaV1.2 L-type VOCC trafficking has
been studied in cerebral arteries where insertion of activated
CaV1.2 into the cell membrane enhanced basal and myogenic tone (1058). Stretch-induced influx of Ca2⫹ generated
IP3 and released Ca2⫹ from the sarcoplasmic reticulum.
Increased myogenic tone and elevated [Ca2⫹]i were greater
in rat brain parenchymal arterioles than in middle cerebral
arteries (264), likely because of a higher density of VOCC
and a lower BKCa activity.
PI3K/Akt signaling promoted voltage-dependent trafficking
of Ca2⫹ channels to the plasma membrane that increased
Ca2⫹ influx via L-type VOCC and thereby was involved in
the myogenic response in mouse second-order mesenteric
arteries (197, 910). Genetic knockdown of Dicer for 5– 6
wk to reduce the global role of microRNAs (miRNAs) lowered BP and myogenic tone in mesenteric arteries by reducing PI3K/Akt signaling and L-type VOCC activity (specifically the ␣2␦1 subunit) (101). Interestingly, myogenic tone
was restored by stimulation of PI3K signaling by short-term
ANG II exposure and by activation of L-type VOCC channels by BAY K8644. Thus noncoding miRNAs have been
identified as regulators of vascular contractility and the
myogenic response. Future studies are warranted to inves-
tigate the role of this novel mechanism in renal autoregulatory responses.
In summary, L-type VOCCs play an essential role in the
transmembrane Ca2⫹ flux that is required to initiate a myogenic contraction, but the role of T-type VOCCs remains
controversial and likely depends on the experimental conditions.
2. Calcium mobilization mediated by IP3Rs
Ca2⫹ is stored in the sarcoplasmic reticulum in VSMCs at
high millimolar concentrations from where it can be released by either IP3 or ryanodine acting on specific receptors/channels (FIGURE 8). Type 1 IP3Rs were expressed on
VSMCs of intrarenal arteries, afferent arterioles, and glomerular mesangial cells, and type 3 IP3Rs were expressed on
VSMCs and mesangial cells (596, 1018). The mRNAs for
IP3 type 1 and 2 receptors have been identified in VSMCs of
small renal arteries (1197).
Increased PP in renal arteries and afferent arterioles activated PLC which enhanced the production of IP3 and subsequent Ca2⫹ mobilization (1059). On the other hand, inhibition of IP3 receptor-mediated [Ca2⫹]i mobilization in
the dog perfused kidney did not affect steady-state RBF
autoregulation, although it attenuated NE-induced renal
vasoconstriction (1111). Autoregulatory adjustments of afferent arterioles of JMN were attenuated by reducing Ca2⫹
mobilization from internal stores (688). Inhibition of Ca2⫹
uptake into the sarcoplasmic reticulum by Ca2⫹-ATPase
prevented myogenic contraction of afferent arterioles in the
JMN preparation (688), but the relative involvement of
IP3Rs and RyRs in release of stored Ca2⫹ was not investigated. Inhibition of PLC and thus of IP3 generation in the
cortical radial artery attenuated myogenic contractions by
50%, but this was limited to the proximal arterial segments
(1447). Thus there are different mechanisms for mobilization of Ca2⫹ for the myogenic response along the cortical
radial artery of JMN and perhaps the preglomerular vasculature in general. Nevertheless, Ca2⫹ mobilization contributes importantly to the myogenic mechanism, and perhaps
also to MD-TGF in the renal microvasculature of at least
some nephron populations.
The myogenic response of VSMCs can be engaged after
activation of PLC, generation of IP3, and activation of IP3sensitive receptors on the sarcoplasmic reticulum to release
Ca2⫹ from internal stores. Alternatively, DAG produced by
PLC in VSMCs can activate ion channels including TRPC6
in the plasma membrane to increase [Ca2⫹]i by transmembrane flux. Stretch- or pressure-induced recruitment of
GPCR may increase PLC production of IP3 and DAG during the myogenic response (1137). Inhibition of PLC in
nonrenal vessels hyperpolarized the EM from ⫺44 to ⫺53
mV, diminished Ca2⫹ entry, reduced [Ca2⫹]i, and attenuated the myogenic response (301, 724). The PLC/DAG sig-
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
441
RENAL AUTOREGULATORY MECHANISMS
naling pathway participated in myogenic responses of human subcutaneous arteries (266). DAG may activate
TRPC6 channels to depolarize the plasma membrane to
initiate a rise in [Ca2⫹]i via VOCCs. However, IP3 of skeletal muscle arterioles did not influence Ca2⫹ release during
myogenic responses (827). The specific role of TRPC channel activation by DAG in the myogenic response of the renal
microcirculation is poorly understood and warrants further
investigations.
3. Calcium mobilization mediated by RyRs and
cADP-ribose
Ca2⫹-induced Ca2⫹ release (CICR) of VSMCs refers primarily to activation or sensitization of RyRs, in particular
RyR2 and RyR3, to mobilize intracellular Ca2⫹, thereby
releasing more Ca2⫹ from the sarcoplasmic reticulum (49).
Activation of GPCRs is linked to ADP-ribosyl cyclase that
generates cADP-ribose that sensitizes the RyR to mobilize
Ca2⫹ and enhance vasoconstriction (FIGURE 8) (49, 399,
400, 1456, 1457). This is a relatively newly recognized
VSMC signaling pathway. Ca2⫹ entry through L-type
VOCCs also enhances RyR-mediated Ca2⫹ mobilization
and renal vasoconstriction (401, 1457), thereby implicating
cADP-ribose and RyR in Ca2⫹ signaling in the myogenic
response especially in the maintenance phase. A similar
mechanism may explain the coupling of RyRs to stretchactivated cation channels (1032).
O2⫺
produced in VSMCs by nicotinamide adenine dinucleotide phosphate (NADPH) oxidase may link activation of
GPCRs to stimulation of ADP-ribosyl cyclase production of
cADP-ribose that increases Ca2⫹ signaling and renal vasomotor tone (399, 400, 1651). Stretch of afferent arterioles
stimulated NOX-2/p47phox-dependent production of O2⫺
that enhanced the strength of the myogenic response (864,
866). Moreover, O2⫺ generated by NOX-2 in stretched cardiac myocytes and pulmonary VSMCs sensitized RyRs to
enhance [Ca2⫹]i (349, 1201). The extent to which RyRs
mediate Ca2⫹ mobilization in renal afferent arterioles, and
their involvement in the myogenic response, requires further investigation.
Afferent arteriolar vasomotion following activation of Ltype VOCCs and Ca2⫹ entry, which is an index of intrinsic
myogenic tone, was prevented by inhibition of RyRs, but
not by inhibition of IP3Rs (1443). In fact, caffeine enhanced
the fast oscillations in vasomotor tone, indicating that Ca2⫹
entry via L-type VOCCs was reinforced by Ca2⫹ mobilization by RyRs (1443).
An increase in luminal pressure in cerebral arteries induced
Ca2⫹ sparks that were generated by RyRs acting on the
sarcoplasmic reticulum adjacent to the plasma membrane
(720, 919). The local increase in [Ca2⫹]i activated BKCa in
the VSMC plasma membrane to hyperpolarize the cell, and
thereby inactivate L-type VOCCs. As a result, there was a
442
reduced Ca2⫹ entry that limited [Ca2⫹]i and opposed myogenic tone (16, 720, 810). The same sequence of events was
reported in mesenteric arteries of one study (836), whereas
myogenic tone of mesenteric arterioles was independent of
Ca2⫹ mobilization in another (1577).
Blockade of RyRs in VSMCs of cremaster muscle feed arteries by ryanodine or tetracaine inhibited Ca2⫹ sparks and
Ca2⫹ waves, BKCa activity, global [Ca2⫹]i, and myogenic
tone. On the other hand, blockade of RyRs in smaller cremaster muscle arterioles did not affect Ca2⫹ signaling or
vascular tone (1596, 1597). In fact, Ca2⫹ mobilization mediated by IP3R enhanced global [Ca2⫹]i and myogenic tone
in both vessel types without affecting Ca2⫹ sparks. Other
studies reported that activation of RyRs increased global
[Ca2⫹]i, Ca2⫹ entry through L-type VOCC, and myogenic
tone in isolated rat skeletal muscle gracilis arterioles but not
in mesenteric arterioles (1576, 1577). Depletion of sarcoplasmic reticulum Ca2⫹ stores by ryanodine in isolated skeletal muscle gracilis arterioles slowed the development of
myogenic tone (178). Thus there are striking regional differences in the roles of RyRs in modulating Ca2⫹ signaling
and myogenic responses. RyRs in skeletal muscle arterioles
mobilize Ca2⫹ and induce myogenic constriction, whereas
RyRs in cerebral arteries induce Ca2⫹ sparks that activate
BKCa and inhibit myogenic tone.
As discussed in section IVB, BKCa channels in the renal
preglomerular vessels generally are inactive. Moreover,
global Ca2⫹ mobilization by RyRs lead to marked increases
in [Ca2⫹]i and pronounced vasoconstriction rather than the
vasorelaxation anticipated from activation of BKCa (49).
However, little is known about the function of small Ca2⫹
sparks in the renal microcirculation except that they likely
are triggered by integrin deformation.
Fibronectin in the extracellular matrix surrounding VSMC
is a natural ligand for ␣5␤1-integrin whose mechanical distortion in the rat afferent arteriole can activate Ca2⫹ release
via RyRs on the sarcoplasmic reticulum and trigger local
Ca2⫹ sparks, global Ca2⫹ mobilization, and contraction
2⫹
(FIGURE 8) (64, 214). ␣5␤1 Integrins also initiated Ca
waves that propagated along preglomerular VSMCs (214).
Myogenic constriction was inhibited by integrin-specific
binding peptides or by blockade of Ca2⫹ mobilization by
RyRs. Integrins in pulmonary VSMCs activated ADP-ribosyl cyclase to produce cADP-ribose that activated RyR and
released Ca2⫹ from the sarcoplasmic reticulum (1500). Repetitive cycles of RyR-mediated Ca2⫹ mobilization and
Ca2⫹ entry through plasma membrane SOC of vasa recta
pericytes may be responsible for oscillations in membrane
currents and global [Ca2⫹]i (1443, 1656).
In summary, pressure-induced membrane depolarization
and Ca2⫹ entry through VOCCs are fundamental to the
arteriolar myogenic response. The initial mechanisms me-
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
diating mechanotransduction and triggering the depolarization of the plasma membrane of VSMCs are poorly understood. Ca2⫹ sensitivity of actin/myosin contractile complex
is central to the maintenance phase of myogenic contractions. RyR-mediated Ca2⫹ sparks and BKCa activation can
oppose myogenic vasoconstriction in some vascular beds
such as the cerebral circulation, whereas RyR-induced
CICR and amplification of [Ca2⫹]i can potentiate myogenic
constriction in other circulations, likely including the afferent arteriole. Increased [Ca2⫹]i may initiate membrane depolarization by activating ClCa channels or by offsetting the
effects of membrane hyperpolarization by activation of
BKCa channels. The precise role of RyR-mediated Ca2⫹ mobilization in the renal myogenic response requires further
investigation, as does its participation in MD-TGF activity.
Overall, CICR provides an important amplification of the
increase in [Ca2⫹]i initiated by Ca2⫹ entry through L-type
VOCCs.
B. Mechanisms Increasing Calcium
Sensitivity
Although an elevated [Ca2⫹]i is a key requirement for initiating the myogenic response in arteries and arterioles,
[Ca2⫹]i may wane over time despite a sustained myogenic
vasoconstriction (290). This dissociation implies a change
in the balance between the MLCK and MLCP and the phosphorylation state of the actin/myosin proteins to favor contraction (59, 1328). The Ca2⫹ sensitivity of the contractile
apparatus of VSMCs is defined by the amount of active
phosphorylated 20-kDa MLC (MLC20) at a given level of
[Ca2⫹]i and is regulated by PKC and the Rho/Rho kinase
(FIGURE 8) (1392). Increased vascular transmural pressure
facilitated Ca2⫹ sensitization and contraction mediated by
the RhoA/Rho-associated kinase pathway that inhibited
MLCP by phosphorylating the MLCP targeting regulatory
subunit (MYPT1) and/or the 17-kDa PKC (743). The 17kDa form of PKC in VSMCs inhibited MLCP that potentiated protein phosphatase 1 inhibitor protein (CPI-17). This
increased MLC20 phosphorylation and thereby favored
VSMC contraction. However, other mechanisms contribute to Ca2⫹ sensitivity. Thus PKC in VSMCs also phosphorylates MLC20 via MLCK that magnifies Ca2⫹-induced actin-myosin interaction and force development. Actin polymerization enhanced connections between the actin
cytoskeleton, the plasma membrane, and the extracellular
matrix which promoted force transmission (1548). Both
cytochrome P-450 – 4A/20-HETE and ROS enhanced Ca2⫹
sensitivity (1328). Endothelial factors such as ET-1 and
TxA2 also increased Ca2⫹ sensitivity and myogenic vasoconstriction in gracilis muscle arterioles (1502). Inhibition
of PKC, but not the cytochrome P450 inhibitor 17-ODYA,
in isolated arterioles from rat skeletal muscle reduced myogenic tone (63). Moreover, smoothelin-like proteins have
been implicated in VSMC contractility and vascular adaptations during hypertension. Smoothelin-like 1 protein at-
tenuated the myogenic response in cerebral arteries possibly
by inhibiting the expression of MYPT1 (1496).
Thus Ca2⫹ sensitivity is a complex process involving PKC
and Rho/Rho kinase signaling and other mechanisms that
regulate the phosphorylation state of the contractile proteins. These processes play an important role in myogenic
constriction.
1. PKC
PP and GPCR agonists stimulate phospholipid turnover
that generates DAG that in turn activates PKC (FIGURE 8).
Translocation of PKC to the plasma membrane triggered a
cascade of signaling that enhanced Ca2⫹ sensitivity (1275).
Both the Ca2⫹-dependent PKC␣ and the Ca2⫹-independent
PKC⑀ enhanced the sensitivity of myofilaments to [Ca2⫹]i,
whereas the cytoskeletal PKC␦ and the nuclear PKC␨ were
not effective (324). Stretch of rabbit basilar arteries triggered translocation of PKC␣ (but not PKC⑀) and RhoA
from the cytosol to the plasma membrane which increased
phosphorylation of MLC and contributed to the Ca2⫹ sensitivity and the contractile response (1634).
PKC is considered crucial for renal vasoconstriction produced by ANG II and ET-1 (22, 68, 1265, 1322, 1659). The
myogenic tone of afferent arterioles of the rat HNK preparation was mediated in part by the basal level of PKC that
enhanced Ca2⫹ sensitivity, independent of voltage-dependent activation (802).
Inhibition of PKC attenuates myogenic tone in basilar arteries (1634), cerebral arteries (63, 474, 603, 725, 855,
1136, 1136, 1164), coronary arteries (324, 1006), mesenteric arteries (380), cremaster and skeletal muscle arterioles
(63, 977), and subcutaneous arteries (266). Inhibition of
PKC in skeletal muscle arteries prevented a myogenic response to a PP increase without affecting the increase in
[Ca2⫹]i (768). This was specific for PKC␣ in coronary arteries (324). Pressure-induced constriction of cerebral arteries was dependent on increased Ca2⫹ sensitivity due to PKC
and Rho kinase (63, 603, 855).
In summary, PKC activation is a central component of Ca2⫹
sensitivity and the myogenic response in most arteries studied, including the renal preglomerular vessels.
2. Rho/Rho-kinase
As mentioned above, Rho kinase can enhance contractility
by inhibiting MLC phosphatase by targeting MYPT1 (FIGURE 8). Mechanical stimulation of VSMCs translocates the
GTPase Rho and its associated Rho kinase to caveolin-1 in
the plasma membrane. The myogenic response was enhanced by RhoA/Rho-kinase phosphorylation that inactivateed MLCP and augmented Ca2⫹ sensitivity (350). Inhi-
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
443
RENAL AUTOREGULATORY MECHANISMS
bition of this system, or disruption of caveolin-1, reduced
myogenic force in mesenteric and ophthalmic arteries (380,
694). Rat cerebral artery myogenic tone depended on the
independent effects of Rho-kinase and PKC mediating increased Ca2⫹ sensitivity, but Rho kinase predominated
(694, 725, 855). The myogenic response of gracilis and
cremaster muscle arteries involved Rho-kinase and PKCevoked Ca2⫹ sensitization leading to inhibition of MLCP
and cytoskeletal reorganization of actin polymerization
(1027). Membrane depolarization in pulmonary artery
VSMCs generated O2⫺ that activated RhoA and led to Ca2⫹
sensitization (1233). RhoA activation and interaction with
caveolin-1 were critical for pressure-induced myogenic tone
in rat mesenteric resistance arteries (350).
Juxtamedullary preglomerular arterioles expressed Rho-kinase ␣ and ␤ (691), whose inhibition completely abolished
autoregulation (691), thereby implying inhibition of both
myogenic and MD-TGF mechanisms. Indeed, recent work
showed that Rho-kinase participates in afferent arteriolar
vasoconstriction mediated by both myogenic and MD-TGF
mechanisms in response to elevated RPP in the superficial
cortex of WKY rats (632). Rho kinase in the rat HNK
preparation participated in the myogenic vasoconstriction
of cortical radial arteries and afferent arterioles to changes
in increased PP, ANG II, and reduced EM (1259, 1358).
Inhibition of Rho-kinase blunted pressure-induced myogenic tone and depolarization-induced constriction of isolated afferent arterioles (1054) and also almost abolished
the dynamic RPP-RBF transfer functions representative of
both the myogenic and MD-TGF responses (1358).
In summary, activation of Rho/Rho-kinase is a quantitatively important step in increasing the Ca2⫹ sensitivity that
maintains or enhances both the myogenic and MD-TGF
autoregulatory mechanisms in the kidney.
VI. MODULATING AGENTS
The myogenic response of renal (915) and nonrenal (301)
vessels was generally intrinsic to their VSMCs and independent of endothelial input. Thus myogenic constriction
clearly was triggered within VSMCs, although it could be
modulated by various endothelial factors. Despite the constancy of afferent arteriolar blood flow during increased PP
by effective autoregulatory adjustments in vascular resistance, the narrowed vascular lumen will increase vessel wall
shear stress that could activate endothelial pathways producing vasoactive agents. For example, NO was released by
eNOS in response to ANG II (699, 1160), ET-1 (701, 757),
or shear stress (696, 747, 1184). Other potential modulators of myogenic tone in the kidney can originate from
structures involved in MD-TGF or CT-GF with signaling
from MD or CT tubular cells (FIGURE 6).
444
A. ANG II
JG granular cells in the terminal afferent arteriole increase
renin release and thereby ANG II production (FIGURE 1) in
response to reduced luminal pressure at the end of the afferent arteriole, low NaCl delivery to the MD, or increased
sympathetic nerve activity activation of ␤-adrenoceptors
(204, 846). Proximal tubular and CD cells also contribute
to intrarenal renin and ANG II generation (812, 1076).
Most studies have concluded that the RAS was not essential
for efficient steady-state autoregulation of RBF. ANG II
infusion, or inhibition of AT1 receptors or ACE, did not
modify highly efficient whole kidney autoregulation in dogs
(5, 27, 90, 97, 103, 194, 548, 550, 755, 1003, 1047, 1072,
1171, 1261) or rats (45, 329, 750) even during low-salt diet
producing strong activation of the RAS and did not usually
affect the lower limit of RBF autoregulation. ANG II also
did not affect renal medullary blood flow autoregulation
(288). However, ANG II was more important in steadystate autoregulation of the GFR than the RBF (488, 550,
774, 1261), especially at lower RPP, most likely by maintaining efferent arteriolar tone. However, a few studies have
reported that the RAS was essential for effective steadystate autoregulation of RBF and GFR (764, 1479). Moreover, autoregulation of isolated perfused rats kidneys was
improved by ANG II (523).
Micropuncture studies of superficial nephrons demonstrated that MD-TGF responses were enhanced by ANG II
(133, 1012, 1308, 1309) and diminished by blockade of
ACE (133, 648, 1189, 1422, 1495), or AT1 receptors (781,
1309, 1590), or by deletion of the genes for ACE or AT1A
receptors (582, 1310, 1318, 1481). AT1 receptors may
modulate MD-TGF responses by actions on the vasculature
or on NKCC2 cotransport activity in MD cells (829, 1012,
1556). These effects were countered by NO production by
MD nNOS (656).
One mechanism of MD-TGF enhancement by ANG II is
synergistic potentiation of the actions of adenosine on the
afferent arteriole (440, 556, 862, 863, 1159, 1480, 1580).
Thus ANG II appears to be an important modulator of
MD-TGF activity, but its presence or absence usually has
little impact on highly efficient steady-state autoregulation
of RBF, at least during the plateau phase.
The effects of ANG II on the myogenic mechanism in the
kidney are controversial. ANG II contracted the afferent
arteriole of the JMN preparation without affecting its autoregulatory response (690). However, low levels of ANG II
enhanced the myogenic response in the HNK preparation
by potentiating PKC-induced increases in Ca2⫹ sensitivity
(802). ANG-(1–7) had no effect on the myogenic tone of
preglomerular vessels in the HNK preparation (1524).
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
Short-term infusion of ANG II at pressor rates caused renal
vasoconstriction without major effects on the myogenic response, the MD-TGF, or the third mechanism analyzed by
time-dependent changes in RVR following a rapid, step
increase in RPP (750, 755). ACE inhibition has no effect on
the myogenic response, and the MD-TGF was slightly attenuated (755). Short-term administration of vasoconstrictor amounts of ANG II to the conscious dog slightly augmented the admittance gain of the transfer functions representing the myogenic and MD-TGF responses, but without
affecting their relative contributions (755). Prolonged infusions of ANG II, but not PE, into conscious or anesthetized
rats potentiated the admittance gain of the myogenic response but not the MD-TGF (1193). Whereas ANG II
slightly increased the slope of the admittance gain reduction, consistent with increased strength of the myogenic
mechanism, it did not modify the signature resonance peak
of the myogenic mechanism (at ⬃0.23 Hz), which is an
indicator of the speed of contraction. The ANG II effect is
attributed to activation of AT1 receptors rather than increased RPP. Other frequency analysis studies indicated
that increased endogenous ANG II enhanced the myogenic
and MD-TGF responses in rats (329, 597), but reduced
ANG II activity produced by high salt diet had little effect
(1271, 1273). Other studies employing dynamic frequency
analysis reported that acute hypertension produced by PE
does not affect the myogenic or the MD-TGF mechanism in
Wistar (1358) and Brown-Norway (BN) rats (1570).
Staircase changes in RPP can produce hysteresis loops
where there are different RVR responses to increases versus
decreases in RPP in the anesthetized (284) and conscious rat
(423). Autoregulation curves for RBF had a lower plateau
level when measured during sequential increases compared
with subsequent decreases in RPP. This was attributed to
activation of the RAS by hypotension. The outcome was
that the RAS enhanced RVR at low RPP where it engaged
the MD-TGF for RBF autoregulation (284).
Early micropuncture studies in the rat reported reductions
in RPP below the normal lower limit of autoregulation
abolished MD-TGF, but this was slowly restored by ANG II
generation after 20 – 60 min of maintained hypoperfusion
(1343). The increased RAS activity reduced the lower limit
for RBF autoregulation (284, 624, 707, 1397).
The putative third mechanism may be modulated by ANG
II (1342). ANG II can enhance CT-GF leading to a paradoxical dilation of the afferent arteriole by activating ENaC
and stimulating Na⫹ reabsorption in the CT (1227).
Sex differences have been reported for expression of AT1
and AT2 receptors and their functional impact on renal
autoregulation, MD-TGF, and pressure-natriuresis. Kidneys of female rats express a higher density of vasodilator
AT2 receptors that can increase the lower limit of RPP for
RBF autoregulation and shift the pressure-natriuresis relation to excrete more Na⫹ at a lower RPP (619). Augmented
AT2 receptor signaling in female rats also accounted for a
blunted effect of ANG II to enhance MD-TGF (156).
In summary, ANG II is a potent vasoconstrictor of the afferent and efferent arterioles. Reductions in RPP release
renin and generate ANG II within the kidney which preferentially activates AT1 receptors on the efferent arteriole,
thereby maintaining PGC and GFR. This accounts for the
particular effect of ANG II in maintaining autoregulation of
GFR, relative to RBF, at reduced RPP. Additionally, agonist-independent, stretch activation of AT1 receptors can
interact with TRP channels as mechanosensors (see sect.
IIIC). ANG II clearly magnifies MD-TGF responses, but
most studies suggest a rather weak effect on the myogenic
response. Yet, although the individual components of renal
autoregulation may be modified by acute ANG II, this generally does not cause major changes in overall steady-state
RBF autoregulation. The exception is that ANG II generated by hypotension can modulate the lower limit of PP
steady-state autoregulation, perhaps through its ability to
enhance MD-TGF activity. As noted in section VIIB5,
chronic ANG II infusion causes hypertension, but the effects on the efficiency of RBF autoregulation are inconsistent.
B. Eicosanoids
Vascular production of vasodilatory PGs such as PGE2 and
PGI2 commonly buffer the actions of vasoconstrictor agents
(671), but renal hemodynamics also are modulated by vasoconstrictor PGs and by tubular production of eicosanoids. A reduction in RPP reduced the release of PGE2
and PGI2 from the kidney. However, PGs did not affect
steady-state RBF autoregulation in salt-replete rats (422) or
dogs (1531) or autoregulation of the PGC in superficial
nephrons of normal Munich-Wistar rats (1167). Moreover,
arachidonate metabolites of COX or cytochrome P-450
were not required for steady-state RBF autoregulation in
isolated perfused kidneys of the dog (763, 1107) or rat
(1285) or for myogenic contractions of afferent arterioles
(588, 747).
Nevertheless, other studies have implicated PGs under certain circumstances. Thus the poor autoregulation of papillary blood flow in the rat kidney was improved by COX
inhibition (1127), while whole kidney RBF autoregulation
was slowly enhanced following COX inhibition (833). Efficient autoregulation of GFR in superficial nephrons depended on COX metabolites contributing to MD-TGF
(1313). COX products also participated in conventional
MD-TGF-mediated adjustments of afferent arteriolar tone.
Micropuncture studies of superficial nephrons demonstrated that COX inhibition reduced MD-TGF sensitivity to
changes in tubular flow (1034, 1307, 1319), but this was
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
445
RENAL AUTOREGULATORY MECHANISMS
restored by tubular perfusion with PGE2 or PGI2 but not
PGF2␣ (1319). However, other investigators reported that
MD-TGF sensitivity was enhanced by tubular perfusion
with PGE2 and PGF2␣, whereas perfusion with PGI2 markedly attenuated feedback sensitivity (1170).
omega hydroxylase (445, 446). It augmented autoregulation of the afferent arteriole in the JMN preparation with
participation of both myogenic and MD-TGF mechanisms
(673, 676) and enhanced the myogenic tone of rabbit and
mouse isolated afferent arterioles (464).
More consistent is the observation that MD-TGF is enhanced by MD generation of prostaglandin endoperoxide
(PGH2) and TxA2 (137, 1589, 1593) or by isoprostanes, all
of which activate TP receptors that can constrict afferent
arterioles and also can enhance Cl⫺ reabsorption from the
TAL of Henle’s loop (38, 437, 932, 1319, 1581, 1591).
Thus activation of TP receptors may augment MD-TGF by
enhancing both the MD signal generated by NaCl reabsorption and the afferent arteriolar responsiveness. TxA2 in gracilis muscle arteries also increased myogenic tone (1502),
which was ascribed to enhanced Ca2⫹ sensitivity of the
contractile apparatus.
Increased transmural pressure in cerebral arteries increased
20-HETE production that activated L-type VOCC currents, myogenic vasoconstriction, and autoregulation of
blood flow (468, 1181). 20-HETE also augmented the myogenic responses of cerebral, skeletal muscle, and mesenteric
arterioles (243, 468). These vasoconstrictor actions of 20HETE were ascribed to blockade of BKca channels on
VSMCs which caused depolarization and activation of
Ca2⫹ entry via VOCCs (677, 943, 1666). 20-HETE also has
been implicated in RBF autoregulation in vivo in some
(1668) but not all studies (1199). It enhanced MD-TGF
activity independent of NKCC2-mediated MD reabsorption, presumably by magnifying the vasoconstrictor response to MD-derived adenosine and/or ATP (464, 1668).
In contrast, vasodilator EETs generated by arachidonate
cytochrome P-450 epoxygenase in endothelial cells in the
JMN preparation blunted the afferent arteriolar autoregulatory vasoconstriction (673). It is not yet clear how cytochrome P-450 enzymes are modulated by RPP is not yet
clear.
Sex differences have been noted in the control of blood
perfusion within the kidney by COX metabolites (1152,
1412). Blood flow autoregulation in cortical and outer medullary regions of kidneys of male rats was efficient and
independent of PGs, whereas poor medullary autoregulation in female rats was attributed to overproduction of PGs.
COX-2 generated vasodilatory metabolites in the JMN
preparation during increased activation of MD-TGF, which
counteracted MD-TGF-mediated constriction of the afferent arteriole and blunted its constrictor response to increased RPP (657, 658). In contrast, COX-1 knockout mice
have a normal MD-TGF response (39). These actions of
COX-2 metabolites were attributed to activation of MD
(39, 232, 577, 657). Thus the rather inconsistent effects of
COX inhibition on renal autoregulation in health may relate to offsetting effects of MD release of vasodilator and
vasoconstrictor prostanoids, confounding effects of sex
differences between COX-1 and-2, and moderating effects
of NO.
Products of COX-2 and nNOS produced in MD cells of the
rat JMN preparation attenuated autoregulation (657, 658).
Inhibition of nNOS no longer enhanced MD-TGF after
blockade of COX-2. Therefore, COX-2 products appear to
blunt autoregulation secondary to MD nNOS generation of
NO. COX inhibition did not affect the myogenic response
of the rat HNK lacking MD-TGF (588), consistent with a
primary effect of vasodilator PGs on MD signaling.
Global inhibition of cytochrome P-450 production of 20HETE and EETs in the JMN preparation (469) or the dog
arcuate artery (777) attenuated the myogenic response. In
contrast, myogenic responses were not affected by global
inhibition of COX-1 and -2 activity. 20-HETE was generated during increased PP in the preglomerular vasculature
and the gracilis muscle arteries by a cytochrome P-450
446
In summary, arachidonate metabolites have complex effects
on both myogenic and MD-TGF mechanisms. PGs generally are not required for overall renal autoregulation. However, MD-TGF response can be enhanced by vasoconstrictor prostanoids and TxA2 and vasoconstriction can be
blunted by vasodilator PGs. In most studies, myogenic responses are increased by 20-HETE and reduced by EETs.
C. NO
Endogenous NO has complex effects on renal autoregulation. It can blunt both myogenic and MD-TGF mechanisms. Whereas NO contributes importantly to the regulation of RVR in many preparations, it does not affect the
overall efficiency of steady-state whole kidney RBF autoregulation over a defined range of RPP. Short-term inhibition
of NOS or administration of an NO donor does not perturb
steady-state RBF autoregulation in the normotensive rat,
the SHR, or the dog (69, 81, 815, 853, 952, 953, 960, 963,
1052, 1109, 1492) nor does it perturb renal cortical or
medullary blood flow autoregulation in the dog (960). NOS
inhibition improved the poor papillary autoregulation
without affecting efficient RBF autoregulation in one study
of the hydropenic Munich-Wistar rat (1127).
Although not affecting steady-state RBF autoregulation in
the plateau phase, NO modulated the range of RPP over
which RBF was autoregulated by shifting the inflection
point to a lower RPP in some (831, 832, 1203), but not all
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
studies (69, 81, 815, 853, 952, 953, 960, 963, 1109, 1492,
1494).
On the other hand, increased NO release in the JMN preparation attenuated PP-induced diameter adjustments of the
cortical radial artery and afferent arterioles (675, 1184) by
reducing myogenic and MD-TGF responses. Inhibition of
NOS improved the myogenic response of afferent arterioles
of JMN preparation perfused with a cell-free artificial solution (675, 1184). Autoregulation of this preparation was
improved with the use of erythrocytes or albumin in the
perfusate that scavenge NO (192, 1168, 1184). Thus endogenous generation of NO can blunt both the MD-TGF
and myogenic mechanisms and yet has little effect on
steady-state autoregulation of RBF.
Moreover, global inhibition of NOS, or selective blockade
of nNOS, magnifies MD-TGF responses in superficial
nephrons whether assessed by micropuncture and microperfusion (38, 116, 133, 188, 661, 761, 1052, 1228,
1463, 1493, 1515, 1553, 1586, 1587, 1602, 1608, 1609) or
by wavelet analysis of blood flow oscillations of superficial
nephrons (1402). Oscillations in RBF and RVR often become readily apparent after NOS inhibition. Comparison
of oscillations in blood flow in the efferent arteriole and the
whole kidney led to the conclusion that NO attenuates
the strength of arteriolar oscillations mediated by both the
myogenic and the MD-TGF mechanisms while improving
their synchronization (1402).
Dynamic studies have demonstrated that NO slows and/or
attenuates renal myogenic vasoconstriction. Inhibition of
NOS in the rat augments the strength and speed of the
myogenic response to a rapid, step increase in RPP (FIGURE
7A). This increased the contribution of the myogenic mechanism from 35 to 80% (299, 750) or 45 to 100% (1626). As
is shown in FIGURE 7A, NOS inhibition also shortens the
time to completion of a myogenic response from ⬃8 to 4 s.
Indeed, inhibition of NOS augmented the myogenic response sufficiently to induce complete autoregulation
thereby providing transmission of little error signal to MD
cells to activate MD-TGF.
Vascular admittance transfer function analysis in conscious
dogs (754) or in some (750, 1357, 1358, 1402, 1573), but
not all (1340), conscious or anesthetized rats confirmed that
endogenous NO can attenuate the strength of the renal
myogenic response. NOS inhibition primarily enhanced the
myogenic mechanism by increasing the slope of reduction in
admittance gain between 0.05– 0.20 Hz (FIGURE 7B) and
the height of the associated phase peak in normotensive
(Long-Evans, SD, and Wistar) rats but not in SHR (1341,
1358, 1563, 1569), which could possibly be explained by
reduced NO signaling or bioavailability in their kidney due
to excessive ROS. The slope of the decline in admittance
gain increased from ⬃30 to ⬃55 dB/decade, suggesting the
emergence of a second-order component responding to the
rate of change in BP in addition to the actual level of BP.
NOS inhibition increased the fractional compensation from
40 to 58%, which indicates an enhanced vigor of the myogenic response. However, in this study, the corner frequency where admittance began to decline was largely unaffected by NOS inhibition, which indicates a maintained
rate of myogenic contraction. Moreover, the markedly attenuated myogenic response in normotensive BN rat with
forced increases in RPP (1563) was normalized by inhibition of NOS (1563, 1569, 1570) (see sect. VIIB8).
There is controversy about the origin of the NO that modulates the renal myogenic response. Early studies of rabbit
isolated perfused afferent arterioles or afferent arterioles of
the rat HNK indicated that the NO that attenuated the
myogenic responses was generated by the endothelium
(586, 593, 696, 702, 747). Moreover, stimulation of endothelial-derived NO by activation of vascular ETB receptors
blunted the myogenic response of the rat isolated renal and
mesenteric arteries (1096) and the whole kidney (1357). In
contrast, a unique renal source was implicated by the finding that NO modulated the myogenic response in the rat
kidney but not in the hindlimb (750). This was ascribed to
blunting of myogenic tone by nNOS-derived NO in the MD
rather than by NO produced by the vascular endothelium
(FIGURE 7). However, NOS inhibition in some other studies
of nonrenal vessels, such as skeletal muscle arteries, mesenteric arteries, and splenic arterioles, accelerated or magnified myogenic constriction, thereby implicating eNOS in
these extrarenal vascular beds (154, 377, 519, 944, 1339,
1530).
Selective inhibition of nNOS in the JMN preparation enhanced the autoregulatory constrictor responses to increased PP of the afferent arteriole (FIGURE 7D) (657, 658,
662). The NO signal arose from MD cells since the effect of
nNOS inhibition was abolished by elimination of MD-TGF
by papillectomy. However, it is not clear from these JMN
studies whether NO affects primarily the myogenic or the
MD-TGF mechanism. Support for the view that MD-derived NO attenuates myogenic tone derives from the finding
that NOS inhibition did not enhance myogenic constriction
in preparations lacking functional MD cells, i.e., mouse
isolated perfused afferent arterioles with endothelium present (866) or rat afferent arterioles in the HNK preparation
(FIGURE 7C) (1358). Moreover, isolated perfused afferent
arterioles from eNOS-deficient mice retained normal myogenic response (866). In addition, NOS inhibition did not
affect afferent arteriolar constrictor responses in the artificial fluid-perfused JMN preparation to increased RPP
(675). These studies appear to exclude a major role for NO
derived from eNOS in afferent arterioles and implicate
nNOS-derived NO in the MD as the primary source that
attenuates the strength of the myogenic response in intact
kidneys.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
447
RENAL AUTOREGULATORY MECHANISMS
Collectively, there are several lines of evidence to support
the notion that NO generated by MD cells attenuates myogenic response of the preglomerular vasculature by buffering the speed and/or the strength of the myogenic mechanism (750, 1358, 1569). Support derives from studies of
renal vascular dynamics based on responses to a rapid, step
increase in RPP, on frequency analysis of transfer function
admittance gain and a dependence on MD-TGF activity. In
some studies, the myogenic response after acute global NOS
inhibition became so rapid and robust that the autoregulatory response to a step increase in RPP was complete within
10 s without any evidence of participation of a MD-TGF or
other mechanisms. The exaggerated myogenic system during NOS inhibition appeared to have become sufficiently
potent to prevent an error signal from reaching the MD to
activate MD-TGF and other mechanisms. Global NOS inhibition may prevent modulation of myogenic responses by
NO generated from any source. However, the enhancing
effect of L-NAME on the myogenic response was linked to
MD nNOS since it was reduced (1358) or abolished (750)
by furosemide.
Frequency analysis reported that global inhibition of NOS
in Long-Evans, SD, and Wistar rats (FIGURE 7B) (1569) or
selective intrarenal inhibition of nNOS, but not iNOS, augmented the myogenic contribution to dynamic autoregulation in SD rats (1358). Selective inhibition of nNOS in vivo
increased the slope of admittance gain decline from 24 to 42
dB/decade, a change similar to that observed with global
NOS inhibition with L-NAME (1358). These results indicate that MD-derived NO is an important pool to attenuate
the myogenic responses of the preglomerular vasculature in
the kidney.
A recent study questioned this view. Pharmacological inhibition of NOS, but not nNOS or iNOS, increased the speed
and strength of the myogenic response to a step increase in
RPP without perturbing overall steady-state RBF autoregulation in rats (299). In agreement, global NOS inhibition
magnified the myogenic component of RBF autoregulation
in wild-type mice and nNOS mutants, but not in mice lacking eNOS (299). This suggests an NO-independent role for
eNOS that warrants further investigation.
The cellular events by which NO attenuates VSMC myogenic responses are incompletely understood. In one study
of the renal microcirculation, exogenous NO signaled via
cGMP-dependent and -independent pathways depending
on its concentration (1486). Low renal NO concentrations
(⬍1 ␮M) that appear to be in the physiological range signal
primarily through cGMP (887, 1486). Kurtz and colleagues
(1283) demonstrated that NO generated cGMP that caused
renal vasodilation primarily by inhibiting phosphodiesterase 3 that degrades cAMP. NO donors, and high levels of
8-bromo-cGMP, completely abolished autoregulation of
preglomerular vessels in the JMN preparation (132), indi-
448
cating inhibition of both myogenic and MD-TGF mechanisms. Since stimulation of cGMP by atrial natriuretic peptide (306, 904, 1644) failed to restore RBF or reverse the
exaggerated myogenic response detected by transfer function analysis following NOS inhibition (1573), there must
be different pools of renal NO that regulate vascular responses.
There are multiple sites where NO, cGMP, or its target
PKG and cAMP/PKA could dampen Ca2⫹ signaling or sensitivity, and thereby moderate arteriolar tone (434). NO in
the renal vasculature reduced agonist- or stretch-induced
increases in Ca2⫹ influx by inhibiting VOCCs or TRPC
channels or by activating BKCa channels (403, 408, 832,
898, 1011, 1642, 1645). NO in the afferent arteriole is
reported to inhibit both L- and T-type VOCCS (408), which
could account for its ability to antagonize both MD-TGF
and myogenic responses. Moreover, NO in VSMCs suppressed ADP-ribosyl cyclase activity, thereby reducing
RyR-mediated Ca2⫹ mobilization (403, 898, 1642). Both
NO and ROS in VSMCs regulated RyR-mediated Ca2⫹
mobilization (384, 1645), but by distinct signaling pathways (752, 861, 866, 1224). NO dilated gracilis muscle
arteries and afferent arterioles by activating MLCP (129,
1499, 1503), and attenuated myogenic tone of nonrenal
vessels by inhibiting Rho kinase in VSMCs (475, 1214,
1458). This may also occur in the renal microcirculation
(1358). NO inhibition of 20-HETE formation or action
diminished myogenic constriction of coronary arteries (21,
645, 921, 1144). Thus NO has multiple mechanisms that
can reduce vascular tone and/or Ca2⫹ sensitivity.
NO also could impact the myogenic response indirectly by
attenuating MD-TGF signaling. NO from nNOS can inhibit reabsorption of NaCl by activated MD cells, thereby
blunting the signal for MD-TGF responses (702, 1460,
1602). NO and GMP signaling in the MD phosphorylates
and inhibits NKCC2 and thereby reduces luminal NaCl
uptake (828, 1228). Moreover, NO inhibited ecto-5’-nucleotidase that is the final enzyme in the conversion of ATP to
adenosine, which can limit MD-TGF activation (1287). NO
and O2⫺ interact in MD cells where O2⫺ inactivates NO and
enhances MD-TGF responses during oxidative stress (913,
1393). The effect of NOS inhibition to enhance the renal
myogenic response can be dissociated from renal or systemic vasoconstriction in most (750, 754, 1358, 1573) but
not all studies (1572). The enhancement of myogenic responses by NOS inhibition in vivo can be prevented by
furosemide, thereby implicating MD-TGF (755, 1193,
1563).
In summary, NO has complex interactive effects on renal
autoregulation. It modulates both the dynamics and the
steady-state contribution of individual mechanisms. There
are mixed results for a role of NO derived from the endothelium of the afferent arteriole in blunting the myogenic
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
response. More convincing evidence demonstrates that NO
generated by nNOS in the MD not only attenuates MDTGF but also restricts the speed and strength of the myogenic response. During NOS inhibition, the myogenic tone
can become sufficiently strong to account for nearly all of
the highly efficient renal autoregulation (FIGURE 7, A AND
B). The possible role of NO in vasomotion of isolated arteries/arterioles was discussed in section IIE.
D. Carbon Monoxide
Constitutively active heme oxygenase 2 (HO-2) and inducible HO-1 catalayze heme to CO, iron, and bilverdin.
HO-2, but not HO-1, is highly expressed in endothelial and
VSMC. CO produced a biphasic response in the renal vasculature, with vasodilation at low concentrations (100 nM)
and vasoconstriction at high concentrations (1–10 ␮M)
(1445). These vasoactive effects were ascribed to stimulation of NO production by endothelial cells and inhibition of
eNOS, respectively. CO attenuated the actions of vasoconstrictors in isolated interlobar arteries by activating VSMC
K⫹ channels (1653).
Induction of HO-1 in rat kidneys increased basal RBF but
did not impair the efficiency of RBF autoregulation (130).
However, induction of HO-1 in tubular epithelial cells in
the rat JMN preparation increased CO production that
blunted the myogenic tone of afferent arterioles, whereas
biliverdin was without effect (131). The reason for discrepancy between whole kidney autoregulation in vivo and that
recorded for myogenic responses of the afferent arteriole of
JMN in vitro requires further investigation.
The MD expressed HO-1 and HO-2 that generated CO and
inhibited MD-TGF in vivo via activation of sGC and cGMP
signaling (1557). Biliverdin appeared to attenuate MDTGF activity by scavenging O2⫺ (1557). Both CO and biliverdin inhibited MD-TGF in the rabbit isolated JGA preparation (FIGURE 6) (1222, 1225, 1226).
The HO inhibitor chromium mesoporphyrin decreased renal HO-1 and CO but did not affect RBF or GFR or NO
production in the rat (718). In another study, HO inhibition
increased RVR in the rat kidney which was more pronounced when NOS was inhibited (1242). On the other
hand, HO and CO in the renal medulla may participate in
the regulation of medullary blood flow, Na⫹ excretion, and
BP (897, 1665). CO blunted rat gracilis arteriolar myogenic
responses (830) and attenuated VSMC myogenic tone by
activating BKCa channels to increase EM (881, 1653).
In summary, CO at low concentrations in vitro in isolated
preparations generates NO that activates sGC signaling to
produce vasodilation and inhibit both the myogenic and
MD-TGF mechanisms. However, CO has little effect on
RBF autoregulation in vivo. Renal HO-1 is often induced
during pathophysiological conditions (6, 1062), which
might be a pathway to attenuate autoregulation and accelerate barotrauma. Thus, despite this potential, the roles or
the underlying mechanisms of CO and HO in autoregulation are not clearly established.
E. Endothelin-1
Endothelin is produced primarily by vascular endothelial
cells and by tubular cells of the TAL of Henle’s loop and the
CD. Renal vascular actions of endothelin (ET)-1 are mediated by vasoconstrictor ETA and ETB receptors on VSMCs
and vasodilator ETB receptors on vascular endothelial cells
linked to NO production. Steady-state RBF autoregulation
was unaffected by ETA receptor blockade (97, 140) or by
combined blockade of ETA⫹ETB receptors (831). However, blockade of ETB, but not ETA, receptors enhanced the
myogenic response in vivo and intrarenal arteries in vitro by
inhibiting NO production (453, 1357). For actions of NO
on RBF autoregulation and individual components, see section VIIC. A low concentration of ET-1 that does not contract afferent arterioles in the HNK potentiated myogenic
reactivity by increasing PKC and its modulation of K⫹
channels (802). Mechanistically, Kirton and Loutzenhiser
(802) proposed activation of PKC and modulation of K⫹
channel activity, but as discussed in next section ET-1 may
also activate NADPH oxidases.
F. Reactive Oxygen Species
Reactive oxygen species (ROS), such as O2⫺, H2O2, and
hydroxyl anion (OH.⫺), are reactive byproducts of mitochondrial respiration, uncoupled NOS, or specific oxidases,
notably NADPH oxidase, xanthine oxidoreductase, and
certain arachidonic acid oxygenases. NADPH oxidase is a
major source of renal and vascular ROS (494, 895, 1016,
1086, 1151). It is a multiunit enzyme comprising the membrane-bound neutrophil oxidase (NOX) unit, either
NOX-1, NOX-2 (gp91phox), NOX-4, or NOX-5. All vascular NOXs generate ROS, but their intracellular distribution, their requirement for interaction with the NOX subunits p22phox, p47phox (or NOXO1), p67phox (or NOXA1),
and p40phox and the small G protein Rac-1 or Rac-2, and
their modes of regulation and generated ROS (O2⫺ or H2O2)
differ (78, 136, 1086). For activation of the enzyme, participation of Rac 2 (or Rac 1) and Rap 1A (128), but also PKC
are considered crucial (136).
The kidney has a rich, diverse, and cell-type specific distribution of NADPH oxidases (471). ROS were produced in
all renal vascular and tubular cells studied (40). NOX-1
was expressed primarily in large vessels, whereas NOX-2
and NOX-4 were expressed abundantly together with their
chaperone protein (p22phox) in resistance arterioles.
NOX-5 was expressed in human, but not rodent, vessels
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
449
RENAL AUTOREGULATORY MECHANISMS
(78, 182, 873, 1150, 1477). The afferent arteriole and MD
cells both expressed NOX-2 and NOX-4 and the subunits
capable of generating O2⫺ or H2O2 (471, 1603, 1657). The
rat and rabbit afferent arteriole expressed p22phox that was
required for activation of NOX-1, NOX-2, and NOX-4
(471). NADPH oxidase-mediated O2⫺ generation in MD
cells enhanced MD-TGF in large part by limiting NO bioavailability (912, 913, 1608). Flow-dependent NaCl reabsorption, or stimulation by ANG II, primarily activated MD
NOX-2 and enhanced MD-TGF signaling, whereas NOX-4
contributed to basal ROS production (448, 912).
Oxidative stress implies an increased production, or a
decreased scavenging or metabolism of ROS. Many effects of oxidative stress can be attributed to NO deficiency. Increased NADPH oxidase-derived O2⫺ production and vascular dysfunction have been reported in hypertension, diabetes, atherosclerosis, and CKD (40, 895,
1604). This section primarily focuses on the roles of vascular O2⫺ and H2O2 in renal autoregulation in health,
whereas section VII considers the pathophysiological
contribution of ROS.
Preliminary studies suggested that renal autoregulation in
vivo to a step increase in RPP was dependent on NADPH
oxidase whose inhibition with apocynin attenuated myogenic response, but had little effect on MD-TGF and the
third mechanism. The enhanced myogenic response following inhibition of NOS was largely prevented by inhibition
of NADPH oxidase or by tempol (752). Thus the enhancement of myogenic responses by NADPH oxidase-derived
O2⫺ was offset by interactions with NO and by metabolism
of O2⫺ by superoxide dismutase (SOD). Indeed, many studies in the renal vasculature have concluded that ROS can
regulate vasomotor tone by direct effects on VSMC and by
quenching and inactivating NO (758, 1301, 1603, 1604).
O2⫺ in the JGA enhanced vasoconstriction by scavenging
and inactivating NO (1217, 1301, 1588, 1592, 1603,
1608), but also acted directly on the afferent arteriole. The
inhibition of MD-TGF by tempol was enhanced by endothelium removal (913), suggesting a primary action of ROS
and modulatory role for NO.
Activation of GPCRs in VSMCs rapidly increased O2⫺
that potentiated Ca2⫹ signaling in afferent arterioles and
renal vasoconstriction (183, 399, 400, 864). Thus renal
vasoconstriction with ANG II, ET-1, NE, or U-46619
was blunted by inhibition of NADPH oxidase, genetic
deletion of p47phox or NOX-2, or by metabolism of ROS
by SOD or tempol (182, 183, 758, 759, 864, 1037,
1301). O2⫺ generated by NADPH oxidase in afferent arteriolar VSMC activated ADP-ribosyl cyclase to increase
Ca2⫹ mobilization by RyRs (399, 400, 1456), whereas
NO suppressed ADP-ribosyl cyclase activity to reduce
[Ca2⫹]i (403, 898, 1642).
450
The enhancement of myogenic responses by ROS may
involve activation of GPCR coupled to second messengers, cytosolic proteins, and [Ca2⫹]i (FIGURE 8) (1328)
linked to redox signaling by vascular NADPH oxidases
(1086). Moreover, stretch activation of AT1 receptors, in
the absence of ANG II, enhanced the myogenic tone of
cerebral and renal cortical radial arteries via ROS generation (999).
Several studies have demonstrated potentiation of renal vasoconstriction by ROS generation. Thus contraction of rabbit isolated afferent arterioles to U-46619 was enhanced by
vascular ROS and blunted by NO (1301, 1303). The sensitivity of afferent arterioles to ANG II was enhanced markedly in mice with SOD1 (Cu-Zn-SOD) deficiency (183).
This effect was primarily attributed to enhanced O2⫺ inactivation of NO. Conversely, prevention of NADPH oxidase-derived O2.⫺ generation in afferent arterioles from
p47phox knockout mice blunted ANG II and myogenic contractions (183, 864, 1097), whereas myogenic contractions
were preserved in afferent arterioles from eNOS-deficient
mice (864). Thus O2⫺ generated by ANG II enhances afferent arteriolar contractions by inactivation of NO, whereas
O2.⫺ generated by increased PP enhances myogenic contractions independently of NO. PP increased the generation
of O2⫺ from NADPH oxidase in large conduit vessels (1621)
and in afferent arterioles from rats (1224) or mice (702,
861, 866, 1224). Cell-permeable polyethylene glycol
(PEG)-SOD and tempol both blunted the myogenic responses of mouse isolated afferent arterioles, whereas PEGcatalase was ineffective, thereby implicating O2⫺ rather than
H2O2 (866). Tempol also attenuated myogenic responses in
the rat isolated perfused HNK (1146). Moreover, prolonged infusion of ANG II increased ROS production that
enhanced myogenic response of isolated afferent arteriole
(864).
Taken together, many studies have implicated O2⫺ generated by NADPH oxidase/p22phox/p47phox NOX-2 during
stretch of afferent arterioles in potentiating the myogenic
response. However, there are contrary findings. For example, the myogenic response of isolated afferent arterioles of
normotensive WKY rats did not require ROS generation,
although O2⫺ enhanced the myogenic response of SHR afferent arterioles (1224). Transforming growth factor-␤1
(TGF-␤1) perfused into the JMN preparation antagonized
myogenic constriction by generation of ROS from NADPH
oxidase (1354). However, this growth factor caused prominent basal constriction of the afferent arteriole, which may
have obscured an additional myogenic contractile response.
Moreover, TGF-␤1 activity and its ability to increase BP
and myogenic tone of mesenteric arteries were normally
suppressed by Emilin-1, a protein of elastic extracellular
matrix (196). Intrarenal infusion of hypoxanthine plus xanthine oxidase increased O2⫺ and caused vasoconstriction as
well as natriuresis, but steady-state RBF autoregulation was
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
well maintained (1204). A high-salt diet that activated NADPH oxidase and increased ROS production in kidneys
(471, 579) failed to affect the myogenic transfer function in
rats and, when given with ANG II, actually blunted the
myogenic response (1273). Another study of rats fed 8%
NaCl for 2 wk reported that RBF autoregulation was abolished (AI increased from 0.0 to 1.4) and the autoregulation
of afferent arterioles of JMN was blunted by an apparent
increase in ROS which was restored in part by prolonged
treatment with apocynin (396). The reason for these discrepancies is presently unclear but may relate to different
effects of O2⫺ and H2O2.
Increased ROS formation enhanced myogenic responses in
other vascular beds including mesenteric resistance arteries
(1530) and tail arterioles (1097). Elevated PP in gracilis
muscle arteries activated sphingosine kinase-1 to stimulate
production of ROS from NADPH oxidase that enhanced
Ca2⫹ sensitivity and myogenic responses likely by activation of Rac and RhoA/Rho kinase and inhibition of MLCP
(783). Activation of NADPH oxidases by increased PP in
bovine coronary and rat small cerebral arteries enhanced
myogenic contractions and/or autoregulation of blood flow
by activation of the ERK/MAPK cascade (1104), or by activating PI3K/Akt signaling that reduced BKCa activity, and
increased EM and VSMC [Ca2⫹]i (468, 470, 1454). However, low levels of O2⫺ generated by subdural perfusion of
xanthine plus xanthine oxidase decreased cerebral blood
flow, but did not affect autoregulation (1646), whereas high
levels of ROS increased blood flow and inhibited blood flow
autoregulation and myogenic constriction by activating
BKCa channels that hyperpolarized the VMSC plasma membrane. This suggests an effect of H2O2, which is known to
hyperpolarize VSMCs (1646). ROS-mediated vasodilation
counteracted myogenic tone in rat ophthalmic arteries, but
this effect was partially dependent on intact endothelium,
which is a source of H2O2 as well as NO (1543). Indeed,
ROS, probably H2O2, impaired endothelium-dependent
vasodilation by inhibiting BKCa channels in renal artery
endothelial cells (135). H2O2 can increase endothelial NO
production (721) and decrease Ca2⫹ sensitivity in aortic
VSMCs (663).
In summary, ROS impacts renal autoregulation by modulating both myogenic and MD-TGF mechanisms. Most
studies of isolated afferent arterioles concluded that the
NADPH oxidase/p47phox/NOX-2 signaling pathway is activated by increased RPP and is the major source of O2⫺ that
enhances the myogenic constrictor response, whereas H2O2
may inhibit this response. The opposing effects of these two
ROS may account for some inconsistences in the reported
effects of ROS on autoregulation, but this requires further
study. The mechanisms whereby O2⫺ sensitizes contractions
involve reduced NO bioavailability, but the effects on the
myogenic response likely involve direct VSMC effects to
mobilize [Ca2⫹] or increase Ca2⫹ sensitivity.
G. Endothelium-Derived Hyperpolarizing
Factor and Cxs
Endothelial cells control vascular tone by changes in EM
transmitted to VSMC and by the release of factors that
relax or constrict VSMCs. Endothelial dilator factors include NO, PGs, and EDHF. ACh and bradykinin can elicit
vasodilation and buffer vasoconstriction in part via EDHF
that hyperpolarizes VSMCs and relaxes intrarenal arterioles independent of the vasodilators NO, PGE2, and PGI2.
EDHF stimulated Na⫹-K⫹-ATPase and enhanced hyperpolarization of the renal interlobar artery (172). EDHF in mice
was estimated to contribute ⬍20% of the renal vasodilation
produced by ACh or bradykinin (299).
Little is known about the influence of EDHF on RBF autoregulation. Although the myogenic response is not mediated by endothelial factors, it is likely that endothelial cells
can modulate VSMC tone by electrical coupling and production of paracrine agents such as NO as well as EDHF
and EETs. ACh-induced EDHF production transiently
blunted myogenic vasoconstriction of afferent arterioles in
the isolated perfused rat HNK (592) where it stimulated K⫹
channels to modulate afferent but not efferent arteriolar
tone, presumably by affecting EM and VOCC activity
(1571). Interestingly, the inhibition was transient, reversing
after 5–10 min.
EETs mediated EDHF actions in the rabbit isolated, perfused afferent arteriole. They were generated by a cytochrome P-450 epoxygenase and elicited vasodilation by
opening BKCa channels and activating Na⫹-K⫹-ATPase on
VSMCs (1551). A metabolite of cytochrome P-450 epoxygenase, presumably an EET, blunted myogenic constriction
of afferent arterioles in the JMN preparation (673). Vasodilatory EETs attenuated the myogenic response of isolated
gracilis muscle arterioles (1424). Metabolism of EETs by
soluble epoxide hydrolase regulated myogenic response and
BP via effect on Ca2⫹ entry via TRPC1 and TRPC4 channels (1209). Other candidates for an EDHF include H2O2
and hydrogen sulfide (H2S) (914, 1449, 1652).
Gap junctions composed of Cxs may contribute to endothelium-dependent vasorelaxation via conducted vascular responses. The extent to which cell-to-cell communication
depends on direct signaling through myoendothelial gap
junctions is controversial. Isolated arteries and arterioles
displayed bidirectional electrical coupling between endothelial cells and VSMCs. Vasodilator waves passing via cellto-cell gap junctions can spread to upstream feed arteries
that control blood flow into arteriolar networks (60). Cx
may mediate this spreading vasodilation elicited by EDHF
by transfer of hyperpolarization from endothelium via gap
junction Cx to adjacent VSMCs to conduct along the vessels and thereby close their VOCCs. However, such coupling between endothelium and VSMC cells has not been
verified in intact animals where myoendothelial gap junc-
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
451
RENAL AUTOREGULATORY MECHANISMS
tions may be controlled by yet unknown mechanisms that
impede such signaling (313, 314). In nonrenal vascular beds
including aorta, pulmonary, and coronary arteries, Cx40
and Cx37 interacted to form heterotypic channels, and they
were codependent on each other, or mutually dependent,
for optimal expression and function in vascular endothelium (1368, 1369, 1599, 1633).
ACh acting on cremaster arteries in vitro elicits local vasodilation that was conducted upstream via Cx40, but not on
Cx37 (738). However, vasodilation produced by ACh in
mouse cremaster arterioles in vivo was largely independent
of this mechanism (125, 314, 315, 1362). Moreover, ACh
produced potent vasodilation in Cx40-deficient gracilis arteries in vitro or in endothelium-specific Cx40 deficient vessels, but this was detected only under isobaric conditions
(125). ACh-induced EDHF-like dilation observed in vivo
was independent of the presence of Cx40 (756). Thus ACh
responses appear to be conducted along VSMCs, but not
between endothelial cells.
Cxs at myoendothelial gap junctions of small arteries in
vitro modulated the effects of NO on Ca2⫹ signal propagation which prolonged increases in [Ca2⫹]i in endothelial
cells and accelerated Ca2⫹ signal transmission in VSMCs
(1192), but this was independent of Cx37 (419).
In summary, the roles of Cxs in the EDHF response and in
myogenic tone are unclear. The role of gap junctions and
Cx in the conduction of vasoactive signals along the vascular tree of the microcirculation requires further investigation and resolution concerning differences in results obtained in vitro and in vivo.
H. Modulation of Myogenic Responses by
Agents Released From Activated MD or
CT Cells
Both MD, and some CT, cells abut the afferent arteriole
supplying the parent glomerulus. As discussed earlier in
section IIA3, MD cell signaling is activated by reabsorption
via luminal NKCC2 and Na⫹/H⫹ exchange. As summarized in FIGURE 6, MD-derived metabolites can either enhance (ATP, adenosine, PGH2, TxA2, and O2⫺) or inhibit
(NO, CO, and PGE2) afferent arteriolar tone. As discussed
in section IIB2, CT cell signaling can be activated by reabsorption via luminal ENaC which release EETs and PGE2
that attenuate afferent arteriolar tone (FIGURE 6). The resultant vasodilation from CT-GF activation may contribute
to resetting of MD-TGF during salt loading (1554). Such
rich paracrine signaling in the JGA provides for complex
interactions between the tubular-glomerular systems and
myogenic mechanism of vascular control and contributes to
the positive and negative interactions that were described in
earlier sections.
452
VII. RENAL AUTOREGULATION IN
DISEASE: HYPERTENSION, RENAL
DISEASE, AND DIABETES
A. Introduction
Studies in the 1930s demonstrated that hypertension is associated with arteriolosclerosis of the small renal arterioles
and renal tubulointerstitial inflammation, fibrosis, and glomerulosclerosis (795, 1033). There were two contrasting
theories concerning the principal causes and the secondary
compensations (388, 1033, 1541). One held that hypertension is primary and induces secondary renal and vascular
abnormalities. In 1937, Moritz and Old (1033) reported
that renal afferent arteriopathy was present in 98% of autopsied kidneys from subjects who had hypertension but in
only 15% of these who were normotensive. They concluded
that the renal arterial lesions are a consequence of the hypertension. The earliest structural changes with hypertension are in the smallest resistance arterioles, a site where a
myogenic response plays a fundamental role. Hypertension
is associated with structural remodeling of the vascular wall
and vascular hyalinization, but the contribution of wall
remodeling to functional changes in myogenic responses is
uncertain. The second theory held that renal inflammation
and afferent arteriopathy are primary and induce hypertension (1033). In 1934, Goldblatt (476) showed convincingly
that creating a renal artery stenosis to reduce the RPP leads
to hypertension. Renal artery stenosis, whether caused by
atherosclerosis or fibromuscular disease, is detected in
5–10% of patients with hypertension (1455). However, recent controlled clinical trials of renal artery angioplasty and
stenting to correct a stenosis have concluded that hypertension, although often lessened in severity, is only rarely cured
(72). Thus the debate remains incompletely resolved. Structural renal arterial lesions that limit RPP clearly raise BP.
However, the renal arteriolar remodeling that accompanies
sustained hypertension likely impairs regulation of RBF and
RPP and increases RVR and may thereby contribute to
sustaining the hypertension. Thus the structural renal arterial changes can be viewed as a consequence of hypertension, but in some settings, likely also contribute to the development or severity of the hypertension. Therefore, they
are of considerable importance.
Glomerular hyperfiltration refers to an increase in the
SNGFR, but this is not necessarily accompanied by glomerular hypertension, i.e., an increase in the PGC (601). Schmieder and co-workers (1297, 1298) suggested that glomerular
hyperfiltration occurs early in hypertensive patients with
target-organ damage, and Campese et al. (177) concluded
that PGC is likely increased in a group of salt-sensitive hypertensive patients that included many African-Americans.
An impaired renal autoregulatory response in hypertensive
African-Americans (FIGURE 2A) may contribute to an increased PGC (826).
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
Many rodent hypertensive models eventually develop CKD
(104). However, the development of CKD, unless associated with malignant hypertension, is uncommon in humans
except in African-Americans. Recent studies report that the
primary problem in African-Americans with CKD is inheritance of an abnormal genotype that confers increased risk
of renal damage. Hypertension is usually a secondary consequence (909). Nevertheless, it is clear that once renal
damage has been sustained, hypertension is one factor predicting the rate of further decline in renal function, especially in those with heavy proteinuria (803).
Autoregulation normally protects the glomerular and tubulointerstitial tissues from hypertensive injury (108, 112,
933). However, although autoregulation usually adapts to
higher BPs in hypertensive models, its efficiency often declines over time. The responsible mechanisms are poorly
understood. Some, but not all, models of hypertension,
CKD, or diabetes mellitus (DM) have impaired renal autoregulation that permits transmission of more of the BP fluctuations to the glomeruli where it increases the PGC. The
degree of glomerular injury in hypertensive models generally increases in proportion to the impairment in renal autoregulation (106, 107, 110). Bidani, Kriz, Byrom, and colleagues (103–110, 173, 839, 840, 1162) proposed that hypertension combined with impaired renal autoregulation
cause renal barotrauma and enhance the progression of
CKD. Kriz et al. (838) demonstrated that an initial podocyte injury from barotrauma in the context of FSGS misdirects filtration into the interstitium, thereby initiating a
spreading renal tubulointerstitial inflammation and later
fibrosis with degeneration or fibrosis of the glomerulus and
its attached tubule.
It follows that the ideal therapy for hypertension should
lower the BP while preserving, or improving, the effectiveness of the renal autoregulation (105, 106, 110, 500). Ideally, the preglomerular vasculature should protect the kidney from both sustained hypertension and from rapid oscillatory fluctuations in BP (108, 109, 927, 934) since both
increased BP and increased pulse pressure predict hypertensive renal damage (108, 927, 933).
Examples of impaired autoregulation in disease models include rodent remnant kidney models of CKD with prolonged RRM, some, but not all, rodent insulinopenic or
obesity-associated models of DM, or Dahl salt-sensitive or
Goldblatt ischemic renovascular hypertensive models. In
contrast, SHR retain highly efficient autoregulation
throughout life that likely contributes to their resistance to
proteinuria and glomerulosclerosis even with established
and prolonged hypertension. Renal cross-transplantation
between BN rats and SHR proved that the protection from
glomerular damage in the SHR was intrinsic to their kidneys (260), which corresponds to the excellent autoregulatory adjustments of the preglomerular vasculature of the
SHR but the much weaker adjustments of BN rats. A similar kidney cross-transplantation study demonstrated that
the kidneys from stroke-prone SHR (SHR-SP) exhibit more
hypertensive damage and malignant nephrosclerosis than
those of SHR during a high-salt diet (261, 498). Moreover,
poor RBF autoregulation in Fawn-Hooded (FH) rats is improved by replacing chromosome 1 genetic loci from BN
rats (922, 1610). Roman, Provoost, and colleagues have
provided extensive information concerning the mechanisms, and their genetic components, that underlie autoregulation in many hypertensive models, as described in detail
under the individual models below. These studies are important because they have established that the efficiency of
renal autoregulation and the associated protection of the
kidneys from hypertensive damage are heritable.
B. Hypertension
1. SHR
SHR are considered a model of human essential hypertension (1115, 1483). The strain was developed from inbreeding of selected rats that developed hypertension spontaneously. Hypertension develops progressively over 4 – 8 wk
while fed a normal-salt diet (161, 386, 744, 867, 1009,
1389, 1467). The hypertension is exacerbated by a high salt
intake (34, 256, 325, 326, 447, 498, 880, 917, 978, 1406).
The degree of elevation in BP was quite variable among
studies. Factors likely contributing to this variability are the
methodology for BP measurement, the animals initial age
(6 –14 wk), duration of treatment (2–28 wk), the amount of
salt in the diet (ranging from 1 to 8%), as well as the genetic
background of the SHR and control WKY colonies (848,
849, 1428). Cross-transplantation of kidneys between SHR
and normotensive controls established that the kidney of
SHR is the primary cause of the development of hypertension (506, 575, 779, 1158, 1235, 1236).
Renal hemodynamic abnormalities are important in the initiation of hypertension in the SHR. Basal GFR and RBF are
reduced during the development of hypertension in young
SHR (4 –10 wk old) (335, 386, 567, 574, 1038, 1252, 1507,
1526), but are restored to the levels of normotensive WKY
rats during the established phase at 12–20 wk (42, 418,
443, 574, 809, 1246, 1252). A rightward shift in the pressure-natriuresis curve to a higher level of RPP (43, 443, 997,
1093, 1246) led to an early period of relative Na⫹ retention
during the development of hypertension (79, 570, 571, 605,
940). A genetic link between abnormal renal hemodynamics and the development of hypertension was established in
cross-breeding studies between SHR and WKY. Renal vasoconstriction (reduced RBF and GFR) and elevated BP
cosegregated (572, 710) and was evident during the early
development phase of hypertension, suggesting causality.
This link has been investigated extensively. Cosegregation
of the renin allele of SHR with increased BP has been noted
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
453
RENAL AUTOREGULATORY MECHANISMS
(850, 1427, 1641), and expression of the SA region on
chromosome 1 is enhanced in hypertensive SHR (693, 711,
712, 906, 1405), in particular in the proximal tubule
(1631). However, studies using congenic SHR carrying a
smaller segment of BN chromosome 1 but without the SA
gene concluded that molecular variation in this gene was
not required for the effect of this region of chromosome 1
on blood pressure (1405).
The RAS is especially important in the developmental phase
of hypertension. ACE inhibition in young SHR for 4 wk
between 2 and 10 wk of age normalized the exaggerated
renal vasoconstriction and has long lasting effects in preventing full expression of hypertension later in life (250,
354, 417, 472, 567, 568, 573, 574, 939), beneficial effects
independent of changes in renal vasculature structure (786).
Administration of an AT1 receptor blocker from 3 to 8 wk
of age led to a sustained reduction in BP in adult SHR (95,
1103). Although less well studied, blockade of the RAS may
be efficacious in reversing established hypertension (355,
393, 939).
Renal nerves have an important role in the pathogenesis of
hypertension in the SHR. Renal nerve activity was elevated
in SHR at the age of 8 –10 wk (746, 941). Denervation of
the kidneys in young (4 – 8 wk old) SHR delayed the development of hypertension (606, 807, 1092, 1617) (775), due
at least in part to increased Na⫹ excretion (606, 1268).
Young (4 – 6 wk old) SHR had excessive preglomerular vasoconstriction that reduced RBF and GFR (43, 276, 1252,
1507). The lower GFR was ascribed to reductions in the
glomerular blood flow and the glomerular hydrostatic filtration coefficient (KUF) while the PGC was maintained by
increased afferent and efferent arteriolar resistances (335).
Hypertensive SHR between 12 and 16 wk of age have increased preglomerular vascular resistance that parallels the
increase in BP (42, 58, 358, 639, 704, 709, 1471), mediates
efficient autoregulation of RBF and GFR (8, 41, 107, 276,
640, 709, 1246, 1507), and maintains PGC in the normal
range (42, 58, 358, 704, 852, 1471) and normal glomerular
structure (787). The total number of glomeruli and mean
glomerular volume were similar in kidneys of SHR and
WKY at 4 wk of age (115). Albuminuria developed only
after ⬃20 wk of age.
Although the PGC was normal in 12- to 16-wk-old SHR
(1471), it tended to increase in some (1471), but not all
(394), studies by 36 mo, while the vascular and glomerular
morphology remained normal (1471). Thus glomerular hypertension can precede glomerular injury in SHR. Remarkably, in another study, the PGC of 15-mo-old SHR was
lower than age-matched WKY (394).
The administration of low doses of a CCB or an ACE inhibitor to SHR reduced their BP, increased their RBF and
454
GFR (412– 414, 567, 574, 765, 1056, 1098) and SNGFR
(435), and normalized their preglomerular and efferent arteriolar resistances (1056, 1098). Higher doses of a CCB
that blocked both L- and T-type VOCC abolished renal
autoregulation and rendered PGC highly dependent on BP
(365, 500, 852, 1056). These data suggest that the use of
CCBs to treat hypertension in patients with CKD is less
protective of renal function than ACE inhibitors, as confirmed in the African-American Study of Kidney Disease
(AASK) trial (909) and in experimental studies in rats (29,
365, 500, 503). Moreover, administration of an ACE inhibitor or an AT1 receptor blocker to young SHR caused a
predominant reduction in efferent arteriolar resistance,
thereby reducing PGC and the lower limit of renal autoregulation (852).
Young SHR have exaggerated renal vascular reactivity to
ANG II (44, 220, 222, 311, 787, 824, 825, 1409, 1542),
ATP (1434), catecholamines (787, 813, 1385, 1498), and
AVP (404) and attenuated vasodilation to PGE2, PGI2 (219,
223, 224, 716, 717), and agonists of dopamine D1 receptors
(225). Thus vasodilation and reduced buffering of vasoconstriction by GPCRs signaling and activation of the cAMP/
PKA pathway are defective in the renal vasculature of SHR
and contribute to the exaggerated vasoconstriction. The
exaggerated renal vascular reactivity to ANG II was largely
due to abnormal blunting by vasodilator PGs (217, 219,
221, 224, 716, 717), since the density of AT1 receptors in
the renal vasculature was similar in SHR and WKY (216,
220). However, vasodilation to ACh, sodium nitroprusside,
or bradykinin that is mediated by the cGMP/PKG pathway
was preserved or even exaggerated in young SHR (174,
225). Indeed, inhibition of NOS led to stronger increases in
RVR and reductions in RBF in adult SHR (10 –16 wk old)
than in WKY in some studies (96, 432, 853, 1203), implying enhanced regulation of the renal circulation by NO in
the SHR with established hypertension, whereas others reported a normal degree of vasoconstriction in adult and
aged (70 wk old) SHR (311, 432, 853).
Renal autoregulation operates normally in young and mature SHR. Unlike most other hypertensive models, the RBF
and GFR of SHR are very efficiently autoregulated with
maintained, or even an exaggerated, with strong influence
of myogenic and MD-TGF components throughout the established phase of hypertension (707, 1567). Steady-state
experiments demonstrated that SHR autoregulate their
RBF effectively with a normal plateau phase over a defined
range of RPP as they age from 4 – 6 wk (1252) to 10 –16 wk
(8, 42, 598, 640, 853, 1203, 1246, 1252, 1567) or even up
to 40 (708, 709, 852, 1567) to 70 wk (853). The plateau
phase of RBF autoregulation was unaffected by inhibition
of NO (853, 1203), the RAS (352, 1565), ET-1 and ETA
receptors (140), COX metabolites, ROS (815), or renal
nerve activity (708). Whereas autoregulation was not affected by endogenous ET-1 in normotensive animals (97,
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
140, 831), it was improved in SHR after short-term ETAreceptor blockade (140). However, in animals with established hypertension, these systems contribute to switching
the autoregulatory range to higher levels of RPP (705–707,
709, 852, 997, 1188, 1203, 1565–1567). The lower limit of
RPP for autoregulation was preserved in young hypertensive SHR (41, 1252).
Whereas SHR exhibited excellent autoregulation of renal
and outer cortical blood flow (704), JMN have increased
PGC with less efficient and more sluggish autoregulation.
Thus the PGC of JMN of SHR was 5 mmHg higher than
cortical nephrons at 10 wk of age and increased to 10
mmHg higher at 70 wk. The higher PGC corresponded with
an earlier development of glomerulosclerosis and with
larger diameters of glomerular capillaries in JMN (704).
Moreover, an abrupt step increase in RPP produced a rapid
adaptive increase in vascular resistance in the superficial
renal cortex but a slower, less complete response in the
juxtamedullary region (1240).
Autoregulation of medullary blood flow was well established in young SHR (1252). By 12–16 wk, the medullary
blood flow and the RIHP were reduced in SHR, accompanied by a rightward shift in the pressure-natriuresis relation
(412). At that age, autoregulation of medullary blood flow
was diminished during acute volume expansion, although
excellent autoregulation of whole kidney RBF persists
(352). Since medullary blood flow contributes ⬍10% to
total RBF, small changes may not be detected in measurements of total RBF. These findings are consistent with the
hypothesis of Cowley et al. (274) that an attenuation of
medullary blood flow autoregulation during volume expansion enhances the strength of the pressure-natriuresis relation and thereby modulates the level of BP, as was discussed
in section IIA2 and is illustrated in FIGURE 4B.
Medullary blood flow was especially sensitive to ANG II in
SHR during the development phase of hypertension (353).
ACE inhibition from 4 to 14 wk of age in volume-expanded
SHR attenuated the autoregulation of the medullary blood
flow despite a maintained autoregulation of RBF (352).
Administration of an AT1 receptor blocker to SHR between
4 and 12 wk of age reduced BP, increased medullary blood
flow, and reduced the number of pericytes surrounding vasa
recta capillaries (70). These treatments also impaired medullary blood flow autoregulation and increased the RIHP
which may contribute to the resetting of the pressure-natriuresis relationship to a lower level of RPP, and limiting the
development of hypertension (997, 1567). Indeed, most
studies employing infusions of CCBs, ACE inhibitors, tempol, lovastatin, or L-arginine concluded that SHR have a
reduced NO and a reduced medullary perfusion secondary
to increased medullary oxidative stress and that could account for a rightward resetting of the pressure-natriuresis
relationship and a higher ambient BP (276, 406, 412, 733,
872). These studies indicate NO is important in maintaining the medullary blood flow in SHR (96), and a reduction
in NO contributes to the lower medullary blood flow and
reduced pressure-natriuresis relation in SHR (872). Other
studies, however, suggested that NOS activity was elevated
in the renal medulla of SHR (585) and its inhibition produced a greater reduction in medullary blood flow in SHR
than in WKY (96). Collectively, these data suggest an important role for ANG II and the balance between ROS and
NO in the renal medulla during the development of hypertension in SHR. Indeed, several studies have reported enhanced interaction between ANG II, NO, and ROS in renal
autoregulation of SHR.
A strong MD-TGF response may contribute to the resilient
renal autoregulation in SHR. Young (6 wk old) SHR had
exaggerated MD-TGF response (333) that was attributed to
ANG II acting on AT1 receptors and to a vasoconstrictor
eicosanoid acting on TP receptors (138, 139). MD-TGF
activity reverted towards normal during established hypertension in adult SHR from the NIH/Chapel Hill colony
(333) and from some commercial suppliers (231), but remained exaggerated in adult (12–18 wk of age) SHR from
some other sources (894, 1587, 1592). However, there is
genetic heterogeneity between colonies of SHR, and especially WKY (848). Increased MD-TGF response in adult
SHR was attributed to AT1 receptor activation and oxidative stress (894, 1587, 1588, 1592) and weak buffering by
NO. Although SHR had increased renal cortical expression
of endothelial NOS and of MD nNOS, neither global inhibition of NOS nor relatively selective inhibition of nNOS
with 7-nitroindazole increased the MD-TGF in SHR as it
did in WKY (1464, 1587, 1588). On the other hand, peritubular capillary perfusion of tempol blunted MD-TGF activity more in SHR than WKY and restored the effect of
nNOS inhibition to enhance MD-TGF in SHR (1588). This
suggests that the JGA of hypertensive SHR produces excess
O2⫺ that bioinactivates MD-derived NO. O2⫺ in the JGA was
dependent on tubular fluid reabsorption activating NADPH oxidase NOX-2 in MD cells (448, 1592). Excess O2⫺
diminished dilation of juxtamedullary afferent arterioles by
neuronal NOS-derived NO in SHR but not in WKY (655).
Thus SHR generally have enhanced MD-TGF responses
that are related to oxidative stress in the JGA that inactivate
NO, thereby contributing to their efficient renal autoregulation.
Vasoconstriction from activation of MD-TGF can be propagated upstream along the afferent arteriole to reduce the
perfusion and the PGC in adjacent nephrons supplied by a
common cortical radial artery. The magnitude of such a
MD-TGF-initiated nephron-nephron interaction was stronger in SHR, implicating greater signal transduction (231).
Chon, Holstein-Rathlou, Marsh, and associates demonstrated that MD-TGF signaling to the parent glomerulus
caused oscillations in preglomerular vascular tone and PPT
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
455
RENAL AUTOREGULATORY MECHANISMS
that were more time-variant and irregular in SHR, indicative of more complex coupling between MD-TGF and myogenic mechanisms (230, 244, 247, 625, 626, 1205, 1635,
1672). Thus studies of transfer function analysis generally
confirmed an enhanced MD-TGF, and perhaps myogenic
component, of autoregulation in SHR, which was consistent with increased ROS and reduced NO signaling in the
JGA.
Myogenic constriction of isolated perfused afferent arterioles was enhanced in both young and adult SHR (700,
1224). Thus exaggerated myogenic tone seems to be an
intrinsic functional abnormality evident during the development of hypertension in young SHR and not a consequence of long-standing hypertension in adult SHR.
The PP threshold for a myogenic contraction in the afferent
arteriole of the HNK was shifted 40 mmHg higher, while
the maximum myogenic constriction was preserved (593).
NOS inhibition lowered the threshold PP by 20 mmHg in
WKY and by 40 mmHg in SHR (15 wk old), whereas the
maximum myogenic response was augmented to a similar
degree in both strains (593). Thus NO can modulate the
myogenic response with different effects in the renal vasculature of SHR and WKY, perhaps due to different actions
on pre- and postglomerular arterioles. Perfusion of cortical
radial arteries and afferent arterioles of JMN with a physiological salt solution revealed enhanced myogenic contractions in SHR as young as 3– 4 wk of age (465, 672), which
were related to enhanced generation of a cytochrome P-450
metabolite (469, 672), presumably the vasoconstrictor 20HETE.
Oxidative stress is frequently associated with increased vascular reactivity, hypertension, and renal failure (1330,
1529). Prolonged tempol administration reduced BP in
SHR and shifted the pressure-natriuresis relationship to the
left (406, 1300, 1302, 1585). Tempol increased basal RBF
in hypertensive SHR, reduced exaggerated renal vascular
reactivity to ANG II, independent of NO, suggesting a direct effect of O2.⫺ on vascular responsiveness (311). Tempol quenching of O2.⫺ increased medullary blood flow
without perturbing RBF autoregulation in adult SHR (311,
406). Thus ROS apparently restrains renal and medullary
blood flow in the SHR without major effects on autoregulation.
The afferent arteriole of adult SHR has a fourfold increase
in ROS production by NOX-2 (1224). Thus, bioinactivation of NO by O2.⫺ in the JGA of SHR is sufficient to
prevent endogenous NO from blunting MD-TGF. It also
may explain the lack of effect of NOS inhibition to magnify
the renal myogenic response of SHR as it did in normotensive Long-Evans, Sprague-Dawley and Wistar rats (1569).
The strong myogenic response of young SHR was sufficient
456
to maintain RBF autoregulation after complete inhibition
of MD-TGF by ureteral obstruction (297).
Nevertheless, whereas NO is considered a strong modulator of the myogenic component of renal autoregulation in
normotensive rat strains, the modulation was weak or absent in the SHR (1569). This indicates a separation of the
role of NO to blunt vasoconstriction (evident in SHR and
normotensive strains) and to blunt the myogenic response
(absent in SHR).
Nevertheless, efficient RBF autoregulation and the myogenic response in SHR was preserved in the absence of
MD-TGF input during ureteral obstruction (297) and persisted after damage to mesangial cells by antithymocyte
antibodies in both SHR and WKY, whereas autoregulation
of GFR was attenuated (1564). Thus, although mesangial
cells can participate in the regulation of GFR, likely via
MD-TGF signaling, they do not seem to contribute to RBF
autoregulation, which, in their absence, is maintained by
enhanced myogenic tone.
SHR also have enhanced myogenic tone in nonrenal vessels
including cremaster muscle (389, 643, 644, 1359), basilar
(18), cerebral (127, 484, 564, 566, 725, 1469), mesenteric
(472, 714), and coronary arteries (458). Enhanced myogenic tone in these nonrenal vessels of SHR has been related
to increased Ca2⫹ entry due to depolarization of the plasma
membrane (564) and elevated VOCC activity (1359), perhaps secondary to reduced EM and reduced K⫹ channel
activity (1469). Also contributing were increased Ca2⫹ mobilization and Ca2⫹ sensitivity due to augmented Rho kinase activity (18, 725, 1164). Enhanced myogenic tone was
dependent on increased endothelial-derived ET-1 and
PGH2 (643, 644) and reduced NO (484), or other endothelial factors (458). The enhanced myogenic responses occurred without vascular remodeling (389) and were not
secondary to hypertension since they predated its onset
(127, 714, 1359). Mechanisms that enhanced myogenic
tone in these nonrenal vessels of SHR may also occur in the
renal microcirculation. However, these effects were specific
for resistance vessels since normal myogenic responses are
reported in larger conduit vessels from SHR (166, 167,
1084).
Remodeling (reorganization of the vessel wall components)
of small arteries can limit organ blood flow (425, 1044,
1322) and may contribute to exaggerated renal vasoconstriction. An increased cross-sectional area of the preglomerular arterial media was apparent in young SHR before
the development of hypertension (1385). Furthermore,
some of the second generation offspring of rats crossbred
between SHR and normotensive WKY had a narrowed afferent arteriole lumen diameter at 7 wk which predicted the
later development of high BP at 21–23 wk (1094, 1380).
This suggests that a narrowed renal afferent arteriole, as
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
from remodeling, can contribute to the development of sustained hypertension. Most, but not all (1381), studies reported hypertrophic remodeling of the media with luminal
narrowing of cortical radial arteries of SHR that may persist
(33) or reverse with prolonged ACE inhibition (1379) and
may originate from enhanced sympathetic nerve activity
(1474). Thus a variable degree of genetically and neurally
determined remodeling of the afferent arteriole of the SHR
could restrict the transmission of BP into the kidneys,
thereby protecting the glomeruli from barotrauma and
maintaining the set-point for pressure-natriuresis at ambient level of RPP.
Other studies have identified increased wall thickness in
the arcuate and cortical radial arteries of 5- or 10-wk-old
SHR but not in the afferent arteriole at either age (33,
1385). The ratio of wall thickness to radius of the afferent arterioles in the inner and outer cortex was not different between SHR and WKY at 6 and 12 wk of age
(461). Another morphological study reported that afferent arterioles of 12-wk-old SHR had a smaller, not
larger, media cross-sectional area than age-matched
WKY (1381). Aged SHR have hypertrophy of the afferent arteriolar wall (882).
A hallmark of the SHR is excellent autoregulation of whole
kidney and outer cortical blood flow and PGC, which may
account for the paucity of glomerulosclerosis until up to 70
wk of age, despite increasing hypertension. Excessive NADPH oxidase-mediated O2⫺ production and downregulation of SOD and other antioxidant enzymes in the kidney of
SHR (12, 114, 212, 1649) likely contributes to the enhancement of myogenic and MD-TGF mechanisms. Much, but
not all, of the effects of O2⫺ in the JGA were secondary to
reduced NO activity (1588). However, as SHR age, their
cortical autoregulation eventually becomes slightly less
efficient and their outer cortical glomeruli develop modest glomerulosclerosis. The earlier impairment of autoregulation in JMN may account for their earlier development of glomerulosclerosis. Nonrenal resistance arterioles of SHR also generally display enhanced myogenic
tone, suggesting that it is a fundamental property of
VSMCs in this rat strain. Since augmented myogenic responses of renal and systemic microvessels of young SHR
precede the development of hypertension, they are likely
primary and genetically determined. A primary increase
in O2⫺ with bioinactivation of NO in the renal cortex and
medulla of the SHR could contribute to the excellent
autoregulation and reduction of transmission of BP into
the renal parenchyma. This should buffer the effects of
increased BP to raise medullary blood flow and RIHP in
the SHR, thereby attenuating the pressure natriuresis and
sustaining the hypertension on the one hand, yet protecting the renal parenchyma from the effects of hypertensive
barotrauma on the other.
2. Stroke-prone SHR
Stroke-prone SHR (SHR-SP) have a genetic predisposition
to stroke, a more rapid rise in BP, and a more pronounced
salt-sensitivity than SHR (1048, 1116). They were normotensive at 4 wk of age and developed hypertension before 9
wk (908, 1049). Kidneys of SHR-SP required a higher BP
than WKY to excrete a given amount of salt and water
(1049). Marked proteinuria was evident by 12–16 wk of
age (1055, 1478). Glomerulosclerosis was present by 24 wk
of age (1055). At 15 mo, the kidneys of SHR-SP on a normal-salt diet displayed reduced GFR, arterial wall hypertrophy and sclerosis, glomerulopathy, and tubular interstitial
injury, all of which were prevented by normalization of the
BP during ACE inhibition since 4 wk of age (908). Remarakbly, this nephroprotection was associated with almost a
doubling of the lifespan.
When given 1– 4% NaCl to drink, they develop severe hypertension and usually die from stroke within 14 –20 wk.
Transplantation of a kidney from an SHR-SP into an SHR
accelerated the rise of BP, the reduction of RBF and GFR
(1049), and the development of renal damage (498). Thus
SHR-SP kidneys have a genetic propensity to cause hypertension and also an enhanced susceptibility to hypertensive
injury (261).
The basal RBF and GFR were reduced and the RVR increased in hypertensive 9- or 12-wk-old SHR-SP relative to
normotensive controls (1049, 1055). The BP and the RBF
increased with a high-salt diet (8). The renal vasculature of
4-, 8-, and 16-wk-old SHR-SP was hyperresponsive to ANG
II, AVP, NE (94, 1050), and ET-1 but not to AVP (457).
Dopamine D1 receptor stimulation produced enhanced vasodilation via PKC/cAMP signaling in intrarenal arteries of
SHR-SP (456). Acute CCB reduced BP and increased RBF
without affecting GFR (1031). Six-month-old SHR-SP
given CCB for 3 mo had reduced BP with attenuation of the
development of glomerulosclerosis, reduced vascular wall
thickness, less glomerular and interstitial fibrosis, and increased glomerular volume (692). Similar beneficial structural effects were observed when BP was reduced by blockade of adrenoceptors. ETA receptor blockade starting at 8
wk of age attenuated progressive decline in renal function
(measured as creatinine clearance), delayed proteinuria,
and doubled the life-span of both male and female SHR-SP
on a 4% NaCl intake (1478). Such functional changes took
place even though antagonism of ETA receptors had little or
no effect on BP, as systolic BP (tail-cuff) rose to 240 mmHg
in both treated and untreated animals. AT1 receptor blockade starting at 12 wk reduced BP, normalized RBF and GFR
in SHR-SP to WKY levels, and reduced the index of glomerulosclerosis (1055). Prolonged TP receptor antagonism
from either 7 or 16 wk of age had no effect on BP, development of proteinuria, cerebrovascular lesions, or stroke in
SHR SP drinking 1% NaCl (741).
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
457
RENAL AUTOREGULATORY MECHANISMS
The steady-state values and the autoregulation of total and
cortical blood flow of SHR-SP were well maintained during
normal salt diet (8, 94, 646), but the autoregulation of
medullary blood flow was weak (646). As expected, a CCB
abolished autoregulation of renal cortical blood flow (646).
Another study reported moderate efficiency of RBF autoregulation with an AI of 0.5 in 9-wk-old SHR-SP (1049).
Transfer function analysis confirmed a well-preserved renal
autoregulation in conscious 12-wk-old SHR-SP consuming
a normal-salt diet, but less efficient autoregulation after an
8% NaCl diet for 3–5 days (8). The sharp increase in BP and
the loss of renal autoregulation may contribute to the severe
salt-dependent renal damage (8). Enhanced renal vasoconstriction and MD-TGF activity in SHR-SP nephrons were
normalized by acute renal denervation (1562).
Cortical radial arteries of the HNK preparation retained a
normal myogenic response in young SHR-SP but an increased PP threshold (586). Hayashi et al. (587) reported
that the myogenic reactivity of the cortical radial artery of
the HNK increased along the length of the vessel and that
SHR-SP exhibited augmented responsiveness of the intermediate but not the distal segment. This was evident by
constriction at a lower threshold PP and greater maximum
reductions in diameter compared with WKY.
Cerebral arteries of 12-wk-old SHR-SP displayed medial
hypertrophy and remodeling, together with attenuated
myogenic response (715). Moreover, loss of cerebral blood
flow autoregulation was observed prior to the event of
stroke (1387). Interestingly, RAS inhibition preserved cerebral autoregulation and the myogenic response and delayed
the onset of stroke (1133, 1134, 1559), largely independently of BP (1384). Thus a high-salt diet impairs autoregulation of the renal and cerebral blood flow of SHR-SP that
likely contributes to the severe hypertensive renal damage
and the propensity for stroke.
3. Dahl salt-sensitive rat
Three-week-old Dahl salt-sensitive (DS) and Dahl salt-resistant (DR) rats, when nephrogenesis is still ongoing, were
normotensive when receiving a low-salt diet (742), but this
has not been studied by telemetry in conscious rats. Renal
injuries developed before the loss of dynamic renal autoregulation (769). Indeed, prehypertensive DS rats of the JR (J
Rapp) strain already had albuminuria, and glomerular
damage accelerated once hypertension was established
(1418). Moreover, the GFR of DS rats, even when fed a
low-salt diet, was reduced by 6 wk (1526). When fed a
high-salt diet, the BP of DS rats increased and both RBF and
GFR decreased (642) leading to glomerular hypertension,
progressive proteinuria, and renal insufficiency (229, 1208,
1418). This was accompanied by intimal thickening and
fibrinoid necrosis (deposits of immune complexes and fibrin) with hyperplasia and necrotizing arteritis (i.e., inflammation of the vascular wall) of the large preglomerular
458
vessels (551, 788), all characteristic for human malignant
hypertension. Medullary interstitial fibrosis and tubular necrosis were also severe (278). Many DS rats died after 4 – 8
wk treatment with a high-salt diet (229, 1208). Thus the DS
rat has severe structural and functional changes in its renal
vasculature that are apparent very early in life even when
fed a low-salt diet to prevent hypertension. A switch to a
high-salt diet leads to malignant hypertension with pronounced renal damage.
Kidney cross-transplantation between DS and DR rats demonstrated that the genotype of the donor kidney determined
the BP of the recipient rat (259, 292). Single chromosome
substitutions from normotensive BN into DS rats demonstrated that the development of hypertension and proteinuria were determined by genes on chromosomes 1, 5, 7, 8,
13, and 18 in males and on chromosomes 1 and 5 in females, although several chromosomes in males and females
affected the development of albuminuria without affecting
BP (988). Transfer of the region of chromosome 5 that
contained the gene for cytochrome P-450 – 4A from Lewis
into male DS rats increased 20-HETE in the outer medulla
and attenuated hypertension and renal injury (1611). These
responses may relate to the natriuretic actions of 20-HETE
(1250). Thus multiple sex-specific genes contribute to hypertension in DS rats. Since some of these differ from those
determining CKD, there is a genetically determined dissociation of renal damage from hypertension, but the specific
genes involved have yet to be identified. This is somewhat
analogous to African-Americans who have a much increased risk of developing hypertensive renal damage (nephrosclerosis), yet lowering their BP below normal failed to
slow the progressive loss of their GFR (13), perhaps because
of inherited alleles from genes such as apolipoprotein L-1
(APOL-1) or the nonmuscle myosin heavy chain 9 (MYH9)
that confer genetic risk for renal damage in African-Americans specifically (442, 909, 1156).
RBF and GFR were similar initially in prehypertensive and
moderately hypertensive DS and DR rats (1255), but became reduced in DS rats with the development of hypertension and proteinuria during salt loading (229, 259, 769,
1244, 1445). The renal vasculature of DS rats was hyperresponsive to vasoconstriction by ANG II, ET-1, and NE
during a high-salt diet (517, 1366) but failed to dilate with
atrial natriuretic peptide (ANP) or NO (1366). This vascular phenotype was genetically determined since it was corrected in consomic DS rats receiving chromosome 13 from
BN rats (278). The development of hypertension in DS rats
was preceded by a shift in the pressure-natriuresis relationship toward a higher BP (1244, 1251), implying a renal
mechanism and Na⫹ retention. Normotensive DR and DS
rats fed a low-salt diet have effective steady-state whole
kidney and outer cortical and medullary blood flow autoregulation (1251) which were maintained in DS receiving a
high-salt diet in some (1244, 1445) but not all studies
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
(1251). The plateau phase of RBF autoregulation was maintained, although the inflection point for cortical blood flow
autoregulation was elevated ⬃30 – 40 mmHg (1244, 1251,
1445). However, others studies reported that neither the
renal medullary nor the papillary blood flow was autoregulated in either DS or DR rats whether fed low-salt (1244,
1251, 1366) or high-salt diets (1251), which suggests that
the salt-sensitive hypertension in DS rats is independent of
changes in medullary autoregulation. Nevertheless, another
study reported a high-salt diet fed to DS rats impaired RBF
autoregulation and related this to AT1 receptor-mediated
ROS production (1286). Moreover, frequency analysis of
dynamic RBF autoregulation in DS rats revealed normal
autoregulatory pattern of transfer gain when normotensive
on a low-salt diet, but a progressive decline in the efficacy of
autoregulation when hypertension was induced by a highsalt diet, especially when initiated 2 wk after weaning (769).
The absence of dynamic RBF autoregulation in the DS rats
with early-onset hypertension was evident as both the myogenic and TGF contributions were defective. The observation that vascular injury preceded the loss of whole kidney
autoregulation led to the conclusion that the impaired autoregulation was the result, not the cause, of the renal failure (769). This may explain the variable reports of renal
autoregulation described above.
Oxidative stress contributes to vascular damage and abnormal renal hemodynamics and Na⫹ excretion in DS rats.
NADPH oxidase was overexpressed in the renal cortex of
DS rats fed a high-salt diet (449). DS rats have enhanced O2⫺
production in the outer medullary TAL of Henle’s loop that
reduced the bioavailability of medullary NO (1030, 1453).
This could reduce medullary perfusion and attenuate the
pressure-natriuresis (424). Indeed, restoration of NO in DS
rats normalized the medullary blood flow and prevented
hypertension (1014).
MD-TGF was normal in DS and DR rats even on a high-salt
diet (770). A mathematical model led to the conclusion that
hypertensive DS rats have a severely reduced efficiency of
the myogenic mechanism, whereas the MD-TGF response
was normal (811). However, hypertensive DS rats also have
an enhanced CT-GF response that buffers renal vasoconstriction (1555). This may contribute to vasodilation, increased PGC, and glomerular damage in hypertensive DS
rats fed with high-salt diet. Afferent arterioles in the HNK
preparation of both normotensive and hypertensive DS rats
have greatly diminished myogenic responsiveness and a
higher threshold PP (589, 1436). The preglomerular resistance vessels of normotensive DS rats fed a low-salt diet
have an enhanced sensitivity to inhibition of myogenic responses by a CCB, suggesting reduced activity or density of
L-type VOCCs (1436). Both NO- and EDHF-mediated
components of ACh-induced afferent arteriolar dilation
were attenuated in DS rats by enhanced ROS (1145).
Reduced renal vascular production of 20-HETE accounted
for the impaired myogenic and MD-TGF responses of the
afferent arteriole and the weak autoregulation of RBF and
PGC (463, 1220).
Cerebral arteries from salt-fed hypertensive DS rats normally have a fast myogenic response. This was maintained
in one study (56), but attenuated in another report, thereby
leading to hypertensive encephalopathy, seizures, and disruption of the blood-brain barrier (600, 1164, 1386). The
myogenic response of skeletal muscle arterioles that is normally slow and modest was reduced, but the gracilis arteries
of hypertensive DS rats maintained myogenic tone related
to 20-HETE production that inhibited KCa channels and
reduced EM (446).
Knockdown of TGF-␤1 expression slowed the progression
of proteinuria, glomerulosclerosis, and renal interstitial fibrosis in DS rats fed a high-salt diet independently of
changes in BP (226). Prolonged blockade of TGF-␤1 increased the medullary blood flow and reduced BP modestly
(293). This suggests that the development of hypertension,
glomerular and tubular injury, and vasoconstriction all may
depend on overproduction of this growth factor in the renal
medulla of DS rats.
In summary, renal cortical blood flow autoregulation is well
maintained in normotensive DS and DR rats fed a low-salt
diet and in early hypertensive DS rats during increased salt
intake. Since the impairment of autoregulation follows the
development of renal injury, which indeed is apparent soon
after birth, it is likely the result, not the cause, of the renal
failure. The eventual loss of renal autoregulation is related
to impaired myogenic responses in hypertensive DS rats
consuming salt since MD-TGF activity is maintained. These
rats tend to have impaired cerebral arteriolar myogenic
tone, which probably contributes to their development of
hypertensive encephalopathy. The finding of preserved
myogenic responses in skeletal muscle arterioles suggests
that the severely impaired responses in most studies of renal
vessels is not due to a generalized failure of myogenic mechanisms and is consistent with a secondary response to renal
vascular and parenchymal damage.
4. Goldblatt renovascular hypertension
In 1937, Goldblatt (476) reported that hypertension commonly developed after constriction of one main renal artery
to reduce RPP (two-kidney, one-clip or 2K,1C model). This
models renovascular hypertension from unilateral renal artery stenosis. The GFR was similar in the two kidneys of
rats with moderate, early 2K,1C hypertension with no obvious renal pathology (1410). Hypertension and renal vasoconstriction are related to an activated RAS since AT1A
receptor deficient mice develop little hypertension after the
clipping of one renal artery (206, 208), and blockade of the
RAS or AT1 receptors in rats reduces the BP and produces
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
459
RENAL AUTOREGULATORY MECHANISMS
marked vasodilation of the nonclipped kidney (209, 210,
649, 650, 669, 1186, 1566), similar to CCB blockade
(1186). In contrast, the clipped kidney sustains a further
reduction in its RBF and GFR, suggesting either that its
autoregulation is severely impaired or that the RPP beyond
the clip is reduced below the autoregulatory range. An ACE
inhibitor given to rats with 2K,1C hypertension reduced the
BP, RBF, PGC, and afferent and efferent arteriolar resistances in the clipped kidneys (310), whereas a CCB reduced
afferent arteriolar resistance selectively (310).
Micropuncture studies of superficial nephrons in moderately hypertensive 2K,1C rats reported that the nonclipped
kidney has well maintained or increased SNGBF, SNGFR,
and PGC, findings consistent with a predominant increase in
preglomerular vascular resistance (308, 452). However, in
others, the PGC was elevated and the SNGBF and SNGFR
were reduced, suggesting a reduced KUF (452, 1410). The
SNGBF, SNGFR, and PGC were reduced in the clipped kidney of 2K,1C rats, implying increased preglomerular vascular resistance from the clip and/or the vasculature (1336 –
1338).
Constriction of one main renal artery and removal of the
other kidney (one-kidney, one-clip or 1K,1C model) also
causes hypertension, but this model is less dependent on the
RAS and is distinctly salt sensitive. The 1K,1C is a model of
renal transplant stenosis or bilateral renal artery stenosis.
These clinical conditions resemble the 1K,1C model since
all lack a “normal” contralateral kidney to correct renal salt
retention by the clipped kidney. The clipped kidney of
1K,1C rats had reduced SNGBF and SNGFR with unchanged PGC due to a preferential increase in efferent arteriolar resistance (1336).
Dogs (1047) and some rats (705, 707, 1365, 1566) with
2K,1C hypertension have maintained RBF and GFR autoregulation in the nonclipped kidney, although other studies
in rats have reported weak autoregulation (1188, 1404,
1494) that was improved by NOS inhibition (1494). Blockade of AT1 receptors in 2K,1C hypertensive rats did not
affect RBF autoregulation in the nonclipped kidney but
abolished GFR autoregulation (1566), presumably because
of a reduced efferent arteriolar resistance. Autoregulation
of renal cortical perfusion of the nonclipped kidney of
2K,1C hypertensive rats was relatively resistant to blockade
by a CCB (646), suggesting upregulation of L-type VOCCs.
The clipped kidney has a weak, or even nonexistent, autoregulation of RBF (705). The post-clip kidney and its blood
vessels were protected from barotrauma, whereas there
were extensive glomerular and vascular injuries in the contralateral nonclipped kidney (1365).
yet maintained NO vasodilation (684), indicating a selective impairment of autoregulation. The myogenic response
of afferent arterioles in HNK nonclipped kidneys of rats
with 2K,1C hypertension was blunted and shifted to higher
levels of PP (589).
Increased NO production by the nonclipped kidney (318,
1363) impaired its RBF autoregulation since this was normalized by NOS inhibition (1494), whereas reduced EET
production contributed to renal vasoconstriction. However, restoring EETs did not restore autoregulation (1404),
indicating that they were not the cause of defective renal
autoregulation. Similarly, 20-HETE participated in basal
vasoconstriction, but did not account for the attenuated
autoregulation in the unclipped kidney of 2K,1C rats. Thus
increased levels of NO could account for some of the
changes in autoregulatory efficiency, whereas altered production of EETs and 20-HETE are not implicated.
The MD-TGF response in the nonclipped kidney of 2K,1C
rats generally was normal or enhanced (134, 648). However, MD-TGF responses in the nonclipped kidneys of
2K,1C rats early after developing hypertension were attenuated despite an activated RAS (1191, 1314, 1317). The
MD-TGF response in the nonclipped kidneys was enhanced
greatly after NOS inhibition (1491, 1494). MD-TGF in the
clipped kidney was very hard to measure, but was normalized shortly after removing the renal artery clip (1191,
1317). MD-TGF was attenuated severely in the solitary
kidney of 1K,1C rats (1187).
Dynamic analysis of proximal tubular pressure in the nonclipped kidney of 2K,1C rats during the first week of hypertension revealed normal MD-TGF oscillations. As the hypertension progresses, these oscillations transitioned into
deterministic chaos, similar to that found in adult SHR
(1635). To our knowledge, renal vascular admittance has
not been analyzed across the full-frequency spectrum in
animal models of renovascular hypertension.
In summary, results on the efficacy of steady-state autoregulation of RBF in the nonclipped kidney of 2K,1C hypertensive rats are quite mixed, with roughly equal numbers
reporting that it is well-maintained or impaired. The reason
for this disparity is not known. Likewise, reports of MDTGF responses in nonclipped kidneys have varied from normal to markedly impaired. The MD-TGF activity appears
to be the outcome of an enhancement due to ATI-receptor
activation from renin released by the clipped kidney that is
offset by increased NO. Autoregulation is weak or absent in
the clipped kidneys, perhaps because the RPP is at, or below, the lower limit of autoregulation.
5. ANG II-induced hypertension
Medullary blood flow autoregulation in the nonclipped kidney was impaired (646), and its preglomerular juxtamedullary blood vessels failed to vasodilate during reduced RPP
460
Prolonged infusion of ANG II at an initially subpressor rate
gradually increases BP, oxidative stress, RVR, inflamma-
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
tion, and renal parenchymal injury in association with renal
vasoconstriction and an early phase of Na⫹ retention (492).
The degree of hypertension and renal injuries are related to
the dose of ANG II used, choice of species (rats vs mice), and
the duration of treatment. Cross-transplantation studies in
AT1 receptor-deficient mice demonstrated that the major
prohypertensive actions of ANG II are on renal AT1 receptors (281). Development of hypertension was prevented by
an AT1 receptor blocker (1550). ANG II receptors, in particular AT1, in the vasculature are usually downregulated
by high levels of ANG II (15, 77, 91, 526, 1266). Renal
vascular AT1 receptors expression was reduced in ANG
II-induced hypertension (39).
Most studies of renal function have been conducted during
the established phase of hypertension after 1 or 2 wk of
ANG II infusion. At this time point, the Na⫹ excretion,
RBF, and GFR were normal or modestly reduced (207, 652,
1088). RVR is consistently increased. The SNGBF, SNGFR,
and KUF were reduced and afferent and efferent arteriolar
resistances increased (436). PGC was increased during albuminuria (1008). Pressure natriuresis in ANG II-induced hypertension was reset to a higher RPP despite efficient autoregulation of medullary blood flow (1516, 1549).
Blockade of AT1 receptors in ANG II-induced hypertensive
rats reduced BP and increased RBF and cortical blood flow,
GFR, and Na⫹ excretion (1550). Reactivity of afferent arterioles of JMN to ANG II, but not PE, was increased after
1 or 2 wk of ANG II infusion (670). The exaggerated response depended on the presence of renal nerves (660) and
increased inactivation of EETs by soluble epoxide hydrolase (1662). Reactivity of afferent arterioles of JMN to ATP
by P2X1 receptors was reduced after 1 or 2 wk of ANG II
infusion (1661).
ANG II-induced hypertension and renal vasoconstriction
are dependent on enhanced ROS production by NADPH
oxidase (1015, 1128, 1584), impaired renal cortical NOS
activity (241, 242, 1025), and reduced renal medullary NO
production (1432). Although COX inhibition has little impact (483), hypertension and renal vasoconstriction were
less in TP receptor gene-deleted mice (441, 780), implicating involvement of a vasoconstrictor prostanoid, perhaps
TxA2, acting on TP receptors.
Prolonged infusion of ANG II increased NOS activity in the
renal cortex, with enhanced expression of eNOS and
nNOS, but no change in NOS activity in the renal medulla
(241). Accordingly, NOS inhibition produced greater renal
vasoconstriction with greater reductions in RBF and cortical blood flow (242). The authors concluded that a compensatory increase in NO helps to counteract the vasoconstrictor influence of ANG II and thus maintains RBF and
cortical perfusion.
The normal renal vasoconstriction after inhibition of nNOS
was lost in ANG II-infused rats (207). A CCB normalized
BP, RVR, GFR, and Na⫹ excretion in rats with ANG IIinduced hypertension (652). Vascular remodeling may contribute to ANG II-induced hypertension (373). ANG II produced less hypertension in adenosine A1 receptor-deficient
mice (455). Enhanced afferent arteriolar responses to ANG
II in hypertensive wild-type mice were reduced by tempol.
Contractions in arterioles lacking the A1 receptor were
weaker and not attenuated by tempol, suggesting that A1
signaling may modulate ANG II-mediated contraction or
oxidative stress.
RBF autoregulation was impaired, whereas medullary
autoregulation was preserved in nondiuretic ANG II-infused rats (1549). One study of ANG II-induced hypertension in UNx rats reported normal autoregulatory responses of the afferent arteriole of JMN whether or not
the kidney was denervated (660). Other investigators reported that the autoregulatory response of the cortical
radial artery and afferent arteriole in the JMN preparation was attenuated 14 days after ANG II-induced hypertension (198, 670, 689). This was related to vascular
injury (198, 1029) and activation of ET-1 receptors. Normalization of BP by prolonged AT1 receptor blockade or
by triple therapy (hydralazine, hydrochlorothiazide, and
reserpine) restored autoregulation of afferent arterioles
(689). RBF autoregulation was impaired during ANG
II-induced hypertension (AI increased from 0.04 to 0.66),
which was normalized by blockade of P2Y12 receptors
(1132). Administration of the anti-inflammatory agent
pentosan polysulfate for 14 days restored autoregulation
of afferent arterioles of JMN in ANG II hypertensive rats
via normalization of P2X1 receptor activation by ATP
during sustained hypertension (520). Other studies on
superficial nephrons reported augmented P2X1-mediated
constriction of afferent and efferent arterioles in rats
with ANG II-induced hypertension while P2X1 receptor
density was normal, perhaps due to inhibition of NO
production (436).
In contrast, UNx rats infused with ANG II for 2 wk have
normal afferent arteriolar myogenic responses of JMN
(660), as do cortical afferent arterioles from ANG II-infused
mice studied ex vivo (864). Afferent arterioles from ANG
II-infused rabbits have enhanced contractions to ANG II
despite a twofold downregulation of mRNA for AT1 receptors (1552). This was offset by a fivefold upregulation of
mRNAs for p22phox (1552). In agreement, enhanced arteriolar reactivity to ANG II was absent in mice with NOX-2
deficiency (182, 184). Moreover, isolated cortical radial
arteries of ANG II-infused mice lacking regulator of G protein signaling-2 had stronger myogenic tone and constrictor
responses to ANG II, probably by sensitizing AT1-receptor
signaling (604).
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
461
RENAL AUTOREGULATORY MECHANISMS
Admittance transfer function in anesthetized ANG II-infused rats demonstrated impaired myogenic and MD-TGF
components of renal autoregulation independent of ET-1
(1272) but magnified by a high-salt diet that increased ROS
(1273). However, a careful study of dynamic autoregulation in conscious ANG II-infused rats reported a clearly
enhanced myogenic mechanism and reduced transmission
of BP fluctuations to the kidney (1193), whereas the MDTGF signature was unaltered. Likewise, long-term ANG II
infusion in conscious dogs potentiated the myogenic resonance peak (755). Thus these studies in conscious animals
with prolonged ANG II infusion may explain the relatively
modest degree of hypertensive glomerular injury since increased myogenic responses were lost under anesthesia.
ANG II is known to stimulate oxidative stress and increase
NADPH oxidase expression in the kidney (211, 455, 578).
Tempol normalized the vascular O2⫺ production and reduced BP without affecting the RBF of ANG II-infused rats
(1088). Oxidative stress is thought to play an important
role in the regulation of medullary blood flow and pressure
natriuresis (273, 964, 1099, 1670).
ANG II can activate T cells that release cytokines to promote oxidative stress, inflammation, renal vasoconstriction, and Na⫹ retention, all of which are prohypertensive
(534, 579). As observed with an AT1 receptor blocker, administration of the anti-inflammatory agent mycophenolate
mofetil during ANG II infusion reduced BP, renal vasoconstriction, and renal interstitial inflammation (439).
Although the MD-TGF was augmented by short-term ANG
II infusions (1012, 1077) and usually attenuated by pharmacological or genetic inhibition of the RAS and AT1 receptors (see sect. VIA), prolonged slow-pressor infusion of
ANG II did not change MD-TGF at normal tubular fluid
flow (1095). During chronic ANG II infusion, there was a
complex regulation of MD-TGF by a vasoconstrictor prostanoid that activated TP receptors and by ROS that impaired NO signaling in the JGA and by reduced expression
of AT1 receptors (39). COX-1 products enhanced MD-TGF
by activation of TP receptors that elevated ROS in the kidney (137, 1082, 1083, 1581, 1593), whereas COX-2 products attenuated MD-TGF in ANG II-infused mice (39).
ANG II-induced hypertension depended in part on activation of TP receptors (780). Increased MD-TGF during prolonged ANG II infusion was related to a reduced blunting
by MD-derived NO because of enhanced O2⫺ (133, 184,
659, 1608). Small interference RNA knockdown of p22phox
reduced ROS and MD-TGF activity by 50% in ANG IIinfused rats, thereby linking ROS to enhanced MD-TGF
(1095). P2X1 receptor signaling in the kidney (436) and
preglomerular vessels of JMN was impaired by inflammation caused by ANG II infusion (520, 686, 1661). Prevention of inflammation improved autoregulation, apparently
by normalizing MD-TGF activity (686, 1661).
462
In summary, steady-state RBF autoregulation is attenuated
in most studies of ANG II-induced hypertension, although
medullary blood flow and its autoregulation are generally
well maintained. Myogenic tone is enhanced in intact kidneys of conscious animals, but not in isolated arterioles,
suggesting that a reduced MD-derived NO may enhance
responses in intact kidneys. MD-TGF is normal or exaggerated, likely the outcome of enhanced O2⫺ generated in the
JGA in response to vasoconstrictor COX-1 metabolites that
activate TP receptors and decrease NO activity. However,
MD-TGF is attenuated in JMN, perhaps related to inflammation impairing P2X1 receptor signaling. The reasons for
the differential effects of ANG II on MD-TGF responses of
superficial and deep nephrons and the variable effects of
ROS to dampen or enhance myogenic responses are not yet
resolved. Thus there are major hemodynamic differences in
the response to chronic versus acute activation of the RAS
and perhaps the balance of NO and oxidative stress. Longterm ANG II stimulates ROS production that impairs the
bioavailability of NO and enhances production of a vasoconstrictor prostanoid that activates TP receptors. These
are offset by adaptive reductions of AT receptor expression
and structural changes of the vascular wall.
6. DOCA-salt hypertension
Prolonged administration of the mineralocorticosteroid deoxycorticosterone acetate (DOCA) and a high-salt diet
(e.g., 1% NaCl) have usually been combined with UNx to
model volume-expanded, low plasma renin activity (PRA)
hypertension similar to human primary hyperaldosteronism. This model is dependent on increased ET-1 production
in blood vessels and kidneys (1291). DOCA-salt mice have
a rightward shift in the pressure-natriuresis relation (512).
UNx rats with DOCA-salt hypertension have increased PGC
and develop proteinuria and focal glomerulosclerosis (361,
363). Nonspecific antihypertensive therapy (hydrochlorothiazide, hydralazine, and reserpine) reduced BP in
DOCA-salt rats but failed to prevent elevated PGC, proteinuria, or glomerular injury (361). Combined blockade of Tand L-type VOCCs decreased BP and PGC and reduced proteinuria and glomerular lesions without affecting the RAS
(75, 767), whereas selective antagonists of L-type VOCCs
also decreased the BP but stimulated the RAS and failed to
prevent proteinuria or glomerular lesions (75, 365, 767).
This finding in hypertensive animals is perplexing in view of
the reported absence of T-type channels in glomerular arterioles of normotensive rats (1388). However, another study
reported that an L-type CCB was more effective than ACE
inhibition in reducing the BP, proteinuria, and glomerular
pathology, although both treatments reduced the PGC
(364). A sharp fall in BP sufficient to reduce PGC may account for the protective effects of the CCB in this study.
Dietary Ca2⫹ supplementation reduced PGC, proteinuria,
and renal injury without changing the BP (359). Thus an
elevated PGC and the RAS are more important than hyper-
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
tension in predicting proteinuria and renal damage in the
DOCA-salt model.
Basal RBF and GFR are normal or reduced in DOCA-salt
rats and dogs (45, 540, 540, 674, 969, 1047, 1047, 1253).
There was a maintained plateau phase of RBF autoregulation, with either a normal (45) or a 20 mmHg increase in its
lower limit (674). RBF autoregulation in dogs with DOCAsalt hypertension was well preserved and independent of
renal sympathetic nerves or PGs (1420), but GFR was
poorly autoregulated (550). Neither RBF nor GFR was autoregulated in the isolated perfused kidney of DOCA-salt
dogs (764), but autoregulation generally is less efficient in
this preparation. The RBF was less efficiently autoregulated
in DOCA-salt hypertensive Yucatan miniature swine, but
again there was even less complete autoregulation of GFR
(1647). RBF and renal cortical blood flow were well autoregulated in DOCA-salt mice, and surprisingly, medullary
blood flow was more efficiently autoregulated in DOCAsalt than in control mice (512).
Some studies of DOCA-salt rats reported attenuated MDTGF (1040, 1315) and autoregulation of SNGFR (1022),
whereas others reported that the autoregulation of GFR
and SNGFR were well maintained despite inhibition the
MD-TGF (540). The findings in several studies of DOCAsalt hypertensive rats, dogs, or swine that the GFR was less
well autoregulated than the RBF suggest a reduced efferent
arteriolar resistance due to the strongly suppressed RAS.
Vascular O2⫺ production was increased in DOCA-salt hypertensive rats (1391). Antioxidants (100, 968) or inhibition of NADPH oxidase (641, 735) reduced BP, vascular
O2⫺, and renal damage. ET-1 production was increased in
the vasculature of DOCA-salt hypertensive rats (227,
1292), and selective blockade of ETA receptors reduced BP,
vascular O2⫺ generation, and RVR, but did not always improve renal damage (14, 175, 938, 1264). ETB receptors
may provide a protective function. Rats lacking ETB receptors have exaggerated vascular and renal pathology (986)
accompanied by reduced NO production (620) that may
underlie the enhanced vasoconstriction of their afferent arterioles (1470). However, others have ascribed the renal
vasoconstriction and hypertension to enhanced activity of
the renal sympathetic nerves (719, 776), since renal denervation delayed the development of hypertension and increased Na⫹ excretion (775) or to inflammation since immunosuppression preserved GFR and reduced proteinuria
in DOCA-salt hypertension (124). Another study implicated purinergic signaling since P2X7 receptor-deficient
DOCA-salt mice were less hypertensive and infiltration of
immune cells and renal injuries were blunted (732).
In summary, DOCA-salt is a model of low-renin, volumeexpanded hypertension that resembles human primary hyperaldosteronism. Factors contributing to hypertension, re-
nal vasoconstriction, and renal damage include oxidative
stress, NO deficiency, inflammation, enhanced P2X7 and
ETA but reduced ETB receptor signaling, and enhanced renal nerve activity. Despite the renal vasoconstriction and
abnormal receptor signaling, RBF autoregulation is well
maintained in most studies, albeit GFR autoregulation is
less efficient, likely reflecting a diminished efferent arteriolar resistance in this low renin-angiotensin model. The preservation of RBF autoregulation despite a markedly attenuated MD-TGF response suggests an intact or enhanced
myogenic mechanism, but this has not been studied specifically.
7. FH hypertensive rat
The FH rat is a genetic, renin-dependent, model of systemic
and glomerular hypertension that develops progressive albuminuria and spontaneous focal and segmental glomerulosclerosis (FSGS) that shortens its lifespan (841). It resembles progressive loss of function in humans with CKD. Mild
hypertension and increased PGC developed by 8 wk (843,
1371), proteinuria and enhanced myogenic contraction of
small mesenteric artery by 20 wk (1527), FSGS and proteinuria by 2– 6 mo, and severe hypertension and CKD with
malignant nephrosclerosis by 1 yr (294, 840, 842, 843,
1202, 1527, 1579). The FH rats are reported to die of renal
failure at sea level, whereas at altitude (e.g., Denver-raised
FH rats) they die of pulmonary hypertension, which has
been characterized by a deficiency of pulmonary eNOS and
abnormal vascular growth (878).
Studies in FH rats, before and after appearance of renal
injury, demonstrated that attenuated renal myogenic responses in 7-wk-old FH rats preceded glomerulosclerosis
and proteinuria (1102). Increased renal cortical and medullary blood flow and PGC were related to a low ratio of
afferent to efferent arteriolar resistance. Autoregulation of
RBF, cortical perfusion, and PGC were impaired markedly
in adult FH rats (1522, 1534). Prolonged ACE inhibition
largely prevented the systemic and glomerular hypertension
and renal damage (1534, 1535).
Isolated, perfused cortical radial arteries from young FH
rats constricted weakly or even dilated with increased PP
(1102, 1522, 1527). This severe defect, or even absence, of
myogenic contraction was considered renal specific since
those contracted normally to PE (1534) and the myogenic
responses of isolated mesenteric arteries and the aorta were
preserved (1102, 1527). This suggested a profound defect in
renal arteriolar mechanosensitivity or mechanotransduction. On the other hand, cortical radial arteries and afferent
arterioles in the HNK preparation of FH rats had nonselective impairment in reactivity to PP and to ANG II (1523).
Micropuncture studies of FH rats with mild FSGS reported
normal MD-TGF activity that increased with age and was
independent of ANG II (1534). An increased expression of
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
463
RENAL AUTOREGULATORY MECHANISMS
neuronal NOS and COX-2 in the MD and of renin in the
afferent arterioles of FH rats preceded the development of
glomerulosclerosis (1579). Thus preglomerular vasodilation by NO, perhaps driven by a COX-2 metabolite, may
combine with hypertension caused by high ANG II to impair renal autoregulation and increase PGC that culminates
in barotrauma manifest as FSGS.
Linkage analysis showed that genetic loci on chromosome 1
of FH rats cause the hypertension, impaired autoregulation,
and proteinuria (1518 –1521). Transfer of this region from
BN into FH rats improved RBF autoregulation, normalized
the elevated PGC, and reduced protein excretion and the
progression of renal disease (922, 1610), whereas transfer
of this chromosomal region from FH rats into normotensive
controls attenuated renal autoregulation and led to FSGS
(1518). This provides strong evidence that the selective impairment of the renal myogenic response and the hypertension in this model are genetically determined and are causally related to glomerulosclerosis. The particular genes involved have yet to be identified.
Thus the selective impairment of renal myogenic responsiveness in FH rats reported in most studies despite preserved MD-TGF, PE-induced vasoconstriction, and myogenic responses in other vascular beds, suggests that the
defect in renal autoregulation is the outcome of genetically
determined changes in the preglomerular vasculature. The
weak myogenic response likely accounts for the attenuated
renal autoregulation that, in the context of hypertension,
leads to barotrauma and CKD. This conclusion is supported by recent studies of the FH hypertensive, consomic
and congenic rat strains, which demonstrate the importance
of autoregulation in preserving glomerular structure and
function.
8. BN rat
The BN rat is of special interest because it develops renal
failure despite a low BP, likely related to weak and inconsistent renal autoregulation because of excessive renal NO
generation that can prevent full vasoconstrictor responses
(260, 916). Although the BN kidney is susceptible to hypertension-induced kidney disease (260), the BN lives a long
life, perhaps due to its low BP (916). Cross-transplantation
studies demonstrated that the kidney of the normotensive
BN rat was inherently more susceptible to hypertensioninduced damage than the SHR (260, 916), thereby locating
the deficit within the kidney itself. Surprisingly, the kidneys
of BN rats have efficient steady-state RBF autoregulation
with a basal AI of 0.19 (922), but this is unusually fragile
since they are susceptible to injury with even moderate hypertension or enhanced oscillations in RPP. Dynamic admittance transfer function analysis identified an apparently
normal myogenic response with spontaneous changes in
RPP, but a weaker myogenic response to larger, forced
changes in RPP (1563). Thus incomplete autoregulation
464
could underlie the susceptibility to hypertensive renal injury. Cupples and co-workers (1569, 1570) reported that
the attenuated myogenic autoregulation was improved by
nonselective NOS inhibition or by inhibition of iNOS but
not of nNOS, implicating high NO generation from iNOS
as the cause of the blunted myogenic response. In contrast,
the renal autoregulation of normal Wistar rats was unaffected by inhibition of iNOS, but the strength of the myogenic response is enhanced by inhibition of nNOS (1358).
Since these BN rats also have exaggerated depressor response to ACh, which was normalized by global NOS inhibition, they also may have an enhanced eNOS-derived NO
(1570). Wang and Cupples (1570) concluded that the normal buffering of the myogenic response by MD nNOSderived NO was dominated in BN rats by an inappropriate
release of NO by iNOS, and perhaps by eNOS, that attenuated the renal myogenic response. The roles of NO in renal
autoregulatory mechanisms were discussed in detail earlier
in section VIC.
BN rats have a preserved myogenic response of cremaster
arteries and gracilis muscle resistance arteries that depended on 20-HETE production (445). Impaired myogenic
tone has been reported of cerebral arterioles (322) and may
contribute to their susceptibility to cerebral hemorrhage
and stroke.
Recent studies show that a 2.4-Mbp region of chromosome
1 was critical for normal RBF autoregulation since its substitution from BN into FH rats improved their RBF autoregulatory efficiency and afferent arteriolar myogenic responses, reduced BK channel activity (168), and improved
the myogenic tone of cerebral arteries (1147).
In summary, BN rats have labile renal autoregulation that
breaks down during hypertension. The weak link is identified as an impaired dynamic myogenic response related to
dysregulated overproduction of NO, but the molecular
mechanisms have not yet been defined. The defective myogenic response is considered genetically determined by sites
on chromosome 1 that may impair both renal and cerebral
vasoconstriction during increased PP.
C. CKD and RRM
A fundamental problem with human CKD is its propensity
to undergo a detrimental switch from beneficial adaptations
that restore GFR to pathological proteinuria and declining
renal function (637). This sequence is modeled by a 50 –
85% RRM that is characterized in rodents by initial increases in glomerular size, PGC, and SNGFR and hypertrophy in remnant nephrons and followed by proteinuria and
progressive tubulointerstitial fibrosis and glomerulosclerosis usually accompanied by hypertension (29, 144, 147).
The rate and the magnitude of these changes depends on the
species, age, amount of renal mass removed, and the tech-
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
nique (surgical ablation or infarction). The infarction
method causes ischemic, renin-dependent hypertension and
a more rapid decline of GFR, but is a less representative
model for human CKD where renal tissue is not clearly
ischemic and PRA is generally suppressed. These defects
were accelerated by high dietary intakes of protein or salt
(147, 803). Bidani and colleagues related the renal damage
to hypertension coupled with impaired autoregulation that
allowed transmission of enhanced BP fluctuations to the
glomeruli to initiate barotrauma (103, 104, 108, 112, 496,
504).
Nephron adaptations have been studied extensively after
either UNx or 5/6 Nx and generally produce similar results.
The SNGFR of Munich-Wistar rats doubles by day two of
RRM and triples by 2 wk (319, 636, 637, 1165), restoring
whole kidney GFR and RBF almost to normal (497). The
early renal vasodilatation is attributed to enhanced PG production (497, 1166, 1167). Proteinuria and mesangial expansion are apparent by 2 wk and are accompanied by an
increased glomerular volume and capillary diameter (518,
637, 1119). There is a ⬃10 mmHg increase in PGC as a
result of a preferential decrease in afferent arteriolar resistance (357, 876, 1166). Rats with RRM had persistent systemic hypertension, progressive proteinuria, and glomerular mesangial expansion and segmental sclerosis by 8 wk
(31). The decline in renal function was considered to be
secondary to an initial increase in PGC, SNGFR, and
SNGBF (29, 144, 147). Inhibition of the RAS lowered BP
and PGC and lessened the proteinuria and glomerulosclerosis in the RRM infarction rat model. This was accompanied
by a normalization of SNGBF and SNGFR by a relatively
selective reduction in the efferent arteriolar resistance (32,
1004). Impaired autoregulatory adjustments of the afferent
arteriole in rats with RRM likely contributed to the reduced
afferent arteriolar resistance despite hypertension, the increased PGC, and the glomerulosclerosis (160, 876, 1167).
The renal adaptations and impaired autoregulation are dependent on the strain and the age of the rat. Perinatal UNx
in normal rats resulted in salt-sensitive hypertension, exaggerated MD-TGF, proteinuria, increased glomerular volume, and renal injury by 3 mo (186, 1611) which were
ascribed to increased ROS and NO deficiency (181, 185).
UNx in neonatal SD rats led to renal vasodilation and complete loss of steady-state RBF autoregulation at 5– 6 wk
attributed to overproduction of PGs (235–237). However,
UNx in adult SD rats increased basal RBF but with little
impact on whole kidney RBF autoregulation (237). Weak
autoregulation in UNx FH rats accelerated the process of
progressive renal failure with increased PGC and SNGFR
evident by 12 wk, followed by severe proteinuria and a fall
in GFR (1371).
SHR kidneys have highly efficient steady-state RBF autoregulation with normal PGC, as discussed in section VIIB1
(42, 58, 358, 394, 707, 1471) and minimal glomerular
damage (358, 394, 707, 1471). After UNx, the remnant
kidney of young SHR has elevated PGC and they develop
proteinuria (93, 358, 360). However, adult SHR retained
efficient overall steady-state RBF autoregulation after UNx
(709). The renal adaptations were more severe in 5/6 Nx
rats. Within 2 wk of 5/6 Nx produced by surgical ablation
in adult SHR, hypertension worsened and RBF autoregulation was abolished with evident injury to preglomerular
arteries, arterioles, and glomeruli (107). Normal rats subjected to 5/6 Nx also developed hypertension and, by 3 wk,
had impaired steady-state RBF autoregulation (503). Their
juxtamedullary afferent arterioles became hypertrophic and
their arteriolar autoregulatory response was impaired
(903).
Hypertension and elevated PGC are considered key elements
in renal damage in UNx SHR and 5/6 Nx Munich-Wistar
rats since injury was mitigated by antihypertensive treatment with hydralazine, hydrochlorothiazide, and reserpine
(501); salt restriction (357, 876); or with ACE inhibition or
even with a CCB in some studies (360, 362). However,
impaired renal autoregulation also was a critical element in
renal damage since a low-protein diet improved steadystate RBF autoregulation and prevented proteinuria or renal damage without prominent effects on BP in 5/6 Nx rats
(112, 502). Moreover, loss of autoregulation by blockade
of VOCCs by CCBs in some other studies exacerbated renal
damage in rats with RRM despite lowering their BP (499,
500, 502, 503).
The best method to evaluate autoregulation and its mechanisms in RRM is not clear. Frequency analysis of dynamic
RBF autoregulation suggests preserved autoregulation in
rats with RRM. However, this contrasts with defective RBF
autoregulation observed in steady-state studies (109). Bidani et al. (109) concluded that the staircase steady-state
method provides a better measure of susceptibility to hypertensive glomerular damage, but the reason for the preserved
dynamic RBF autoregulation is unresolved.
Most studies have concluded that renal damage requires, or
is accelerated by, a high BP. Thus rodent strains that do not
develop hypertension after RRM, such as WKY rats (110)
or C57Bl/6 mice (861), are relatively resistant to renal damage after RRM despite impaired autoregulation and myogenic responses. However, after the induction of hypertension, WKY rats with RRM indeed sustain renal damage
(121, 609, 1282, 1451, 1452) as did C57Bl/6 mice infused
with ANG II or treated with DOCA and salt (111, 180,
1237).
Some studies have dissociated worsening autoregulation
from progression of CKD. Thus the myogenic response of
renal afferent arterioles isolated from C57Bl/6 mice was
attenuated progressively over 3 mo after RRM (861). While
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
465
RENAL AUTOREGULATORY MECHANISMS
a high-salt diet enhanced the defects in myogenic responses,
there was no glomerulosclerosis and only mild tubulointerstitial fibrosis (861), perhaps reflecting the absence of hypertension even with a high-salt diet (861). Rho-kinase inhibition in SHR after subtotal Nx attenuated renal autoregulation but paradoxically reduced the renal damage (594).
Blockade of PG production in rats with established RRM
increased the afferent and efferent arteriolar resistances, but
did not improve renal autoregulatory efficiency or prevent
renal damage (497, 1063). ACE inhibition after RRM failed
to fully restore RBF autoregulation but was protective of kidney function (503). Anti-inflammatory agents given to rats
with RRM prevented the rises in SNGFR and PGC and the
development of proteinuria and decreased the mesangial proliferation but increased the hypertrophy of afferent arterioles
(121, 609, 1451, 1452). Blockade of ROS generation by tempol or by inhibition of xanthine oxidase protected kidneys of
rats (1282) or mice (861) with RRM without affecting BP.
Nevertheless, the weight of the data generally supports the
proposal of Bidani et al. that both hypertension and impaired
renal autoregulation are required after RRM for kidneys to
develop glomerulosclerosis. However, changes in Rho/Rho kinase, PGs, ANG II, and ROS can impact RRM progression
apparently independent of effects on autoregulation or BP.
Anti-inflammatory or antioxidant drugs might maintain renal
microvascular function after RRM (903, 1112), but this requires further study.
Impaired myogenic constriction of mesenteric arteries of
hypertensive 5/6 Nx rats was restored by long-term inhibition of ACE or antagonism of AT1 receptors that reduced
BP and renal injury (1528, 1537). Tempol plus catalase
improved myogenic constrictor response of mesenteric arteries of rats with RRM but not those with prolonged AT1
receptor blockade (1528), suggesting that ANG II-induced
ROS can impair autoregulation of nonrenal arteries. This
could provide a link between damaged kidneys that release
renin and impaired systemic microvascular function via
ROS generation.
The response to RRM has been studied most extensively in
rodents, and similar patterns are not always reported for other
species. Cats with RRM developed increased SNGFR (158)
and PGC (158, 159), hypertension, and proteinuria (9, 157).
Their remnant nephrons exhibited glomerular hypertrophy
and impaired renal autoregulation (160). Unlike rodents, the
feline renal parenchyma was quite resistant to structural damage (1598), and glomerular hyperfiltration was not exacerbated by increasing the protein intake (159, 1598) even when
studied over several years (1110 –1112). Thus the cat dissociates hemodynamics, and even hypertension, from structural
adaptations and the development of CKD.
Functional adaptations to RRM in dogs are less well characterized. The progression of CKD was accompanied by
glomerular hypertension and glomerular hypertrophy com-
466
bined with proteinuria, tubulointerstitial inflammation,
and oxidative stress (17, 965, 1326). Initially, 3/4 and 7/8
RRM led to increased SNGBF, SNGFR, and PGC (158), but
RBF autoregulation was impaired severely only in dogs
with 7/8 Nx (160). ANG II mediated the progression of
CKD in dogs with RRM since prolonged ACE inhibition
reduced PGC, proteinuria, and glomerular and tubulointerstitial lesions (1326), although it also reduced BP which
may have made a critical impact. Protein loading increased
GFR in normal dogs (420) and in 7/8 RRM dogs with
increased renal growth but failed to worsen the severity of
renal lesions (1598) as was established in rats with RRM
(87, 389). However, a low-protein diet in dogs with 11/12
RRM reduced the severity of proteinuria, glomerulosclerosis, and renal interstitial lesions (1196). Compared with
rodents, a higher degree and longer dosing of RRM are
required for dogs to develop renal damage. Thus rats with
even modest degrees of RRM have impaired renal autoregulation, with glomerular and systemic hypertension, and
develop renal injury quite rapidly. Some strains of mice
(notably the C57BL/6), like cats and dogs, are quite resistant to the detrimental effects of RRM. This raises concern
about whether the rat is a valid model for progressive CKD
in humans. The hypothesis derived from the rodent studies
that a very low protein intake would slow the loss of GFR in
patients with moderate or advanced CKD was disproved by
the results of a controlled clinical trial which also reported
that a low BP goal did not slow loss of GFR unless there was
⬎3 g daily of proteinuria (803). This was confirmed in a
trial of BP reduction in African-Americans with hypertensive nephrosclerosis where an ACE inhibitor was not clearly
superior to a ␤-adrenoceptor blocker in preventing loss of
GFR, unlike the reports in rats (147), although a CCB did
accelerate GFR loss, as in most animal models (13). Again,
these results of human studies challenge the relevance of the
rodent models, at least for a complete understanding of
human CKD progression.
The rapid and substantial increases in SNGFR that occur in
rats within 1 day after RRM indicate that MD-TGF is less
active, or reset, to accommodate the likely considerable
increase in tubular fluid delivery to the MD. Indeed, MDTGF responses were effectively abolished within 7 days of
RRM in most, but not all, rat nephrons (1372, 1373). The
immediate response of MD-TGF is less clear. One view is
that MD-TGF is reset within 2 h of UNx to higher levels of
tubular flow (1041) by vasodilator PGs (546). Another
thought is that MD-TGF is essentially normal shortly after
UNx despite the increase in SNGFR (117, 1372). Efficient
GFR autoregulation persisted in volume-expanded UNx
rats in the absence of a demonstrable MD-TGF, implicating
a strong myogenic response accounting for complete autoregulation (945). One week of administration of an AT1
receptor blocker attenuated the MD-TGF activity in shamoperated rats, but paradoxically normalized the previously
nonexistent MD-TGF response in 5/6 Nx rats (1372). Thus
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
there is a complex interplay of time, volume status, and
ANG II signaling after RRM that can impact MD-TGF in a
rather unpredictable manner.
Hypertensive rats with 5/6 Nx lost myogenic responses of
their mesenteric arteries, but this was preserved after prolonged inhibition of the RAS or after tempol plus catalase
(1528, 1537). Thus ANG II and ROS can impair myogenic
tone in nonrenal vessels from models of CKD in contrast to
the opposite effects in normal renal afferent arterioles (658,
660). However, cremaster and femoral arteries from rats
with RRM displayed enhanced myogenic contractions before the development of hypertension that were normalized
by ACE inhibition (1288, 1289). Human subcutaneous arteries from patients with end-stage renal disease (ESRD)
had an unchanged myogenic tone and lacked vascular remodeling (937). These different patterns of myogenic response in vessels from models of CKD remain unexplained.
In summary, rodent models of UNx or RRM generally display glomerular hypertrophy, increased SNGFR, and PGC.
A combination of systemic hypertension and attenuated
renal autoregulation generally are followed by proteinuria
and glomerulosclerosis. Initially, a vasodilator COX metabolite and NO participate in afferent arteriolar vasodilation and impair renal autoregulation while efferent arteriolar tone is maintained by ANG II, thereby contributing to
the increased PGC and SNGFR. Oxidative stress and inflammation impair renal autoregulation in chronic RRM models. The myogenic response is generally attenuated, but
there are complex, time-dependent effects on MD-TGF.
However, very different patterns of response and renal
damage have been reported between studies and between
different nonrenal vessels. These may depend on the degree,
method of RRM, the age of the animal, and especially its
species, which cautions against drawing uncompromising
conclusions concerning autoregulation in human CKD.
D. Diabetes Mellitus
Diabetic nephropathy entails a combination of albuminuria, proliferative glomerulonephritis, and glomerular and
tubulointerstitial fibrosis and is the leading cause of endstage renal disease (ESRD) (508, 1517). Patients with type 1
diabetes mellitus (DMT1, insulin deficient) and type 2 diabetes mellitus (DMT2, insulin resistant, usually complicating obesity or metabolic syndrome) begin to develop proteinuria and a decline in GFR after an interval of ⬃12–15 yr
(280, 942). Insulin lack or resistance and high blood glucose and hypertension are commonly associated with oxidative stress and can interact synergistically in the pathogenesis of diabetic nephropathy. Diabetic renal disease is
commonly initiated by an increase in GFR (termed hyperfiltration) and microalbuminuria (i.e., daily albumin excretion of 30 –300 mg), especially in those with DMT1 (149)
and hypertension (29, 30). Usually, the increase in GFR in
DMT1 exceeds that of RBF. Studies in rat and some humans reported that inhibition of the RAS reduces oxidative
stress and slows, but does not halt, the development of
albuminuria and glomerular sclerosis (28, 1121, 1619).
However, other studies have failed to disclose a renal benefit from inhibiting the RAS or lowering the BP in preventing DM (17, 113, 966, 967, 1326). However, ACE inhibitors or AT1 receptor blockers reduce proteinuria and slow
the rate of loss of GFR after the development of overt nephropathy with proteinuria ⬎3 g daily (146, 891, 1157).
Dietary Na⫹ restriction corrected increased SNGFR in a rat
study (66), but exacerbated abnormalities in humans with
early DM (1007). Thus fresh insight is required to develop
novel strategies to slow the progression of DM to ESRD.
1. DMT1
Streptozotocin (STZ) damages pancreatic beta cells and
thereby provides a model of DMT1 with insulinopenia and
hyperglycemia. As demonstrated and discussed in several
studies, there is renal vasodilation, increased PGC, and
SNGFR and glomerular enlargement (145, 189, 638, 1321,
1648). Control of hyperglycemia with insulin normalized
PGC, whereas SNGFR remained elevated (1321). RBF increased by 20 –50% by 2 wk after STZ if hyperglycemia
was restrained by insulin treatment, and this renal vasodilation depended on increased NO (88) but was diminished
by severe hyperglycemia (638). Impaired function of
VOCCs may explain the relatively selective afferent arteriolar vasodilatation and increase in PGC since the efferent
arteriole lack functional VOCCs (1487). Afferent arterioles
from STZ rats were remodeled independent of BP (1497).
Indeed, most models are normotensive.
Studies of renal autoregulation in rat models of DMT1
reported quite variable results. The plateau phase of steadystate RBF autoregulation was maintained in one study
(993), but the limit of autoregulation was shifted to a lower
RPP in many studies (307, 612, 874, 993, 1365, 1472),
independent of PGs (1472). However, other studies reported that RBF autoregulation was impaired within hours
of induction of hyperglycemia by STZ and persisted into the
established phase of DM with attenuated myogenic (588)
and MD-TGF responses (1509). Hyperglycemia itself is
considered important since it can attenuate steady-state autoregulation of RBF and GFR in normal dogs (1624).
Rat and rabbit models of DMT1 were accompanied by
severe endothelial dysfunction of isolated renal afferent arterioles and impaired NO-mediated vasodilation despite an
increased basal RBF (821, 1303, 1575). The reduced NO
availability may result from increased production of ROS
(821, 1121) by renal NADPH oxidase (55). However, these
studies on isolated vessels are at variance with reports of
increased GFR, persistent vasodilation (1472), and increased glomerular blood flow and PGC in intact kidneys, all
of which are attributed to excess NO (309). Thus the NO in
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
467
RENAL AUTOREGULATORY MECHANISMS
question may derive elsewhere from the afferent arteriole,
likely the MD cells.
Increased proximal tubular Na⫹ reabsorption in rats with
early DMT1 via Na⫹-glucose cotransport reduces Na⫹ delivery to the MD and thus reduces the signal for MD-TGF.
An inactivated MD-TGF may contribute to an increased
SNGFR in STZ rats (1509), perhaps by exaggerated MD
generation of NO (1461, 1462, 1509). Thus hyperglycemia
in normal dogs increases their RBF and GFR when MDTGF is active but fails to do so in a nonfiltering kidney
(1624). An increased MD-TGF can occur despite increased
proximal Na⫹ reabsorption and reduced Na⫹ delivery to
the MD (1194). However, other factors must participate in
the increased SNGFR since adenosine A1 receptor-deficient
mice that lack functional MD-TGF (155, 896, 1426) retained a modest (1514), full (1276), or exaggerated increase
in GFR (392) in models with DMT1.
Some studies of frequency analysis of the RBF dynamics
revealed strong contributions to autoregulation by both the
myogenic and the MD-TGF mechanisms that were not impaired by diabetes (89, 874). However, another study reported an attenuated RBF autoregulation with weak contributions of myogenic and MD-TGF responses largely due
to excess NO (88). Diminished afferent arteriolar vasodilatation in diabetic models can occur despite maintained NO
production (1023) probably because of quenching of NO
by enhanced O2⫺ generation (1113, 1303, 1325).
The afferent arteriolar autoregulatory response in the JMN
preparation was normal in rats with STZ-induced DMT1
(993), but attenuated in the HNK preparation where it was
restored by blockade of COX or control of hyperglycemia
by insulin (588). However, insulin has no effect on myogenic constriction of normal afferent arterioles despite increasing NO and reducing vasoconstrictor responses to
ANG II and NE (590). Moreover, acute hyperglycemia in
normal rats dilated posterior cerebral arteries and diminished myogenic tone by an endothelium-dependent mechanism involving NO and a COX-derived PG (263), suggesting different regulation of renal and systemic microvessels
in DMT1. Cremaster arterioles from STZ-induced diabetic
rats initially autoregulated poorly, but eventually reverted
to a normal response (616). Thus the variability of autoregulatory responses may be a function of the degree of hyperglycemia and its effects to increase both NO and ROS, the
experimental preparation and the duration of the disease
process.
Defective L-type VOCC function could contribute to the
weakened myogenic and MD-TGF mechanisms reported in
STZ rats. Afferent arterioles of JMN had a reduced contraction to high KCl-induced membrane depolarization or to
the L-type VOCC agonist BAY K 8644 that was normalized
by reducing the glucose concentration (195). Impaired
468
VOCC function appeared to result from hyperpolarization
of the VSMC plasma membrane due to increased activity of
KATP and KIR channels during oxidative stress and hyperglycemia (665, 1484, 1485).
Galactose feeding to inhibit aldose reductase and polyol
metabolism models DMT1 with a normal insulin level and
a normal afferent arteriolar myogenic response. Thus hyperglycemia can inhibit myogenic tone independent from
insulin lack (426, 427).
Some studies have demonstrated enhanced myogenic responses in nonrenal vessels from diabetic animals. Whereas
myogenic regulation of cerebral cortical blood flow in STZinduced diabetic rats was normal in one study (1267), another reported enhanced cerebral myogenic constrictor
tone with reduced NO and activation of KATP channels
(1664). Femoral and skeletal muscle arterioles of rats with
STZ-induced diabetes had enhanced PKC activity that
phosphorylated p47phox and augmented NADPH oxidase
production of O2⫺ (1501, 1504), which increased speed and
strength by upregulation of VOCC activity (1504). However, one should be cautious in extrapolating these findings
to the kidney since there are clearly differences in the myogenic response and its mechanisms between vascular beds.
Moreover, careful assessment of BP shows it to fall with the
onset of DMT1 or DMT2 and to remain reduced for some
time in most studies.
In summary, the kidneys of rats with early DMT1 generally
are vasodilated with increased RBF and GFR attributed to
excess NO, which is impaired by diminishing hyperglycemia with insulin. In contrast, later stages of DMT1 are
associated with vasoconstriction, exaggerated ROS, and reduced NO. The available evidence does not allow a definitive conclusion as to whether RBF autoregulation and the
myogenic and MD-TGF mechanisms are normal or attenuated. Variability in results likely reflects differences in the
disease models or its progression. Attenuated myogenic or
MD-TGF components could contribute to afferent arteriolar dilation, glomerular hypertension, and increased
SNGFR which some studies have ascribed to attenuated
activity of VOCCs. Nevertheless, several studies of nonrenal vessels have demonstrated an enhanced myogenic response. Hyperglycemia and lack of insulin might exert differential effects on the renal and systemic arterioles. Thus
impaired autoregulation and vasodilation could result from
enhanced proximal tubule fluid reabsorption secondary to
activated Na⫹-glucose cotransport leading to reduced NaCl
delivery to the MD and attenuated MD-TGF. However,
increased SNGFR can occur in models lacking a MD-TGF
response. Regardless of the mechanism, a combination of a
prolonged increase in SNGFR and PGC is ominous since it
predisposes to barotrauma and nephropathy (948). Some of
the variability in reports between STZ studies likely relates
to insulin replacement and prevention of severe hyperglyce-
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
mia. Most studies before 2000 did not use insulin such that
animals lost weight and often suffered from prolonged glucose osmotic diuresis.
2. DMT2
Albuminuria is a marker for risk of CKD in patients with
DMT2 (316). The RAS may be a major mediator of glomerular adaptation, renal inflammation, and renal injury during established diabetic nephropathy with heavy proteinuria since inhibition of the RAS at this stage reduces albuminuria and affords some renal protection (316).
There are several rodent models of DMT2, as discussed
below. The obese Zucker diabetic fat rat (ZDF) fed a modestly enriched fat diet develops insulin resistance and hyperglycemia followed by proteinuria and glomerular injury
(272, 772, 773) related to increased SNGFR either without
(1101) or with increased PGC (1299). Renal ANG II was
elevated in ZDF rats (1440), and inhibition of RAS reduced
systemic and glomerular hypertension, albuminuria, and
renal damage (73, 73, 1100, 1299, 1299). ZDF rats have
increased GFR but, impaired GFR and RBF autoregulation
despite reduced NO production in the renal cortex and
medulla attributed to increased O2⫺ (450, 1440). They are
modestly hypertensive with a resetting of the pressure-natriuresis relation to favor Na⫹ reabsorption at a lower RPP.
Hyperglycemia and oxidative stress can activate PKC and
MAPK signaling (271, 279, 1429), which could phosphorylate Cxs. Indeed, kidneys from ZDF rats displayed increased Cx43 phosphorylation and reduced staining of
Cx37 in JGC (1441). Inhibitory peptides against Cx37 and
Cx40/Cx43, but not Cx43, attenuated RBF autoregulation
in control Zucker rats, but did not further decrement the
weak autoregulation in kidneys of ZDF rats (1441). Similar
changes in Cxs in VSMCs may contribute to abnormal
myogenic and MD-TGF responses during diabetes.
Myogenic constrictor and ACh-induced vasodilation of afferent arterioles of the HNK of ZDF rats were both blunted
but were improved by sensitization to insulin release with
troglitazone (591). Insulin buffered ANG II and NE-induced vasoconstriction in the HNK of normal rats that was
attributed to increased NO, but these effects were attenuated in ZDF rats (591), perhaps reflecting oxidative stress.
A hybrid cross of ZDF with spontaneously hypertensive
heart failure rats developed hypertension, proteinuria, and
glomerulosclerosis (495) despite efficient steady-state RBF
autoregulation (AI ⬍0.1) and dynamic transfer function
with normal myogenic and MD-TGF responses (495). Thus
factors other than impaired renal autoregulation are responsible for kidney damage in this model of DMT2.
The myogenic tone of skeletal muscle arterioles from ZDF
rats was normal initially but became enhanced in older
hypertensive animals by a ROS-dependent mechanism that
diminished NO signaling (444, 885). Cerebral arteries of
ZDF rats also had reduced vasodilation to ACh or hypoxia
and enhanced ROS-dependent myogenic constriction
(1180). Although generation of ROS in the ZDF rat may
underlie enhanced myogenic responses in nonrenal vascular
beds, it is unclear whether this occurs in the renal microcirculation.
Otsuka Long-Evans Tokushima rats develop late-onset,
non-insulin-dependent hyperglycemia, moderate obesity,
mild hypertension, and increased GFR and RBF (1506). An
impaired steady-state autoregulation of total and deep cortical blood flow and MD-TGF preceded the diabetic phase
and persisted into the established phase (584). The PGC in
glomeruli of deep versus superficial nephrons was increased
by 20 mmHg (584). This interesting model with insulin
resistance and later development of DMT2 has a weakened
MD-TGF mechanism that may underlie increased RBF and
PGC, impaired autoregulation, and subsequent renal injury.
Goto-Kakizaki rats are a lean Wistar substrain that develops early-onset DMT2 with hyperglycemia despite normal
insulin levels followed by renal and glomerular hypertrophy, glomerular basement membrane thickening, mesangial
expansion, and albuminuria (1327) despite normal BP and
reduced GFR (784, 1327). Hypertension induced by
DOCA-salt accelerated the development of proteinuria and
nephropathy (234). Thus this is a model of a lean, hyperglycemic, euinsulinemic, normotensive DMT2 with prominent renal functional and structural changes characteristic
of human diabetic nephropathy. The myogenic responses
were maintained or augmented in cerebral and mesenteric
arteries of diabetic Goto-Kakizaki rats (784, 785, 1269),
with normalization by glycemic control (784, 1269). Renal
hemodynamics and autoregulatory mechanisms have not
been characterized in this intriguing model.
The leptin receptor-deficient db⫺/db⫺ mouse model of DMT2
develops dyslipidemia, obesity, hyperglycemia, glomerular hypertrophy, and albuminuria (1355) with increased GFR while
BP is normal (888, 889). MD-TGF activity was normal (888,
889), but after correction for the depleted ECV by the osmatic
diuresis from glucosuria the SNGFR increased further and the
MD-TGF activity was abolished (889). The increase in GFR in
volume-replete db⫺/db⫺ mice was ascribed to overproduction
of NO by nNOS in the MD (889). Afferent arterioles in the
JMN preparation of db⫺/db⫺mice were dilated, but had normal autoregulation to increased RPP and normal contractions
to ANG II (1155), suggesting a normal myogenic response.
MD-TGF was normal in JMN but impaired in superficial
nephrons, which might reflect differences in A1 receptor signaling. Myogenic tone of the preglomerular vasculature in the
absence of MD-TGF has not been investigated to date.
Mixed results have been reported for nonrenal vessels from
db⫺/db⫺ mice. Myogenic tone was increased in mesenteric
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
469
RENAL AUTOREGULATORY MECHANISMS
and coronary arteries but was normalized by inhibition of
epidermal growth factor receptor tyrosine kinase (92, 983)
and was related to activation of a vasoconstrictor prostanoid activating TP receptors and to reduce NO (858). Likewise, myogenic tone of gracilis skeletal muscle arterioles
was enhanced by a constrictor prostanoid produced by
COX-2 that activated TP receptors secondary to increased
H2O2 production (61, 382). However, other studies of coronary arteries reported preserved myogenic tone and normal TP responses (1017) despite reduced NO and increased
ROS production (62).
of MD-TGF, following enhanced proximal reabsorption, may
contribute to afferent arteriolar dilation and increased
SNGFR. Glomerular damage can be dissociated from hypertension and impaired renal autoregulation in some models,
thereby highlighting the importance of other, as yet undiscovered, mechanism of renal damage in DMT2.
3. Clinical studies of renal autoregulation in diabetes
and nephropathy
Christensen and co-workers (251–255) studied GFR from
measurements of inulin clearance in human diabetics before
and after an acute 15–20 mmHg reduction in mean BP
produced by oral clonidine (FIGURE 9). The dotted line in
FIGURE 9 indicates the GFR response without any autoregulation (AI ⫽ 1.0). Normal subjects (FIGURE 9F) had excellent autoregulation with only a 1–2% reduction in GFR
when BP was reduced by 18% (253). GFR autoregulation
In summary, studies in rodent models of hyperglycemic, insulin-resistant DMT2 have generally demonstrated renal vasodilation, increased PGC, and GFR. Renal autoregulation and
the myogenic and MD-TGF mechanisms are often normal or
modestly attenuated. Autoregulation is generally improved by
glycemic control or by an insulin-sensitizing agent. Resetting
A
DMT1 (10 yrs); no nephropathy
ΔMAP (%)
-15
-20
-10
-5
0
B
-20
0
C
DMT1 (27 yrs); no nephropathy
ΔMAP (%)
-15
-10
-5
0
-20
0
-15
-10
-5
-5
Normal blood
glucose
-10
-5
Control
ΔGFR
(%)
-15
-10
Spironolactone
-20
DMT2 (14 yrs); no nephropathy
ΔMAP (%)
-20
-15
-10
-5
0
E
0
-5
ΔGFR
(%)
-15
-10
-5
-10
-15
-15
-20
-20
DMT2 (15 yrs); with or without nephropathy
ΔMAP (%)
-20
0
Control
F
Non-diabetic nephropathy
ΔMAP (%)
-20
0
0
ΔGFR
(%)
-15
-10
-5
0
0
No nephropathy
Irsadipine
-5
Control
-5
-5
Nephropathy
-10
ΔGFR
(%)
-10
ΔGFR
(%)
Nephropathy
-10
-15
-15
-15
-20
-20
-20
FIGURE 9. Data have been redrawn from studies of autoregulation of glomerular filtration rate (GFR) during
short-term reductions in mean arterial pressure (MAP) produced by oral clonidine administration to patients
with diabetes mellitus type 1 (DMT1) or type 2 (DMT2), and/or nephropathy (broken line and open circle) and
appropriate control subjects (solid line and closed circle). Perfect autoregulation (AI ⫽ 0) is equated with no
change in GFR with MAP. An absence of autoregulation (AI ⫽ 1.0) with equal changes in GFR and MAP is
depicted by the dotted lines in each panel. A: patients with DMT1 but without nephropathy at an average of 10
yr after diagnosis. Those with high blood glucose had a similar autoregulation as those with normal blood
glucose (control) (255). B: patients with DMT1 but without nephropathy at an average of 27 yr after diagnosis.
Autoregulation was impaired in subjects with DMT1 (control), but was not altered by mineralocorticosteroid
blockade with spironolactone (1293). C: patients with DMT2 but without nephropathy at an average of 13 yr
after diagnosis. Autoregulation was mildly impaired (control), but was not affected by blockade of AT1 receptors
with cansdesartan (254). D: patients with DMT2 without nephropathy at ⬃14 yr after diagnosis. Autoregulation was quite well maintained (control), but was impaired severely by blockade of L-type VOCCs with isradipine
(251). E: patients with DMT2 at 15 yr after diagnosis. Autoregulation was quite well maintained in the group
without nephropathies (control) but was impaired in those with nephropathy (252). F: autoregulation in normal
subjects was excellent (control), but was impaired in those with nondiabetic nephropathy (253).
470
0
Control
Blood
glucose
D
DMT2 (13 yrs); no nephropathy
ΔMAP (%)
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
ΔGFR
(%)
CARLSTRÖM ET AL.
also was well maintained after ⬃10 yr of DMT1, even with
poor glycemic control (FIGURE 9A) (255), but became impaired after ⬃27 yr and was not improved by reduction in
BP by a mineralocorticoid receptor blocker (FIGURE 9B)
(1293). GFR autoregulation was excellent after ⬃13 yr of
DMT2 and was not affected by AT1 receptor blockade (FIGURE 9C) (254), but was almost abolished by VOCC blockade with isradipine (FIGURE 9D) (251). GFR autoregulation
was impaired severely with DMT2 and nephropathy (FIGURE 9E) (252) and in other patients with nondiabetic nephropathy (FIGURE 9F) (253). These data demonstrate that
GFR autoregulation is well preserved in subjects with early
DMT1 or 2, but becomes quite severely impaired later in
the disease, with the development of nephropathy or after
treatment with a CCB.
VIII. CONCLUSIONS
Autoregulation relates preglomerular vascular resistance to
changes in RPP. It thereby stabilizes the RBF, PGC, and GFR
and protects glomerular capillaries from barotrauma during hypertension. Autoregulation in the normal kidney is
highly efficient. It involves a unique combination of tubular
and vascular control mechanisms (FIGURE 6). The major
mechanisms entail a rapid myogenic response and a more
delayed MD-TGF response within the JGA. In addition,
there appear to be contributions from a poorly characterized third mechanism and modulation by a CT-GF system.
Although the myogenic phenomenon was described more
than a century ago, the precise mechanisms for sensing mechanical distension and promoting the response in VSMCs
are still not fully understood. Increased [Ca2⫹]i in arteriolar
VSMCs following mechanical distension is obligate for the
initiation of the response, but prolonged, maintained vasoconstriction also entails enhanced Ca2⫹ sensitivity involving PKC, RhoA/Rho-kinase, and/or ROS signaling (FIGURE
8). Initiation can be triggered by stretch-sensitive cation
channels, integrin-cytoskeleton interactions, and/or GPCR
signaling. VSMC membrane depolarization and Ca2⫹ entry
via VOCCs are central to Ca2⫹ signaling for the myogenic
and MD-TGF-induced vasoconstrictor components of autoregulation. Impaired renal autoregulation is observed in
many, but not all, models of hypertension, diabetes, or intrinsic renal disease. Further investigations are warranted to
advance our understanding of the underlying cellular are
molecular events in these important basic regulatory mechanisms, since they likely contribute to progression of renal
damage.
ACKNOWLEDGMENTS
Address for reprint requests and other correspondence: M.
Carlström, Dept. of Physiology & Pharmacology, Karolinska Institutet, Nanna Svartz Väg 2, S-171 77, Stockholm,
Sweden (e-mail: mattias.carlstrom@ki.se).
GRANTS
The authors’ work was supported by National Heart, Lung,
and Blood Institute Grants HL68686 (to C. S. Wilcox) and
HL02334 (to W. J. Arendshorst); National Institute of Diabetes and Digestive and Kidney Diseases Grants DK036079 and DK-049870 (to C. S. Wilcox); Swedish Research Council Grant K2012-99X-21971-01-3 (to M. Carlström); Swedish Heart and Lung Foundation Grants
20110589 and 20140448 (to M. Carlström); the Swedish
Society for Medical Research (to M. Carlström); Jeanssons
Foundation Grants JS20130064 and JS2011-0212 (to M.
Carlström); the Wenner-Gren Foundation (to M. Carlström); the Swedish Society of Medicine (to M. Carlström);
Karolinska Institutet Research Foundations Grant
2014fobi41264 (to M. Carlström); the George E. Schreiner
Chair of Nephrology (to C. S. Wilcox); the Hypertension,
Kidney, and Vascular Research Center of Georgetown University (to C. S. Wilcox); the Smith Family Trust Fund (to
C. S. Wilcox); as well as the Joseph Gildenhorn/Spiesman
Family Foundation Incorporated and by Ms. Alma Gilde
(to C. S. Wilcox).
DISCLOSURES
No conflicts of interest, financial or otherwise, are declared
by the authors.
REFERENCES
1. Aalkjaer C, Boedtkjer D, Matchkov V. Vasomotion: what is currently thought? Acta
Physiol 202: 253–269, 2011.
2. Aalkjaer C, Nilsson H. Vasomotion: cellular background for the oscillator and for the
synchronization of smooth muscle cells. Br J Pharmacol 144: 605– 616, 2005.
3. Abboud HE. Mesangial cell biology. Exp Cell Res 318: 979 –985, 2012.
4. Abd El-Rahman RR, Harraz OF, Brett SE, Anfinogenova Y, Mufti RE, Goldman D,
Welsh DG. Identification of L- and T-type Ca2⫹ channels in rat cerebral arteries: role
in myogenic tone development. Am J Physiol Heart Circ Physiol 304: H58 –H71, 2013.
5. Abe Y, Kishimoto T, Yamamoto K. Effect of angiotensin II antagonist infusion on
autoregulation of renal blood flow. Am J Physiol 231: 1267–1271, 1976.
6. Abraham NG, Cao J, Sacerdoti D, Li X, Drummond G. Heme oxygenase: the key to
renal function regulation. Am J Physiol Renal Physiol 297: F1137–F1152, 2009.
7. Abu-Amarah I, Ajikobi DO, Bachelard H, Cupples WA, Salevsky FC. Responses of
mesenteric and renal blood flow dynamics to acute denervation in anesthetized rats.
Am J Physiol Regul Integr Comp Physiol 275: R1543–R1552, 1998.
8. Abu-Amarah I, Bidani AK, Hacioglu R, Williamson GA, Griffin KA. Differential effects
of salt on renal hemodynamics and potential pressure transmission in stroke-prone
and stroke-resistant spontaneously hypertensive rats. Am J Physiol Renal Physiol 289:
F305–F313, 2005.
9. Adams LG, Polzin DJ, Osborne CA, O’Brien TD, Hostetter TH. Influence of dietary
protein/calorie intake on renal morphology and function in cats with 5/6 nephrectomy. Lab Invest 70: 347–357, 1994.
10. Adebiyi A, Thomas-Gatewood CM, Leo MD, Kidd MW, Neeb ZP, Jaggar JH. An
elevation in physical coupling of type 1 inositol 1,4,5-trisphosphate (IP3) receptors to
transient receptor potential 3 (TRPC3) channels constricts mesenteric arteries in
genetic hypertension. Hypertension 60: 1213–1219, 2012.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
471
RENAL AUTOREGULATORY MECHANISMS
11. Adebiyi A, Zhao G, Cheranov SY, Ahmed A, Jaggar JH. Caveolin-1 abolishment
attenuates the myogenic response in murine cerebral arteries. Am J Physiol Heart Circ
Physiol 292: H1584 –H1592, 2007.
31. Anderson S, Meyer TW, Rennke HG, Brenner BM. Control of glomerular hypertension limits glomerular injury in rats with reduced renal mass. J Clin Invest 76: 612–
619, 1985.
12. Adler S, Huang H. Oxidant stress in kidneys of spontaneously hypertensive rats
involves both oxidase overexpression and loss of extracellular superoxide dismutase. Am J Physiol Renal Physiol 287: F907–F913, 2004.
32. Anderson S, Rennke HG, Brenner BM. Antihypertensive therapy must control glomerular hypertension to limit glomerular injury. J Hypertens Suppl 4: S242–S244,
1986.
13. Agodoa LY, Appel L, Bakris GL, Beck G, Bourgoignie J, Briggs JP, Charleston J, Cheek
D, Cleveland W, Douglas JG, Dowie D, Faulkner M, Gabriel A, Gassman J, Greene T,
Hall Y, Hebert L, Hiremath L, Jamerson K, Johnson CJ, Kopple J, Kusek J, Lash J, Lea
J, Lewis JB, Lipkowitz M, Massry S, Middleton J, Miller ER, Norris K, O’Connor D,
Ojo A, Phillips RA, Pogue V, Rahman M, Randall OS, Rostand S, Schulman G, Smith
W, Thornley-Brown D, Tisher CC, Toto RD, Wright JT, Xu S. Effect of ramipril vs
amlodipine on renal outcomes in hypertensive nephrosclerosis: a randomized controlled trial. JAMA 285: 2719 –2728, 2001.
33. Anderson WP, Kett MM, Alcorn D, Bertram JF. Angiotensin II antagonism and preglomerular arterial wall dimensions in the kidney of the spontaneously hypertensive
rat. Clin Exp Hypertens 19: 965–979, 1997.
14. Agrawal S, Temsamani J, Tang JY. Pharmacokinetics, biodistribution, and stability of
oligodeoxynucleotide phosphorothioates in mice. Proc Natl Acad Sci USA 88: 7595–
7599, 1991.
15. Aguilera G, Catt KJ. Regulation of vascular angiotensin II receptors in the rat during
altered sodium intake. Circ Res 49: 751–758, 1981.
16. Ahmed A, Waters CM, Leffler CW, Jaggar JH. Ionic mechanisms mediating the
myogenic response in newborn porcine cerebral arteries. Am J Physiol Heart Circ
Physiol 287: H2061–H2069, 2004.
17. Ahmed AK, Kamath NS, El KM, El Nahas AM. The impact of stopping inhibitors of
the renin-angiotensin system in patients with advanced chronic kidney disease.
Nephrol Dial Transplant 25: 3977–3982, 2010.
18. Ahn DS, Choi SK, Kim YH, Cho YE, Shin HM, Morgan KG, Lee YH. Enhanced
stretch-induced myogenic tone in the basilar artery of spontaneously hypertensive
rats. J Vasc Res 44: 182–191, 2007.
19. Ajikobi DO, Novak P, Salevsky FC, Cupples WA. Pharmacological modulation of
spontaneous renal blood flow dynamics. Can J Physiol Pharmacol 74: 964 –972, 1996.
20. Almeida JB, Saragoca MA, Tavares A, Cezareti ML, Draibe SA, Ramos OL. Severe
hypertension induces disturbances of renal autoregulation. Hypertension 19: II279 –
II283, 1992.
21. Alonso-Galicia M, Drummond HA, Reddy KK, Falck JR, Roman RJ. Inhibition of
20-HETE production contributes to the vascular responses to nitric oxide. Hypertension 29: 320 –325, 1997.
22. Amann K, Wiest G, Zimmer G, Gretz N, Ritz E, Mall G. Reduced capillary density in
the myocardium of uremic rats–a stereological study. Kidney Int 42: 1079 –1085,
1992.
23. Ambudkar IS, Brazer SC, Liu X, Lockwich T, Singh B. Plasma membrane localization
of TRPC channels: role of caveolar lipid rafts. Novartis Found Symp 258: 63–70, 2004.
24. Ances BM, Greenberg JH, Detre JA. Interaction between nitric oxide synthase inhibitor induced oscillations and the activation flow coupling response. Brain Res
1309: 19 –28, 2010.
25. Anderson M, Kim EY, Hagmann H, Benzing T, Dryer SE. Opposing effects of podocin on the gating of podocyte TRPC6 channels evoked by membrane stretch or
diacylglycerol. Am J Physiol Cell Physiol 305: C276 –C289, 2013.
26. Anderson RJ, Pluss RG, Pluss WT, Bell J, Zerbe GG. Effect of hypoxia and hypercapnic acidosis on renal autoregulation in the dog: role of renal nerves. Clin Sci 65:
533–538, 1983.
27. Anderson RJ, Taher MS, Cronin RE, McDonald KM, Schrier RW. Effect of ␤-adrenergic blockade and inhibitors of angiotensin II and prostaglandins on renal autoregulation. Am J Physiol 229: 731–736, 1975.
28. Anderson S. Antihypertensive therapy in experimental diabetes. J Am Soc Nephrol 3:
S86 –S90, 1992.
29. Anderson S, Brenner BM. Intraglomerular hypertension: implications and drug treatment. Annu Rev Med 39: 243–253, 1988.
30. Anderson S, Brenner BM. Pathogenesis of diabetic glomerulopathy: hemodynamic
considerations. Diabetes Metab Rev 4: 163–177, 1988.
472
34. Ando K, Ono A, Sato Y, Fujita T. Involvement of prostaglandins and renal haemodynamics in salt-sensitivity of young spontaneously hypertensive rats. J Hypertens 11:
373–377, 1993.
35. Andreasen D, Friis UG, Uhrenholt TR, Jensen BL, Skott O, Hansen PB. Coexpression of voltage-dependent calcium channels Cav1.2, 21a, and 21b in vascular myocytes. Hypertension 47: 735–741, 2006.
36. Anfinogenova Y, Brett SE, Walsh MP, Harraz OF, Welsh DG. Do TRPC-like currents
and G protein-coupled receptors interact to facilitate myogenic tone development?
Am J Physiol Heart Circ Physiol 301: H1378 –H1388, 2011.
37. Ankorina-Stark I, Haxelmans S, Schlatter E. Receptors for bradykinin and prostaglandin E2 coupled to Ca2⫹ signalling in rat cortical collecting duct. Cell Calcium 22:
269 –275, 1997.
38. Araujo M, Welch WJ. Cyclooxygenase 2 inhibition suppresses tubuloglomerular
feedback: roles of thromboxane receptors and nitric oxide. Am J Physiol Renal Physiol
296: F790 –F794, 2009.
39. Araujo M, Welch WJ. Tubuloglomerular feedback is decreased in COX-1 knockout
mice after chronic angiotensin II infusion. Am J Physiol Renal Physiol 298: F1059 –
F1063, 2010.
40. Araujo M, Wilcox CS. Oxidative stress in hypertension: role of the kidney. Antioxid
Redox Signal 20: 74 –101, 2013.
41. Arendshorst WJ. Autoregulation of renal blood flow in spontaneously hypertensive
rats. Circ Res 44: 344 –349, 1979.
42. Arendshorst WJ, Beierwaltes WH. Renal and nephron hemodynamics in spontaneously hypertensive rats. Am J Physiol Renal Fluid Electrolyte Physiol 236: F246 –F251,
1979.
43. Arendshorst WJ, Beierwaltes WH. Renal tubular reabsorption in spontaneously hypertensive rats. Am J Physiol Renal Fluid Electrolyte Physiol 237: F38 –F47, 1979.
44. Arendshorst WJ, Chatziantoniou C, Daniels FH. Role of angiotensin in the renal
vasoconstriction observed during the development of genetic hypertension. Kidney
Int Suppl 30: S92–S96, 1990.
45. Arendshorst WJ, Finn WF. Renal hemodynamics in the rat before and during inhibition of angiotensin II. Am J Physiol Renal Fluid Electrolyte Physiol 233: F290 –F297,
1977.
46. Arendshorst WJ, Finn WF, Gottschalk CW. Autoregulation of blood flow in the rat
kidney. Am J Physiol 228: 127–133, 1975.
47. Arendshorst WJ, Gottschalk CW. Glomerular ultrafiltration dynamics: euvolemic
and plasma volume-expanded rats. Am J Physiol Renal Fluid Electrolyte Physiol 239:
F171–F186, 1980.
48. Arendshorst WJ, Navar LG. Renal Circulation and Glomerular Hemodynamics. In:
Diseases of the Kidney & Urinary Tract, edited by Schrier RW, Nielsen EH, Molitoris
BA, Coffman , Falk RJ. Philadelphia, PA: Lippincott Williams & Wilkins, 2012.
49. Arendshorst WJ, Thai TL. Regulation of the renal microcirculation by ryanodine
receptors and calcium-induced calcium release. Curr Opin Nephrol Hypertens 18:
40 – 49, 2009.
50. Arensbak B, Mikkelsen HB, Gustafsson F, Christensen T, Holstein-Rathlou NH Holstein-Rathlou NH. Expression of connexin 37, 40, and 43 mRNA and protein in renal
preglomerular arterioles. Histochem Cell Biol 115: 479 – 487, 2001.
51. Arima S, Ito S. Isolated juxtaglomerular apparatus as a tool for exploring glomerular
hemodynamics: application of microperfusion techniques. Exp Nephrol 8: 304 –311,
2000.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
52. Arima S, Ito S, Omata K, Tsunoda K, Yaoita H, Abe K. Diverse effects of calcium
antagonists on glomerular hemodynamics. Kidney Int Suppl 55: S132–S134, 1996.
53. Arima S, Kohagura K, Xu HL, Sugawara A, Abe T, Satoh F, Takeuchi K, Ito S.
Nongenomic vascular action of aldosterone in the glomerular microcirculation. J Am
Soc Nephrol 14: 2255–2263, 2003.
54. Arima S, Kohagura K, Xu HL, Sugawara A, Uruno A, Satoh F, Takeuchi K, Ito S.
Endothelium-derived nitric oxide modulates vascular action of aldosterone in renal
arteriole. Hypertension 43: 352–357, 2004.
55. Asaba K, Tojo A, Onozato ML, Goto A, Quinn MT, Fujita T, Wilcox CS. Effects of
NADPH oxidase inhibitor in diabetic nephropathy. Kidney Int 67: 1890 –1898, 2005.
56. Aukes AM, Vitullo L, Zeeman GG, Cipolla MJ. Pregnancy prevents hypertensive
remodeling and decreases myogenic reactivity in posterior cerebral arteries from
Dahl salt-sensitive rats: a role in eclampsia? Am J Physiol Heart Circ Physiol 292:
H1071–H1076, 2007.
57. Aukland K. Myogenic mechanisms in the kidney. J Hypertens Suppl 7: S71–S76, 1989.
58. Azar S, Johnson MA, Scheinman J, Bruno L, Tobian L. Regulation of glomerular
capillary pressure and filtration rate in young Kyoto hypertensive rats. Clin Sci 56:
203–209, 1979.
59. Baek EB, Kim SJ. Mechanisms of myogenic response: Ca2⫹-dependent and -independent signaling. J Smooth Muscle Res 47: 55– 65, 2011.
60. Bagher P, Segal SS. Regulation of blood flow in the microcirculation: role of conducted vasodilation. Acta Physiol 202: 271–284, 2011.
61. Bagi Z, Erdei N, Toth A, Li W, Hintze TH, Koller A, Kaley G. Type 2 diabetic mice
have increased arteriolar tone and blood pressure: enhanced release of COX-2derived constrictor prostaglandins. Arterioscler Thromb Vasc Biol 25: 1610 –1616,
2005.
62. Bagi Z, Koller A, Kaley G. Superoxide-NO interaction decreases flow- and agonistinduced dilations of coronary arterioles in Type 2 diabetes mellitus. Am J Physiol
Heart Circ Physiol 285: H1404 –H1410, 2003.
63. Bakker EN, Kerkhof CJ, Sipkema P. Signal transduction in spontaneous myogenic
tone in isolated arterioles from rat skeletal muscle. Cardiovasc Res 41: 229 –236,
1999.
64. Balasubramanian L, Ahmed A, Lo CM, Sham JS, Yip KP. Integrin-mediated mechanotransduction in renal vascular smooth muscle cells: activation of calcium sparks.
Am J Physiol Regul Integr Comp Physiol 293: R1586 –R1594, 2007.
65. Balasubramanian L, Lo CM, Sham JS, Yip KP. Remanent cell traction force in renal
vascular smooth muscle cells induced by integrin-mediated mechanotransduction.
Am J Physiol Cell Physiol 304: C382–C391, 2013.
66. Bank N, Lahorra G, Aynedjian HS, Wilkes BM. Sodium restriction corrects hyperfiltration of diabetes. Am J Physiol Renal Fluid Electrolyte Physiol 254: F668 –F676, 1988.
67. Barajas L, Powers K, Carretero OA, Scicli AG, Inagami T. Immunocytochemical
localization of renin and kallikrein in the rat renal cortex. Kidney Int 29: 965–970,
1986.
73. Baylis C, Atzpodien EA, Freshour G, Engels K. Peroxisome proliferator-activated
receptor [gamma] agonist provides superior renal protection versus angiotensinconverting enzyme inhibition in a rat model of type 2 diabetes with obesity. J Pharmacol Exp Ther 307: 854 – 860, 2003.
74. Baylis C, Ichikawa I, Willis WT, Wilson CB, Brenner BM. Dynamics of glomerular
ultrafiltration. IX. Effects of plasma protein concentration. Am J Physiol Renal Fluid
Electrolyte Physiol 232: F58 –F71, 1977.
75. Baylis C, Qiu C, Engels K. Comparison of L-type and mixed L- and T-type calcium
channel blockers on kidney injury caused by deoxycorticosterone-salt hypertension
in rats. Am J Kidney Dis 38: 1292–1297, 2001.
76. Bayliss WM. On the local reactions of the arterial wall to changes of internal pressure.
J Physiol 28: 220 –231, 1902.
77. Beaufils M, Sraer J, Lepreux C, Ardaillou R. Angiotensin II binding to renal glomeruli
from sodium-loaded and sodium-depleted rats. Am J Physiol 231: 1187–1193, 1976.
78. Bedard K, Krause KH. The NOX family of ROS-generating NADPH oxidases: physiology and pathophysiology. Physiol Rev 87: 245–313, 2007.
79. Beierwaltes WH, Arendshorst WJ, Klemmer PJ. Electrolyte and water balance in
young spontaneously hypertensive rats. Hypertension 4: 908 –915, 1982.
80. Beierwaltes WH, Schryver S, Britton SL, Romero JC. Renal pressure-flow relationships in severely hypertensive rabbits. Proc Soc Exp Biol Med 164: 495– 499, 1980.
81. Beierwaltes WH, Sigmon DH, Carretero OA. Endothelium modulates renal blood
flow but not autoregulation. Am J Physiol Renal Fluid Electrolyte Physiol 262: F943–
F949, 1992.
82. Bell PD. Calcium antagonists and intrarenal regulation of glomerular filtration rate.
Am J Nephrol 7 Suppl 1: 24 –31, 1987.
83. Bell PD, Komlosi P, Zhang ZR. ATP as a mediator of macula densa cell signalling.
Purinergic Signal 5: 461– 471, 2009.
84. Bell PD, LaPointe JY, Peti-Peterdi J. Macula densa cell signaling. Annu Rev Physiol 65:
481–500, 2003.
85. Bell PD, LaPointe JY, Sabirov R, Hayashi S, Peti-Peterdi J, Manabe K, Kovacs G,
Okada Y. Macula densa cell signaling involves ATP release through a maxi anion
channel. Proc Natl Acad Sci USA 100: 4322– 4327, 2003.
86. Bell PD, Navar LG. Stop-flow pressure feedback responses during reduced renal
vascular resistance in the dog. Am J Physiol Renal Fluid Electrolyte Physiol 237: F204 –
F209, 1979.
87. Bell PD, Reddington M, Ploth D, Navar LG. Tubuloglomerular feedback-mediated
decreases in glomerular pressure in Munich-Wistar rats. Am J Physiol Renal Fluid
Electrolyte Physiol 247: F877–F880, 1984.
88. Bell TD, DiBona GF, Biemiller R, Brands MW. Continuously measured renal blood
flow does not increase in diabetes if nitric oxide synthesis is blocked. Am J Physiol
Renal Physiol 295: F1449 –F1456, 2008.
68. Bauer J, Parekh N. Variations in cell signaling pathways for different vasoconstrictor
agonists in renal circulation of the rat. Kidney Int 63: 2178 –2186, 2003.
89. Bell TD, DiBona GF, Wang Y, Brands MW. Mechanisms for renal blood flow control
early in diabetes as revealed by chronic flow measurement and transfer function
analysis. J Am Soc Nephrol 17: 2184 –2192, 2006.
69. Baumann JE, Persson PB, Ehmke H, Nafz B, Kirchheim HR. Role of endotheliumderived relaxing factor in renal autoregulation in conscious dogs. Am J Physiol Renal
Fluid Electrolyte Physiol 263: F208 –F213, 1992.
90. Belleau LJ, Earley LE. Autoregulation of renal blood flow in the presence of angiotensin infusion. Am J Physiol 213: 1590 –1595, 1967.
70. Baumann M, Janssen BJ, Rob Hermans JJ, Bartholome R, Smits JF, Struijker-Boudier
HA. Renal medullary effects of transient prehypertensive treatment in young spontaneously hypertensive rats. Acta Physiol 196: 231–237, 2009.
71. Baumgartl H, Zimelka W, Lubbers DW. Effects of puncturing on the measurement of
local oxygen pressure using polarographic microelectrodes. Oxygen Transport Tissue
18: 527–549, 1997.
72. Bax L, Woittiez AJ, Kouwenberg HJ, Mali WP, Buskens E, Beek FJ, Braam B, Huysmans FT, Schultze Kool LJ, Rutten MJ, Doorenbos CJ, Aarts JC, Rabelink TJ, Plouin
PF, Raynaud A, van Montfrans GA, Reekers JA, van den Meiracker AH, Pattynama
PM, van de Ven PJ, Vroegindeweij D, Kroon AA, de Haan MW, Postma CT, Beutler
JJ. Stent placement in patients with atherosclerotic renal artery stenosis and impaired
renal function: a randomized trial. Ann Intern Med 150: 840 – 841, 2009.
91. Bellucci A, Wilkes BM. Mechanism of sodium modulation of glomerular angiotensin
receptors in the rat. J Clin Invest 74: 1593–1600, 1984.
92. Belmadani S, Palen DI, Gonzalez-Villalobos RA, Boulares HA, Matrougui K. Elevated
epidermal growth factor receptor phosphorylation induces resistance artery dysfunction in diabetic db/db mice. Diabetes 57: 1629 –1637, 2008.
93. Benstein JA, Feiner HD, Parker M, Dworkin LD. Superiority of salt restriction over
diuretics in reducing renal hypertrophy and injury in uninephrectomized SHR. Am J
Physiol Renal Fluid Electrolyte Physiol 258: F1675–F1681, 1990.
94. Berecek KH, Schwertschlag U, Gross F. Alterations in renal vascular resistance and
reactivity in spontaneous hypertension of rats. Am J Physiol Heart Circ Physiol 238:
H287–H293, 1980.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
473
RENAL AUTOREGULATORY MECHANISMS
95. Bergstrom G, Johansson I, Wickman A, Gan L, Thorup C. Brief losartan treatment in
young spontaneously hypertensive rats abates long-term blood pressure elevation
by effects on renal vascular structure. J Hypertens 20: 1413–1421, 2002.
116. Blantz RC, Deng A, Lortie M, Munger K, Vallon V, Gabbai FB, Thomson SC. The
complex role of nitric oxide in the regulation of glomerular ultrafiltration. Kidney Int
61: 782–785, 2002.
96. Bergstrom G, Rudenstam J, Taghipour K, Gothberg G, Karlstrom G. Effect of nitric
oxide and renal nerves on renomedullary haemodynamics in SHR and Wistar rats,
studied with laser Doppler technique. Acta Physiol Scand 156: 27–35, 1995.
117. Blantz RC, Gabbai FB, Peterson OW, Thomson SC. Tubuloglomerular feedback
activity after acute reductions in renal mass. Kidney Int Suppl 32: S102–S105, 1991.
97. Berthold H, Munter K, Just A, Kirchheim HR, Ehmke H. Contribution of endothelin
to renal vascular tone and autoregulation in the conscious dog. Am J Physiol Renal
Physiol 276: F417–F424, 1999.
98. Bertuglia S, Colantuoni A, Intaglietta M. Capillary reperfusion after L-arginine, LNMMA, and L-NNA treatment in cheek pouch microvasculature. Microvasc Res 50:
162–174, 1995.
99. Berwick ZC, Moberly SP, Kohr MC, Morrical EB, Kurian MM, Dick GM, Tune JD.
Contribution of voltage-dependent K⫹ and Ca2⫹ channels to coronary pressureflow autoregulation. Basic Res Cardiol 107: 264 – 0264, 2012.
100. Beswick RA, Zhang H, Marable D, Catravas JD, Hill WD, Webb RC. Long-term
antioxidant administration attenuates mineralocorticoid hypertension and renal inflammatory response. Hypertension 37: 781–786, 2001.
101. Bhattachariya A, Dahan D, Turczynska KM, Sward K, Hellstrand P, Albinsson S.
Expression of microRNAs is essential for arterial myogenic tone and pressure-induced activation of the PI3-kinase/Akt pathway. Cardiovasc Res 101: 288 –296, 2014.
102. Bidani AK, Griffin KA. Calcium channel blockers and renal protection: is there an
optimal dose? J Lab Clin Med 125: 553–555, 1995.
103. Bidani AK, Griffin KA. Long-term renal consequences of hypertension for normal
and diseased kidneys. Curr Opin Nephrol Hypertens 11: 73– 80, 2002.
104. Bidani AK, Griffin KA. Pathophysiology of hypertensive renal damage: implications
for therapy. Hypertension 44: 595– 601, 2004.
105. Bidani AK, Griffin KA, Bakris G, Picken MM. Lack of evidence of blood pressureindependent protection by renin-angiotensin system blockade after renal ablation.
Kidney Int 57: 1651–1661, 2000.
106. Bidani AK, Griffin KA, Picken M, Lansky DM. Continuous telemetric blood pressure
monitoring and glomerular injury in the rat remnant kidney model. Am J Physiol Renal
Fluid Electrolyte Physiol 265: F391–F398, 1993.
107. Bidani AK, Griffin KA, Plott W, Schwartz MM. Renal ablation acutely transforms
“benign” hypertension to “malignant” nephrosclerosis in hypertensive rats. Hypertension 24: 309 –316, 1994.
108. Bidani AK, Griffin KA, Williamson G, Wang X, Loutzenhiser R. Protective importance of the myogenic response in the renal circulation. Hypertension 54: 393–398,
2009.
109. Bidani AK, Hacioglu R, Abu-Amarah I, Williamson GA, Loutzenhiser R, Griffin KA.
“Step” vs “dynamic” autoregulation: implications for susceptibility to hypertensive
injury. Am J Physiol Renal Physiol 285: F113–F120, 2003.
110. Bidani AK, Mitchell KD, Schwartz MM, Navar LG, Lewis EJ. Absence of glomerular
injury or nephron loss in a normotensive rat remnant kidney model. Kidney Int 38:
28 –38, 1990.
111. Bidani AK, Polichnowski AJ, Loutzenhiser R, Griffin KA. Renal microvascular dysfunction, hypertension and CKD progression. Curr Opin Nephrol Hypertens 22: 1–9,
2013.
112. Bidani AK, Schwartz MM, Lewis EJ. Renal autoregulation and vulnerability to hypertensive injury in remnant kidney. Am J Physiol Renal Fluid Electrolyte Physiol 252:
F1003–F1010, 1987.
113. Bilous R, Chaturvedi N, Sjolie AK, Fuller J, Klein R, Orchard T, Porta M, Parving HH.
Effect of candesartan on microalbuminuria and albumin excretion rate in diabetes:
three randomized trials. Ann Intern Med 151: 11–14, 2009.
114. Biswas SK, de Faria JB. Which comes first: renal inflammation or oxidative stress in
spontaneously hypertensive rats? Free Radic Res 41: 216 –224, 2007.
115. Black MJ, Wang Y, Bertram JF. Nephron endowment and renal filtration surface area
in young spontaneously hypertensive rats. Kidney Blood Press Res 25: 20 –26, 2002.
474
118. Blantz RC, Konnen KS, Tucker BJ. Glomerular filtration response to elevated ureteral pressure in both the hydropenic and the plasma-expanded rat. Circ Res 37:
819 – 829, 1975.
119. Blaustein MP, Hamlyn JM. Signaling mechanisms that link salt retention to hypertension: endogenous ouabain, the Na⫹ pump, the Na⫹/Ca2⫹ exchanger and TRPC
proteins. Biochim Biophys Acta 1802: 1219 –1229, 2010.
120. Blaustein MP, Lederer WJ. Sodium/calcium exchange: its physiological implications.
Physiol Rev 79: 763– 854, 1999.
121. Bobadilla NA, Tack I, Tapia E, Sanchez-Lozada LG, Santamaria J, Jimenez F, Striker
LJ, Striker GE, Herrera-Acosta J. Pentosan polysulfate prevents glomerular hypertension and structural injury despite persisting hypertension in 5/6 nephrectomy
rats. J Am Soc Nephrol 12: 2080 –2087, 2001.
122. Bock HA, Hermle M, Fiallo A, Osgood RW, Fried TA. Measurement of renin secretion in single perfused rabbit glomeruli. Am J Physiol Renal Fluid Electrolyte Physiol 258:
F1460 –F1465, 1990.
123. Boedtkjer DM, Matchkov VV, Boedtkjer E, Nilsson H, Aalkjaer C. Vasomotion has
chloride-dependency in rat mesenteric small arteries. Pflügers Arch 457: 389 – 404,
2008.
124. Boesen EI, Williams DL, Pollock JS, Pollock DM. Immunosuppression with mycophenolate mofetil attenuates the development of hypertension and albuminuria in
deoxycorticosterone acetate-salt hypertensive rats. Clin Exp Pharmacol Physiol 37:
1016 –1022, 2010.
125. Boettcher M, de Wit C. Distinct endothelium-derived hyperpolarizing factors
emerge in vitro and in vivo and are mediated in part via connexin 40-dependent
myoendothelial coupling. Hypertension 57: 802– 808, 2011.
126. Bogeski I, Kappl R, Kummerow C, Gulaboski R, Hoth M, Niemeyer BA. Redox
regulation of calcium ion channels: chemical and physiological aspects. Cell Calcium
50: 407– 423, 2011.
127. Bohlen HG. Enhanced cerebral vascular regulation occurs by age 4 to 5 weeks in
spontaneously hypertensive rats. Hypertension 9: 325–331, 1987.
128. Bokoch GM, Zhao T. Regulation of the phagocyte NADPH oxidase by Rac GTPase.
Antioxid Redox Signal 8: 1533–1548, 2006.
129. Bolz SS, Vogel L, Sollinger D, Derwand R, de Wit C, Loirand G, Pohl U. Nitric
oxide-induced decrease in calcium sensitivity of resistance arteries is attributable to
activation of the myosin light chain phosphatase and antagonized by the RhoA/Rho
kinase pathway. Circulation 107: 3081–3087, 2003.
130. Botros FT, Dobrowolski L, Navar LG. Renal heme oxygenase-1 induction with
hemin augments renal hemodynamics, renal autoregulation, and excretory function.
Int J Hypertens 2012: 1– 8, 2012.
131. Botros FT, Prieto-Carrasquero MC, Martin VL, Navar LG. Heme oxygenase induction attenuates afferent arteriolar autoregulatory responses. Am J Physiol Renal
Physiol 295: F904 –F911, 2008.
132. Bouriquet N, Casellas D. Interaction between cGMP-dependent dilators and autoregulation in rat preglomerular vasculature. Am J Physiol Renal Fluid Electrolyte Physiol
268: F338 –F346, 1995.
133. Braam B, Koomans HA. Nitric oxide antagonizes the actions of angiotensin II to
enhance tubuloglomerular feedback responsiveness. Kidney Int 48: 1406 –1411,
1995.
134. Braam B, Navar LG, Mitchell KD. Modulation of tubuloglomerular feedback by
angiotensin II type 1 receptors during the development of Goldblatt hypertension.
Hypertension 25: 1232–1237, 1995.
135. Brakemeier S, Eichler I, Knorr A, Fassheber T, Kohler R, Hoyer J. Modulation of
Ca2⫹-activated K⫹ channel in renal artery endothelium in situ by nitric oxide and
reactive oxygen species. Kidney Int 64: 199 –207, 2003.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
136. Brandes RP, Kreuzer J. Vascular NADPH oxidases: molecular mechanisms of activation. Cardiovasc Res 65: 16 –27, 2005.
137. Brannstrom K, Arendshorst WJ. Thromboxane A2 contributes to the enhanced
tubuloglomerular feedback activity in young SHR. Am J Physiol Renal Physiol 276:
F758 –F766, 1999.
138. Brannstrom K, Morsing P, Arendshorst WJ. Exaggerated tubuloglomerular feedback
activity in genetic hypertension is mediated by ANG II and AT1 receptors. Am J
Physiol Renal Fluid Electrolyte Physiol 270: F749 –F755, 1996.
139. Brannstrom K, Morsing P, Arendshorst WJ. Candesartan normalizes exaggerated
tubuloglomerular feedback activity in young spontaneously hypertensive rats. J Am
Soc Nephrol 10: S213–S219, 1999.
140. Braun C, Lang C, Hocher B, van der Woude FJ, Rohmeiss P. Influence of the renal
endothelin A system on the autoregulation of renal hemodynamics in SHRs and WKY
rats. J Cardiovasc Pharmacol 31: 643– 648, 1998.
141. Brayden JE, Earley S, Nelson MT, Reading S. Transient receptor potential (TRP)
channels, vascular tone and autoregulation of cerebral blood flow. Clin Exp Pharmacol
Physiol 35: 1116 –1120, 2008.
142. Brayden JE, Li Y, Tavares MJ. Purinergic receptors regulate myogenic tone in cerebral parenchymal arterioles. J Cereb Blood Flow Metab 33: 293–299, 2013.
143. Brekke JF, Jackson WF, Segal SS. Arteriolar smooth muscle Ca2⫹ dynamics during
blood flow control in hamster cheek pouch. J Appl Physiol 101: 307–315, 2006.
144. Brenner BM. Nephron adaptation to renal injury or ablation. Am J Physiol Renal Fluid
Electrolyte Physiol 249: F324 –F337, 1985.
145. Brenner BM, Anderson S. Glomerular function in diabetes mellitus. Adv Nephrol
Necker Hosp 19: 135–144, 1990.
158. Brown SA, Finco DR, Crowell WA, Choat DC, Navar LG. Single-nephron adaptations to partial renal ablation in the dog. Am J Physiol Renal Fluid Electrolyte Physiol 258:
F495–F503, 1990.
159. Brown SA, Finco DR, Crowell WA, Navar LG. Dietary protein intake and the glomerular adaptations to partial nephrectomy in dogs. J Nutr 121: S125–S127, 1991.
160. Brown SA, Finco DR, Navar LG. Impaired renal autoregulatory ability in dogs with
reduced renal mass. J Am Soc Nephrol 5: 1768 –1774, 1995.
161. Bruno L, Azar S, Weller D. Absence of a pre-hypertensive stage in post-natal Kyoto
hypertensive rats. Jpn Heart J 90 –92, 1997.
162. Buhrle CP, Hackenthal E, Helmchen U, Lackner K, Nobiling R, Steinhausen M,
Taugner R. The hydronephrotic kidney of the mouse as a tool for intravital microscopy and in vitro electrophysiological studies of renin-containing cells. Lab Invest 54:
462– 472, 1986.
163. Buhrle CP, Hackenthal E, Nobiling R, Skott O, Baumbach L, Taugner R. Tachyphylaxis of juxtaglomerular epithelioid cells to angiotensin II. Differences between the
electrical membrane response and renin secretion. Pflügers Arch 410: 55– 62, 1987.
164. Buhrle CP, Nobiling R, Mannek E, Schneider D, Hackenthal E, Taugner R. The
afferent glomerular arteriole: immunocytochemical and electrophysiological investigations. J Cardiovasc Pharmacol 6 Suppl 2: S383–S393, 1984.
165. Bulley S, Neeb ZP, Burris SK, Bannister JP, Thomas-Gatewood CM, Jangsangthong
W, Jaggar JH. TMEM16A channels contribute to the myogenic response in cerebral
arteries. Circ Res 111: 1027–1036, 2012.
166. Bund SJ. Contractility of resistance arteries of spontaneously hypertensive rats related to their media: lumen ratio. J Hypertens 18: 1223–1231, 2000.
167. Bund SJ. Spontaneously hypertensive rat resistance artery structure related to myogenic and mechanical properties. Clin Sci 101: 385–393, 2001.
146. Brenner BM, Cooper ME, de Zeeuw D, Keane WF, Mitch WE, Parving HH, Remuzzi
G, Snapinn SM, Zhang Z, Shahinfar S. Effects of losartan on renal and cardiovascular
outcomes in patients with type 2 diabetes and nephropathy. N Engl J Med 345:
861– 869, 2001.
168. Burke M, Pabbidi M, Fan F, Ge Y, Liu R, Williams JM, Sarkis A, Lazar J, Jacob HJ,
Roman RJ. Genetic basis of the impaired renal myogenic response in FHH rats. Am J
Physiol Renal Physiol 304: F565–F577, 2013.
147. Brenner BM, Lawler EV, Mackenzie HS. The hyperfiltration theory: a paradigm shift
in nephrology. Kidney Int 49: 1774 –1777, 1996.
169. Burke M, Pabbidi MR, Farley J, Roman RJ. Molecular mechanisms of renal blood flow
autoregulation. Curr Vasc Pharmacol 12: 845– 858, 2014.
148. Brenner BM, Troy JL, Daugharty TM. The dynamics of glomerular ultrafiltration in
the rat. J Clin Invest 50: 1776 –1780, 1971.
170. Burke TJ, Duchin KL. Glomerular filtration during furosemide diuresis in the dog.
Kidney Int 16: 672– 680, 1979.
149. Breyer MD, Bottinger E, Brosius FC, Coffman TM, Fogo A, Harris RC, Heilig CW,
Sharma K. Diabetic nephropathy: of mice and men. Adv Chronic Kidney Dis 12:
128 –145, 2005.
171. Busse R, Fleming I. Vascular endothelium and blood flow. Handb Exp Pharmacol 176:
43–78, 2006.
150. Breyer MD, Breyer RM. Prostaglandin E receptors and the kidney. Am J Physiol Renal
Physiol 279: F12–F23, 2000.
151. Breyer MD, Zhang Y, Guan YF, Hao CM, Hebert RL, Breyer RM. Regulation of renal
function by prostaglandin E receptors. Kidney Int Suppl 67: S88 –S94, 1998.
172. Bussemaker E, Wallner C, Fisslthaler B, Fleming I. The Na-K-ATPase is a target for
an EDHF displaying characteristics similar to potassium ions in the porcine renal
interlobar artery. Br J Pharmacol 137: 647– 654, 2002.
173. Byrom FB. Angiotensin and renal vascular damage. Br J Exp Pathol 45: 7–12, 1964.
152. Briet M, Schiffrin EL. Aldosterone: effects on the kidney and cardiovascular system.
Nat Rev Nephrol 6: 261–273, 2010.
174. Cachofeiro V, Nasjletti A. Increased vascular responsiveness to bradykinin in kidneys
of spontaneously hypertensive rats. Effect of N-nitro-L-arginine. Hypertension 18:
683– 688, 1991.
153. Broegger T, Jacobsen JC, Secher D, V, Boedtkjer DM, Kold-Petersen H, Pedersen
FS, Aalkjaer C, Matchkov VV. Bestrophin is important for the rhythmic but not the
tonic contraction in rat mesenteric small arteries. Cardiovasc Res 91: 685– 693, 2011.
175. Callera GE, Touyz RM, Teixeira SA, Muscara MN, Carvalho MH, Fortes ZB, Nigro
D, Schiffrin EL, Tostes RC. ETA receptor blockade decreases vascular superoxide
generation in DOCA-salt hypertension. Hypertension 42: 811– 817, 2003.
154. Brookes ZL, Kaufman S. Myogenic responses and compliance of mesenteric and
splenic vasculature in the rat. Am J Physiol Regul Integr Comp Physiol 284: R1604 –
R1610, 2003.
176. Campean V, Theilig F, Paliege A, Breyer M, Bachmann S. Key enzymes for renal
prostaglandin synthesis: site-specific expression in rodent kidney (rat, mouse). Am J
Physiol Renal Physiol 285: F19 –F32, 2003.
155. Brown R, Ollerstam A, Johansson B, Skott O, Gebre-Medhin S, Fredholm B, Persson
AE. Abolished tubuloglomerular feedback and increased plasma renin in adenosine
A1 receptor-deficient mice. Am J Physiol Regul Integr Comp Physiol 281: R1362–
R1367, 2001.
177. Campese VM, Parise M, Karubian F, Bigazzi R. Abnormal renal hemodynamics in
black salt-sensitive patients with hypertension. Hypertension 18: 805– 812, 1991.
156. Brown RD, Hilliard LM, Head GA, Jones ES, Widdop RE, Denton KM. Sex differences in the pressor and tubuloglomerular feedback response to angiotensin II.
Hypertension 59: 129 –135, 2012.
157. Brown SA, Brown CA. Single-nephron adaptations to partial renal ablation in cats.
Am J Physiol Regul Integr Comp Physiol 269: R1102–R1108, 1995.
178. Cancela JM, Gerasimenko OV, Gerasimenko JV, Tepikin AV, Petersen OH. Two
different but converging messenger pathways to intracellular Ca2⫹ release: the roles
of nicotinic acid adenine dinucleotide phosphate, cyclic ADP-ribose and inositol
trisphosphate. EMBO J 19: 2549 –2557, 2000.
179. Carlson BE, Secomb TW. A theoretical model for the myogenic response based on
the length-tension characteristics of vascular smooth muscle. Microcirculation 12:
327–338, 2005.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
475
RENAL AUTOREGULATORY MECHANISMS
180. Carlstrom M, Brown RD, Sallstrom J, Larsson E, Zilmer M, Zabihi S, Eriksson UJ,
Persson AE. SOD1-deficiency causes salt-sensitivity and aggravates hypertension in
hydronephrosis. Am J Physiol Regul Integr Comp Physiol 297: R82–R92, 2009.
200. Casellas D, Moore LC. Autoregulation and tubuloglomerular feedback in juxtamedullary glomerular arterioles. Am J Physiol Renal Fluid Electrolyte Physiol 258: F660 –
F669, 1990.
181. Carlstrom M, Brown RD, Yang T, Hezel M, Larsson E, Scheffer PG, Teerlink T,
Lundberg JO, Persson AE. L-Arginine or tempol supplementation improves renal and
cardiovascular function in rats with reduced renal mass and chronic high salt intake.
Acta Physiol 207: 732–741, 2013.
201. Casellas D, Moore LC. Autoregulation of intravascular pressure in preglomerular
juxtamedullary vessels. Am J Physiol Renal Fluid Electrolyte Physiol 264: F315–F321,
1993.
182. Carlstrom M, Lai EY, Ma Z, Patzak A, Brown RD, Persson AE. Role of NOX2 in the
regulation of afferent arteriole responsiveness. Am J Physiol Regul Integr Comp Physiol
296: R72–R79, 2009.
183. Carlstrom M, Lai EY, Ma Z, Steege A, Patzak A, Eriksson UJ, Lundberg JO, Wilcox
CS, Persson AE. Superoxide dismutase 1 limits renal microvascular remodeling and
attenuates arteriole and blood pressure responses to angiotensin II via modulation of
nitric oxide bioavailability. Hypertension 56: 907–913, 2010.
184. Carlstrom M, Persson AE. Important role of NAD(P)H oxidase 2 in the regulation of
the tubuloglomerular feedback. Hypertension 53: 456 – 457, 2009.
185. Carlstrom M, Persson AE, Larsson E, Hezel M, Scheffer PG, Teerlink T, Weitzberg
E, Lundberg JO. Dietary nitrate attenuates oxidative stress, prevents cardiac and
renal injuries, and reduces blood pressure in salt-induced hypertension. Cardiovasc
Res 89: 574 –585, 2011.
186. Carlstrom M, Sallstrom J, Skott O, Larsson E, Persson AE. Uninephrectomy in young
age or chronic salt loading causes salt-sensitive hypertension in adult rats. Hypertension 49: 1342–1350, 2007.
187. Carlstrom M, Wilcox CS, Welch WJ. Adenosine A2 receptors modulate the tubuloglomerular feedback. Am J Physiol Renal Physiol 299: F412–F417, 2010.
188. Carlstrom M, Wilcox CS, Welch WJ. Adenosine A2A receptor activation attenuates
tubuloglomerular feedback responses by stimulation of endothelial nitric oxide synthase. Am J Physiol Renal Physiol 300: F457–F464, 2011.
189. Carmines PK. The renal vascular response to diabetes. Curr Opin Nephrol Hypertens
19: 85–90, 2010.
190. Carmines PK, Bell PD, Roman RJ, Work J, Navar LG. Prostaglandins in the sodium
excretory response to altered renal arterial pressure in dogs. Am J Physiol Renal Fluid
Electrolyte Physiol 248: F8 –F14, 1985.
191. Carmines PK, Fowler BC, Bell PD. Segmentally distinct effects of depolarization on
intracellular [Ca2⫹] in renal arterioles. Am J Physiol Renal Fluid Electrolyte Physiol 265:
F677–F685, 1993.
192. Carmines PK, Inscho EW. Perfusate composition modulates in vitro renal microvascular pressure responsiveness in a segment-specific manner. Microvasc Res 43: 347–
351, 1992.
193. Carmines PK, Inscho EW, Gensure RC. Arterial pressure effects on preglomerular
microvasculature of juxtamedullary nephrons. Am J Physiol Renal Fluid Electrolyte
Physiol 258: F94 –F102, 1990.
194. Carmines PK, Mitchell KD, Navar LG. Effects of calcium antagonists on renal hemodynamics and glomerular function. Kidney Int 41: 43– 48, 1992.
202. Casellas D, Navar LG. In vitro perfusion of juxtamedullary nephrons in rats. Am J
Physiol Renal Fluid Electrolyte Physiol 246: F349 –F358, 1984.
203. Castrop H. Mediators of tubuloglomerular feedback regulation of glomerular filtration: ATP and adenosine. Acta Physiol 189: 3–14, 2007.
204. Castrop H, Hocherl K, Kurtz A, Schweda F, Todorov V, Wagner C. Physiology of
kidney renin. Physiol Rev 90: 607– 673, 2010.
205. Castrop H, Huang Y, Hashimoto S, Mizel D, Hansen P, Theilig F, Bachmann S, Deng
C, Briggs J, Schnermann J. Impairment of tubuloglomerular feedback regulation of
GFR in ecto-5’-nucleotidase/CD73-deficient mice. J Clin Invest 114: 634 – 642, 2004.
206. Cervenka L, Horácek V, Vanecková I, Hubácek JA, Oliverio MI, Coffman TM, Navar
LG. Essential role of AT1A receptor in the development of 2K1C hypertension.
Hypertension 40: 735–741, 2002.
207. Cervenka L, Kramer HJ, Malý J, Heller J. Role of nNOS in regulation of renal function
in angiotensin II-induced hypertension. Hypertension 38: 280 –285, 2001.
208. Cervenka L, Vaneckova I, Huskova Z, Vanourkova Z, Erbanova M, Thumova M,
Skaroupkova P, Opocensky M, Maly J, Chabova VC, Tesar V, Burgelova M, Viklicky
O, Teplan V, Zelizko M, Kramer HJ, Navar LG. Pivotal role of angiotensin II receptor
subtype 1A in the development of two-kidney, one-clip hypertension: study in angiotensin II receptor subtype 1A knockout mice. J Hypertens 26: 1379 –1389, 2008.
209. Cervenka L, Wang CT, Mitchell KD, Navar LG. Proximal tubular angiotensin II levels
and renal functional responses to AT1 receptor blockade in nonclipped kidneys of
Goldblatt hypertensive rats. Hypertension 33: 102–107, 1999.
210. Cervenka L, Wang CT, Navar LG. Effects of acute AT1 receptor blockade by candesartan on arterial pressure and renal function in rats. Am J Physiol Renal Physiol 274:
F940 –F945, 1998.
211. Chabrashvili T, Kitiyakara C, Blau J, Karber A, Aslam S, Welch WJ, Wilcox CS. Effects
of ANG II type 1 and 2 receptors on oxidative stress, renal NADPH oxidase and SOD
expression. Am J Physiol Regul Integr Comp Physiol 285: R117–R124, 2003.
212. Chabrashvili T, Tojo A, Onozato M, Kitiyakara C, Quinn MT, Fujita T, Welch WJ,
Wilcox CS. Expression and cellular localization of classic NADPH oxidase subunits in
the spontaneously hypertensive rat kidney. Hypertension 39: 269 –274, 2002.
213. Champion D, Simatos D, Kalogianni EP, Cayot P, Le MM. Ascorbic acid oxidation in
sucrose aqueous model systems at subzero temperatures. J Agric Food Chem 52:
3399 –3404, 2004.
214. Chan WL, Holstein-Rathlou NH, Yip KP. Integrin mobilizes intracellular Ca2⫹ in
renal vascular smooth muscle cells. Am J Physiol Cell Physiol 280: C593–C603, 2001.
195. Carmines PK, Ohishi K, Ikenaga H. Functional impairment of renal afferent arteriolar
voltage-gated calcium channels in rats with diabetes mellitus. J Clin Invest 98: 2564 –
2571, 1996.
215. Chaston DJ, Baillie BK, Grayson TH, Courjaret RJ, Heisler JM, Lau KA, Machaca K,
Nicholson BJ, Ashton A, Matthaei KI, Hill CE. Polymorphism in endothelial connexin40 enhances sensitivity to intraluminal pressure and increases arterial stiffness.
Arterioscler Thromb Vasc Biol 33: 962–970, 2013.
196. Carnevale D, Lembo G. PI3Kgamma in hypertension: a novel therapeutic target
controlling vascular myogenic tone and target organ damage. Cardiovasc Res 95:
403– 408, 2012.
216. Chatziantoniou C, Arendshorst WJ. Angiotensin and thromboxane in genetically
hypertensive rats: renal blood flow and receptor studies. Am J Physiol Renal Fluid
Electrolyte Physiol 261: F238 –F247, 1991.
197. Carnevale D, Vecchione C, Mascio G, Esposito G, Cifelli G, Martinello K, Landolfi A,
Selvetella G, Grieco P, Damato A, Franco E, Haase H, Maffei A, Ciraolo E, Fucile S,
Frati G, Mazzoni O, Hirsch E, Lembo G. PI3Kgamma inhibition reduces blood pressure by a vasorelaxant Akt/L-type calcium channel mechanism. Cardiovasc Res 93:
200 –209, 2012.
217. Chatziantoniou C, Arendshorst WJ. Impaired ability of prostaglandins to buffer renal
vasoconstriction in genetically hypertensive rats. Am J Physiol Renal Fluid Electrolyte
Physiol 263: F573–F580, 1992.
198. Casellas D, Bouriquet N, Herizi A. Bosentan prevents preglomerular alterations
during angiotensin II hypertension. Hypertension 30: 1613–1620, 1997.
199. Casellas D, Carmines PK, Navar LG. Microvascular reactivity of in vitro blood perfused juxtamedullary nephrons from rats. Kidney Int 28: 752–759, 1985.
476
218. Chatziantoniou C, Arendshorst WJ. Prostaglandin interactions with angiotensin, norepinephrine, and thromboxane in rat renal vasculature. Am J Physiol Renal Fluid
Electrolyte Physiol 262: F68 –F76, 1992.
219. Chatziantoniou C, Arendshorst WJ. Renal vascular reactivity to vasodilator prostaglandins in genetically hypertensive rats. Am J Physiol Renal Fluid Electrolyte Physiol
262: F124 –F130, 1992.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
220. Chatziantoniou C, Arendshorst WJ. Angiotensin receptor sites in renal vasculature of
rats developing genetic hypertension. Am J Physiol Renal Fluid Electrolyte Physiol 265:
F853–F862, 1993.
240. Chilton L, Smirnov SV, Loutzenhiser K, Wang X, Loutzenhiser RD. Segment-specific
differences in the inward rectifier K⫹ current along the renal interlobular artery.
Cardiovasc Res 92: 169 –177, 2011.
221. Chatziantoniou C, Arendshorst WJ. Vascular interactions of prostaglandins with
thromboxane in kidneys of rats developing hypertension. Am J Physiol Renal Fluid
Electrolyte Physiol 265: F250 –F256, 1993.
241. Chin SY, Pandey KN, Shi SJ, Kobori H, Moreno C, Navar LG. Increased activity and
expression of Ca2⫹-dependent NOS in renal cortex of ANG II-infused hypertensive
rats. Am J Physiol Renal Physiol 277: F797–F804, 1999.
222. Chatziantoniou C, Daniels FH, Arendshorst WJ. Exaggerated renal vascular reactivity to angiotensin and thromboxane in young genetically hypertensive rats. Am J
Physiol Renal Fluid Electrolyte Physiol 259: F372–F382, 1990.
242. Chin SY, Wang CT, Majid DSA, Navar LG. Renoprotective effects of nitric oxide in
angiotensin II-induced hypertension in the rat. Am J Physiol Renal Physiol 274: F876 –
F882, 1998.
223. Chatziantoniou C, Ruan X, Arendshorst WJ. Impaired ability of prostaglandins to
buffer renal vasoconstriction in genetically hypertensive rats. Am J Physiol Renal Fluid
Electrolyte Physiol 263: F573–F580, 1992.
243. Chlopicki S, Nilsson H, Mulvany MJ. Initial and sustained phases of myogenic response of rat mesenteric small arteries. Am J Physiol Heart Circ Physiol 281: H2176 –
H2183, 2001.
224. Chatziantoniou C, Ruan X, Arendshorst WJ. Interactions of cAMP-mediated vasodilators with angiotensin II in rat kidney during hypertension. Am J Physiol Renal Fluid
Electrolyte Physiol 265: F845–F852, 1993.
244. Chon KH, Chen YM, Holstein-Rathlou NH, Marmarelis VZ. Nonlinear system analysis of renal autoregulation in normotensive and hypertensive rats. IEEE Trans
Biomed Eng 45: 342–353, 1998.
225. Chatziantoniou C, Ruan X, Arendshorst WJ. Defective G protein activation of the
cAMP pathway in rat kidney during genetic hypertension. Proc Natl Acad Sci USA 92:
2924 –2928, 1995.
245. Chon KH, Chen YM, Marmarelis VZ, Marsh DJ, Holstein-Rathlou NH. Detection of
interactions between myogenic and TGF mechanisms using nonlinear analysis. Am J
Physiol Renal Fluid Electrolyte Physiol 267: F160 –F173, 1994.
226. Chen CC, Geurts AM, Jacob HJ, Fan F, Roman RJ. Heterozygous knockout of transforming growth factor-beta1 protects Dahl S rats against high salt-induced renal
injury. Physiol Genomics 45: 110 –118, 2013.
246. Chon KH, Raghavan R, Chen YM, Marsh DJ, Yip KP. Interactions of TGF-dependent
and myogenic oscillations in tubular pressure. Am J Physiol Renal Physiol 288: F298 –
F307, 2005.
227. Chen DD, Dong YG, Yuan H, Chen AF. Endothelin 1 activation of endothelin A
receptor/NADPH oxidase pathway and diminished antioxidants critically contribute
to endothelial progenitor cell reduction and dysfunction in salt-sensitive hypertension. Hypertension 59: 1037–1043, 2012.
247. Chon KH, Zhong Y, Moore LC, Holstein-Rathlou NH, Cupples WA. Analysis of
nonstationarity in renal autoregulation mechanisms using time-varying transfer and
coherence functions. Am J Physiol Regul Integr Comp Physiol 295: R821–R828, 2008.
228. Chen J, Sgouralis I, Moore LC, Layton HE, Layton AT. A mathematical model of the
myogenic response to systolic pressure in the afferent arteriole. Am J Physiol Renal
Physiol 300: F669 –F681, 2011.
229. Chen PY, St. John PL, Kirk KA, Abrahamson DR, Sanders PW. Hypertensive nephrosclerosis in the Dahl/Rapp rat Initial sites of injury and effect of dietary L-arginine
supplementation. Lab Invest 68: 174 –184, 1993.
230. Chen YM, Holstein-Rathlou NH. Differences in dynamic autoregulation of renal
blood flow between SHR and WKY rats. Am J Physiol Renal Fluid Electrolyte Physiol
264: F166 –F174, 1993.
231. Chen YM, Yip KP, Marsh DJ, Holstein-Rathlou NH. Magnitude of TGF-initiated
nephron-nephron interactions is increased in SHR. Am J Physiol Renal Fluid Electrolyte
Physiol 269: F198 –F204, 1995.
232. Cheng HF, Zhang MZ, Harris RC. Nitric oxide stimulates cyclooxygenase-2 in cultured cTAL cells through a p38-dependent pathway. Am J Physiol Renal Physiol 290:
F1391–F1397, 2006.
2⫹
233. Cheng X, Jaggar JH. Genetic ablation of caveolin-1 modifies Ca spark coupling in
murine arterial smooth muscle cells. Am J Physiol Heart Circ Physiol 290: H2309 –
H2319, 2006.
234. Cheng ZJ, Vaskonen T, Tikkanen I, Nurminen K, Ruskoaho H, Vapaatalo H, Muller
D, Park JK, Luft FC, Mervaala EM. Endothelial dysfunction and salt-sensitive hypertension in spontaneously diabetic Goto-Kakizaki rats. Hypertension 37: 433– 439,
2001.
235. Chevalier RL, Carey RM, Kaiser DL. Endogenous prostaglandins modulate autoregulation of renal blood flow in young rats. Am J Physiol Renal Fluid Electrolyte Physiol
253: F66 –F75, 1987.
236. Chevalier RL, Kaiser DL. Autoregulation of renal blood flow in the rat: effects of
growth and uninephrectomy. Am J Physiol Renal Fluid Electrolyte Physiol 244: F483–
F487, 1983.
237. Chevalier RL, Kaiser DL. Effects of acute uninephrectomy and age on renal blood
flow autoregulation in the rat. Am J Physiol Renal Fluid Electrolyte Physiol 249: F672–
F679, 1985.
248. Chou CL, Marsh DJ. Role of proximal convoluted tubule in pressure diuresis in the
rat. Am J Physiol Renal Fluid Electrolyte Physiol 251: F283–F289, 1986.
249. Chou CL, Marsh DJ. Time course of proximal tubule response to acute arterial
hypertension in the rat. Am J Physiol Renal Fluid Electrolyte Physiol 254: F601–F607,
1988.
250. Christensen HR, Nielsen H, Christensen KL, Baandrup U, Jespersen LT, Mulvany MJ.
Long-term hypotensive effects of an angiotensin converting enzyme inhibitor in
spontaneously hypertensive rats: is there a role for vascular structure? J Hypertens 6:
S27–S31, 1988.
251. Christensen PK, Akram K, Konig KB, Parving HH. Autoregulation of glomerular
filtration rate in patients with type 2 diabetes during isradipine therapy. Diabetes Care
26: 156 –162, 2003.
252. Christensen PK, Hansen HP, Parving HH. Impaired autoregulation of GFR in hypertensive non-insulin dependent diabetic patients. Kidney Int 52: 1369 –1374, 1997.
253. Christensen PK, Hommel EE, Clausen P, Feldt-Rasmussen B, Parving HH. Impaired
autoregulation of the glomerular filtration rate in patients with nondiabetic nephropathies. Kidney Int 56: 1517–1523, 1999.
254. Christensen PK, Lund S, Parving HH. Autoregulated glomerular filtration rate during
candesartan treatment in hypertensive type 2 diabetic patients. Kidney Int 60: 1435–
1442, 2001.
255. Christensen PK, Lund S, Parving HH. The impact of glycaemic control on autoregulation of glomerular filtration rate in patients with non-insulin dependent diabetes.
Scand J Clin Lab Invest 61: 43–50, 2001.
256. Chrysant SG, Walsh GM, Frohlich ED, Townsend SM. Hemodynamic changes induced by prolonged NaCl and DOCA administration in spontaneously hypertensive
rats. Angiology 29: 303–309, 1978.
257. Chung WS, Farley JM, Drummond HA. ASIC-like currents in freshly isolated cerebral
artery smooth muscle cells are inhibited by endogenous oxidase activity. Cell Physiol
Biochem 27: 129 –138, 2011.
258. Chung WS, Weissman JL, Farley J, Drummond HA. betaENaC is required for whole
cell mechanically gated currents in renal vascular smooth muscle cells. Am J Physiol
Renal Physiol 304: F1428 –F1437, 2013.
238. Chilton L, Loutzenhiser K, Morales E, Breaks J, Kargacin GJ, Loutzenhiser RD.
Inward rectifier K⫹ currents Kir2.1 expression in renal afferent and efferent arterioles. J Am Soc Nephrol 19: 69 –76, 2008.
259. Churchill PC, Churchill MC, Bidani AK. Kidney cross transplants in Dahl salt-sensitive
and salt-resistant rats. Am J Physiol Heart Circ Physiol 262: H1809 –H1817, 1992.
239. Chilton L, Loutzenhiser RD. Functional evidence for an inward rectifier potassium
current in rat renal afferent arterioles. Circ Res 88: 152–158, 2001.
260. Churchill PC, Churchill MC, Bidani AK, Griffin KA, Picken M, Pravenec M, Kren V, St
Lezin E, Wang JM, Wang N, Kurtz TW. Genetic susceptibility to hypertension-
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
477
RENAL AUTOREGULATORY MECHANISMS
induced renal damage in the rat. Evidence based on kidney-specific genome transfer.
J Clin Invest 100: 1373–1382, 1997.
283. Cupples WA. Renal medullary blood flow: Its measurement and physiology. Can J
Physiol Pharmacol 64: 873– 870, 1985.
261. Churchill PC, Churchill MC, Griffin KA, Picken M, Webb RC, Kurtz TW, Bidani AK.
Increased genetic susceptibility to renal damage in the stroke-prone spontaneously
hypertensive rat. Kidney Int 61: 1794 –1800, 2002.
284. Cupples WA. Angiotensin II conditions the slow component of autoregulation of
renal blood flow. Am J Physiol Renal Fluid Electrolyte Physiol 264: F515–F522, 1993.
262. Cipolla MJ, Gokina NI, Osol G. Pressure-induced actin polymerization in vascular
smooth muscle as a mechanism underlying myogenic behavior. FASEB J 16: 72–76,
2002.
263. Cipolla MJ, Porter JM, Osol G. High glucose concentrations dilate cerebral arteries
and diminish myogenic tone through an endothelial mechanism. Stroke 28: 405– 410,
1997.
264. Cipolla MJ, Sweet JG, Chan SL, Tavares MJ, Gokina NI, Brayden JE. Increased pressure-induced tone in rat parenchymal arterioles vs. middle cerebral arteries: role of
ion channels and calcium sensitivity. J Appl Physiol 117: 53–59, 2014.
265. Clausen G, Oien AH, Aukland K. Myogenic vasoconstriction in the rat kidney elicited
by reducing perirenal pressure. Acta Physiol Scand 144: 277–290, 1992.
266. Coats P, Johnston F, MacDonald J, McMurray JJ, Hillier C. Signalling mechanisms
underlying the myogenic response in human subcutaneous resistance arteries. Cardiovasc Res 49: 828 – 837, 2001.
267. Cohen HJ, Marsh DJ, Kayser B. Autoregulation in vasa recta of the rat kidney. Am J
Physiol Renal Fluid Electrolyte Physiol 245: F32–F40, 1983.
268. Colantuoni A, Bertuglia S, Intaglietta M. Quantitation of rhythmic diameter changes
in arterial microcirculation. Am J Physiol Heart Circ Physiol 246: H508 –H517, 1984.
269. Conger JD, Burke TJ. Effects of anesthetic agents on autoregulation of renal hemodynamics in the rat and dog. Am J Physiol 230: 652– 657, 1976.
270. Conrad KP, Brinck-Johnsen T, Gellai M, Valtin H. Renal autoregulation in chronically
catheterized conscious rats. Am J Physiol Renal Fluid Electrolyte Physiol 247: F229 –
F233, 1984.
271. Cooper ME. Interaction of metabolic and haemodynamic factors in mediating experimental diabetic nephropathy. Diabetologia 44: 1957–1972, 2001.
272. Corsetti JP, Sparks JD, Peterson RG, Smith RL, Sparks CE. Effect of dietary fat on the
development of non-insulin dependent diabetes mellitus in obese Zucker diabetic
fatty male and female rats. Atherosclerosis 148: 231–241, 2000.
273. Cowley AW Jr. Renal medullary oxidative stress, pressure-natriuresis, and hypertension. Hypertension 52: 777–786, 2008.
274. Cowley AW Jr, Mattson DL, Lu S, Roman RJ. The renal medulla and hypertension.
Hypertension 25: 663– 673, 1995.
275. Cowley AW Jr, Mori T, Mattson D, Zou AP. Role of renal NO production in the
regulation of medullary blood flow. Am J Physiol Regul Integr Comp Physiol 284:
R1355–R1369, 2003.
276. Cowley AW Jr, Roman RJ. Abnormal pressure-diuresis-natriuresis response in spontaneously hypertensive rats. Am J Physiol Renal Fluid Electrolyte Physiol 248: F199 –
F205, 1985.
277. Cowley AW Jr, Roman RJ, Fenoy FJ, Mattson DL. Effect of renal medullary circulation
on arterial pressure. J Hypertens 10: S187–S193, 1992.
278. Cowley AW Jr, Roman RJ, Kaldunski ML, Dumas P, Dickhout JG, Greene AS, Jacob
HJ. Brown Norway chromosome 13 confers protection from high salt to consomic
Dahl S rat. Hypertension 37: 456 – 461, 2001.
279. Craven PA, Studer RK, Negrete H, DeRubertis FR. Protein kinase C in diabetic
nephropathy. J Diabetes Complications 9: 241–245, 1995.
280. Creager MA, Luscher TF, Cosentino F, Beckman JA. Diabetes and vascular disease:
pathophysiology, clinical consequences, and medical therapy: Part I. Circulation 108:
1527–1532, 2003.
281. Crowley SD, Coffman TM. In hypertension, the kidney rules. Curr Hypertens Rep 9:
148 –153, 2007.
282. Crowley SD, Zhang J, Herrera M, Griffiths RC, Ruiz P, Coffman TM. Role of AT1
receptor-mediated salt retention in angiotensin II-dependent hypertension. Am J
Physiol Renal Physiol 301: F1124 –F1130, 2011.
478
285. Cupples WA. Interactions contributing to kidney blood flow autoregulation. Curr
Opin Nephrol Hypertens 16: 39 – 45, 2007.
286. Cupples WA, Braam B. Assessment of renal autoregulation. Am J Physiol Renal Physiol
292: F1105–F1123, 2007.
287. Cupples WA, Loutzenhiser RD. Dynamic autoregulation in the in vitro perfused
hydronephrotic rat kidney. Am J Physiol Renal Physiol 275: F126 –F130, 1998.
288. Cupples WA, Marsh DJ. Autoregulation of blood flow in renal medulla of the rat: no
role for angiotensin II. Can J Physiol Pharmacol 66: 833– 836, 1988.
289. Cupples WA, Novak P, Novak V, Salevsky FC. Spontaneous blood pressure fluctuations and renal blood flow dynamics. Am J Physiol Renal Fluid Electrolyte Physiol 270:
F82–F89, 1996.
290. D’Angelo G, Davis MJ, Meininger GA. Calcium and mechanotransduction of the
myogenic response. Am J Physiol Heart Circ Physiol 273: H175–H182, 1997.
291. D’Angelo G, Mogford JE, Davis GE, Davis MJ, Meininger GA. Integrin-mediated
reduction in vascular smooth muscle [Ca2⫹]i induced by RGD-containing peptide.
Am J Physiol Heart Circ Physiol 272: H2065–H2070, 1997.
292. Dahl LK, Heine M, Thompson K. Genetic influence of the kidneys on blood pressure.
Evidence from chronic renal homografts in rats with opposite predispositions to
hypertension. Circ Res 34: 94 –101, 1974.
293. Dahly AJ, Hoagland KM, Flasch AK, Jha S, Ledbetter SR, Roman RJ. Antihypertensive
effects of chronic anti-TGF-beta antibody therapy in Dahl S rats. Am J Physiol Regul
Integr Comp Physiol 283: R757–R767, 2002.
294. Dalle-Donne I, Aldini G, Carini M, Colombo R, Rossi R, Milzani A. Protein carbonylation, cellular dysfunction, and disease progression. J Cell Mol Med 10: 389 – 406,
2006.
295. Dam VS, Boedtkjer DM, Nyvad J, Aalkjaer C, Matchkov V. TMEM16A knockdown
abrogates two different Ca-activated Cl currents and contractility of smooth muscle
in rat mesenteric small arteries. Pflügers Arch 466: 1391–1409, 2013.
296. Daniels FH, Arendshorst WJ. Tubuloglomerular feedback kinetics in spontaneously
hypertensive and Wistar-Kyoto rats. Am J Physiol Renal Fluid Electrolyte Physiol 259:
F529 –F534, 1990.
297. Daniels FH, Arendshorst WJ, Roberds RG. Tubuloglomerular feedback and autoregulation in spontaneously hypertensive rats. Am J Physiol Renal Fluid Electrolyte Physiol
258: F1479 –F1489, 1990.
298. Dautzenberg M, Just A. Temporal characteristics of nitric oxide-, prostaglandin-, and
EDHF-mediated components of endothelium-dependent vasodilation in the kidney.
Am J Physiol Regul Integr Comp Physiol 305: R987–R998, 2013.
299. Dautzenberg M, Keilhoff G, Just A. Modulation of the myogenic response in renal
blood flow autoregulation by NO depends on endothelial nitric oxide synthase
(eNOS), but not neuronal or inducible NOS. J Physiol 589: 4731– 4744, 2011.
300. Davis MJ. Perspective: physiological role(s) of the vascular myogenic response. Microcirculation 19: 99 –114, 2012.
301. Davis MJ, Hill MA. Signaling mechanisms underlying the vascular myogenic response.
Physiol Rev 79: 387– 423, 1999.
302. Davis MJ, Hill MA, Kuo L. Local regulation of microvascular perfusion. In: Handbook
of Physiology: Microcirculation. Boston: Academic, 2008.
303. Davis MJ, Sikes PJ. Myogenic responses of isolated arterioles: test for a rate-sensitive
mechanism. Am J Physiol Heart Circ Physiol 259: H1890 –H1900, 1990.
304. Davis MJ, Wu X, Nurkiewicz TR, Kawasaki J, Davis GE, Hill MA, Meininger GA.
Integrins and mechanotransduction of the vascular myogenic response. Am J Physiol
Heart Circ Physiol 280: H1427–H1433, 2001.
305. Davis MJ, Wu X, Nurkiewicz TR, Kawasaki J, Gui P, Hill MA, Wilson E. Regulation of
ion channels by integrins. Cell Biochem Biophys 36: 41– 66, 2002.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
306. De Leon H, Gauquelin G, Thibault G, Garcia R. Characterization of receptors for the
atrial natriuretic factor in rat renal microvessels. J Hypertens 11: 499 –508, 1993.
329. DiBona GF, Sawin LL. Effect of endogenous angiotensin II on the frequency response
of the renal vasculature. Am J Physiol Renal Physiol 287: F1171–F1178, 2004.
307. De Micheli AG, Forster H, Duncan RC, Epstein M. A quantitative assessment of renal
blood flow autoregulation in experimental diabetes. Nephron 68: 245–251, 1994.
330. Dietrich A, Kalwa H, Gudermann T. TRPC channels in vascular cell function. Thromb
Haemost 103: 262–270, 2010.
308. De Nicola L, Blantz RC, Gabbai FB. Renal functional reserve in treated and untreated
hypertensive rats. Kidney Int 40: 406 – 412, 1991.
331. Dietrich A, Kalwa H, Storch U, Mederos y Schnitzler M, Salanova B, Pinkenburg O,
Dubrovska G, Essin K, Gollasch M, Birnbaumer L, Gudermann T Mederos y Schnitzler M, Salanova B, Pinkenburg O, Dubrovska G, Essin K, Gollasch M, Birnbaumer L,
Gudermann T. Pressure-induced and store-operated cation influx in vascular
smooth muscle cells is independent of TRPC1. Pflügers Arch 455: 465– 477, 2007.
309. De Nicola L, Blantz RC, Gabbai FB. Renal functional reserve in the early stage of
experimental diabetes. Diabetes 41: 267–273, 1992.
310. De Nicola L, Keiser JA, Blantz RC, Gabbai FB. Angiotensin II and renal functional
reserve in rats with Goldblatt hypertension. Hypertension 19: 790 –794, 1992.
311. De Richelieu LT, Sorensen CM, Holstein-Rathlou NH, Salomonsson M. NO-independent mechanism mediates tempol-induced renal vasodilation in SHR. Am J Physiol
Renal Physiol 289: F1227–F1234, 2005.
312. De Souza Md, Bouskela E. Arteriolar diameter and spontaneous vasomotion: importance of potassium channels and nitric oxide. Microvasc Res 90: 121–127, 2013.
313. De Wit C, Boettcher M, Schmidt VJ. Signaling across myoendothelial gap junctions–
fact or fiction? Cell Commun Adhes 15: 231–245, 2008.
314. De Wit C, Griffith TM. Connexins and gap junctions in the EDHF phenomenon and
conducted vasomotor responses. Pflügers Arch 459: 897–914, 2010.
315. De Wit C, Wolfle SE. EDHF and gap junctions: important regulators of vascular tone
within the microcirculation. Curr Pharm Biotechnol 8: 11–25, 2007.
316. De Zeeuw D, Remuzzi G, Parving HH, Keane W, Zhang Z, Shahinfar S, Snapinn S,
Cooper ME, Mitch WE, Brenner BM. Proteinuria, a target for renoprotection in
patients with type 2 diaetic nephropathy:lessons from RENAAL. Kidney Int 65:
2309 –2320, 2004.
317. DeClue JW, Guyton AC, Cowley AW Jr, Coleman TG, Norman RA Jr, McCaa RE.
Subpressor angiotensin infusion, renal sodium handling, and salt-induced hypertension in the dog. Circ Res 43: 503–512, 1978.
318. Dedeoglu IO, Springate JE. Effect of nitric oxide inhibition on kidney function in
experimental renovascular hypertension. Clin Exp Hypertens 23: 267–275, 2001.
319. Deen WM, Maddox DA, Robertson CR, Brenner BM. Dynamics of glomerular ultrafiltration in the rat. VII. Response to reduced renal mass. Am J Physiol 227: 556 –
562, 1974.
320. Deen WM, Troy JL, Robertson CR, Brenner BM. Dynamics of glomerular ultrafiltration in the rat. IV. Determination of the ultrafiltration coefficient. J Clin Invest 52:
1500 –1508, 1973.
321. Dehmel B, Mervaala E, Lippoldt A, Gross V, Bohlender J, Ganten D, Luft FC. Pressure-natriuresis and -diuresis in transgenic rats harboring both human renin and
human angiotensinogen genes. J Am Soc Nephrol 9: 2212–2222, 1998.
322. Delaney PJ, Burnham MP, Heagerty AM, Izzard AS. Impaired myogenic properties of
cerebral arteries from the Brown Norway rat. J Hypertens 30: 926 –931, 2012.
323. Denton KM, Anderson WP. Glomerular ultrafiltration in rabbits with superficial
glomeruli. Pflügers Arch 419: 235–242, 1991.
324. Dessy C, Matsuda N, Hulvershorn J, Sougnez CL, Sellke FW, Morgan KG. Evidence
for involvement of the PKC-alpha isoform in myogenic contractions of the coronary
microcirculation. Am J Physiol Heart Circ Physiol 279: H916 –H923, 2000.
325. Di Nicolantonio R, Silvapulle MJ. Blood pressure, salt appetite and mortality of
genetically hypertensive and normotensive rats maintained on high and low salt diets
from weaning. Clin Exp Pharmacol Physiol 15: 741–751, 1988.
326. Di Nicolantonio R, Spargo S, Morgan TO. Prenatal high salt diet increases blood
pressure and salt retention in the spontaneously hypertensive rat. Clin Exp Pharmacol
Physiol 14: 233–235, 1987.
327. Dibb KM, Graham HK, Venetucci LA, Eisner DA, Trafford AW. Analysis of cellular
calcium fluxes in cardiac muscle to understand calcium homeostasis in the heart. Cell
Calcium 42: 503–512, 2007.
328. DiBona GF, Kopp UC. Neural control of renal function. Physiol Rev 77: 75–197, 1997.
332. Dietrich A, Mederos YS, Gollasch M, Gross V, Storch U, Dubrovska G, Obst M,
Yildirim E, Salanova B, Kalwa H, Essin K, Pinkenburg O, Luft FC, Gudermann T,
Birnbaumer L. Increased vascular smooth muscle contractility in TRPC6⫺/⫺ mice.
Mol Cell Biol 25: 6980 – 6989, 2005.
333. Dilley JR, Arendshorst WJ. Enhanced tubuloglomerular feedback activity in rats developing spontaneous hypertension. Am J Physiol Renal Fluid Electrolyte Physiol 247:
F672–F679, 1984.
334. Dilley JR, Corradi A, Arendshorst WJ. Glomerular ultrafiltration dynamics during
increased renal venous pressure. Am J Physiol Renal Fluid Electrolyte Physiol 244:
F650 –F658, 1983.
335. Dilley JR, Stier CT Jr, Arendshorst WJ. Abnormalities in glomerular function in rats
developing spontaneous hypertension. Am J Physiol Renal Fluid Electrolyte Physiol 246:
F12–F20, 1984.
336. Dirnagl U, Lindauer U, Villringer A. Nitric oxide synthase blockade enhances vasomotion in the cerebral microcirculation of anesthetized rats. Microvasc Res 45: 318 –
323, 1993.
337. Diz DI, Nasjletti A, Baer PG. Renal denervation at weaning retards development of
hypertension in New Zealand genetically hypertensive rats. Hypertension 4: 361–
368, 1982.
338. Dopico AM, Kirber MT, Singer JJ, Walsh JV Jr. Membrane stretch directly activates
large conductance Ca2⫹-activated K⫹ channels in mesenteric artery smooth muscle
cells. Am J Hypertens 7: 82– 89, 1994.
339. Dorup J, Morsing P, Rasch R. Tubule-tubule and tubule-arteriole contacts in rat
kidney distal nephrons. A morphologic study based on computer-assisted threedimensional reconstructions. Lab Invest 67: 761–769, 1992.
340. Dos Santos EA, Dahly-Vernon AJ, Hoagland KM, Roman RJ. Inhibition of the formation of EETs and 20-HETE with 1-aminobenzotriazole attenuates pressure natriuresis. Am J Physiol Regul Integr Comp Physiol 287: R58 –R68, 2004.
341. Drenckhahn D, Schnittler H, Nobiling R, Kriz W. Ultrastructural organization of
contractile proteins in rat glomerular mesangial cells. Am J Pathol 137: 1343–1351,
1990.
342. Dresser TP, Lynch RE, Schneider EG, Knox FG. Effect of increases in blood pressure
on pressure and reabsorption in the proximal tubule. Am J Physiol 220: 444 – 447,
1971.
343. Drummond HA. B-ENaC is a molecular component of a VSMC mechanotransducer
that contributes to renal blood flow regulation, protection from renal injury, and
hypertension. Front Physiol 3: 1–11, 2012.
344. Drummond HA, Gebremedhin D, Harder DR. Degenerin/epithelial Na⫹ channel
proteins: components of a vascular mechanosensor. Hypertension 44: 643– 648,
2004.
345. Drummond HA, Grifoni SC, Abu-Zaid A, Gousset M, Chiposi R, Barnard JM, Murphey B, Stec DE. Renal inflammation and elevated blood pressure in a mouse model
of reduced ␤-ENaC. Am J Physiol Renal Physiol 301: F443–F449, 2011.
346. Drummond HA, Grifoni SC, Jernigan NL. A new trick for an old dogma: ENaC
proteins as mechanotransducers in vascular smooth muscle. Physiology 23: 23–31,
2008.
347. Drumond MC, Deen WM. Analysis of pulsatile pressures and flows in glomerular
filtration. Am J Physiol Renal Fluid Electrolyte Physiol 261: F409 –F419, 1991.
348. Dryer SE, Reiser J. TRPC6 channels and their binding partners in podocytes: role in
glomerular filtration and pathophysiology. Am J Physiol Renal Physiol 299: F689 –F701,
2010.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
479
RENAL AUTOREGULATORY MECHANISMS
349. Du W, Stiber JA, Rosenberg PB, Meissner G, Eu JP. Ryanodine receptors in muscarinic receptor-mediated bronchoconstriction. J Biol Chem 280: 26287–26294, 2005.
369. Earley S, Resta TC, Walker BR. Disruption of smooth muscle gap junctions attenuates myogenic vasoconstriction of mesenteric resistance arteries. Am J Physiol Heart
Circ Physiol 287: H2677–H2686, 2004.
350. Dubroca C, Loyer X, Retailleau K, Loirand G, Pacaud P, Feron O, Balligand JL, Levy
BI, Heymes C, Henrion D. RhoA activation and interaction with Caveolin-1 are
critical for pressure-induced myogenic tone in rat mesenteric resistance arteries.
Cardiovasc Res 73: 190 –197, 2007.
370. Earley S, Straub SV, Brayden JE. Protein kinase C regulates vascular myogenic tone
through activation of TRPM4. Am J Physiol Heart Circ Physiol 292: H2613–H2622,
2007.
351. Duchin KL, Peterson LN, Burke TJ. Effect of furosemide on renal autoregulation.
Kidney Int 12: 379 –386, 1977.
371. Earley S, Waldron BJ, Brayden JE. Critical role for transient receptor potential channel TRPM4 in myogenic constriction of cerebral arteries. Circ Res 95: 922–929, 2004.
352. Dukacz SA, Adams MA, Kline RL. Short- and long-term enalapril affect renal medullary hemodynamics in the spontaneously hypertensive rat. Am J Physiol Regul Integr
Comp Physiol 276: R10 –R16, 1999.
372. Eckel J, Lavin PJ, Finch EA, Mukerji N, Burch J, Gbadegesin R, Wu G, Bowling B, Byrd
A, Hall G, Sparks M, Zhang ZS, Homstad A, Barisoni L, Birbaumer L, Rosenberg P,
Winn MP. TRPC6 enhances angiotensin II-induced albuminuria. J Am Soc Nephrol 22:
526 –535, 2011.
353. Dukacz SA, Feng MG, Yang LF, Lee RM, Kline RL. Abnormal renal medullary response to angiotensin II in SHR is corrected by long-term enalapril treatment. Am J
Physiol Regul Integr Comp Physiol 280: R1076 –R1084, 2001.
354. Dukacz SA, Kline RL. Differing effects of enalapril and losartan on renal medullary
blood flow and renal interstitial hydrostatic pressure in spontaneously hypertensive
rats. J Hypertens 17: 1345–1352, 1999.
355. Duke LM, Paull JR, Widdop RE. Cardiovascular status following combined angiotensin-converting enzyme and AT1 receptor inhibition in conscious spontaneously hypertensive rats. Clin Exp Pharmacol Physiol 30: 317–323, 2003.
356. Durao V, Martins Prata M, Pires Goncalves LM. Modification of antihypertensive
effect of b-adrenoceptor-blocking agents by inhibition of endogenous prostaglandin
synthesis. Lancet 12: 1005–1007, 1977.
357. Dworkin LD, Benstein JA, Tolbert E, Feiner HD. Salt restriction inhibits renal growth
and stabilizes injury in rats with established renal disease. J Am Soc Nephrol 7: 437–
442, 1996.
358. Dworkin LD, Feiner HD. Glomerular injury in uninephrectomized spontaneously
hypertensive rats. A consequence of glomerular capillary hypertension. J Clin Invest
77: 797– 809, 1986.
359. Dworkin LD, Feiner HD, Parker M, Randazzo J. Calcium carbonate exacerbates
glomerular capillary hypertension and injury in rats with desoxycorticosterone-salt
hypertension. Am J Hypertens 3: 444 – 450, 1990.
360. Dworkin LD, Feiner HD, Parker M, Tolbert E. Effects of nifedipine and enalapril on
glomerular structure and function in uninephrectomized SHR. Kidney Int 39: 1112–
1117, 1991.
361. Dworkin LD, Feiner HD, Randazzo J. Glomerular hypertension and injury in desoxycorticosterone-salt rats on antihypertensive therapy. Kidney Int 31: 718 –724,
1987.
362. Dworkin LD, Grosser M, Feiner HD, Ullian M, Parker M. Renal vascular effects of
antihypertensive therapy in uninephrectomized SHR. Kidney Int 35: 790 –798, 1989.
363. Dworkin LD, Hostetter TH, Rennke HG, Brenner BM. Hemodynamic basis for
glomerular injury in rats with desoxycorticosterone-salt hypertension. J Clin Invest
73: 1448 –1461, 1984.
364. Dworkin LD, Levin RI, Benstein JA, Parker M, Ullian ME, Kim Y, Feiner HD. Effects
of nifedipine and enalapril on glomerular injury in rats with deoxycorticosterone-salt
hypertension. Am J Physiol Renal Fluid Electrolyte Physiol 259: F598 –F604, 1990.
365. Dworkin LD, Tolbert E, Recht PA, Hersch JC, Feiner H, Levin RI. Effects of amlodipine on glomerular filtration, growth, and injury in experimental hypertension.
Hypertension 27: 245–250, 1996.
366. Earley S, Brayden JE. Transient receptor potential channels and vascular function.
Clin Sci 119: 19 –36, 2010.
367. Earley S, Pauyo T, Drapp R, Tavares MJ, Liedtke W, Brayden JE. TRPV4-dependent
dilation of peripheral resistance arteries influences arterial pressure. Am J Physiol
Heart Circ Physiol 297: H1096 –H1102, 2009.
368. Earley S, Reading S, Brayden JE. Functional significance of transient receptor potential channels in vascular function. In: TRP Ion Channel Function in Sensory Transduction
and Cellular Signaling Cascades., edited by Liedtke WB, Heller S. Boca Raton, FL:
CRC, 2007.
480
373. Edgley AJ, Kett MM, Anderson WP. Evidence for renal vascular remodeling in angiotensin II-induced hypertension. J Hypertens 21: 1401–1406, 2003.
374. Edwards A, Layton AT. Calcium dynamics underlying the afferent arteriole myogenic
response. Am J Physiol Renal Physiol 306: F34 –F48, 2014.
375. Edwards RM. Segmental effects of norepinephrine and angiotensin II on isolated
renal microvessels. Am J Physiol Renal Fluid Electrolyte Physiol 244: F526 –F534, 1983.
376. Ehmke H, Persson PB, Seyfarth M, Kirchheim HR. Neurogenic control of pressure
natriuresis in conscious dogs. Am J Physiol Renal Fluid Electrolyte Physiol 259: F466 –
F473, 1990.
377. Ekelund U, Bjornberg J, Grande PO, Albert U, Mellander S. Myogenic vascular
regulation in skeletal muscle in vivo is not dependent of endothelium-derived nitric
oxide. Acta Physiol Scand 144: 199 –207, 1992.
378. Elger M, Drenckhahn D, Nobiling R, Mundel P, Kriz W. Cultured rat mesangial cells
contain smooth muscle alpha-actin not found in vivo. Am J Pathol 142: 497–509,
1993.
379. Eliahou HE, Cohen D, Hellberg B, Ben-David A, Herzog D, Shechter P, Kapuler S,
Kogan N. Effect of the calcium channel blocker nisoldipine on the progression of
chronic renal failure in man. Am J Nephrol 8: 285–290, 1988.
380. Enouri S, Monteith G, Johnson R. Characteristics of myogenic reactivity in isolated
rat mesenteric veins. Am J Physiol Regul Integr Comp Physiol 300: R470 –R478, 2011.
381. Eppel GA, Bergstrom G, Anderson WP, Evans RG. Autoregulation of renal medullary
blood flow in rabbits. Am J Physiol Regul Integr Comp Physiol 284: R233–R244, 2003.
382. Erdei N, Bagi Z, Edes I, Kaley G, Koller A. H2O2 increases production of constrictor
prostaglandins in smooth muscle leading to enhanced arteriolar tone in Type 2
diabetic mice. Am J Physiol Heart Circ Physiol 292: H649 –H656, 2007.
383. Eskildsen-Helmond YE, Mulvany MJ. Pressure-induced activation of extracellular
signal-regulated kinase 1/2 in small arteries. Hypertension 41: 891– 897, 2003.
384. Eu JP, Xu L, Stamler JS, Meissner G. Regulation of ryanodine receptors by reactive
nitrogen species. Biochem Pharmacol 57: 1079 –1084, 1999.
385. Evans RG, Szenasi G, Anderson WP. Effects of NG-nitro-L-arginine on pressure
natriuresis in anaesthetized rabbits. Clin Exp Pharmacol Physiol 22: 94 –101, 1995.
386. Evenwel RT, Kasbergen CM, Struyker-Boudier HA. Central and regional hemodynamics and plasma volume distribution during the development of spontaneous
hypertension in rats. Clin Exp Hypertens A 5: 1511–1536, 1983.
387. Facemire CS, Mohler PJ, Arendshorst WJ. Expression and relative abundance of
short transient receptor potential channels in the rat renal microcirculation. Am J
Physiol Renal Physiol 286: F546 –F551, 2004.
388. Fahr T. Handbuch der speziellen pathologischen Anatomie und Histologie. Berlin:
Springer, 1934.
389. Falcone JC, Granger HJ, Meininger GA. Enhanced myogenic activation in skeletal
muscle arterioles from spontaneously hypertensive rats. Am J Physiol Heart Circ
Physiol 265: H1847–H1855, 1993.
390. Fallet RW, Bast JP, Fujiwara K, Ishii N, Sansom SC, Carmines PK. Influence of
Ca2⫹-activated K⫹ channels on rat renal arteriolar responses to depolarizing agonists. Am J Physiol Renal Physiol 280: F583–F591, 2001.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
391. Farrugia E, Lockhart JC, Larson TS. Relation between vasa recta blood flow and renal
interstitial hydrostatic pressure during pressure natriuresis. Circ Res 71: 1153–1158,
1992.
413. Fenoy FJ, Milicic I, Mistry M, Mecca TE, Roman RJ. Effect of clentiazem on arterial
pressure and renal function in normotensive and hypertensive rats. J Pharmacol Exp
Ther 261: 470 – 475, 1992.
392. Faulhaber-Walter R, Chen L, Oppermann M, Kim SM, Huang Y, Hiramatsu N, Mizel
D, Kajiyama H, Zerfas P, Briggs JP, Kopp JB, Schnermann J. Lack of A1 adenosine
receptors augments diabetic hyperfiltration and glomerular injury. J Am Soc Nephrol
19: 722–730, 2008.
414. Fenoy FJ, Milicic I, Smith RD, Wong PC, Timmermans PB, Roman RJ. Effects of DuP
753 on renal function of normotensive and spontaneously hypertensive rats. Am J
Hypertens 4: 321S–326S, 1991.
393. Feld LG, Cachero S, Van Liew JB, Zamlauski-Tucker M, Noble B. Enalapril and renal
injury in spontaneously hypertensive rats. Hypertension 16: 544 –554, 1990.
394. Feld LG, Zamlauski-Tucker MJ, Springate J, Van Liew JB. Single nephron hemodynamics in spontaneously hypertensive rats. Proc Soc Exp Biol Med 209: 185–189,
1995.
395. Feldberg R, Colding-Jorgensen M, Holstein-Rathlou NH. Analysis of interaction between TGF and the myogenic response in renal blood flow autoregulation. Am J
Physiol Renal Fluid Electrolyte Physiol 269: F581–F593, 1995.
396. Fellner RC, Cook AK, O’Connor PM, Zhang S, Pollock DM, Inscho EW. High salt diet
blunts renal autoregulation by a reactive oxygen species-dependent mechanism. Am
J Physiol Renal Physiol 307: F33–F40, 2014.
397. Fellner SK, Arendshorst WJ. Capacitative calcium entry in smooth muscle cells from
preglomerular vessels. Am J Physiol Renal Physiol 277: F533–F542, 1999.
398. Fellner SK, Arendshorst WJ. Ryanodine receptor and capacitative Ca2⫹ entry in fresh
preglomerular vascular smooth muscle cells. Kidney Int 58: 1686 –1694, 2000.
399. Fellner SK, Arendshorst WJ. Angiotensin II, reactive oxygen species, and Ca2⫹ signaling in afferent arterioles. Am J Physiol Renal Physiol 289: F1012–F1019, 2005.
400. Fellner SK, Arendshorst WJ. Endothelin-A and -B receptors, superoxide, and Ca2⫹
signaling in afferent arterioles. Am J Physiol Renal Physiol 292: F175–F184, 2007.
401. Fellner SK, Arendshorst WJ. Voltage-gated Ca2⫹ entry and ryanodine receptor
Ca2⫹-induced Ca2⫹ release in preglomerular arterioles. Am J Physiol Renal Physiol
292: F1568 –F1572, 2007.
402. Fellner SK, Arendshorst WJ. Angiotensin II-stimulated Ca2⫹ entry mechanisms in
afferent arterioles: role of transient receptor potential canonical channels and reverse Na⫹/Ca2⫹ exchange. Am J Physiol Renal Physiol 294: F212–F219, 2008.
403. Fellner SK, Arendshorst WJ. Complex interactions of NO/cGMP/PKG systems on
Ca2⫹ signaling in afferent arteriolar vascular smooth muscle. Am J Physiol Heart Circ
Physiol 298: H144 –H151, 2010.
404. Feng JJ, Arendshorst WJ. Enhanced renal vasoconstriction induced by vasopressin in
SHR is mediated by V1 receptors. Am J Physiol Renal Fluid Electrolyte Physiol 271:
F304 –F313, 1996.
405. Feng L, Siu K, Moore LC, Marsh DJ, Chon KH. A robust method for detection of
linear and nonlinear interactions: application to renal blood flow dynamics. Ann
Biomed Eng 34: 339 –353, 2006.
406. Feng MG, Dukacz SAW, Kline RL. Selective effect of tempol on renal medullary
hemodynamics in spontaneously hypertensive rats. Am J Physiol Regul Integr Comp
Physiol 281: R1420 –R1425, 2001.
407. Feng MG, Li M, Navar LG. T-type calcium channels in the regulation of afferent and
efferent arterioles in rats. Am J Physiol Renal Physiol 286: F331–F337, 2004.
2⫹
408. Feng MG, Navar LG. Nitric oxide synthase inhibition activates L- and T-type Ca
channels in afferent and efferent arterioles. Am J Physiol Renal Physiol 290: F873–
F879, 2006.
409. Feng MG, Navar LG. Adenosine A2 receptor activation attenuates afferent arteriolar
autoregulation during adenosine receptor saturation in rats. Hypertension 50: 744 –
749, 2007.
410. Fenger-Gron J, Mulvany MJ, Christensen KL. Mesenteric blood pressure profile of
conscious, freely moving rats. J Physiol 488: 753–760, 1995.
411. Fenoy FJ, Ferrer P, Carbonell L, Garcia-Salom M. Role of nitric oxide on papillary
blood flow and pressure natriuresis. Hypertension 25: 408 – 414, 1995.
412. Fenoy FJ, Kauker ML, Milicic I, Roman RJ. Normalization of pressure-natriuresis by
nisoldipine in spontaneously hypertensive rats. Hypertension 19: 49 –55, 1992.
415. Fenoy FJ, Roman RJ. Effect of volume expansion on papillary blood flow and sodium
excretion. Am J Physiol Renal Fluid Electrolyte Physiol 260: F813–F822, 1991.
416. Fenoy FJ, St Lezin E, Kurtz TW, Roman RJ. Genetic heterogeneity and differences in
glomerular hemodynamics between inbred colonies of Munich-Wistar rats. J Am Soc
Nephrol 3: 66 –72, 1992.
417. Ferrone RA, Antonaccio MJ. Prevention of the development of spontaneous hypertension in rats by captopril (SQ 14,225). Eur J Pharmacol 60: 131–137, 1979.
418. Ferrone RA, Heran CL, Antonaccio MJ. Comparison of the acute and chronic hemodynamic effects of captopril and guanethidine in spontaneously hypertensive rats.
Clin Exp Hypertens 2: 247–272, 1980.
419. Figueroa XF, Duling BR. Dissection of two Cx37-independent conducted vasodilator mechanisms by deletion of Cx40: electrotonic versus regenerative conduction.
Am J Physiol Heart Circ Physiol 295: H2001–H2007, 2008.
420. Finco DR, Cooper TL. Soy protein increases glomerular filtration rate in dogs with
normal or reduced renal function. J Nutr 130: 745–748, 2000.
421. Finke R, Gross R, Hackenthal E, Huber J, Kirchheim HR. Threshold pressure for the
pressure-dependent renin release in the autoregulating kidney of conscious dogs.
Pflügers Arch 399: 102–110, 1983.
422. Finn WF, Arendshorst WJ. Effect of prostaglandin synthetase inhibitors on renal
blood flow in the rat. Am J Physiol 231: 1541–1545, 1976.
423. Flemming B, Arenz N, Seeliger E, Wronski T, Steer K, Persson PB. Time-dependent
autoregulation of renal blood flow in conscious rats. J Am Soc Nephrol 12: 2253–2262,
2001.
424. Fliser D, Schroter M, Neubeck M, Ritz E. Coadministration of thiazides increases the
efficacy of loop diuretics even in patients with advanced renal failure. Kidney Int 46:
482– 488, 1994.
425. Folkow B, Grimby G, Thulesius O. Adaptive structural changes of the vascular walls
in hypertension and their relation to the control of the peripheral resistance. Acta
Physiol Scand 44: 255–272, 1958.
426. Forster HG, Ter Wee PM, Hohman TC, Epstein M. Impairment of afferent arteriolar
myogenic responsiveness in the galactose-fed rat is prevented by tolrestat. Diabetologia 39: 907–914, 1996.
427. Forster HG, Ter Wee PM, Takenaka T, Hohman TC, Epstein M. Impairment of
afferent arteriolar myogenic responsiveness in the galactose-fed rat. Proc Soc Exp Biol
Med 206: 365–374, 1994.
428. Forster RP, Maes JP. Effect of experimental neurogenic hypertension on renal blood
flow and glomerular filtration rates in intact denervated kidneys of unanesthetized
rabbits with adrenal glands demedullated. Am J Physiol 150: 534 –540, 1947.
429. Fortepiani LA, Rodrigo E, Ortiz MC, Cachofeiro V, Atucha NM, Ruilope LM, Lahera
V, Garcia-Estan J. Pressure natriuresis in nitric oxide-deficient hypertensive rats:
effect of antihypertensive treatments. J Am Soc Nephrol 10: 21–27, 1999.
430. Fowler BC, Carmines PK, Nelson LD, Bell PD. Characterization of sodium-calcium
exchange in rabbit renal arterioles. Kidney Int 50: 1856 –1862, 1996.
431. Fowler BC, Chang YS, Laamarti A, Higdon M, LaPointe JY, Bell PD. Evidence for
apical sodium proton exchange in macula densa cells. Kidney Int 47: 746 –751, 1995.
432. Fozard JR, Part ML. Haemodynamic responses to NG-monomethyl-L-arginine in
spontaneously hypertensive and normotensive Wistar-Kyoto rats. Br J Pharmacol
102: 823– 826, 1991.
433. Franchini KG, Mattson DL, Cowley AW Jr. Vasopressin modulation of medullary
blood flow and pressure-natriuresis-diuresis in the decerebrated rat. Am J Physiol
Regul Integr Comp Physiol 272: R1472–R1479, 1997.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
481
RENAL AUTOREGULATORY MECHANISMS
434. Francis SH, Busch JL, Corbin JD, Sibley D. cGMP-dependent protein kinases and
cGMP phosphodiesterases in nitric oxide and cGMP action. Pharmacol Rev 62: 525–
563, 2010.
453. Gandley RE, Conrad KP, McLaughlin MK. Endothelin and nitric oxide mediate reduced myogenic reactivity of small renal arteries from pregnant rats. Am J Physiol
Regul Integr Comp Physiol 280: R1–R7, 2001.
435. Francischetti A, Ono H, Frohlich ED. Renoprotective effects of felodipine and/or
enalapril in spontaneously hypertensive rats with and without L-NAME. Hypertension
31: 795– 801, 1998.
454. Gannon KP, VanLandingham LG, Jernigan NL, Grifoni SC, Hamilton G, Drummond
HA. Impaired pressure-induced constriction in mouse middle cerebral arteries of
ASIC2 knockout mice. Am J Physiol Heart Circ Physiol 294: H1793–H1803, 2008.
436. Franco M, Bautista R, Tapia E, Soto V, Santamaria J, Osorio H, Pacheco U, SanchezLozada LG, Kobori H, Navar LG. Contribution of renal purinergic receptors to renal
vasoconstriction in angiotensin II-induced hypertensive rats. Am J Physiol Renal Physiol
300: F1301–F1309, 2011.
455. Gao X, Patzak A, Sendeski M, Scheffer PG, Teerlink T, Sallstrom J, Fredholm BB,
Persson AE, Carlstrom M. Adenosine A1 receptor deficiency diminishes afferent
arteriolar and blood pressure responses during nitric oxide inhibition and angiotensin
II treatment. Am J Physiol Regul Integr Comp Physiol 301: R16669 –R16681, 2011.
437. Franco M, Bell PD, Navar LG. Evaluation of prostaglandins as mediators of tubuloglomerular feedback. Am J Physiol Renal Fluid Electrolyte Physiol 254: F642–F649,
1988.
456. Gao YJ, Nishimura Y, Suzuki A. Dopamine-induced relaxation in isolated intrarenal
arteries from adult stroke-prone spontaneously hypertensive rats. Clin Exp Pharmacol Physiol 22 Suppl 1: S96 –S98, 1995.
438. Franco M, Bell PD, Navar LG. Effect of adenosine A1 analogue on tubuloglomerular
feedback mechanism. Am J Physiol Renal Fluid Electrolyte Physiol 257: F231–F236,
1989.
457. Gao YJ, Nishimura Y, Suzuki A, Higashino H. Reactivity of intrarenal arteries to
vasoconstrictor and vasorelaxant polypeptides in adult stroke-prone spontaneously
hypertensive rats. J Smooth Muscle Res 35: 125–134, 1999.
439. Franco M, Martinez F, Rodriguez-Iturbe B, Johnson RJ, Santamaria J, Montoya A,
Nepomuceno T, Bautista R, Tapia E, Herrera-Acosta J. Angiotensin II, interstitial
inflammation, and the pathogenesis of salt-sensitive hypertension. Am J Physiol Renal
Physiol 291: F1281–F1287, 2006.
458. Garcia SR, Izzard AS, Heagerty AM, Bund SJ. Myogenic tone in coronary arteries
from spontaneously hypertensive rats. J Vasc Res 34: 109 –116, 1997.
440. Franco M, Perez-Mendez O, Martinez F. Interaction of intrarenal adenosine and
angiotensin II in kidney vascular resistance. Curr Opin Nephrol Hypertens 18: 63– 67,
2009.
441. Francois H, Athirakul K, Mao L, Rockman H, Coffman TM. Role for thromboxane
receptors in angiotensin-II-induced hypertension. Hypertension 43: 364 –369, 2004.
442. Freedman BI, Hicks PJ, Bostrom MA, Cunningham ME, Liu Y, Divers J, Kopp JB,
Winkler CA, Nelson GW, Langefeld CD, Bowden DW. Polymorphisms in the nonmuscle myosin heavy chain 9 gene (MYH9) are strongly associated with end-stage
renal disease historically attributed to hypertension in African Americans. Kidney Int
75: 736 –745, 2009.
443. Frey BA, Grisk O, Bandelow N, Wussow S, Bie P, Rettig R. Sodium homeostasis in
transplanted rats with a spontaneously hypertensive rat kidney. Am J Physiol Regul
Integr Comp Physiol 279: R1099 –R1104, 2000.
444. Frisbee JC, Maier KG, Stepp DW. Oxidant stress-induced increase in myogenic
activation of skeletal muscle resistance arteries in obese Zucker rats. Am J Physiol
Heart Circ Physiol 283: H2160 –H2168, 2002.
445. Frisbee JC, Roman RJ, Falck JR, Krishna UM, Lombard JH. 20-HETE contributes to
myogenic activation of skeletal muscle resistance arteries in Brown Norway and
Sprague-Dawley rats. Microcirculation 8: 45–55, 2001.
446. Frisbee JC, Roman RJ, Krishna UM, Falck JR, Lombard JH. 20-HETE modulates
myogenic response of skeletal muscle resistance arteries from hypertensive Dahl-SS
rats. Am J Physiol Heart Circ Physiol 280: H1066 –H1074, 2001.
447. Frohlich ED, Chien Y, Sesoko S, Pegram BL. Relationship between dietary sodium
intake, hemodynamics, and cardiac mass in SHR and WKY rats. Am J Physiol Regul
Integr Comp Physiol 264: R30 –R34, 1993.
448. Fu Y, Zhang R, Lu D, Liu H, Chandrashekar K, Juncos LA, Liu R. NOX2 is the primary
source of angiotensin II-induced superoxide in the macula densa. Am J Physiol Regul
Integr Comp Physiol 298: R707–R712, 2010.
449. Fujii S, Zhang L, Igarashi J, Kosaka H. L-Arginine reverses p47phox and gp91phox
expression induced by high salt in Dahl rats. Hypertension 42: 1014 –1020, 2003.
450. Fujiwara K, Hayashi K, Matsuda H, Kubota E, Honda M, Ozawa Y, Saruta T. Altered
pressure-natriuresis in obese Zucker rats. Hypertension 33: 1470 –1475, 1999.
451. Fuller AJ, Hauschild BC, Gonzalez-Villalobos R, Awayda MS, Imig JD, Inscho EW,
Navar LG. Calcium and chloride channel activation by angiotensin II-AT1 receptors
in preglomerular vascular smooth muscle cells. Am J Physiol Renal Physiol 289: F760 –
F767, 2005.
452. Gabbai FB, Gushwa LC, Peterson OW, Wilson CB, Blantz RC. Analysis of renal
function in the two-kidney Goldblatt model. Am J Physiol Renal Fluid Electrolyte Physiol
252: F131–F137, 1987.
482
459. Garcia-Estan J, Roman RJ. Role of renal interstitial hydrostatic pressure in the pressure diuresis response. Am J Physiol Renal Fluid Electrolyte Physiol 256: F63–F70,
1989.
460. Garvin JL, Herrera M, Ortiz PA. Regulation of renal NaCl transport by nitric oxide,
endothelin, and ATP: clinical implications. Annu Rev Physiol 73: 359 –376, 2011.
461. Gattone VH, Evan AP, Willis LR, Luft FC. Renal afferent arteriole in the spontaneously hypertensive rat. Hypertension 5: 8 –16, 1983.
462. Ge Y, Gannon KP, Gousset M, Liu R, Murphey B, Drummond HA. Impaired myogenic constriction of the renal afferent arteriole in a mouse model of reduced betaENaC expression. Am J Physiol Renal Physiol 302: F1486 –F1493, 2012.
463. Ge Y, Murphy SR, Fan F, Williams JM, Falck JR, Liu R, Roman RJ. Role of 20-HETE in
the impaired myogenic and TGF responses of the Af-Art of Dahl Salt-sensitive rats.
Am J Physiol Renal Physiol 307: F509 –F515, 2014.
464. Ge Y, Murphy SR, Lu Y, Falck J, Liu R, Roman RJ. Endogenously produced 20-HETE
modulates myogenic and TGF response in microperfused afferent arterioles. Prostaglandins Other Lipid Mediat 102–103: 42– 48, 2013.
465. Gebremedhin D, Fenoy FJ, Harder DR, Roman RJ. Enhanced vascular tone in the
renal vasculature of spontaneously hypertensive rats. Hypertension 16: 648 – 654,
1990.
466. Gebremedhin D, Gopalakrishnan S, Harder DR. Endogenous events modulating
myogenic regulation of cerebrovascular function. Curr Vasc Pharmacol 12: 810 – 817,
2014.
467. Gebremedhin D, Kaldunski ML, Jacobs ER, Harder DR, Roman RJ. Coexistence of
two types of Ca2⫹-activated K⫹ channels in rat renal arterioles. Am J Physiol Renal
Fluid Electrolyte Physiol 270: F69 –F81, 1996.
468. Gebremedhin D, Lange AR, Lowry TF, Taheri MR, Birks EK, Hudetz AG, Narayanan
J, Falck JR, Okamoto H, Roman RJ, Nithipatikom K, Campbell WB, Harder DR.
Production of 20-HETE and its role in autoregulation of cerebral blood flow. Circ Res
87: 60 – 65, 2000.
469. Gebremedhin D, Ma YH, Imig JD, Harder DR, Roman RJ. Role of cytochrome P-450
in elevating renal vascular tone in spontaneously hypertensive rats. J Vasc Res 30:
53– 60, 1993.
470. Gebremedhin D, Terashvili M, Wickramasekera N, Zhang DX, Rau N, Miura H,
Harder DR. Redox signaling via oxidative inactivation of PTEN modulates pressuredependent myogenic tone in rat middle cerebral arteries. PLoS One 8: e68498, 2013.
471. Gill PS, Wilcox CS. NADPH oxidases in the kidney. Antioxid Redox Signal 8: 1597–
1607, 2006.
472. Gillies LK, Lu M, Wang H, Lee RMKW. AT1 receptor antagonist treatment caused
persistent arterial functional changes in young spontaneously hypertensive rats. Hypertension 30: 1471–1478, 1997.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
473. Gilmore JP, Cornish KG, Rogers SD, Joyner WL. Direct evidence for myogenic
autoregulation of the renal microcirculation in the hamster. Circ Res 47: 226 –230,
1980.
494. Griendling KK, Minieri CA, Ollerenshaw JD, Alexander RW. Angiotensin II stimulates
NADH and NADPH oxidase activity in cultured vascular smooth muscle cells. Circ
Res 74: 1141–1148, 1994.
474. Gokina NI, Knot HJ, Nelson MT, Osol G. Increased Ca2⫹ sensitivity as a key mechanism of PKC-induced constriction in pressurized cerebral arteries. Am J Physiol
Heart Circ Physiol 277: H1178 –H1188, 1999.
495. Griffin KA, Abu-Naser M, Abu-Amarah I, Picken M, Williamson GA, Bidani AK.
Dynamic blood pressure load and nephropathy in the ZSF1 (fa/facp) model of type 2
diabetes. Am J Physiol Renal Physiol 293: F1605–F1613, 2007.
475. Gokina NI, Park KM, Elroy-Yaggy K, Osol G. Effects of Rho kinase inhibition on
cerebral artery myogenic tone and reactivity. J Appl Physiol 98: 1940 –1948, 2005.
496. Griffin KA, Bidani AK. Hypertensive renal damage: insights from animal models and
clinical relevance. Curr Hypertens Rep 6: 145–153, 2004.
476. Goldblatt H, Lynch J, Hanzal RF, Summerville WW. Studies on experimental hypertension. I. The production of persistent elevation of systolic blood pressure by means
of renal ischemia. J Exp Med 59: 347–379, 1934.
497. Griffin KA, Bidani AK, Picken M, Ellis VR, Churchill PC. Prostaglandins do not mediate impaired autoregulation or increased renin secretion in remnant rat kidneys. Am
J Physiol Renal Fluid Electrolyte Physiol 263: F1057–F1062, 1992.
477. Goligorsky MS, Colflesh D, Gordienko DV, Moore LC. Branching points of renal
resistance arteries are enriched in L-type calcium channels and initiate vasoconstriction. Am J Physiol Renal Fluid Electrolyte Physiol 268: F251–F257, 1995.
478. Gollasch M, Lohn M, Furstenau M, Nelson MT, Luft FC, Haller H. Ca2⫹ channels,
“quantized” Ca2⫹ release, and differentiation of myocytes in the cardiovascular
system. J Hypertens 18: 989 –998, 2000.
479. Gomez RA, Norwood VF, Tufro-McReddie A. Development of the kidney vasculature. Microscopy Res Tech 39: 254 –260, 1997.
480. Gonzales AL, Amberg GC, Earley S. Ca2⫹ release from the sarcoplasmic reticulum is
required for sustained TRPM4 activity in cerebral artery smooth muscle cells. Am J
Physiol Cell Physiol 299: C279 –C288, 2010.
481. Gonzales AL, Earley S. Regulation of cerebral artery smooth muscle membrane
potential by Ca2⫹-activated cation channels. Microcirculation 20: 337–347, 2012.
482. Gonzales AL, Yang Y, Sullivan MN, Sanders L, Dabertrand F, Hill-Eubanks DC,
Nelson MT, Earley S. A PLCgamma1-dependent, force-sensitive signaling network
in the myogenic constriction of cerebral arteries. Sci Signal 7: ra49, 2014.
483. Gonzalez JD, Llinas MT, Moreno C, Rodriguez F, Salazar FJ. Renal effects of prolonged cyclooxygenase inhibition when angiotensin II levels are elevated. J Cardiovasc
Pharmacol 36: 236 –241, 2000.
484. Gonzalez JM, Somoza B, Conde MV, Fernandez-Alfonso MS, Gonzalez MC, Arribas
SM. Hypertension increases middle cerebral artery resting tone in spontaneously
hypertensive rats: role of tonic vasoactive factor availability. Clin Sci 114: 651– 659,
2008.
485. Gonzalez R, Fernandez-Alfonso MS, Rodriguez-Martinez MA, Fuertes E, Angulo J,
Sanchez-Ferrer CF, Marin J. Pressure-induced contraction of the juxtamedullary
afferent arterioles in spontaneously hypertensive rats. Gen Pharmacol 25: 333–339,
1994.
486. Gonzalez-Campoy JM, Long C, Roberts D, Berndt TJ, Romero JC, Knox FG. Renal
interstitial hydrostatic pressure and PGE2 in pressure natriuresis. Am J Physiol Renal
Fluid Electrolyte Physiol 260: F643–F649, 1991.
487. Goormaghtigh N. L’appareil neuro-myo-arteriel juxta-glmerulaire du rein: ses reactions en pathologie et ses rapports avec le tube urinifere. CR Seances Soc Biol Fil 124:
293–296, 1937.
488. Goransson A, Isaksson B, Sjoquist M. Whole kidney response to reduced arterial
pressure during converting enzyme inhibition in the rat. Renal Physiol 9: 287–301,
1986.
498. Griffin KA, Churchill PC, Picken M, Webb RC, Kurtz TW, Bidani AK. Differential
salt-sensitivity in the pathogenesis of renal damage in SHR and stroke prone SHR. Am
J Hypertens 14: 311–320, 2001.
499. Griffin KA, Hacioglu R, Abu-Amarah I, Loutzenhiser R, Williamson GA, Bidani AK.
Effects of calcium channel blockers on “dynamic” and “steady-state step” renal
autoregulation. Am J Physiol Renal Physiol 286: F1136 –F1143, 2004.
500. Griffin KA, Picken M, Bakris GL, Bidani AK. Comparative effects of selective T- and
L-type calcium channel blockers in the remnant kidney model. Hypertension 37:
1268 –1272, 2001.
501. Griffin KA, Picken M, Bidani AK. Radiotelemetric BP monitoring, antihypertensives
and glomeruloprotection in remnant kidney model. Kidney Int 46: 1010 –1018, 1994.
502. Griffin KA, Picken M, Giobbie-Hurder A, Bidani AK. Low protein diet mediated
renoprotection in remnant kidneys: renal autoregulatory versus hypertrophic mechanisms. Kidney Int 63: 607– 616, 2003.
503. Griffin KA, Picken MM, Bidani AK. Deleterious effects of calcium channel blockade
on pressure transmission and glomerular injury in rat remnant kidneys. J Clin Invest
96: 793– 800, 1995.
504. Griffin KA, Picken MM, Bidani AK. Blood pressure lability and glomerulosclerosis
after normotensive 5/6 renal mass reduction in the rat. Kidney Int 65: 209 –218, 2004.
505. Grifoni SC, Chiposi R, McKey SE, Ryan MJ, Drummond HA. Altered whole kidney
blood flow autoregulation in a mouse model of reduced beta-ENaC. Am J Physiol
Renal Physiol 298: F285–F292, 2010.
506. Grisk O, Rettig R. Renal transplantation studies in genetic hypertension. News Physiol
Sci 16: 262–265, 2001.
507. Groschner K, Rosker C, Lukas M. Role of TRP channels in oxidative stress. Novartis
Found Symp 258: 222–230, 2004.
508. Gross JL, de Azevedo MJ, Silveiro SP, Canani LH, Caramori ML, Zelmanovitz T.
Diabetic nephropathy: diagnosis, prevention, treatment. Diabetes Care 28: 164 –176,
2005.
509. Gross JM, Dwyer JE, Knox FG. Natriuretic response to increased pressure is preserved with COX-2 inhibitors. Hypertension 34: 1163–1167, 1999.
510. Gross R, Kirchheim H, Brandstetter K. Basal vascular tone in the kidney. Evaluation
from the static pressure-flow relationship under normal autoregulation and at maximal dilation in the dog. Circ Res 38: 525–531, 1976.
489. Gordienko DV, Clausen C, Goligorsky MS. Ionic currents and endothelin signaling in
smooth muscle cells form rat renal resistance arterioles. Am J Physiol Renal Fluid
Electrolyte Physiol 266: F325–F341, 1994.
511. Gross R, Kirchheim H, Ruffmann K. Effect of carotid occlusion and of perfusion
pressure on renal function in conscious dogs. Circ Res 48: 777–784, 1981.
490. Gottlieb P, Folgering J, Maroto R, Raso A, Wood TG, Kurosky A, Bowman C, Bichet
D, Patel A, Sachs F, Martinac B, Hamill OP, Honore E. Revisiting TRPC1 and TRPC6
mechanosensitivity. Pflügers Arch 455: 1097–1103, 2008.
512. Gross V, Lippoldt A, Bohlender J, Bader M, Hansson A, Luft FC. Cortical and medullary hemodynamics in deoxycorticosterone acetate-salt hypertensive mice. J Am
Soc Nephrol 9: 346 –354, 1998.
491. Gouedard O, Blanc J, Gaudet E, Ponchon P, Elghozi JL. Contribution of the reninangiotensin system to short-term blood pressure variability during blockade of nitric
oxide synthesis in the rat. Br J Pharmacol 119: 1085–1092, 1996.
513. Gross V, Lippoldt A, Schneider W, Luft FC. Effect of captopril and angiotensin II
receptor blockade on pressure natriuresis in transgenic TGR(mRen-2)27 rats. Hypertension 26: 471– 479, 1995.
492. Granger JP, Schnackenberg CG. Renal mechanisms of angiotensin II-induced hypertension. Semin Nephrol 20: 417– 425, 2000.
514. Gross V, Obst M, Luft FC. Insights into angiotensin II receptor function through AT2
receptor knockout mice. Acta Physiol Scand 181: 487– 494, 2004.
493. Green R, Giebisch G. Luminal hypotonicity: a driving force for fluid absorption from
the proximal tubule. Am J Physiol Renal Fluid Electrolyte Physiol 246: F167–F174, 1984.
515. Gross V, Roman RJ, Cowley AW Jr. Abnormal pressure-natriuresis in transgenic
renin gene rats. J Hypertension 12: 1029 –1034, 1994.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
483
RENAL AUTOREGULATORY MECHANISMS
516. Gross V, Schunck WH, Honeck H, Milia AF, Kargel E, Walther T, Bader M, Inagami
T, Schneider W, Luft FC. Inhibition of pressure natriuresis in mice lacking the AT2
receptor. Kidney Int 57: 191–202, 2000.
537. Haas JA, Khraibi AA, Perrella MA, Knox FG. Role of renal interstitial hydrostatic
pressure in natriuresis of systemic nitric oxide inhibition. Am J Physiol Renal Fluid
Electrolyte Physiol 264: F411–F414, 1993.
517. Grossman E, Hoffman A, Keiser HR. Sodium intake modulates renal vascular reactivity to endothelin-1 in Dahl rats. Clin Exp Pharmacol Physiol 17: 121–128, 1990.
538. Haas JA, Knox FG. Effect of meclofenamate or ketoconazole on the natriuretic
response to increased pressure. J Lab Clin Med 128: 202–207, 1996.
518. Gschwend S, Buikema H, Navis G, Henning RH, de Zeeuw D, Van Dokkum RP.
Endothelial dilatory function predicts individual susceptibility to renal damage in the
5/6 nephrectomized rat. J Am Soc Nephrol 13: 2909 –2915, 2002.
539. Haas JA, Lockhart JC, Larson TS, Henrikson T, Knox FG. Natriuretic response to
renal interstitial hydrostatic pressure during angiotensin II blockade. Am J Physiol
Renal Fluid Electrolyte Physiol 266: F117–F119, 1994.
519. Gschwend S, Henning RH, Pinto YM, de Zeeuw D, van Gilst WH, Buikema H.
Myogenic constriction is increased in mesenteric resistance arteries from rats with
chronic heart failure: instantaneous counteraction by acute AT1 receptor blockade.
Br J Pharmacol 139: 1317–1325, 2003.
540. Haberle DA, Konigbauer B, Davis JM, Kawata T, Mast C, Metz C, Dahlheim H.
Autoregulation of the glomerular filtration rate and the single-nephron glomerular
filtration rate despite inhibition of tubuloglomerular feedback in rats chronically
volume-expanded by deoxycorticosterone acetate. Pflügers Arch 416: 548 –553,
1990.
520. Guan Z, Fuller BS, Yamamoto T, Cook AK, Pollock JS, Inscho EW. Pentosan polysulfate treatment preserves renal autoregulation in Ang II-infused hypertensive tats
via normalization of P2X1 receptor activation. Am J Physiol Renal Physiol 298: F1276 –
F1284, 2010.
541. Hacioglu R, Williamson GA, Abu-Amarah I, Griffin KA, Bidani AK. Characterization
of dynamics in renal autoregulation using volterra models. IEEE Trans Biomed Eng 53:
2166 –2176, 2006.
521. Guan Z, Osmond DA, Inscho EW. P2X receptors as regulators of the renal microvasculature. Trends Pharmacol Sci 28: 646 – 652, 2007.
522. Guan Z, Pollock JS, Cook AK, Hobbs JL, Inscho EW. Effect of epithelial sodium
channel blockade on the myogenic response of rat juxtamedullary afferent arterioles. Hypertension 54: 1062–1069, 2009.
523. Guan Z, Willgoss DA, Matthias A, Manley SW, Crozier S, Gobe G, Endre ZH.
Facilitation of renal autoregulation by angiotensin II is mediated through modulation
of nitric oxide. Acta Physiol Scand 179: 189 –201, 2003.
524. Guarasci GR, Kline RL. Pressure natriuresis following acute and chronic inhibition of
nitric oxide synthase in rats. Am J Physiol Regul Integr Comp Physiol 270: R469 –R478,
1996.
525. Gui P, Chao JT, Wu X, Yang Y, Davis GE, Davis MJ. Coordinated regulation of
vascular Ca2⫹ and K⫹ channels by integrin signaling. Adv Exp Med Biol 674: 69 –79,
2010.
526. Gunther S, Gimbrone MA Jr, Alexander RW. Regulation by angiotensin II of its
receptors in resistance blood vessels. Nature 287: 230 –232, 1980.
527. Gustafsson F, Holstein-Rathlou NH. Conducted vasomotor responses in arterioles:
characteristics, mechanisms and physiological significance. Acta Physiol Scand 167:
11–21, 1999.
528. Gustafsson H. Vasomotion and underlying mechanisms in small arteries. An in vitro
study of rat blood vessels. Acta Physiol Scand Suppl 614: 1– 44, 1993.
529. Gustafsson H, Bulow A, Nilsson H. Rhythmic contractions of isolated, pressurized
small arteries from rat. Acta Physiol Scand 152: 145–152, 1994.
530. Gustafsson H, Mulvany MJ, Nilsson H. Rhythmic contractions of isolated small arteries from rat: influence of the endothelium. Acta Physiol Scand 148: 153–163, 1993.
531. Gustafsson H, Nilsson H. Rhythmic contractions of isolated small arteries from rat:
role of calcium. Acta Physiol Scand 149: 283–291, 1993.
532. Guyton AC, Hall JE, Coleman TG, Manning RD Jr, Norman RA. The dominant role of
the kidneys in long-term arterial pressure regulation in normal and hypertensive
states. In: Hypertension: Pathophysiology, Diagnosis and Management, edited by Laragh JH and Brenner BM. New York: Raven, 1995.
533. Guyton AC, Langston JB, Navar LG. Theory for renal autoregulation by feedback at
the juxtaglomerular apparatus. Circ Res 15: 187–197, 1964.
542. Haddock RE, Grayson TH, Brackenbury TD, Meaney KR, Neylon CB, Sandow SL,
Hill CE. Endothelial coordination of cerebral vasomotion via myoendothelial gap
junctions containing connexins 37 and 40. Am J Physiol Heart Circ Physiol 291: H2047–
H2056, 2006.
543. Haddock RE, Hill CE. Rhythmicity in arterial smooth muscle. J Physiol 566: 645– 656,
2005.
544. Haddock RE, Hirst GD, Hill CE. Voltage independence of vasomotion in isolated
irideal arterioles of the rat. J Physiol 540: 219 –229, 2002.
545. Haefliger JA, Krattinger N, Martin D, Pedrazzini T, Capponi A, Doring B, Plum A,
Charollais A, Willecke K, Meda P. Connexin43-dependent mechanism modulates
renin secretion and hypertension. J Clin Invest 116: 405– 413, 2006.
546. Hahne B, Selen G, Persson AEG. Indomethacin inhibits renal functional adaptation to
nephron loss. Ren Physiol 7: 13–21, 1984.
547. Hall JE. Control of sodium excretion by angiotensin II: intrarenal mechanisms and
blood pressure regulation. Am J Physiol Regul Integr Comp Physiol 250: R960 –R972,
1986.
548. Hall JE, Coleman TG, Guyton AC, Balfe JW, Salgado HC. Intrarenal role of angiotensin II and [des-Asp1]angiotensin II. Am J Physiol Renal Fluid Electrolyte Physiol 236:
F252–F259, 1979.
549. Hall JE, Granger JP, Hester RL, Coleman TG, Smith MJ Jr, Cross RB. Mechanisms of
escape from sodium retention during angiotensin II hypertension. Am J Physiol Renal
Fluid Electrolyte Physiol 246: F627–F634, 1984.
550. Hall JE, Guyton AC, Cowley AW Jr. Dissociation of renal blood flow and filtration
rate autoregulation by renin depletion. Am J Physiol Renal Fluid Electrolyte Physiol 232:
F215–F221, 1977.
551. Hampton JA, Bernardo DA, Khan NA, Lacher DA, Rapp JP, Gohara AF, Goldblatt PJ.
Morphometric evaluation of the renal arterial system of Dahl salt-sensitive and
salt-resistant rats on a high salt diet. II. Interlobular arteries and intralobular arterioles. Lab Invest 60: 839 – 846, 1989.
552. Hanner F, Lam L, Nguyen MT, Yu A, Peti-Peterdi J. Intrarenal localization of the
plasma membrane ATP channel pannexin1. Am J Physiol Renal Physiol 303: F1454 –
F1459, 2012.
553. Hanner F, Sorensen CM, Holstein-Rathlou NH, Peti-Peterdi J. Connexins and the
kidney. Am J Physiol Regul Integr Comp Physiol 298: R1143–R1155, 2010.
534. Guzik TJ, Hoch NE, Brown KA, McCann LA, Rahman A, Dikalov S, Goronzy J,
Weyand C, Harrison DG. Role of the T cell in the genesis of angiotensin II induced
hypertension and vascular dysfunction. J Exp Med 204: 2449 –2460, 2007.
554. Hanner F, von Maltzahn J, Maxeiner S, Toma I, Sipos A, Kruger O, Willecke K,
Peti-Peterdi J. Connexin45 is expressed in the juxtaglomerular apparatus and is
involved in the regulation of renin secretion and blood pressure. Am J Physiol Regul
Integr Comp Physiol 295: R371–R380, 2008.
535. Haas JA, Granger JP, Knox FG. Effect of renal perfusion pressure on sodium reabsorption from proximal tubules of superficial and deep nephrons. Am J Physiol Renal
Fluid Electrolyte Physiol 250: F425–F429, 1986.
555. Hansen PB, Castrop H, Briggs J, Schnermann J. Adenosine induces vasoconstriction
through Gi-dependent activation of phospholipase C in isolated perfused afferent
arterioles of mice. J Am Soc Nephrol 14: 2457–2465, 2003.
536. Haas JA, Granger JP, Knox FG. Effect of meclofenamate on lithium excretion in
response to changes in renal perfusion pressure. J Lab Clin Med 111: 543–547,
1988.
556. Hansen PB, Hashimoto S, Briggs J, Schnermann J. Attenuated renovascular constrictor responses to angiotensin II in adenosine 1 receptor knockout mice. Am J Physiol
Regul Integr Comp Physiol 285: R44 –R49, 2003.
484
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
557. Hansen PB, Hashimoto S, Oppermann M, Huang Y, Briggs JP, Schnermann J. Vasoconstrictor and vasodilator effects of adenosine in the mouse kidney due to preferential activation of A1 or A2 adenosine receptors. J Pharmacol Exp Ther 315: 1150 –
1157, 2005.
579. Harrison DG, Gongora MC, Guzik TJ, Widder J. Oxidative stress and hypertension.
J Am Soc Hypertens 1: 30 – 44, 2007.
580. Harrison-Bernard LM, Navar LG, Cook AK. Renal cortical and medullary microvascular blood flow autoregulation in rat. Kidney Int 50: S23–S29, 1996.
558. Hansen PB, Jensen BL, Andreasen D, Skott O. Differential expression of T- and
L-type voltage-dependent calcium channels in renal resistance vessels. Circ Res 89:
630 – 638, 2001.
581. Harsing L, Fonyodi S, Laszlo K, Takacs G. Effect of hypertonic infusions on renal
haemodynamics. Acta Physiol Hung 12: 351–361, 1957.
559. Hansen PB, Jensen BL, Skott O. Chloride regulates afferent arteriolar contraction in
response to depolarization. Hypertension 32: 1066 –1070, 1998.
582. Hashimoto S, Adams JW, Bernstein KE, Schnermann J. Micropuncture determination of nephron function in mice without tissue angiotensin-converting enzyme. Am
J Physiol Renal Physiol 288: F445–F452, 2005.
560. Hansen PB, Poulsen CB, Walter S, Marcussen N, Cribbs LL, Skott O, Jensen BL.
Functional importance of L- and P/Q-type voltage-gated calcium channels in human
renal vasculature. Hypertension 58: 464 – 470, 2011.
561. Harder DR. Pressure-dependent membrane depolarization in cat middle cerebral
artery. Circ Res 55: 197–202, 1984.
562. Harder DR, Gilbert R, Lombard JH. Vascular muscle cell depolarization and activation in renal arteries on elevation of transmural pressure. Am J Physiol Renal Fluid
Electrolyte Physiol 253: F778 –F781, 1987.
563. Harder DR, Narayanan J, Gebremedhin D. Pressure-induced myogenic tone and
role of 20-HETE in mediating autoregulation of cerebral blood flow. Am J Physiol
Heart Circ Physiol 300: H1557–H1565, 2011.
564. Harder DR, Smeda J, Lombard J. Enhanced myogenic depolarization in hypertensive
cerebral arterial muscle. Circ Res 57: 319 –322, 1985.
565. Harhun MI, Povstyan OV, Gordienko DV. Purinoreceptor-mediated current in myocytes from renal resistance arteries. Br J Pharmacol 160: 987–997, 2010.
566. Harper SL, Bohlen HG. Microvascular adaptation in the cerebral cortex of adult
spontaneously hypertensive rats. Hypertension 6: 408 – 419, 1984.
567. Harrap SB. Angiotensin converting enzyme inhibitors, regional vascular hemodynamics, and the development and prevention of experimental genetic hypertension.
Am J Hypertens 4: 212S–216S, 1991.
568. Harrap SB. Cardiovascular disease and genetics of the renin-angiotensin system.
Heart 76: 13–17, 1996.
569. Harrap SB, Clark SA, Fraser R, Towrie A, Brown AJ, Lever AF. Effects of sodium
intake and aldosterone on the renal pressure-natriuresis. Am J Physiol Renal Fluid
Electrolyte Physiol 254: F697–F703, 1988.
570. Harrap SB, Doyle AE. Total body sodium in immature spontaneously hypertensive
and Wistar Kyoto rats. Clin Exp Pharmacol Physiol 12: 315–318, 1985.
571. Harrap SB, Doyle AE. Renal haemodynamics and total body sodium in immature
spontaneously hypertensive and Wistar-Kyoto rats. J Hypertens Suppl 4: S249 –S252,
1986.
572. Harrap SB, Doyle AE. Genetic co-segregation of renal haemodynamics and blood
pressure in the spontaneously hypertensive rat. Clin Sci 74: 63– 69, 1988.
573. Harrap SB, Mirakian C, Datodi SR, Lever AF. Blood pressure and lifespan following
brief ACE inhibitor treatment in young spontaneously hypertensive rats. Clin Exp
Pharmacol Physiol 21: 125–127, 1994.
574. Harrap SB, Nicolaci JA, Doyle AE. Persistent effects on blood pressure and renal
haemodynamics following chronic angiotensin converting enzyme inhibition with
perindopril. Clin Exp Pharmacol Physiol 13: 753–765, 1986.
575. Harrap SB, Wang BZ, MacLellan DG. Transplantation studies of the role of the
kidney in long-term blood pressure reduction following brief ACE inhibitor treatment in young spontaneously hypertensive rats. Clin Exp Pharmacol Physiol 21: 129 –
131, 1994.
576. Harraz OF, Brett SE, Welsh DG. Nitric oxide suppresses vascular voltage-gated
T-type Ca2⫹ channels through cGMP/PKG signaling. Am J Physiol Heart Circ Physiol
306: H279 –H285, 2014.
583. Hashimoto S, Huang Y, Briggs J, Schnermann J. Reduced autoregulatory effectiveness in adenosine 1 receptor-deficient mice. Am J Physiol Renal Physiol 290: F888 –
F891, 2006.
584. Hashimoto S, Yamada K, Kawata T, Mochizuki T, Schnermann J, Koike T. Abnormal
autoregulation and tubuloglomerular feedback in prediabetic and diabetic OLETF
rats. Am J Physiol Renal Physiol 296: F598 –F604, 2009.
585. Hayakawa H, Raij L. Nitric oxide synthase activity and renal injury in genetic hypertension. Hypertension 31: 266 –270, 1998.
586. Hayashi K, Epstein M, Loutzenhiser R. Pressure-induced vasoconstriction of renal
microvessels in normotensive and hypertensive rats. Studies in the isolated perfused
hydronephrotic kidney. Circ Res 65: 1475–1484, 1989.
587. Hayashi K, Epstein M, Loutzenhiser R. Enhanced myogenic responsiveness of renal
interlobular arteries in spontaneously hypertensive rats. Hypertension 19: 153–160,
1992.
588. Hayashi K, Epstein M, Loutzenhiser R, Forster H. Impaired myogenic responsiveness
of the afferent arteriole in streptozotocin-induced diabetic rats: role of eicosanoid
derangements. J Am Soc Nephrol 2: 1578 –1586, 1992.
589. Hayashi K, Epstein M, Saruta T. Altered myogenic responsiveness of the renal microvasculature in experimental hypertension. J Hypertens 14: 1387–1401, 1996.
590. Hayashi K, Fujiwara K, Oka K, Nagahama T, Matsuda H, Saruta T. Effects of insulin
on rat renal microvessels: studies in the isolated perfused hydronephrotic kidney.
Kidney Int 51: 1507–1513, 1997.
591. Hayashi K, Kanda T, Homma K, Tokuyama H, Okubo K, Takamatsu I, Tatematsu S,
Kumagai H, Saruta T. Altered renal microvascular response in Zucker obese rats.
Metabolism 51: 1553–1561, 2002.
592. Hayashi K, Loutzenhiser RD, Epstein M, Suzuki H, Saruta T. Multiple factors contribute to acetylcholine-induced renal afferent arteriolar vasodilation during myogenic and norepinephrine- and KCl-induced vasoconstriction. Studies in the isolated
perfused hydronephrotic kidney. Circ Res 75: 821– 828, 1994.
593. Hayashi K, Suzuki H, Saruta T. Nitric oxide modulates but does not impair myogenic
vasoconstriction of the afferent arteriole in spontaneously hypertensive rats. Hypertension 25: 1212–1219, 1995.
594. Hayashi K, Wakino S, Kanda T, Homma K, Sugano N, Saruta T. Molecular mechanisms and therapeutic strategies of chronic renal injury: role of rho-kinase in the
development of renal injury. J Pharmacol Sci 100: 29 –33, 2006.
595. Hayashi K, Wakino S, Sugano N, Ozawa Y, Homma K, Saruta T. Ca2⫹ channel
subtypes and pharmacology in the kidney. Circ Res 100: 342–353, 2007.
596. Hayashi M, Monkawa T, Saruta T. Cell-specific expression of the IP3 receptor gene
family in the kidney. Exp Nephrol 8: 215–218, 2000.
597. He J, Marsh DJ. Effect of captopril on fluctuations of blood pressure and renal blood
flow in rats. Am J Physiol Renal Fluid Electrolyte Physiol 264: F37–F44, 1993.
598. Heeg JE, de Jong PE, de Zeeuw D. Additive antiproteinuric effect of angiotensinconverting enzyme inhibition and non-steroidal anti-inflammatory drug therapy: a
clue to the mechanism of action. Clin Sci 81: 367–372, 1991.
577. Harris RC. An update on cyclooxygenase-2 expression and metabolites in the kidney. Curr Opin Nephrol Hypertens 17: 64 – 69, 2008.
599. Heinze C, Seniuk A, Sokolov MV, Huebner AK, Klementowicz AE, Szijarto IA,
Schleifenbaum J, Vitzthum H, Gollasch M, Ehmke H, Schroeder BC, Hubner CA.
Disruption of vascular Ca2⫹-activated chloride currents lowers blood pressure. J
Clin Invest 124: 675– 686, 2014.
578. Harrison DG, Gongora MC. Oxidative stress and hypertension. Med Clin North Am
93: 621– 635, 2009.
600. Heistad DD, Baumbach GL. Cerebral vascular changes during chronic hypertension:
good guys and bad guys. J Hypertens Suppl 10: S71–S75, 1992.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
485
RENAL AUTOREGULATORY MECHANISMS
601. Helal I, Fick-Brosnahan GM, Reed-Gitomer B, Schrier RW. Glomerular hyperfiltration: definitions, mechanisms and clinical implications. Nat Rev Nephrol 8: 293–300,
2012.
622. Hodges GJ, Del Pozzi AT. Noninvasive examination of endothelial, sympathetic, and
myogenic contributions to regional differences in the human cutaneous microcirculation. Microvasc Res 93: 87–91, 2014.
602. Hellebrekers LJ, Liard JF, Laborde AL, Greene AS, Cowley AW Jr. Regional autoregulatory responses during infusion of vasoconstrictor agents in conscious dogs. Am J
Physiol Heart Circ Physiol 259: H1270 –H1277, 1990.
623. Hofmann T, Obukhov AG, Schaefer M, Harteneck C, Gudermann T, Schultz G.
Direct activation of human TRPC6 and TRPC3 channels by diacylglycerol. Nature
397: 259 –263, 1999.
603. Henrion D, Laher I. Effects of staurosporine and calphostin C, two structurally
unrelated inhibitors of protein kinase C, on vascular tone. Can J Physiol Pharmacol 71:
521–524, 1993.
624. Holm L, Morsing P, Casellas D, Persson AE. Resetting of the pressure range for
blood flow autoregulation in the rat kidney. Acta Physiol Scand 138: 395– 401, 1990.
604. Hercule HC, Tank J, Plehm R, Wellner M, da Costa Goncalves AC, Gollasch M,
Diedrich A, Jordan J, Luft FC, Gross V. Regulator of G protein signalling 2 ameliorates
angiotensin II-induced hypertension in mice. Exp Physiol 92: 1014 –1022, 2007.
605. Herlitz H, Lundin S, Ricksten SE, Gothberg G, Aurell M, Hallback M, Berglund G.
Sodium balance and structural vascular changes in the kidney during development of
hypertension in spontaneously hypertensive rats. Acta Med Scand Suppl 625: 111–
115, 1979.
625. Holstein-Rathlou NH. Oscillations and chaos in renal blood flow control. J Am Soc
Nephrol 4: 1275–1287, 1993.
626. Holstein-Rathlou NH, Leyssac PP. TGF-mediated oscillations in the proximal intratubular pressure: differences between spontaneously hypertensive rats and WistarKyoto rats. Acta Physiol Scand 126: 333–339, 1986.
627. Holstein-Rathlou NH, Marsh DJ. Oscillations of tubular pressure, flow, and distal
chloride concentration in rats. Am J Physiol Renal Fluid Electrolyte Physiol 256: F1007–
F1014, 1989.
606. Herlitz H, Ricksten SE, Lundin S, Thoren P, Aurell M, Berglund G. Renal denervation
and sodium balance in young spontaneously hypertensive rats. Renal Physiol 6: 145–
150, 1983.
628. Holstein-Rathlou NH, Marsh DJ. A dynamic model of renal blood flow autoregulation. Bull Math Biol 56: 411– 429, 1994.
607. Herrera GA, Turbat-Herrera EA, Teng J. Mesangial homeostasis and pathobiology:
their role in health and disease. Contrib Nephrol 169: 6 –22, 2011.
629. Holstein-Rathlou NH, Marsh DJ. Renal blood flow regulation and arterial pressure
fluctuations: a case study in nonlinear dynamics. Physiol Rev 74: 637– 681, 1994.
608. Herrera M, Ortiz PA, Garvin JL. Regulation of thick ascending limb transport: role of
nitric oxide. Am J Physiol Renal Physiol 290: F1279 –F1284, 2006.
630. Holstein-Rathlou NH, Sosnovtseva OV, Pavlov AN, Cupples WA, Sorensen CM,
Marsh DJ. Nephron blood flow dynamics measured by laser speckle contrast imaging. Am J Physiol Renal Physiol 300: F319 –F329, 2011.
609. Herrera-Acosta J, Tapia E, Sanchez-Lozada LG, Franco M, Striker LJ, Striker GE,
Rodriguez IB. Restoration of glomerular haemodynamics and renal injury independent of arterial hypertension in rats with subtotal renal ablation. J Hypertens Suppl 20:
S29 –S35, 2002.
631. Holstein-Rathlou NH, Wagner AJ, Marsh DJ. Tubuloglomerular feedback dynamics
and renal blood flow autoregulation in rats. Am J Physiol Renal Fluid Electrolyte Physiol
260: F53–F68, 1991.
610. Hester RL, Granger JP, Williams J, Hall JE. Acute and chronic servo-control of renal
perfusion pressure. Am J Physiol Renal Fluid Electrolyte Physiol 244: F455–F460, 1983.
611. Heyeraas KJ, Aukland K. Interlobular arterial resistance: influence of renal arterial
pressure and angiotensin II. Kidney Int 31: 1291–1298, 1987.
612. Hill JV, Findon G, Appelhoff RJ, Endre ZH. Renal autoregulation and passive pressure-flow relationships in diabetes and hypertension. Am J Physiol Renal Physiol 299:
F837–F844, 2010.
613. Hill MA, Davis MJ, Meininger GA. Cyclooxygenase inhibition potentiates myogenic
activity in skeletal muscle arterioles. Am J Physiol Heart Circ Physiol 258: H127–H133,
1990.
614. Hill MA, Davis MJ, Meininger GA, Potocnik SJ, Murphy TV. Arteriolar myogenic
signalling mechanisms: Implications for local vascular function. Clin Hemorheol Microcirc 34: 67–79, 2006.
615. Hill MA, Falcone JC, Meininger GA. Evidence for protein kinase C involvement in
arteriolar myogenic reactivity. Am J Physiol Heart Circ Physiol 259: H1586 –H1594,
1990.
616. Hill MA, Meininger GA. Impaired arteriolar myogenic reactivity in early experimental
diabetes. Diabetes 42: 1226 –1232, 1993.
617. Hill MA, Meininger GA. Arteriolar vascular smooth muscle cells: mechanotransducers in a complex environment. Int J Biochem Cell Biol 44: 1505–1510, 2012.
2⫹
⫹
618. Hill MA, Yang Y, Ella SR, Davis MJ, Braun AP. Large conductance, Ca -activated K
channels (BKCa) and arteriolar myogenic signaling. FEBS Lett 584: 2033–2042, 2010.
619. Hilliard LM, Nematbakhsh M, Kett MM, Teichman E, Sampson AK, Widdop RE,
Evans RG, Denton KM. Gender differences in pressure-natriuresis and renal autoregulation: role of the angiotensin type 2 receptor. Hypertension 57: 275–282, 2011.
632. Homma K, Hayashi K, Wakino S, Tokuyama H, Kanda T, Tatematsu S, Hasegawa K,
Fujishima S, Hori S, Saruta T, Itoh H. Rho-kinase contributes to pressure-induced
constriction of renal microvessels. Keio J Med 63: 1–12, 2014.
633. Homma K, Hayashi K, Yamaguchi S, Fujishima S, Hori S, Itoh H. Renal microcirculation and calcium channel subtypes. Curr Hypertens Rev 9: 182–186, 2013.
634. Horiguchi S, Watanabe J, Kato H, Baba S, Shinozaki T, Miura M, Fukuchi M, Kagaya
Y, Shirato K. Contribution of Na⫹/Ca2⫹ exchanger to the regulation of myogenic
tone in isolated rat small arteries. Acta Physiol Scand 173: 167–173, 2001.
635. Horvath B, Lenzser G, Benyo B, Nemeth T, Benko R, Iring A, Herman P, Komjati K,
Lacza Z, Sandor P, Benyo Z. Hypersensitivity to thromboxane receptor mediated
cerebral vasomotion and CBF oscillations during acute NO-deficiency in rats. PLoS
One 5: e14477, 2010.
636. Hostetter TH, Meyer TW, Rennke HG, Brenner BM. Chronic effects of dietary
protein in the rat with intact and reduced renal mass. Kidney Int 30: 509 –517, 1986.
637. Hostetter TH, Olson JL, Rennke HG, Venkatachalam MA, Brenner BM. Hyperfiltration in remnant nephrons: a potentially adverse response to renal ablation. Am J
Physiol Renal Fluid Electrolyte Physiol 241: F85–F93, 1981.
638. Hostetter TH, Troy JL, Brenner BM. Glomerular hemodynamics in experimental
diabetes mellitus. Kidney Int 19: 410 – 415, 1981.
639. Hsu CH, Slavicek JH, Kurtz TW. Segmental renal vascular resistance in the spontaneously hypertensive rat. Am J Physiol Heart Circ Physiol 242: H961–H966, 1982.
640. Hsu CH, Slavicek JM. The effect of renal perfusion pressure on renal vascular resistance in the spontaneously hypertensive rat. Pflügers Arch 393: 340 –343, 1982.
641. Hu L, Zhang Y, Lim PS, Miao Y, Tan C, McKenzie KU, Schyvens CG, Whitworth JA.
Apocynin but not L-arginine prevents and reverses dexamethasone-induced hypertension in the rat. Am J Hypertens 19: 413– 418, 2006.
620. Hirata Y, Hayakawa H, Suzuki E, Kimura K, Kikuchi K, Nagano T, Hirobe M, Omata
M. Direct measurements of endothelium-derived nitric oxide release by stimulation
of endothelin receptors in rat kidney and its alteration in salt-induced hypertension.
Circulation 91: 1229 –1235, 1995.
642. Hua JL, Kaskel FJ, Juno CJ, Moore LC, McCaughran JA Jr. Salt intake and renal
hemodynamics in immature and mature Dahl salt-sensitive (DS/JR) and salt-resistant
(DR/JR) rats. Am J Hypertens 3: 268 –273, 1990.
621. Hirst GD, van Helden DF. Ionic basis of the resting potential of submucosal arterioles
in the ileum of the guinea-pig. J Physiol 333: 53– 67, 1982.
643. Huang A, Koller A. Endothelin and prostaglandin H2 enhance arteriolar myogenic
tone in hypertension. Hypertension 30: 1210 –1215, 1997.
486
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
644. Huang A, Sun D, Koller A. Endothelial dysfunction augments myogenic arteriolar
constriction in hypertension. Hypertension 22: 913–921, 1993.
665. Ikenaga H, Bast JP, Fallet RW, Carmines PK. Exaggerated impact of ATP-sensitive K⫹
channels on afferent arteriolar diameter in diabetes mellitus. J Am Soc Nephrol 11:
1199 –1207, 2000.
645. Huang A, Sun D, Yan C, Falck JR, Kaley G. Contribution of 20-HETE to augmented
myogenic constriction in coronary arteries of endothelial NO synthase knockout
mice. Hypertension 46: 607– 613, 2005.
666. Ikenaga H, Suzuki H, Ishii N, Itoh H, Saruta T. Role of NO on pressure-natriuresis in
Wistar-Kyoto and spontaneously hypertensive rats. Kidney Int 43: 205–211, 1993.
646. Huang C, Davis G, Johns EJ. Effect of nitrendipine on autoregulation of perfusion in
the cortex and papilla of kidneys from Wistar and stroke prone spontaneously
hypertensive rats. Br J Pharmacol 111: 111–116, 1994.
667. Ikeoka K, Nishigaki K, Ohyanagi M, Faber JE. In vitro analysis of alpha-adrenoceptor
interactions with the myogenic response in resistance vessels. J Vasc Res 29: 313–
321, 1992.
647. Huang DY, Vallon V, Zimmermann H, Koszalka P, Schrader J, Osswald H. Ecto-5’nucleotidase (cd73)-dependent and -independent generation of adenosine participates in the mediation of tubuloglomerular feedback in vivo. Am J Physiol Renal Physiol
291: F282–F288, 2006.
668. Iliescu R, Cazan R, McLemore GR Jr, Venegas-Pont M, Ryan MJ. Renal blood flow and
dynamic autoregulation in conscious mice. Am J Physiol Renal Physiol 295: F734 –
F740, 2008.
648. Huang WC, Bell PD, Harvey D, Mitchell KD, Navar LG. Angiotensin influences on
tubuloglomerular feedback mechanism in hypertensive rats. Kidney Int 34: 631– 637,
1988.
649. Huang WC, Ploth DW, Bell PD, Work J, Navar LG. Bilateral renal function responses
to converting enzyme inhibitor (SQ 20,881) in two-kidney, one clip Goldblatt hypertensive rats. Hypertension 3: 285–293, 1981.
650. Huang WC, Ploth DW, Navar LG. Effects of saralasin infusion on bilateral renal
function in two-kidney, one-clip Goldblatt hypertensive rats. Clin Sci 62: 573–579,
1982.
651. Huang Y, Cheung KK. Endothelium-dependent rhythmic contractions induced by
cyclopiazonic acid in rat mesenteric artery. Eur J Pharmacol 332: 167–172, 1997.
652. Huelsemann JL, Sterzel RB, McKenzie DE, Wilcox CS. Effects of a calcium entry
blocker on blood pressure and renal function during angiotensin-induced hypertension. Hypertension 7: 374 –379, 1985.
653. Hundley WG, Renaldo GJ, Levasseur JE, Kontos HA. Vasomotion in cerebral microcirculation of awake rabbits. Am J Physiol Heart Circ Physiol 254: H67–H71, 1988.
654. Ibarrola AM, Inscho EW, Vari RC, Navar LG. Influence of adenosine receptor blockade on renal function and renal autoregulation. J Am Soc Nephrol 2: 991–999, 1991.
655. Ichihara A, Hayashi M, Hirota N, Saruta T. Superoxide inhibits neuronal nitric oxide
synthase influences on afferent arterioles in spontaneously hypertensive rats. Hypertension 37: 630 – 634, 2001.
656. Ichihara A, Hayashi M, Koura Y, Tada Y, Sugaya T, Hirota N, Saruta T. Blunted
tubuloglomerular feedback by absence of angiotensin type 1A receptor involves
neuronal NOS. Hypertension 40: 934 –939, 2002.
657. Ichihara A, Imig JD, Inscho EW, Navar LG. Cyclooxygenase-2 participates in tubular
flow-dependent afferent arteriolar tone: interaction with neuronal NOS. Am J Physiol
Renal Physiol 275: F605–F612, 1998.
658. Ichihara A, Imig JD, Navar LG. Cyclooxygenase-2 modulates afferent arteriolar responses to increases in pressure. Hypertension 34: 843– 847, 1999.
659. Ichihara A, Imig JD, Navar LG. Neuronal nitric oxide synthase-dependent afferent
arteriolar function in angiotensin II-induced hypertension. Hypertension 33: 462–
466, 1999.
660. Ichihara A, Inscho EW, Imig JD, Michel RE, Navar LG. Role of renal nerves in afferent
arteriolar reactivity in angiotensin-induced hypertension. Hypertension 29: 442– 449,
1997.
661. Ichihara A, Inscho EW, Imig JD, Navar LG. Neuronal nitric oxide synthase modulates rat renal microvascular function. Am J Physiol Renal Physiol 274: F516 –
F524, 1998.
662. Ichihara A, Navar LG. Neuronal NOS contributes to biphasic autoregulatory response during enhanced TGF activity. Am J Physiol Renal Physiol 277: F113–F120,
1999.
663. Iesaki T, Okada T, Shimada I, Yamaguchi H, Ochi R. Decrease in Ca2⫹ sensitivity as
a mechanism of hydrogen peroxide-induced relaxation of rabbit aorta. Cardiovasc
Res 31: 820 – 825, 1996.
664. Iijima K, Lin L, Nasjletti A, Goligorsky MS. Intracellular signaling pathway of endothelin-1. J Cardiovasc Pharmacol 17 Suppl 7: S146 –S149, 1991.
669. Imamura A, Mackenzie HS, Lacy ER, Hutchison FN, Fitzgibbon WR, Ploth DW.
Effects of chronic treatment with angiotensin converting enzyme inhibitor or an
angiotensin receptor antagonist in two-kidney, one-clip hypertensive rats. Kidney Int
47: 1394 –1402, 1995.
670. Imig JD. Afferent arteriolar reactivity to angiotensin II is enhanced during the early
phase of angiotensin II hypertension. Am J Hypertens 13: 810 – 818, 2000.
671. Imig JD. Eicosanoid regulation of the renal vasculature. Am J Physiol Renal Physiol 279:
F965–F981, 2000.
672. Imig JD, Falck JR, Gebremedhin D, Harder DR, Roman RJ. Elevated renovascular
tone in young spontaneously hypertensive rats. Role of cytochrome P-450. Hypertension 22: 357–364, 1993.
673. Imig JD, Falck JR, Inscho EW. Contribution of cytochrome P450 epoxygenase and
hydroxylase pathways to afferent arteirolar autorregulatory responsiveness. Br J
Pharmacol 127: 1399 –1405, 1999.
674. Imig JD, Passmore JC, Anderson GL, Jimenez AE. Chloride alters renal blood flow
autoregulation in deoxycorticosterone-treated rats. J Lab Clin Med 121: 608 – 613,
1993.
675. Imig JD, Roman RJ. Nitric oxide modulates vascular tone in preglomerular arterioles.
Hypertension 19: 770 –774, 1992.
676. Imig JD, Zou AP, Ortiz de Montellano PR, Sui Z, Roman, RJ. Cytochrome P-450
inhibitors alter afferent arteriolar responses to elevations in pressure. Am J Physiol
Heart Circ Physiol 266: H1879 –H1885, 1994.
677. Imig JD, Zou AP, Stec DE, Harder DR, Falck JR, Roman, RJ. Formation and actions of
20-hydroxyeicosatetraenoic acid in rat renal arterioles. Am J Physiol Regul Integr Comp
Physiol 270: R217–R227, 1996.
678. Ingber DE. Cellular mechanotransduction: putting all the pieces together again.
FASEB J 20: 811– 827, 2006.
679. Inoue R, Hai L, Honda A. Pathophysiological implications of transient receptor potential channels in vascular function. Curr Opin Nephrol Hypertens 17: 193–198, 2008.
680. Inoue R, Jensen LJ, Jian Z, Shi J, Hai L, Lurie AI, Henriksen FH, Salomonsson M,
Morita H, Kawarabayashi Y, Mori M, Mori Y, Ito Y. Synergistic activation of vascular
TRPC6 channel by receptor and mechanical stimulation via phospholipase C/diacylglycerol and phospholipase A2/omega-hydroxylase/20-HETE pathways. Circ Res
104: 1399 –1409, 2009.
681. Inoue R, Jensen LJ, Shi J, Morita H, Nishida M, Honda A, Ito Y. Transient receptor
potential channels in cardiovascular function and disease. Circ Res 99: 119 –131,
2006.
682. Inscho EW. P2 receptors in regulation of renal microvascular function. Am J Physiol
Renal Physiol 280: F927–F944, 2001.
683. Inscho EW. ATP, P2 receptors and the renal microcirculation. Purinergic Signal 5:
447– 460, 2009.
684. Inscho EW, Carmines PK, Cook AK, Navar LG. Afferent arteriolar responsiveness to
altered perfusion pressure in renal hypertension. Hypertension 15: 748 –752, 1990.
685. Inscho EW, Cook AK. P2 receptor-mediated afferent arteriolar vasoconstriction
during calcium blockade. Am J Physiol Renal Physiol 282: F245–F255, 2002.
686. Inscho EW, Cook AK, Imig JD, Vial C, Evans RJ. Physiological role for P2X1 receptors
in renal microvascular autoregulatory behavior. J Clin Invest 112: 1895–1905, 2003.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
487
RENAL AUTOREGULATORY MECHANISMS
687. Inscho EW, Cook AK, Imig JD, Vial C, Evans RJ. Renal autoregulation in P2X1
knockout mice. Acta Physiol Scand 181: 445– 453, 2004.
688. Inscho EW, Cook AK, Mui V, Imig JD. Calcium mobilization contributes to pressuremediated afferent arteriolar vasoconstriction. Hypertension 31: 421– 428, 1998.
689. Inscho EW, Cook AK, Murzynowski JB, Imig JD. Elevated arterial pressure impairs
autoregulation independently of AT(1) receptor activation. J Hypertens 22: 811– 818,
2004.
709. Iversen BM, Sekse I, Ofstad J. Resetting of renal blood flow autoregulation in spontaneously hypertensive rats. Am J Physiol Renal Fluid Electrolyte Physiol 252: F480 –
F486, 1987.
710. Iwai N, Kinoshita M, Shimoike H. Chromosomal mapping of quantitative trait loci
that influence renal hemodynamic functions. Circulation 100: 1923–1929, 1999.
711. Iwai N, Kurtz TW, Inagami T. Further evidence of the SA gene as a candidate gene
contributing to the hypertension in spontaneously hypertensive rat. Biochem Biophys
Res Commun 188: 64 – 69, 1992.
690. Inscho EW, Cook AK, Navar LG. Pressure-mediated vasoconstriction of juxtamedullary afferent arterioles involves P2-purinoceptor activation. Am J Physiol Renal Fluid
Electrolyte Physiol 271: F1077–F1085, 1996.
712. Iwai N, Tsujita Y, Kinoshita M. Isolation of a chromosome 1 region that contributes
to high blood pressure and salt sensitivity. Hypertension 32: 636 – 638, 1998.
691. Inscho EW, Cook AK, Webb RC, Jin LM. Rho-kinase inhibition reduces pressuremediated autoregulatory adjustments in afferent arteriolar diameter. Am J Physiol
Renal Physiol 296: F590 –F597, 2009.
713. Iwamoto T, Kita S, Zhang J, Blaustein MP, Arai Y, Yoshida S, Wakimoto K, Komuro
I, Katsuragi T. Salt-sensitive hypertension is triggered by Ca2⫹ entry via Na⫹/Ca2⫹
exchanger type-1 in vascular smooth muscle. Nat Med 10: 1193–1199, 2004.
692. Irzyniec T, Mall G, Greber D, Ritz E. Beneficial effect of nifedipine and moxonidine
on glomerulosclerosis in spontaneously hypertensive rats. A micromorphometric
study. Am J Hypertens 5: 437– 443, 1992.
714. Izzard AS, Bund SJ, Heagerty AM. Myogenic tone in mesenteric arteries from spontaneously hypertensive rats. Am J Physiol Heart Circ Physiol 270: H1–H6, 1996.
693. Ishinaga Y, Nabika T, Shimada T, Hiraoka J, Nara Y, Yamori Y. Re-evaluation of the
SA gene in spontaneously hypertensive and Wistar-Kyoto rats. Clin Exp Pharmacol
Physiol 24: 18 –22, 1997.
694. Ito I, Jarajapu YP, Grant MB, Knot HJ. Characteristics of myogenic tone in the rat
ophthalmic artery. Am J Physiol Heart Circ Physiol 292: H360 –H368, 2007.
695. Ito S. Characteristics of isolated perfused juxtaglomerular apparatus. Kidney Int Suppl
67: S46 –S48, 1998.
715. Izzard AS, Graham D, Burnham MP, Heerkens EH, Dominiczak AF, Heagerty AM.
Myogenic and structural properties of cerebral arteries from the stroke-prone spontaneously hypertensive rat. Am J Physiol Heart Circ Physiol 285: H1489 –H1494, 2003.
716. Jackson EK, Herzer WA. Angiotensin II/prostaglandin I2 interactions in spontaneously hypertensive rats. Hypertension 22: 688 – 698, 1993.
717. Jackson EK, Herzer WA. Defective modulation of angiotensin II-induced renal vasoconstriction in hypertensive rats. Hypertension 23: 329 –336, 1994.
696. Ito S, Carretero OA, Abe K. Role of nitric oxide in the control of glomerular microcirculation. Clin Exp Pharmacol Physiol 24: 578 –581, 1997.
718. Jackson KE, Jackson DW, Quadri S, Reitzell MJ, Navar LG. Inhibition of heme oxygenase augments tubular sodium reabsorption. Am J Physiol Renal Physiol 300: F941–
F946, 2011.
697. Ito S, Carretero OA, Abe K, Beierwaltes WH, Yoshinaga K. Effect of prostanoids on
renin release from rabbit afferent arterioles with and without macula densa. Kidney
Int 35: 1138 –1144, 1989.
719. Jacob F, Clark LA, Guzman PA, Osborn JW. Role of renal nerves in development of
hypertension in DOCA-salt model in rats: a telemetric approach. Am J Physiol Heart
Circ Physiol 289: H1519 –H1529, 2005.
698. Ito S, Carretero OA, Murray RD. Renin release from isolated afferent arterioles.
Kidney Int 27: 762–767, 1985.
720. Jaggar JH, Wellman GC, Heppner TJ, Porter VA, Perez GJ, Gollasch M, Kleppisch T,
Rubart M, Stevenson AS, Lederer WJ, Knot HJ, Bonev AD, Nelson MT. Ca2⫹ channels, ryanodine receptors and Ca2⫹-activated K⫹ channels: a functional unit for
regulating arterial tone. Acta Physiol Scand 164: 577–587, 1998.
699. Ito S, Johnson CS, Carretero OA. Modulation of angiotensin II-induced vasoconstriction by endothelium-derived relaxing factor in the isolated microperfused rabbit
afferent arteriole. J Clin Invest 87: 1656 –1663, 1991.
700. Ito S, Juncos LA, Carretero OA. Pressure-induced constriction of the afferent arteriole of spontaneously hypertensive rats. Hypertension 19: II164 –II167, 1992.
701. Ito S, Juncos LA, Nushiro N, Johnson CS, Carretero OA. Endothelium-derived
relaxing factor modulates endothelin action in afferent arterioles. Hypertension 17:
1052–1056, 1991.
702. Ito S, Ren Y. Evidence for the role of nitric oxide in macula densa control of glomerular hemodynamics. J Clin Invest 92: 1093–1098, 1993.
703. Itoh S, Carretero OA. An in vitro approach to the study of macula densa-mediated
glomerular hemodynamics. Kidney Int 38: 1206 –1210, 1990.
704. Iversen BM, Amann K, Kvam FI, Wang X, Ofstad J. Increased glomerular capillary
pressure and size mediate glomerulosclerosis in SHR juxtamedullary cortex. Am J
Physiol Renal Physiol 274: F365–F373, 1998.
705. Iversen BM, Heyeraas KJ, Sekse I, Andersen KJ, Ofstad J. Autoregulation of renal
blood flow in two-kidney, one-clip hypertensive rats. Am J Physiol Renal Fluid Electrolyte Physiol 251: F245–F250, 1986.
706. Iversen BM, Kvam FI, Matre K, Morkrid L, Horvei G, Bagchus W, Grond J, Ofstad J.
Effect of mesangiolysis on autoregulation of renal blood flow and glomerular filtration rate in rats. Am J Physiol Renal Fluid Electrolyte Physiol 262: F361–F366, 1992.
721. Jaimes EA, Sweeney C, Raij L. Effects of the reactive oxygen species hydrogen
peroxide and hypochlorite on endothelial nitric oxide production. Hypertension 38:
877– 883, 2001.
722. Janssen BJ, Malpas SC, Burke SL, Head GA. Frequency-dependent modulation of
renal blood flow by renal nerve activity in conscious rabbits. Am J Physiol Regul Integr
Comp Physiol 273: R597–R608, 1997.
723. Janssen BJ, Oosting J, Slaaf DW, Persson PB, Struijker-Boudier HA. Hemodynamic
basis of oscillations in systemic arterial pressure in conscious rats. Am J Physiol Heart
Circ Physiol 269: H62–H71, 1995.
724. Jarajapu YP, Knot HJ. Role of phospholipase C in development of myogenic tone in
rat posterior cerebral arteries. Am J Physiol Heart Circ Physiol 283: H2234 –H2238,
2002.
725. Jarajapu YP, Knot HJ. Relative contribution of Rho kinase and protein kinase C to
myogenic tone in rat cerebral arteries in hypertension. Am J Physiol Heart Circ Physiol
289: H1917–H1922, 2005.
726. Jensen BL, Ellekvist P, Skott O. Chloride is essential for contraction of afferent
arterioles after agonists and potassium. Am J Physiol Renal Physiol 272: F389 –F396,
1997.
727. Jensen BL, Skott O. Blockade of chloride channels by DIDS stimulates renin release
and inhibits contraction of afferent arterioles. Am J Physiol Renal Fluid Electrolyte
Physiol 270: F718 –F727, 1996.
707. Iversen BM, Kvam FI, Matre K, Ofstad J. Resetting of renal blood autoregulation
during acute blood pressure reduction in hypertensive rats. Am J Physiol Regul Integr
Comp Physiol 275: R343–R349, 1998.
728. Jernigan NL, Drummond HA. Vascular ENaC proteins are required for renal myogenic constriction. Am J Physiol Renal Physiol 289: F891–F901, 2005.
708. Iversen BM, Kvam FI, Morkrid L, Sekse I, Ofstad J. Effect of cyclooxygenase inhibition
on renal blood flow autoregulation in SHR. Am J Physiol Renal Fluid Electrolyte Physiol
263: F534 –F539, 1992.
729. Jernigan NL, Drummond HA. Myogenic vasoconstriction in mouse renal interlobar
arteries: role of endogenous beta and gammaENaC. Am J Physiol Renal Physiol 291:
F1184 –F1191, 2006.
488
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
730. Jernigan NL, Lamarca B, Speed J, Galmiche L, Granger JP, Drummond HA. Dietary
salt enhances benzamil-sensitive component of myogenic constriction in mesenteric
arteries. Am J Physiol Heart Circ Physiol 294: H409 –H420, 2008.
751. Just A, Arendshorst WJ. A novel mechanism of renal blood flow autoregulation and
the autoregulatory role of A1 adenosine receptors in mice. Am J Physiol Renal Physiol
293: F1489 –F1500, 2007.
731. Jernigan NL, Speed J, Lamarca B, Granger JP, Drummond HA. Angiotensin II regulation of renal vascular ENaC proteins. Am J Hypertens 22: 593–597, 2009.
752. Just A, Arendshorst WJ. The role of reactive oxygen species in renal blood flow
autoregulation (Abstract). FASEB J 22: 761.F1489 –19, 2008.
732. Ji X, Naito Y, Weng H, Endo K, Ma X, Iwai N. P2X7 deficiency attenuates hypertension and renal injury in deoxycorticosterone acetate-salt hypertension. Am J
Physiol Renal Physiol 303: F1207–F1215, 2012.
753. Just A, Ehmke H, Toktomambetova L, Kirchheim HR. Dynamic characteristics and
underlying mechanisms of renal blood flow autoregulation in the conscious dog. Am
J Physiol Renal Physiol 280: F1062–F1071, 2001.
733. Jiang J, Roman RJ. Lovastatin prevents development of hypertension in spontaneously hypertensive rats. Hypertension 30: 968 –974, 1997.
754. Just A, Ehmke H, Wittmann U, Kirchheim HR. Tonic and phasic influences of nitric
oxide on renal blood flow autoregulation in conscious dogs. Am J Physiol Renal Physiol
276: F442–F449, 1999.
734. Jin C, Hu C, Polichnowski A, Mori T, Skelton M, Ito S, Cowley AW Jr. Effects of renal
perfusion pressure on renal medullary hydrogen peroxide and nitric oxide production. Hypertension 53: 1048 –1053, 2009.
755. Just A, Ehmke H, Wittmann U, Kirchheim HR. Role of angiotensin II in dynamic renal
blood flow autoregulation of the conscious dog. J Physiol 538: 167–177, 2002.
735. Jin L, Beswick RA, Yamamoto T, Palmer T, Taylor TA, Pollock JS, Pollock DM,
Brands MW, Webb RC. Increased reactive oxygen species contributes to kidney
injury in mineralocorticoid hypertensive rats. J Physiol Pharmacol 57: 343–357, 2006.
736. Jin XH, McGrath HE, Gildea JJ, Siragy HM, Felder RA, Carey RM. Renal interstitial
guanosine cyclic 3’,5’-monophosphate mediates pressure-natriuresis via protein kinase G. Hypertension 43: 1133–1139, 2004.
756. Just A, Kurtz L, de Wit C, Wagner C, Kurtz A, Arendshorst WJ. Connexin 40
mediates the tubuloglomerular feedback contribution to renal blood flow autoregulation. J Am Soc Nephrol 20: 1577–1585, 2009.
757. Just A, Olson AJ, Falck JR, Arendshorst WJ. NO and NO-independent mechanisms
mediate ETB receptor buffering of ET-1-induced renal vasoconstriction in the rat.
Am J Physiol Regul Integr Comp Physiol 288: R1168 –R1177, 2005.
737. Jin XH, Siragy HM, Carey RM. Renal interstitial cGMP mediates natriuresis by direct
tubule mechanism. Hypertension 38: 309 –316, 2001.
758. Just A, Olson AJ, Whitten CL, Arendshorst WJ. Superoxide mediates acute renal
vasoconstriction produced by angiotensin II and catecholamines by a mechanism
independent of nitric oxide. Am J Physiol Heart Circ Physiol 292: H83–H92, 2007.
738. Jobs A, Schmidt K, Schmidt VJ, Lubkemeier I, van Veen TA, Kurtz A, Willecke K, de
Wit C. Defective Cx40 maintains Cx37 expression but intact Cx40 is crucial for
conducted dilations irrespective of hypertension. Hypertension 60: 1422–1429,
2012.
759. Just A, Whitten CL, Arendshorst WJ. Reactive oxygen species participate in acute
renal vasoconstrictor responses induced by ETA and ETB receptors. Am J Physiol
Renal Physiol 294: F719 –F728, 2008.
739. Johnson PC. The myogenic response. In: Handbook of Physiology: Microcirculation,
edited by Tuma R, Duran WN, and Ley N. Boston: Academic, 2008.
760. Just A, Wittmann U, Ehmke H, Kirchheim HR. Autoregulation of renal blood flow in
the conscious dog and the contribution of the tubuloglomerular feedback. J Physiol
506: 275–290, 1998.
740. Johnson PC, Intaglietta M. Contributions of pressure and flow sensitivity to autoregulation in mesenteric arterioles. Am J Physiol 231: 1686 –1698, 1976.
741. Johnson RA, Belmonte A, Fan NY, Lavesa M, Nasjletti A, Stier CT Jr. Effect of
ifetroban, a thromboxane A2 receptor antagonist, in stroke-prone spontaneously
hypertensive rats. Clin Exp Hypertens 18: 171–188, 1996.
742. Johnson RJ, Gordon KL, Giachelli C, Kurth T, Skelton MM, Cowley AW Jr. Tubulointerstitial injury and loss of nitric oxide synthases parallel the development of hypertension in the Dahl-SS rat. J Hypertens 18: 1497–1505, 2000.
743. Johnson RP, El-Yazbi AF, Takeya K, Walsh EJ, Walsh MP, Cole WC. Ca2⫹ sensitization via phosphorylation of myosin phosphatase targeting subunit at threonine-855
by Rho kinase contributes to the arterial myogenic response. J Physiol 587: 2537–
2553, 2009.
744. Jones DR, Dowd DA. Development of elevated blood pressure in young genetically
hypertensive rats. Life Sci 9: 247–250, 1970.
745. Jose PA, Slotkoff LM, Montgomery S, Calcagno PL, Eisner G. Autoregulation of renal
blood flow in the puppy. Am J Physiol 229: 983–988, 1975.
746. Judy WV, Watanabe AM, Murphy WR, Aprison BS, Yu PL. Sympathetic nerve activity
and blood pressure in normotensive backcross rats genetically related to the spontaneously hypertensive rat. Hypertension 1: 598 – 604, 1979.
747. Juncos LA, Garvin J, Carretero OA, Ito S. Flow modulates myogenic responses in
isolated microperfused rabbit afferent arterioles via endothelium-derived nitric oxide. J Clin Invest 95: 2741–2748, 1995.
748. Just A. Mechanisms of renal blood flow autoregulation: dynamics and contributions.
Am J Physiol Regul Integr Comp Physiol 292: R1–R17, 2007.
749. Just A, Arendshorst WJ. Dynamics and contribution of mechanisms mediating renal
blood flow autoregulation. Am J Physiol Regul Integr Comp Physiol 285: R619 –R631,
2003.
750. Just A, Arendshorst WJ. Nitric oxide blunts myogenic autoregulation in rat renal but
not skeletal muscle circulation via tubuloglomerular feedback. J Physiol 569: 959 –
974, 2005.
761. Kaiser G, Ackermann R, Sioufi A. Pharmacokinetics of a new angiotensin-converting
enzyme inhibitor, benazepril hydrochloride, in special populations. Am Heart J 117:
746 –751, 1989.
762. Kallskog O, Marsh DJ. TGF-initiated vascular interactions between adjacent
nephrons in the rat kidney. Am J Physiol Renal Fluid Electrolyte Physiol 259: F60 –F64,
1990.
763. Kaloyanides GJ, Ahrens RE, Shepherd JA, DiBona GF. Inhibition of prostaglandin E2
secretion. Failure to abolish autoregulation in the isolated dog kidney. Circ Res 38:
67–73, 1976.
764. Kaloyanides GJ, Bastron RD, DiBona GF. Impaired autoregulation of blood flow and
glomerular filtration rate in the isolated dog kidney depleted of renin. Circ Res 35:
400 – 412, 1974.
765. Kanagawa R, Wada T, Sanada T, Ojima M, Inada Y. Regional hemodynamic effects of
candesartan cilexetil (TCV-116), an angiotensin II AT1-receptor antagonist, in conscious spontaneously hypertensive rats. Jpn J Pharmacol 73: 185–190, 1997.
766. Kang J, Kang N, Lovatt D, Torres A, Zhao Z, Lin J, Nedergaard M. Connexin 43
hemichannels are permeable to ATP. J Neurosci 28: 4702– 4711, 2008.
767. Karam H, Clozel JP, Bruneval P, Gonzalez MF, Menard J. Contrasting effects of
selective T- and L-type calcium channel blockade on glomerular damage in DOCA
hypertensive rats. Hypertension 34: 673– 678, 1999.
768. Karibe A, Watanabe J, Horiguchi S, Takeuchi M, Suzuki S, Funakoshi M, Katoh H,
Keitoku M, Satoh S, Shirato K. Role of cytosolic Ca2⫹ and protein kinase C in
developing myogenic contraction in isolated rat small arteries. Am J Physiol Heart Circ
Physiol 272: H1165–H1172, 1997.
769. Karlsen FM, Andersen CB, Leyssac PP, Holstein-Rathlou NH. Dynamic autoregulation and renal injury in Dahl rats. Hypertension 30: 975–983, 1997.
770. Karlsen FM, Leyssac PP, Holstein-Rathlou NH. Tubuloglomerular feedback in Dahl
rats. Am J Physiol Regul Integr Comp Physiol 274: R1561–R1569, 1998.
771. Kashihara T, Nakayama K, Matsuda T, Baba A, Ishikawa T. Role of Na⫹/Ca2⫹
exchanger-mediated Ca2⫹ entry in pressure-induced myogenic constriction in rat
posterior cerebral arteries. J Pharmacol Sci 110: 218 –222, 2009.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
489
RENAL AUTOREGULATORY MECHANISMS
772. Kasiske BL, Cleary MP, O’Donnell MP, Keane WF. Effects of genetic obesity on renal
structure and function in the Zucker rat. J Lab Clin Med 106: 598 – 604, 1985.
792. Khraibi AA, Knox FG. Effect of renal decapsulation on renal interstitial hydrostatic
pressure and natriuresis. Am J Physiol Regul Integr Comp Physiol 257: R44 –R48, 1989.
773. Kasiske BL, O’Donnell MP, Keane WF. The Zucker rat model of obesity, insulin
resistance, hyperlipidemia, and renal injury. Hypertension 19: I110 –I115, 1992.
793. Kim EC, Ahn DS, Yeon SI, Lim M, Lee YH. Epithelial Na⫹ channel proteins are
mechanotransducers of myogenic constriction in rat posterior cerebral arteries. Exp
Physiol 97: 544 –555, 2012.
774. Kastner PR, Hall JE, Guyton AC. Control of glomerular filtration rate: role of intrarenally formed angiotensin II. Am J Physiol Renal Fluid Electrolyte Physiol 246: F897–
F906, 1984.
775. Katholi RE. Renal nerves in the pathogenesis of hypertension in experimental animals
and humans. Am J Physiol Renal Fluid Electrolyte Physiol 245: F1–F14, 1983.
776. Katholi RE, Naftilan AJ, Bishop SP, Oparil S. Role of the renal nerves in the maintenance of DOCA-salt hypertension in the rat: influence on the renal vasculature and
sodium excretion. Hypertension 5: 427– 435, 1983.
777. Kauser K, Clark JE, Masters BS, Ortiz de Montellano PR, Ma YH, Harder DR, Roman
RJ. Inhibitors of cytochrome P-450 attenuate the myogenic response of dog renal
arcuate arteries. Circ Res 68: 1154 –1163, 1991.
778. Kawabata M, Ogawa T, Takabatake T. Control of rat glomerular microcirculation by
juxtaglomerular adenosine A1 receptors. Kidney Int Suppl 67: S228 –S230, 1998.
779. Kawabe K, Watanabe TX, Shiono K, Sokabe H. Influence on blood pressure of renal
isografts between spontaneously hypertensive and normotensive rats, utilizing the
F1 hybrids. Jpn Heart J 19: 886 – 894, 1978.
780. Kawada N, Dennehy K, Solis G, Modlinger P, Hamel R, Kawada JT, Aslam S,
MOriyama T, Imai E, Welch WJ, Wilcox CS. TP receptors regulate renal hemodynamics during angiotensin II slow pressor response. Am J Physiol Renal Physiol 287:
F753–F759, 2004.
781. Kawata T, Hashimoto S, Koike T. The angiotensin receptor antagonist 2-ethoxy-1
[[2’-(1H-tetrazol-5-yl) biphenyl-4-yl]methyl]-1H-benzimidazole-7-carboxylic acid
(CV11974) attenuates the tubuloglomerular feedback response during NO synthase
blockade in rats. J Pharmacol Exp Ther 277: 572–577, 1996.
782. Kawata T, Mimuro T, Onuki T, Tsuchiya K, Nihei H, Koike T. The K(ATP) channel
opener nicorandil: effect on renal hemodynamics in spontaneously hypertensive and
Wistar Kyoto rats. Kidney Int Suppl 67: S231–S233, 1998.
783. Keller M, Lidington D, Vogel L, Peter BF, Sohn HY, Pagano PJ, Pitson S, Spiegel S,
Pohl U, Bolz SS. Sphingosine kinase functionally links elevated transmural pressure
and increased reactive oxygen species formation in resistance arteries. FASEB J 20:
702–704, 2006.
784. Kelly-Cobbs AI, Elgebaly MM, Li W, Ergul A. Pressure-independent cerebrovascular
remodelling and changes in myogenic reactivity in diabetic Goto-Kakizaki rat in
response to glycaemic control. Acta Physiol 203: 245–251, 2011.
785. Kelly-Cobbs AI, Prakash R, Coucha M, Knight RA, Li W, Ogbi SN, Johnson M, Ergul
A. Cerebral myogenic reactivity and blood flow in type 2 diabetic rats: role of
peroxynitrite in hypoxia-mediated loss of myogenic tone. J Pharmacol Exp Ther 342:
407– 415, 2012.
794. Kim EC, Choi SK, Lim M, Yeon SI, Lee YH. Role of endogenous ENaC and TRP
channels in the myogenic response of rat posterior cerebral arteries. PLoS One 8:
e84194, 2013.
795. Kimmelstiel P, Wilson C. Inflammatory lesions in the glomeruli in pyelonephritis in
relation to hypertension and renal insufficiency. Am J Pathol 12: 99 –106, 1936.
796. Kimura K, Nagai R, Sakai T, Aikawa M, Kuro-o M, Kobayashi N, Shirato I, Inagami T,
Oshi M, Suzuki N. Diversity and variability of smooth muscle phenotypes of renal
arterioles as revealed by myosin isoform expression. Kidney Int 48: 372–382, 1995.
797. Kinoshita Y, Knox FG. Role of prostaglandins in proximal tubule sodium reabsorption: response to elevated renal interstitial hydrostatic pressure. Circ Res 64: 1013–
1018, 1989.
798. Kinoshita Y, Knox FG. Response of superficial proximal convoluted tubule to decreased and increased renal perfusion pressure. In vivo microperfusion study in rats.
Circ Res 66: 1184 –1189, 1990.
799. Kirchheim HR, Ehmke H, Hackenthal E, Lowe W, Persson P. Autoregulation of renal
blood flow, glomerular filtration rate and renin release in conscious dogs. Pflügers
Arch 410: 441– 449, 1987.
800. Kirchner KA. Hypertension in hemodialysis patients: more questions than answers.
Am J Kidney Dis 30: 577–578, 1997.
801. Kirk KL, Bell PD, Barfuss DW, Ribadeneira M. Direct visualization of the isolated and
perfused macula densa. Am J Physiol Renal Fluid Electrolyte Physiol 248: F890 –F894,
1985.
802. Kirton CA, Loutzenhiser R. Alterations in basal protein kinase C activity modulate
renal afferent arteriolar myogenic reactivity. Am J Physiol Heart Circ Physiol 275:
H467–H475, 1998.
803. Klahr S, Levey AS, Beck GJ, Caggiula AW, Hunsicker L, Kusek JW, Striker G. The
effects of dietary protein restriction and blood-pressure control on the progression
of chronic renal disease. Modification of Diet in Renal Disease Study Group. N Engl
J Med 330: 877– 884, 1994.
804. Kleinbongard P, Schleiger A, Heusch G. Characterization of vasomotor responses in
different vascular territories of C57BL/6J mice. Exp Biol Med 238: 1180 –1191, 2013.
805. Kleinstreuer N, David T, Plank MJ, Endre Z. Dynamic myogenic autoregulation in
the rat kidney: a whole-organ model. Am J Physiol Renal Physiol 294: F1453–
F1464, 2008.
806. Kline RL. Renal nerves and experimental hypertension: evidence and controversy.
Can J Physiol Pharmacol 65: 1540 –1547, 1987.
786. Kett MM, Alcorn D, Bertram JF, Anderson WP. Enalapril does not prevent renal
arterial hypertrophy in spontaneously hypertensive rats. Hypertension 25: 335–342,
1995.
807. Kline RL, Kelton PM, Mercer PF. Effect of renal denervation on the development of
hypertension in spontaneously hypertensive rats. Can J Physiol Pharmacol 56: 818 –
822, 1978.
787. Kett MM, Bergstrom G, Alcorn D, Bertram JF, Anderson WP. Renal vascular resistance properties and glomerular protection in early established SHR hypertension. J
Hypertens 19: 1505–1512, 2001.
808. Kline RL, Liu F. Modification of pressure natriuresis by long-term losartan in spontaneously hypertensive rats. Hypertension 24: 467– 473, 1994.
788. Khan NJ, Hampton JA, Lacher DA, Rapp JP, Gohara AF, Goldblatt PJ. Morphometric
evaluation of the renal arterial system of Dahl salt-sensitive and salt-resistant rats on
a high salt diet. I. Interlobar and arcuate arteries. Lab Invest 57: 714 –723, 1987.
789. Khraibi AA, Granger JP, Haas JA, Burnett JC Jr, Knox FG. Intrarenal pressures during
direct inhibition of sodium transport. Am J Physiol Regul Integr Comp Physiol 263:
R1182–R1186, 1992.
790. Khraibi AA, Haas JA, Knox FG. Effect of renal perfusion pressure on renal interstitial
hydrostatic pressure in rats. Am J Physiol Renal Fluid Electrolyte Physiol 256: F165–
F170, 1989.
791. Khraibi AA, Knox FG. Effect of acute renal decapsulation on pressure natriuresis in
SHR and WKY rats. Am J Physiol Renal Fluid Electrolyte Physiol 257: F785–F789, 1989.
490
809. Kline RL, McLennan GP. Chronic hydralazine treatment alters the acute pressurenatriuresis curve in young spontaneously hypertensive rats. Can J Physiol Pharmacol
69: 164 –169, 1991.
810. Knot HJ, Standen NB, Nelson MT. Ryanodine receptors regulate arterial diameter
and wall [Ca2⫹] in cerebral arteries of rat via Ca2⫹-dependent K⫹ channels. J Physiol
508: 211–221, 1998.
811. Knudsen T, Elmer H, Knudsen MH, Holstein-Rathlou NH, Stoustrup J. Dynamic
modeling of renal blood flow in Dahl hypertensive and normotensive rats. IEEE Trans
Biomed Eng 51: 689 – 697, 2004.
812. Kobori H, Nangaku M, Navar LG, Nishiyama A. The intrarenal renin-angiotensin
system: from physiology to the pathobiology of hypertension and kidney disease.
Pharmacol Rev 59: 251–287, 2007.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
813. Kobrin I, Pegram BL, Frohlich ED. Acute pressure increase and intrarenal hemodynamics in conscious WKY and SHR rats. Am J Physiol Heart Circ Physiol 249: H1114 –
H1118, 1985.
834. Kremer SG, Zeng W, Hurst R, Ning T, Whiteside C, Skorecki KL. Chloride is required for receptor-mediated divalent cation entry in mesangial cells. J Cell Physiol
162: 15–25, 1995.
814. Koch KM, Aynedjian HS, Bank N. Effect of acute hypertension on sodium reabsorption by the proximal tubule. J Clin Invest 47: 1696 –1709, 1968.
835. Kreye VA, Bauer PK, Villhauer I. Evidence for furosemide-sensitive active chloride
transport in vascular smooth muscle. Eur J Pharmacol 73: 91–95, 1981.
815. Koeners MP, Racasan S, Koomans HA, Joles JA, Braam B. Nitric oxide, superoxide
and renal blood flow autoregulation in SHR after perinatal L-arginine and antioxidants. Acta Physiol 190: 329 –338, 2007.
836. Krishnamoorthy G, Sonkusare SK, Heppner TJ, Nelson MT. Opposing roles of
smooth muscle BK channels and ryanodine receptors in the regulation of nerveevoked constriction of mesenteric resistance arteries. Am J Physiol Heart Circ Physiol
306: H981–H988, 2014.
816. Koenig JB, Jaffe IZ. Direct role for smooth muscle cell mineralocorticoid receptors in
vascular remodeling: novel mechanisms and clinical implications. Curr Hypertens Rep
16: 427– 0427, 2014.
837. Kriz W. TRPC6: a new podocyte gene involved in focal segmental glomerulosclerosis. Trends Mol Med 11: 527–530, 2005.
817. Koenigsberger M, Sauser R, Beny JL, Meister JJ. Role of the endothelium on arterial
vasomotion. Biophys J 88: 3845–3854, 2005.
818. Koenigsberger M, Sauser R, Beny JL, Meister JJ. Effects of arterial wall stress on
vasomotion. Biophys J 91: 1663–1674, 2006.
819. Koenigsberger M, Sauser R, Seppey D, Beny JL, Meister JJ. Calcium dynamics and
vasomotion in arteries subject to isometric, isobaric, and isotonic conditions. Biophys
J 95: 2728 –2738, 2008.
820. Koltsova SV, Kotelevtsev SV, Tremblay J, Hamet P, Orlov SN. Excitation-contraction
coupling in resistance mesenteric arteries: evidence for NKCC1-mediated pathway.
Biochem Biophys Res Commun 379: 1080 –1083, 2009.
821. Komers R, Anderson S. Paradoxes of nitric oxide in the diabetic kidney. Am J Physiol
Renal Physiol 284: F1121–F1137, 2003.
822. Komlosi P, Bell PD, Zhang ZR. Tubuloglomerular feedback mechanisms in nephron
segments beyond the macula densa. Curr Opin Nephrol Hypertens 18: 57– 62, 2009.
823. Kompanowska-Jezierska E, Berndt TJ, Knox FG. Prostaglandin E2 concentrations in
rat renal cortical and medullary interstitium: effect of volume expansion and renal
perfusion pressure. Acta Physiol Scand 172: 287–289, 2001.
824. Kost CK Jr, Herzer WA, Li P, Jackson EK. Vascular reactivity to angiotensin II is
selectively enhanced in the kidneys of spontaneously hypertensive rats. J Pharmacol
Exp Ther 269: 82– 88, 1994.
825. Kost CK, Li P, Williams DS, Jackson EK. Renal vascular responses to angiotensin II in
conscious spontaneously hypertensive and normotensive rats. J Cardiovasc Pharmacol
31: 854 – 861, 1998.
826. Kotchen TA, Piering AW, Cowley AW, Grim CE, Gaudet D, Hamet P, Kaldunski ML,
Kotchen JM, Roman RJ. Glomerular hyperfiltration in hypertensive African Americans. Hypertension 35: 822– 826, 2000.
827. Kotecha N, Hill MA. Myogenic contraction in rat skeletal muscle arterioles: smooth
muscle membrane potential and Ca2⫹ signaling. Am J Physiol Heart Circ Physiol 289:
H1326 –H1334, 2005.
828. Kovacs G, Komlosi P, Fuson A, Peti-Peterdi J, Rosivall L, Bell PD. Neuronal nitric
oxide synthase: its role and regulation in macula densa cells. J Am Soc Nephrol 14:
2475–2483, 2003.
838. Kriz W, Hartmann I, Hosser H, Hahnel B, Kranzlin B, Provoost A, Gretz N. Tracer
studies in the rat demonstrate misdirected filtration and peritubular filtrate spreading in nephrons with segmental glomerulosclerosis. J Am Soc Nephrol 12: 496 –506,
2001.
839. Kriz W, Hosser H, Hahnel B, Gretz N, Provoost AP. From segmental glomerulosclerosis to total nephron degeneration and interstitial fibrosis: a histopathological
study in rat models and human glomerulopathies. Nephrol Dial Transplant 13: 2781–
2798, 1998.
840. Kriz W, Hosser H, Hahnel B, Simons JL, Provoost AP. Development of vascular
pole-associated glomerulosclerosis in the Fawn-hooded rat. J Am Soc Nephrol 9:
381–396, 1998.
841. Kuijpers MH, de Jong W. Spontaneous hypertension in the fawn-hooded rat: a
cardiovascular disease model. J Hypertens Suppl 4: S41–S44, 1986.
842. Kuijpers MH, Gruys E. Spontaneous hypertension and hypertensive renal disease in
the fawn-hooded rat. Br J Exp Pathol 65: 181–190, 1984.
843. Kuijpers MH, Provoost AP, de Jong W. Development of hypertension and proteinuria with age in fawn-hooded rats. Clin Exp Pharmacol Physiol 13: 201–209, 1986.
844. Kunau RT Jr, Lameire NH. The effect of an acute increase in renal perfusion pressure
on sodium transport in the rat kidney. Circ Res 39: 689 – 695, 1976.
845. Kur J, Bankhead P, Scholfield CN, Curtis TM, McGeown JG. Ca2⫹ sparks promote
myogenic tone in retinal arterioles. Br J Pharmacol 168: 1675–1686, 2013.
846. Kurtz A. Renin release: sites, mechanisms, control. Annu Rev Physiol 73: 377–399,
2011.
847. Kurtz A, Penner R. Angiotensin II induces oscillations of intracellular calcium and
blocks anomalous inward rectifying potassium current in mouse renal juxtaglomerular cells. Proc Natl Acad Sci USA 86: 3423–3427, 1989.
848. Kurtz TW, Montano M, Chan L, Kabra P. Molecular evidence of genetic heterogeneity in Wistar-Kyoto rats: implications for research with the spontaneously hypertensive rat. Hypertension 13: 188 –192, 1989.
849. Kurtz TW, Morris RC Jr. Biological variability in Wistar-Kyoto rats. Implications for
research with the spontaneously hypertensive rat. Hypertension 10: 127–131, 1987.
850. Kurtz TW, Simonet L, Kabra PM, Wolfe S, Chan L, Hjelle BL. Cosegregation of the
renin allele of the spontaneously hypertensive rat with an increase in blood pressure.
J Clin Invest 85: 1328 –1332, 1990.
829. Kovacs G, Peti-Peterdi J, Rosivall L, Bell PD. Angiotensin II directly stimulates macula
densa Na-2Cl-K cotransport via apical AT(1) receptors. Am J Physiol Renal Physiol
282: F301–F306, 2002.
851. Kusche-Vihrog K, Tarjus A, Fels J, Jaisser F. The epithelial Na⫹ channel: a new player
in the vasculature. Curr Opin Nephrol Hypertens 23: 143–148, 2014.
830. Kozma F, Johnson RA, Zhang F, Yu C, Tong X, Nasjletti A. Contribution of endogenous carbon monoxide to regulation of diameter in resistance vessels. Am J Physiol
Regul Integr Comp Physiol 276: R1–R8, 1999.
852. Kvam FI, Ofstad J, Iversen BM. Effects of antihypertensive drugs on autoregulation of
RBF and glomerular capillary pressure in SHR. Am J Physiol Renal Physiol 275: F576 –
F584, 1998.
831. Kramp RA, Fourmanoir P, Caron N. Endothelin resets renal blood flow autoregulatory efficiency during acute blockade of NO in the rat. Am J Physiol Renal Physiol 281:
F1132–F1140, 2001.
853. Kvam FI, Ofstad J, Iversen BM. Role of nitric oxide in the autoregulation of renal
blood flow and glomerular filtration rate in aging spontaneously hypertensive rats.
Kidney Blood Press Res 23: 376 –384, 2000.
832. Kramp RA, Fourmanoir P, Ladriere L, Joly E, Gerbaux C, El HA, Caron N. Effects of
Ca2⫹ channel activity on renal hemodynamics during acute attenuation of NO synthesis in the rat. Am J Physiol Renal Physiol 278: F561–F569, 2000.
854. Kvandal P, Landsverk SA, Bernjak A, Stefanovska A, Kvernmo HD, Kirkeboen KA.
Low-frequency oscillations of the laser Doppler perfusion signal in human skin.
Microvasc Res 72: 120 –127, 2006.
833. Kramp RA, Genard J, Fourmanoir P, Caron N, Laekeman G, Herman A. Renal
hemodynamics and blood flow autoregulation during acute cyclooxygenase inhibition in male rats. Am J Physiol Renal Fluid Electrolyte Physiol 268: F468 –F479, 1995.
855. Lagaud GJ, Gaudreault N, Moore ED, Van Breemen C, Laher I. Pressure-dependent
myogenic constriction of cerebral arteries occurs independently of voltage-dependent activation. Am J Physiol Heart Circ Physiol 283: H2187–H2195, 2002.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
491
RENAL AUTOREGULATORY MECHANISMS
856. Lagaud GJ, Karicheti V, Knot HJ, Christ GJ, Laher I. Inhibitors of gap junctions
attenuate myogenic tone in cerebral arteries. Am J Physiol Heart Circ Physiol 283:
H2177–H2186, 2002.
877. Layton AT, Bowen M, Wen A, Layton HE. Feedback-mediated dynamics in a model
of coupled nephrons with compliant thick ascending limbs. Math Biosci 230: 115–
127, 2011.
857. Lagaud GJ, Lam E, Lui A, Van Breemen C, Laher I. Nonspecific inhibition of myogenic
tone by PD98059, a MEK1 inhibitor, in rat middle cerebral arteries. Biochem Biophys
Res Commun 257: 523–527, 1999.
878. Le Cras TD, Kim DH, Markham NE, Abman AS. Early abnormalities of pulmonary
vascular development in the Fawn-Hooded rat raised at Denver’s altitude. Am J
Physiol Lung Cell Mol Physiol 279: L283–L291, 2000.
858. Lagaud GJ, Masih-Khan E, Kai S, Van Breemen C, Dube GP. Influence of type II
diabetes on arterial tone and endothelial function in murine mesenteric resistance
arteries. J Vasc Res 38: 578 –589, 2001.
879. Le Gal L, Alonso F, Wagner C, Germain S, Haefliger DN, Meda P, Haefliger JA.
Restoration of connexin 40 (Cx40) in renin-producing cells reduces the hypertension
of Cx40 null mice. Hypertension 63: 1198 –1204, 2014.
859. Lagaud GJ, Skarsgard PL, Laher I, Van Breemen C. Heterogeneity of endotheliumdependent vasodilation in pressurized cerebral and small mesenteric resistance arteries of the rat. J Pharmacol Exp Ther 290: 832– 839, 1999.
880. Ledingham JM, Simpson FO, Hamada M. Salt appetite, body sodium, handling of a
NaCl load, renin, and aldosterone in genetically and spontaneously hypertensive
rats. J Cardiovasc Pharmacol 16 Suppl 7: S6 –S8, 1990.
860. Lai EY, Luo Z, Onozato ML, Rudolph EH, Solis G, Jose PA, Wellstein A, Aslam S,
Quinn MT, Griendling K, Le T, Li P, Palm F, Welch WJ, Wilcox CS. Effects of the
antioxidant drug tempol on renal oxygenation in mice with reduced renal mass. Am
J Physiol Renal Physiol 303: F64 –F74, 2012.
881. Leffler CW, Parfenova H, Jaggar JH, Wang R. Carbon monoxide and hydrogen
sulfide: gaseous messengers in cerebrovascular circulation. J Appl Physiol 100: 1065–
1076, 2006.
861. Lai EY, Onozato ML, Solis G, Aslam S, Welch WJ, Wilcox CS. Myogenic responses of
mouse isolated perfused renal afferent arterioles: effects of salt intake and reduced
renal mass. Hypertension 55: 983–989, 2010.
882. Leh S, Hultstrom M, Rosenberger C, Iversen BM. Afferent arteriolopathy and glomerular collapse but not segmental sclerosis induce tubular atrophy in old spontaneously hypertensive rats. Virchows Arch 459: 99 –108, 2011.
862. Lai EY, Patzak A, Persson AE, Carlstrom M. Angiotensin II enhances the afferent
arteriolar response to adenosine through increases in cytosolic calcium. Acta Physiol
196: 435– 445, 2009.
883. Leong CL, Anderson WP, O’Connor PM, Evans RG. Evidence that renal arterialvenous oxygen shunting contributes to dynamic regulation of renal oxygenation. Am
J Physiol Renal Physiol 292: F1726 –F1733, 2007.
863. Lai EY, Patzak A, Steege A, Mrowka R, Brown R, Spielmann N, Persson PB, Fredholm BB, Persson AE. Contribution of adenosine receptors in the control of arteriolar tone and adenosine-angiotensin II interaction. Kidney Int 70: 690 – 698, 2006.
884. Leong PKK, Zhang Y, Yang LE, Holstein-Rathlou NH, McDonough AA. Diuretic
response to acute hypertension is blunted during angiotensin II clamp. Am J Physiol
Regul Integr Comp Physiol 283: R837–R842, 2002.
864. Lai EY, Solis G, Holland S, Luo Z, Carlstrom M, Sandberg K, Wellstein A, Welch WJ,
Wilcox CS. P47phox is required for renal and afferent arteriolar contractile responses
to angiotensin II and perfusion pressure in mice. Hypertension 59: 415– 420, 2012.
885. Lesniewski LA, Donato AJ, Behnke BJ, Woodman CR, Laughlin MH, Ray CA, Delp
MD. Decreased NO signaling leads to enhanced vasoconstrictor responsiveness in
skeletal muscle arterioles of the ZDF rat prior to overt diabetes and hypertension.
Am J Physiol Heart Circ Physiol 294: H1840 –H1850, 2008.
865. Lai EY, Wang Y, Persson AE, Manning D, Liu R. Pressure induces intracellular calcium
changes in juxtaglomerular cells in perfused afferent arterioles. Hypertension Res 34:
942–948, 2011.
866. Lai EY, Wellstein A, Welch WJ, Wilcox CS. Superoxide modulates myogenic contractions of mouse afferent arterioles. Hypertension 58: 650 – 656, 2011.
867. Lais LT, Rios LL, Boutelle S, DiBona GF, Brody MJ. Arterial pressure development in
neonatal and young spontaneously hypertensive rats. Blood Vessels 14: 277–284,
1977.
868. Lamboley M, Schuster A, Beny JL, Meister JJ. Recruitment of smooth muscle cells and
arterial vasomotion. Am J Physiol Heart Circ Physiol 285: H562–H569, 2003.
869. Landsverk SA, Kvandal P, Bernjak A, Stefanovska A, Kirkeboen KA. The effects of
general anesthesia on human skin microcirculation evaluated by wavelet transform.
Anesth Analg 105: 1012–1019, 2007.
870. Langston JB, Guyton AC, Gillespie WJ Jr. Acute effect of changes in renal arterial
pressure and sympathetic blockade on kidney function. Am J Physiol 197: 595– 600,
1959.
871. Langston JB, Guyton AC, Gillespie WJ Jr. Autoregulation absent in normal kidney but
present after renal damage. Am J Physiol 199: 495– 498, 1960.
872. Larson TS, Lockhart JC. Restoration of vasa recta hemodynamics and pressure
natriuresis in SHR by L-arginine. Am J Physiol Renal Fluid Electrolyte Physiol 268:
F907–F912, 1995.
873. Lassegue B, San MA, Griendling KK. Biochemistry, physiology, and pathophysiology
of NADPH oxidases in the cardiovascular system. Circ Res 110: 1364 –1390, 2012.
886. Lessard A, Salevsky FC, Bachelard H, Cupples WA. Incommensurate frequencies of
major vascular regulatory mechanisms. Can J Physiol Pharmacol 77: 293–299, 1999.
887. Levine DZ, Iacovitti M. Real-time measurement of kidney tubule fluid nitric oxide
concentrations in early diabetes: disparate changes in different rodent models. Nitric
Oxide 15: 87–92, 2006.
888. Levine DZ, Iacovitti M, Robertson SJ. Modulation of single-nephron GFR in the db/db
mouse model of type 2 diabetes mellitus. II. Effects of renal mass reduction. Am J
Physiol Regul Integr Comp Physiol 294: R1840 –R1846, 2008.
889. Levine DZ, Iacovitti M, Robertson SJ, Mokhtar GA. Modulation of single-nephron
GFR in the db/db mouse model of type 2 diabetes mellitus. Am J Physiol Regul Integr
Comp Physiol 290: R975–R981, 2006.
890. Levy D, Anderson KM, Christiansen JC, Campanile G, Stokes J. Antihypertensive
drug therapy and arrhythmia risk. Am J Cardiol 62: 147–149, 1988.
891. Lewis EJ, Hunsicker LG, Clark WR, Berl T, Pohl MA, Lewis JB, Ritz E, Atkins RC,
Rohde R, Raz I. Renoprotective effect of the angiotensin-receptor antagonist irbesartan in patients with nephropathy due to type 2 diabetes. N Engl J Med 345:
851– 860, 2001.
892. Leyssac PP. Further studies on oscillating tubulo-glomerular feedback responses in
the rat kidney. Acta Physiol Scand 126: 271–277, 1986.
893. Leyssac PP, Baumbach L. An oscillating intratubular pressure response to alterations
in Henle loop flow in the rat kidney. Acta Physiol Scand 117: 415– 419, 1983.
874. Lau C, Sudbury I, Thomson M, Howard PL, Magil AB, Cupples WA. Salt-resistant
blood pressure and salt-sensitive renal autoregulation in chronic streptozotocin diabetes. Am J Physiol Regul Integr Comp Physiol 296: R1761–R1770, 2009.
894. Leyssac PP, Holstein-Rathlou NH. Tubulo-glomerular feedback response: enhancement in adult spontaneously hypertensive rats and effects of anaesthetics. Pflügers
Arch 413: 267–272, 1989.
875. Laugesen JL, Sosnovtseva OV, Mosekilde E, Holstein-Rathlou NH, Marsh DJ. Coupling-induced complexity in nephron models of renal blood flow regulation. Am J
Physiol Regul Integr Comp Physiol 298: R997–R1006, 2010.
895. Li JM, Shah AM. Endothelial cell superoxide generation: regulation and relevance for
cardiovascular pathophysiology. Am J Physiol Regul Integr Comp Physiol 287: R1014 –
R1030, 2004.
876. Lax DS, Benstein JA, Tolbert E, Dworkin LD. Effects of salt restriction on renal
growth and glomerular injury in rats with remnant kidneys. Kidney Int 41: 1527–
1534, 1992.
896. Li L, Lai EY, Huang Y, Eisner C, Mizel D, Wilcox CS, Schnermann J. Renal afferent
arteriolar and tubuloglomerular feedback reactivity in mice with conditional deletions of adenosine 1 receptors. Am J Physiol Renal Physiol 303: F1166 –F1175, 2012.
492
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
897. Li N, Yi F, Dos Santos EA, Donley DK, Li PL. Role of renal medullary heme oxygenase
in the regulation of pressure natriuresis and arterial blood pressure. Hypertension 49:
148 –154, 2007.
916. Llamas B, Lau C, Cupples WA, Rainville ML, Souzeau E, Deschepper CF. Genetic
determinants of systolic and pulse pressure in an intercross between normotensive
inbred rats. Hypertension 48: 921–926, 2006.
898. Li N, Zou AP, Ge ZD, Campbell WB, Li PL. Effect of nitric oxide on calcium-induced
calcium release in coronary arterial smooth muscle. Gen Pharmacol 35: 37– 45, 2000.
917. Lo M, Liu KL, Clemitson JR, Sassard J, Samani NJ. Chromosome 1 blood pressure
QTL region influences renal function curve and salt sensitivity in SHR. Physiol Genomics 8: 15–21, 2002.
899. Li S, Moon JJ, Miao H, Jin G, Chen BP, Yuan S, Hu Y, Usami S, Chien S. Signal
transduction in matrix contraction and the migration of vascular smooth muscle cells
in three-dimensional matrix. J Vasc Res 40: 378 –388, 2003.
918. Lo M, Liu KL, Lantelme P, Sassard J. Subtype 2 of angiotensin II receptors controls
pressure-natriuresis in rats. J Clin Invest 95: 1394 –1397, 1995.
900. Lieb DC, Kemp BA, Howell NL, Gildea JJ, Carey RM. Reinforcing feedback loop of
renal cyclic guanosine 3’,5’-monophosphate and interstitial hydrostatic pressure in
pressure-natriuresis. Hypertension 54: 1278 –1283, 2009.
919. Lohn M, Jessner W, Furstenau M, Wellner M, Sorrentino V, Haller H, Luft FC,
Gollasch M. Regulation of calcium sparks and spontaneous transient outward currents by RyR3 in arterial vascular smooth muscle cells. Circ Res 89: 1051–1057, 2001.
901. Liebau G, Levine DZ, Thurau K. Micropuncture studies on the dog kidney. I. The
response of the proximal tubule to changes in systemic blood pressure within and
below the autoregulatory range. Pflügers Arch 304: 57– 68, 1968.
920. Lombard JH, Eskinder H, Kauser K, Osborn JL, Harder DR. Enhanced norepinephrine sensitivity in renal arteries at elevated transmural pressure. Am J Physiol Heart
Circ Physiol 259: H29 –H33, 1990.
902. Lin HB, Young DB. Verapamil alters the relationship between renal perfusion pressure and glomerular filtration rate and renin release: the mechanism of the antihypertensive effect. J Cardiovasc Pharmacol 6: S57–S59, 1988.
921. Lopez B, Moreno C, Salom MG, Roman RJ, Fenoy FJ. Role of guanylyl cyclase and
cytochrome P-450 on renal response to nitric oxide. Am J Physiol Renal Physiol 281:
F420 –F427, 2001.
903. Lin JJ, Tonshoff B, Bouriquet N, Casellas D, Kaskel FJ, Moore LC. Insulin-like growth
factor-I restores microvascular autoregulation in experimental chronic renal failure.
Kidney Int Suppl 67: S195–S198, 1998.
922. Lopez B, Ryan RP, Moreno C, Sarkis A, Lazar J, Provoost AP, Jacob HJ, Roman RJ.
Identification of a QTL on chromosome 1 for impaired autoregulation of RBF in
fawn-hooded hypertensive rats. Am J Physiol Renal Physiol 290: F1213–F1221,
2006.
904. Lincoln TM, Cornwell TL. Intracellular cyclic GMP receptor proteins. FASEB J 7:
328 –338, 1993.
923. Lorenz JN. Micropuncture of the kidney: a primer on techniques. Comp Physiol 2:
621– 637, 2012.
905. Linde CI, Karashima E, Raina H, Zulian A, Wier WG, Hamlyn JM, Ferrari P, Blaustein
MP, Golovina VA. Increased arterial smooth muscle Ca2⫹ signaling, vasoconstriction, and myogenic reactivity in Milan hypertensive rats. Am J Physiol Heart Circ Physiol
302: H611–H620, 2012.
924. Lorenz JN, Schnermann J, Brosius FC, Briggs JP, Furspan PB. Intracellular ATP can
regulate afferent arteriolar tone via ATP-sensitive K⫹ channels in the rabbit. J Clin
Invest 90: 733–740, 1992.
906. Lindpaintner K, Hilbert P, Ganten D, Nadal-Ginard B, Inagami T, Iwai N. Molecular
genetics of the SA-gene: cosegregation with hypertension and mapping to rat chromosome 1. J Hypertens 11: 19 –23, 1993.
925. Loutzenhiser K, Loutzenhiser R. Angiotensin II-induced Ca2⫹ influx in renal afferent
and efferent arterioles: differing roles of voltage-gated and store-operated Ca2⫹
entry. Circ Res 87: 551–557, 2000.
907. Lino S, Calcagnini G, Censi F, Congi M, De PF. Cardiovascular neuroregulation and
rhythms of the autonomic nervous system: frequency domain analysis (Article in
Italian)Cardiologia 44: 281–287, 1999.
926. Loutzenhiser R, Aaronson PI. Role of epithelial sodium channels in the renal myogenic response? Hypertension 55: -e6 – e9, 2010.
908. Linz W, Becker RH, Scholkens BA, Wiemer G, Keil M, Langer KH. Nephroprotection by long-term ACE inhibition with ramipril in spontaneously hypertensive stroke
prone rats. Kindey Int 54: 2037–2044, 1998.
909. Lipkowitz MS, Freedman BI, Langefeld CD, Comeau ME, Bowden DW, Kao WH,
Astor BC, Bottinger EP, Iyengar SK, Klotman PE, Freedman RG, Zhang W, Parekh
RS, Choi MJ, Nelson GW, Winkler CA, Kopp JB. Apolipoprotein L1 gene variants
associate with hypertension-attributed nephropathy and the rate of kidney function
decline in African Americans. Kidney Int 83: 114 –120, 2013.
910. Litteri G, Carnevale D, D’Urso A, Cifelli G, Braghetta P, Damato A, Bizzotto D,
Landolfi A, Ros FD, Sabatelli P, Facchinello N, Maffei A, Volpin D, Colombatti A,
Bressan GM, Lembo G. Vascular smooth muscle Emilin-1 is a regulator of arteriolar
myogenic response and blood pressure. Arterioscler Thromb Vasc Biol 32: 2178 –
2184, 2012.
927. Loutzenhiser R, Bidani AK, Chilton L. Renal myogenic response: kinetic attributes
and physiological role. Circ Res 90: 1316 –1324, 2002.
928. Loutzenhiser R, Chilton L, Trottier G. Membrane potential measurements in renal
afferent and efferent arterioles: actions of angiotensin II. Am J Physiol Renal Physiol
273: F307–F314, 1997.
929. Loutzenhiser R, Epstein M. Effects of calcium antagonists on renal hemodynamics.
Am J Physiol Renal Fluid Electrolyte Physiol 249: F619 –F629, 1985.
930. Loutzenhiser R, Epstein M, Hayashi K, Horton C. Direct visualization of effects of
endothelin on the renal microvasculature. Am J Physiol Renal Fluid Electrolyte Physiol
258: F61–F68, 1990.
931. Loutzenhiser R, Epstein M, Horton C. Inhibition by diltiazem of pressure-induced
afferent vasoconstriction in the isolated perfused rat kidney. Am J Cardiol 59: -72A–
75A, 1987.
911. Liu KL, Lo M, Grouzmann E, Mutter M, Sassard J. The subtype 2 of angiotensin II
receptors and pressure-natriuresis in adult rat kidneys. Br J Pharmacol 126: 826 – 832,
1999.
932. Loutzenhiser R, Epstein M, Horton C, Sonke P. Reversal of renal and smooth muscle
actions of the thromboxane mimetic U-44069 by diltiazem. Am J Physiol Renal Fluid
Electrolyte Physiol 250: F619 –F626, 1986.
912. Liu R, Garvin JL, Ren Y, Pagano PJ, Carretero OA. Depolarization of the macula
densa induces superoxide production via NAD(P)H oxidase. Am J Physiol Renal
Physiol 292: F1867–F1872, 2007.
933. Loutzenhiser R, Griffin K, Williamson G, Bidani A. Renal autoregulation: new perspectives regarding the protective and regulatory roles of the underlying mechanisms. Am J Physiol Regul Integr Comp Physiol 290: R1153–R1167, 2006.
913. Liu R, Ren Y, Garvin JL, Carretero OA. Superoxide enhances tubuloglomerular
feedback by constricting the afferent arteriole. Kidney Int 66: 268 –274, 2004.
934. Loutzenhiser R, Griffin KA, Bidani AK. Systolic blood pressure as the trigger for the
renal myogenic response: protective or autoregulatory? Curr Opin Nephrol Hypertens
15: 41– 49, 2006.
914. Liu Y, Bubolz AH, Mendoza S, Zhang DX, Gutterman DD. H2O2 is the transferrable
factor mediating flow-induced dilation in human coronary arterioles. Circ Res 108:
566 –573, 2011.
915. Liu Y, Harder D, Lombard J. Myogenic activation of canine small renal arteries after
nonchemical removal of the endothelium. Am J Physiol Heart Circ Physiol 267: H302–
H307, 1994.
935. Loutzenhiser R, Hayashi K, Epstein M. Divergent effects of KCl induced depolarization on afferent and efferent arterioles. Am J Physiol Renal Fluid Electrolyte Physiol 257:
F561–F564, 1989.
936. Loutzenhiser R, Parker MJ. Hypoxia inhibits myogenic reactivity of renal afferent
arterioles by activating ATP-sensitive K⫹ channels. Circ Res 74: 861– 869, 1994.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
493
RENAL AUTOREGULATORY MECHANISMS
937. Luksha N, Luksha L, Carrero JJ, Hammarqvist F, Stenvinkel P, Kublickiene K. Impaired resistance artery function in patients with end-stage renal disease. Clin Sci
120: 525–536, 2011.
938. Lund-Johansen P. Hemodynamic changes in long-term diuretic therapy of essential
hypertension. Acta Med Scand 187: 509 –518, 1970.
939. Lundie MJ, Friberg P, Kline RL, Adams MA. Long-term inhibition of the renin-angiotensin system in genetic hypertension: analysis of the impact on blood pressure and
cardiovascular structural changes. J Hypertens 15: 339 –348, 1997.
940. Lundin S, Herlitz H, Hallback-Nordlander M, Ricksten SE, Gothberg G, Berglund G.
Sodium balance during development of hypertension in the spontaneously hypertensive rat (SHR). Acta Physiol Scand 115: 317–323, 1982.
941. Lundin S, Rickesten SE, Thoren P. Renal sympathetic activity in spontaneously hypertensive rats with normotensive controls as studied by three different methods.
Acta Physiol Scand 120: 265–272, 1984.
959. Majid DSA, Omoro SA, Chin SY, Navar LG. Intrarenal nitric oxide activity and
pressure natriuresis in anesthetized dogs. Hypertension 32: 262–272, 1998.
960. Majid DSA, Said KE, Omoro SA. Responses to acute changes in arterial pressure on
renal medullary nitric oxide activity in dogs. Hypertension 34: 832– 836, 1999.
961. Majid DSA, Said KE, Omoro SA, Navar LG. Nitric oxide dependency of arterial
pressure-induced changes in renal interstitial hydrostatic pressure in dogs. Circ Res
88: 347–351, 2001.
962. Majid DSA, Williams A, Kadowitz PJ, Navar LG. Renal responses to intra-arterial
administration of nitric oxide donor in dogs. Hypertension 22: 535–541, 1993.
963. Majid DSA, Williams A, Navar LG. Inhibition of nitric oxide synthesis attenuates
pressure-induced natriuretic responses in anesthetized dogs. Am J Physiol Renal Fluid
Electrolyte Physiol 264: F79 –F87, 1993.
964. Makino A, Skelton MM, Zou AP, Roman RJ, Cowley AW Jr. Increased renal medullary oxidative stress produces hypertension. Hypertension 39: 667– 672, 2002.
942. Luscher TF, Creager MA, Beckman JA, Cosentino F. Diabetes and vascular disease:
pathophysiology, clinical consequences, and medical therapy: Part II. Circulation 108:
1655–1661, 2003.
965. Mann JF, Green D, Jamerson K, Ruilope LM, Kuranoff SJ, Littke T, Viberti G. Avosentan for overt diabetic nephropathy. J Am Soc Nephrol 21: 527–535, 2010.
943. Ma YH, Gebremedhin D, Schwartzman ML, Falck JR, Clark JE, Masters BS, Harder
DR, Roman RJ. 20-Hydroxyeicosatetraenoic acid is an endogenous vasoconstrictor
of canine renal arcuate arteries. Circ Res 72: 126 –136, 1993.
966. Mann JF, Schmieder RE, Dyal L, McQueen MJ, Schumacher H, Pogue J, Wang X,
Probstfield JL, Avezum A, Cardona-Munoz E, Dagenais GR, Diaz R, Fodor G, Maillon
JM, Ryden L, Yu CM, Teo KK, Yusuf S. Effect of telmisartan on renal outcomes: a
randomized trial. Ann Intern Med 151: 1–10, 2009.
944. Macedo MP, Lautt WW. Autoregulatory capacity in the superior mesenteric artery is
attenuated by nitric oxide. Am J Physiol Gastrointest Liver Physiol 271: G400 –G404,
1996.
945. Maddox DA, Troy JL, Brenner BM. Autoregulation of filtration rate in the absence of
macula densa-glomerulus feedback. Am J Physiol 227: 123–131, 1974.
946. Madrid MI, Garcia-Salom M, Tornel J, de Gasparo M, Fenoy FJ. Effect of interactions
between nitric oxide and angiotensin II on pressure diuresis and natriuresis. Am J
Physiol Regul Integr Comp Physiol 273: R1676 –R1682, 1997.
947. Madrid MI, Salom MG, Tornel J, Lopez E, Fenoy FJ. Interactions between nitric oxide
and renal nerves on pressure-diuresis and natriuresis. J Am Soc Nephrol 9: 1588 –
1595, 1998.
948. Magee GM, Bilous RW, Cardwell CR, Hunter SJ, Kee F, Fogarty DG. Is hyperfiltration associated with the future risk of developing diabetic nephropathy? A metaanalysis. Diabetologia 52: 691– 697, 2009.
949. Magnusson L, Sorensen CM, Braunstein TH, Holstein-Rathlou NH, Salomonsson M.
Renovascular BK(Ca) channels are not activated in vivo under resting conditions and
during agonist stimulation. Am J Physiol Regul Integr Comp Physiol 292: R345–R353,
2007.
950. Majid DS, Godfrey M, Navar LG. Pressure natriuresis and renal medullary blood flow
in dogs. Hypertension 29: 1051–1057, 1997.
951. Majid DS, Godfrey M, Omoro SA. Pressure natriuresis and autoregulation of inner
medullary blood flow in canine kidney. Hypertension 29: 210 –215, 1997.
952. Majid DS, Inscho EW, Navar LG. P2 purinoceptor saturation by adenosine triphosphate impairs renal autoregulation in dogs. J Am Soc Nephrol 10: 492– 498, 1999.
953. Majid DS, Navar LG. Suppression of blood flow autoregulation plateau during nitric
oxide blockade of canine kidney. Am J Physiol Renal Fluid Electrolyte Physiol 262:
F40 –F46, 1992.
954. Majid DS, Navar LG. Medullary blood flow responses to changes in arterial pressure
in canine kidney. Am J Physiol Renal Fluid Electrolyte Physiol 270: F833–F838, 1996.
955. Majid DS, Navar LG. Nitric oxide in the control of renal hemodynamics and excretory function. Am J Hypertens 14: 74S– 82S, 2001.
967. Mann JF, Schmieder RE, McQueen M, Dyal L, Schumacher H, Pogue J, Wang X,
Maggioni A, Budaj A, Chaithiraphan S, Dickstein K, Keltai M, Metsarinne K, Oto A,
Parkhomenko A, Piegas LS, Svendsen TL, Teo KK, Yusuf S. Renal outcomes with
telmisartan, ramipril, or both, in people at high vascular risk (the ONTARGET
study): a multicentre, randomised, double-blind, controlled trial. Lancet 372: 547–
553, 2008.
968. Manning RD Jr, Tian N, Meng S. Oxidative stress and antioxidant treatment in
hypertension and the associated renal damage. Am J Nephrol 25: 311–317, 2005.
969. Marchand GR, Burke TJ, Haas JA, Romero JC, Knox FG. Regulation of filtration rate
in sodium-depleted and -expanded dogs. Am J Physiol Renal Fluid Electrolyte Physiol
232: F325–F328, 1977.
970. Marchenko SM, Sage SO. Smooth muscle cells affect endothelial membrane potential in rat aorta. Am J Physiol Heart Circ Physiol 267: H804 –H811, 1994.
971. Marsh DJ. Frequency response of autoregulation. Kidney Int 22: S165–S172, 1982.
972. Marsh DJ, Martin CM. Effects of diuretic states on collecting duct fluid flow resistance in the hamster kidney. Am J Physiol 229: 13–17, 1975.
973. Marsh DJ, Sosnovtseva OV, Chon KH, Holstein-Rathlou NH. Nonlinear interactions
in renal blood flow regulation. Am J Physiol Regul Integr Comp Physiol 288: R1143–
R1159, 2005.
974. Marsh DJ, Sosnovtseva OV, Mosekilde E, Holstein-Rathlou NH. Vascular coupling
induces synchronization, quasiperiodicity, and chaos in a nephron tree. Chaos 17:
015114, 2007.
975. Marsh DJ, Sosnovtseva OV, Pavlov AN, Yip KP, Holstein-Rathlou NH. Frequency
encoding in renal blood flow regulation. Am J Physiol Regul Integr Comp Physiol 288:
R1160 –R1167, 2005.
976. Marsh DJ, Toma I, Sosnovtseva OV, Peti-Peterdi J, Holstein-Rathlou NH. Electrotonic vascular signal conduction and nephron synchronization. Am J Physiol Renal
Physiol 296: F751–F761, 2009.
977. Massett MP, Ungvari Z, Csiszar A, Kaley G, Koller A. Different roles of PKC and MAP
kinases in arteriolar constrictions to pressure and agonists. Am J Physiol Heart Circ
Physiol 283: H2282–H2287, 2002.
956. Majid DSA, Navar LG. Blockade of distal nephron sodium transport attenuates
pressure natriuresis in dogs. Hypertension 23: 1040 –1045, 1994.
978. Matavelli LC, Zhou X, Varagic J, Susic D, Frohlich ED. Salt loading produces severe
renal hemodynamic dysfunction independent of arterial pressure in spontaneously
hypertensive rats. Am J Physiol Heart Circ Physiol 292: H814 –H819, 2007.
957. Majid DSA, Navar LG. Nitric oxide in the mediation of pressure natriuresis. Clin Exp
Pharmacol Physiol 24: 595–599, 1997.
979. Matchkov VV. Mechanisms of cellular synchronization in the vascular wall. Mechanisms of vasomotion. Dan Med Bull 57: 1–50, 2010.
958. Majid DSA, Navar LG. Nitric oxide in the control of renal hemodynamics and excretory function. Am J Hypertens 14: 74S– 82S, 2001.
980. Matchkov VV, Larsen P, Bouzinova EV, Rojek A, Boedtkjer DM, Golubinskaya V,
Pedersen FS, Aalkjaer C, Nilsson H. Bestrophin-3 (vitelliform macular dystrophy
494
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
2-like 3 protein) is essential for the cGMP-dependent calcium-activated chloride
conductance in vascular smooth muscle cells. Circ Res 103: 864 – 872, 2008.
1002. Meininger GA, Faber JE. Adrenergic tone selectively augments myogenic constriction in skeletal muscle microcirculation. Am J Physiol Heart Circ Physiol 260: H1424 –
H1432, 1991.
981. Matchkov VV, Rahman A, Bakker LM, Griffith TM, Nilsson H, Aalkjaer C. Analysis of
effects of connexin-mimetic peptides in rat mesenteric small arteries. Am J Physiol
Heart Circ Physiol 291: H357–H367, 2006.
1003. Messick J Jr, Spielman WS, Knox FG. Failure of high doses of Sar1-lle8-angiotensin II
to abolish autoregulation of renal blood flow. Mayo Clin Proc 55: 619 – 622, 1980.
982. Matchkov VV, Secher D, V, Bodtkjer DM, Aalkjaer C. Transport and function of
chloride in vascular smooth muscles. J Vasc Res 50: 69 – 87, 2012.
1004. Meyer TW, Anderson S, Rennke HG, Brenner BM. Reversing glomerular hypertension stabilizes established glomerular injury. Kidney Int 31: 752–759, 1987.
983. Matrougui K. Diabetes and microvascular pathophysiology: role of epidermal growth
factor receptor tyrosine kinase. Diabetes Metab Res Rev 26: 13–16, 2010.
1005. Miller BA, Zhang W. TRP channels as mediators of oxidative stress. Adv Exp Med Biol
704: 531–544, 2011.
984. Matrougui K, Eskildsen-Helmond YE, Fiebeler A, Henrion D, Levy BI, Tedgui A,
Mulvany MJ. Angiotensin II stimulates extracellular signal-regulated kinase activity in
intact pressurized rat mesenteric resistance arteries. Hypertension 36: 617– 621,
2000.
1006. Miller FJ Jr, Dellsperger KC, Gutterman DD. Myogenic constriction of human coronary arterioles. Am J Physiol Heart Circ Physiol 273: H257–H264, 1997.
1007. Miller JA. Renal responses to sodium restriction in patients with early diabetes
mellitus. J Am Soc Nephrol 8: 749 –755, 1997.
985. Matrougui K, Loufrani L, Heymes C, Levy BI, Henrion D. Activation of AT2 receptors
by endogenous angiotensin II is involved in flow-reduced dilation in rat resistance
arteries. Hypertension 34: 659 – 665, 1999.
1008. Miller PL, Rennke HG, Meyer TW. Glomerular hypertrophy accelerates hypertensive glomerular injury in rats. Am J Physiol Renal Fluid Electrolyte Physiol 261: F459 –
F465, 1991.
986. Matsumura Y, Kuro T, Kobayashi Y, Konishi F, Takaoka M, Wessale JL, Opgenorth
TJ, Gariepy CE, Yanagisawa M. Exaggerated vascular and renal pathology in endothelin-B receptor-deficient rats with deoxycorticosterone acetate-salt hypertension. Circulation 102: 2765–2773, 2000.
1009. Minami N, Imai Y, Munakata M, Sasaki S, Sekino H, Abe K, Yoshinaga K. Age-related
changes in blood pressure, heart rate and baroreflex sensitivity in SHR. Clin Exp
Pharmacol Physiol 15 Suppl: 85– 87, 1989.
987. Matsunaga H, Yamashita N, Okuda T, Kurokawa K. Mesangial cell ion transport and
tubuloglomerular feedback. Curr Opin Nephrol Hypertens 3: 518 –522, 1994.
1010. Miriel VA, Mauban JR, Blaustein MP, Wier WG. Local and cellular Ca2⫹ transients in
smooth muscle of pressurized rat resistance arteries during myogenic and agonist
stimulation. J Physiol 518: 815– 824, 1999.
988. Mattson DL, Dwinell MR, Greene AS, Kwitek AE, Roman RJ, Jacob HJ, Cowley AW
Jr. Chromosome substitution reveals the genetic basis of Dahl salt-sensitive hypertension and renal disease. Am J Physiol Renal Physiol 295: F837–F842, 2008.
1011. Mistry DK, Garland CJ. Nitric oxide (NO)-induced activation of large conductance
Ca2⫹-dependent K⫹ channels [BK(Ca)] in smooth muscle cells isolated from the rat
mesenteric artery. Br J Pharmacol 124: 1131–1140, 1998.
989. Mattson DL, Lu S, Cowley AW Jr. Role of nitric oxide in the control of the renal
medullary circulation. Clin Exp Pharmacol Physiol 24: 587–590, 1997.
1012. Mitchell KD, Navar LG. Enhanced tubuloglomerular feedback during peritubular
infusions of angiotensins I and II. Am J Physiol Renal Fluid Electrolyte Physiol 255:
F383–F390, 1988.
990. Mattson DL, Lu S, Roman RJ, Cowley AW Jr. Relationship between renal perfusion
pressure and blood flow in different regions of the kidney. Am J Physiol Regul Integr
Comp Physiol 264: R578 –R583, 1993.
991. Mattson DL, Roman RJ, Cowley AW Jr. Role of nitric oxide in renal papillary blood
flow and sodium excretion. Hypertension 19: 766 –769, 1992.
992. Mauban JR, Wier WG. Essential role of EDHF in the initiation and maintenance of
adrenergic vasomotion in rat mesenteric arteries. Am J Physiol Heart Circ Physiol 287:
H608 –H616, 2004.
993. Mauer SM, Brown DM, Steffes MW, Azar S. Studies of renal autoregulation in pancreatectomized and streptozotocin diabetic rats. Kidney Int 37: 909 –917, 1990.
994. McCurley A, Pires PW, Bender SB, Aronovitz M, Zhao MJ, Metzger D, Chambon P,
Hill MA, Dorrance AM, Mendelsohn ME, Jaffe IZ. Direct regulation of blood pressure
by smooth muscle cell mineralocorticoid receptors. Nat Med 18: 1429 –1433, 2012.
995. McDonough AA. Mechanisms of proximal tubule sodium transport regulation that
link extracellular fluid volume and blood pressure. Am J Physiol Regul Integr Comp
Physiol 298: R851–R861, 2010.
1013. Mitchell KD, Navar LG. Modulation of tubuloglomerular feedback responsiveness by
extracellular ATP. Am J Physiol Renal Fluid Electrolyte Physiol 264: F458 –F466, 1993.
1014. Miyata N, Zou AP, Mattson DL, Cowley AW Jr. Renal medullary interstitial infusion
of L-arginine prevents hypertension in Dahl salt-sensitive rats. Am J Physiol Regul
Integr Comp Physiol 275: R1667–R1673, 1998.
1015. Modlinger P, Chabrashvili T, Gill PS, Mendonca M, Harrison DG, Griendling KK, Li
M, Raggio J, Wellstein A, Chen Y, Welch WJ, Wilcox CS. RNA silencing in vivo reveals
role of p22phox in rat angiotensin slow pressor response. Hypertension 47: 238 –
244, 2006.
1016. Mohazzab KM, Kaminski PM, Wolin MS. NADH oxidoreductase is a major source of
superoxide anion in bovine coronary artery endothelium. Am J Physiol Heart Circ
Physiol 266: H2568 –H2572, 1994.
1017. Moien-Afshari F, Ghosh S, Elmi S, Khazaei M, Rahman MM, Sallam N, Laher I.
Exercise restores coronary vascular function independent of myogenic tone or hyperglycemic status in db/db mice. Am J Physiol Heart Circ Physiol 295: H1470 –H1480,
2008.
996. McDonough AA, Leong PK, Yang LE. Mechanisms of pressure natriuresis: how blood
pressure regulates renal sodium transport. Ann NY Acad Sci 986: 669 – 677, 2003.
1018. Monkawa T, Hayashi M, Miyawaki A, Sugiyama T, Yamamoto-Hino M, Hasegawa M,
Furuichi T, Mikoshiba K, Saruta T. Localization of inositol 1,4,5-trisphosphate receptors in the rat kidney. Kidney Int 53: 296 –301, 1998.
997. McLennan GP, Kline RL, Mercer PF. Effect of enalapril treatment on the pressure-natriuresis curve in spontaneously hypertensive rats. Hypertension 17: 54 –
62, 1991.
1019. Moore LC. Tubuloglomerular feedback and SNGFR autoregulation in the rat. Am J
Physiol Renal Fluid Electrolyte Physiol 247: F267–F276, 1984.
998. McNay JL, Kishimoto T. Association between autoregulation and pressure dependency of renal vascular responsiveness in dogs. Circ Res 24: 599 – 605, 1969.
1020. Moore LC, Casellas D. Tubuloglomerular feedback dependence of autoregulation in
rat juxtamedullary afferent arterioles. Kidney Int 37: 1402–1408, 1990.
999. Mederos y Schnitzler M, Storch U, Gudermann T. AT1 receptors as mechanosensors. Curr Opin Pharmacol 11: 112–116, 2011.
1021. Moore LC, Rich A, Casellas D. Ascending myogenic autoregulation: interactions
between tubuloglomerular feedback and myogenic mechanisms. Bull Math Biol 56:
391– 410, 1994.
1000. Medros y Schnitzler M, Storch U, Meibers S, Nurwakagari P, Breit A, Essin K,
Gollasch M, Gudermann T. Gq-coupled receptors as mechanosensors mediating
myogenic vasoconstriction. EMBO J 27: 3092–3103, 2008.
1022. Moore LC, Schnermann J, Yarimizu S. Feedback mediation of SNGFR autoregulation
in hydropenic and DOCA- and salt-loaded rats. Am J Physiol 6: F63–F74, 1979.
1001. Meininger GA, Davis MJ. Cellular mechanisms involved in the vascular myogenic
response. Am J Physiol Heart Circ Physiol 263: H647–H659, 1992.
1023. Moore LC, Thorup C, Ellinger A, Paccione J, Casellas D, Kaskel FJ. Advanced glycosylation end-products and NO-dependent vasodilation in renal afferent arterioles
from diabetic rats. Acta Physiol Scand 168: 101–106, 2000.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
495
RENAL AUTOREGULATORY MECHANISMS
1024. Moosmang S, Haider N, Bruderl B, Welling A, Hofmann F. Antihypertensive effects
of the putative T-type calcium channel antagonist mibefradil are mediated by the
L-type calcium channel Cav1.2. Circ Res 98: 105–110, 2006.
1046. Murphy TV, Spurrell BE, Hill MA. Cellular signalling in arteriolar myogenic constriction: involvement of tyrosine phosphorylation pathways. Clin Exp Pharmacol Physiol
29: 612– 619, 2002.
1025. Moreno C, Lopez A, Llinas MT, Rodriguez F, Lopez-Farre A, Nava E, Salazar FJ.
Changes in NOS activity and protein expression during acute and prolonged ANG II
administration. Am J Physiol Regul Integr Comp Physiol 282: R31–R37, 2002.
1047. Murray RD, Malvin RL. Intrarenal renin and autoregulation of renal plasma flow and
glomerular filtration rate. Am J Physiol Renal Fluid Electrolyte Physiol 236: F559 –F566,
1979.
1026. Moreno JM, Rodriguez G, I, Wangensteen R, varez-Guerra M, Luna JD, Garcia-Estan
J, Vargas F. Tempol improves renal hemodynamics and pressure natriuresis in hyperthyroid rats. Am J Physiol Regul Integr Comp Physiol 294: R867–R873, 2008.
1048. Nagaoka A, Iwatsuka H, Suzuoki Z, Okamoto K. Genetic predisposition to stroke in
spontaneously hypertensive rats. Am J Physiol 230: 1354 –1359, 1976.
1027. Moreno-Dominguez A, Colinas O, El-Yazbi A, Walsh EJ, Hill MA, Walsh MP, Cole
WC. Ca2⫹ sensitization due to myosin light chain phosphatase inhibition and cytoskeletal reorganization in the myogenic response of skeletal muscle resistance arteries. J Physiol 591: 1235–1250, 2013.
1028. Morgan T. Renin, angiotensin, sodium and organ damage. Hypertension Res 26:
349 –354, 2003.
1029. Mori T, Cowley AW Jr. Role of pressure in angiotensin II-induced renal injury:
chronic servo-control of renal perfusion pressure in rats. Hypertension 43: 752–759,
2004.
1049. Nagaoka A, Kakihana M, Suno M, Hamajo K. Renal hemodynamics and sodium
excretion in stroke-prone spontaneously hypertensive rats. Am J Physiol Renal Fluid
Electrolyte Physiol 241: F244 –F249, 1981.
1050. Nagaoka A, Toyoda S, Iwatsuka H. Increased renal vascular reactivity to norepinephrine in stroke-prone spontaneously hypertensive rat. Life Sci 23: 1159 –1166, 1978.
1051. Nagasawa T, Imig JD. Afferent arteriolar responses to beta,gamma-methylene ATP
and 20-HETE are not blocked by ENaC inhibition. Physiol Rep 1: e00082, 2013.
1052. Naguib RE, Contant C, Cupples WA. Atrial natriuretic factor, angiotensin II, and the
slow component of renal autoregulation. Can J Physiol Pharmacol 72: 1132–1137,
1994.
1030. Mori T, O’Connor PM, Abe M, Cowley AW Jr. Enhanced superoxide production in
renal outer medulla of Dahl salt-sensitive rats reduces nitric oxide tubular-vascular
cross-talk. Hypertension 49: 1336 –1341, 2007.
1053. Nahmod VE, Lanari A. Abolution of autoregulation of renal blood flow by acetylcholine. Am J Physiol 207: 123–127, 1964.
1031. Morimoto S, Ohyama T, Hisaki K, Matsumura Y. Effects of CV-4093, a new dihydropyridine calcium channel blocker, on renal hemodynamics and function in strokeprone spontaneously hypertensive rats (SHRSP). Jpn J Pharmacol 51: 257–265, 1989.
1054. Nakamura A, Hayashi K, Ozawa Y, Fujiwara K, Okubo K, Kanda T, Wakino S, Saruta
T. Vessel- and vasoconstrictor-dependent role of rho/rho-kinase in renal microvascular tone. J Vasc Res 40: 244 –251, 2003.
1032. Morita H, Honda A, Inoue R, Ito Y, Abe K, Nelson MT, Brayden JE. Membrane
stretch-induced activation of a TRPM4-like nonselective cation channel in cerebral
artery myocytes. J Pharmacol Sci 103: 417– 426, 2007.
1055. Nakamura T, Obata J, Kuroyanagi R, Kimura H, Ikeda Y, Takano H, Naito A, Sato T,
Yoshida Y. Involvement of angiotensin II in glomerulosclerosis of stroke-prone spontaneously hypertensive rats. Kidney Int 55 Suppl: S109 –S112, 1996.
1033. Moritz AR, Oldt MR. Arteriolar sclerosis in hypertensive and non-hypertensive individuals. Am J Pathol 13: 679 –728, 1937.
1056. Nakamura Y, Ono H, Frohlich ED. Differential effects of T- and L-type calcium
antagonists on glomerular dynamics in spontaneously hypertensive rats. Hypertension 34: 273–278, 1999.
1034. Morsing P, Persson AE. Effect of prostaglandin synthesis inhibition on the tubuloglomerular feedback control in the rat kidney. Ren Physiol Biochem 15: 66 –72, 1992.
1035. Morsing P, Velazquez H, Ellison D, Wright FS. Resetting of tubuloglomerular feedback by interrupting early distal flow. Acta Physiol Scand 148: 63– 68, 1993.
1036. Moss NG, Kopple TE, Arendshorst WJ. Renal vasoconstriction by vasopressin V1a
receptors is modulated by nitric oxide, prostanoids, and superoxide but not the ADP
ribosyl cyclase CD38. Am J Physiol Renal Physiol 306: F1143–F1154, 2014.
1037. Moss NG, Vogel PA, Kopple TE, Arendshorst WJ. Thromboxane-induced renal
vasoconstriction is mediated by the ADP-ribosyl cyclase CD38 and superoxide anion. Am J Physiol Renal Physiol 305: F830 –F838, 2013.
1038. Moulana M, Hosick K, Stanford J, Zhang H, Roman RJ, Reckelhoff JF. Sex differences
in blood pressure control in SHR: lack of a role for EETs. Physiol Rep 2: e12022, 2014.
1039. Muller JM, Chilian WM, Davis MJ. Integrin signaling transduces shear stress-dependent vasodilation of coronary arterioles. Circ Res 80: 320 –326, 1997.
1040. Muller-Suur R, Gutsche HU, Samwer KF, Oelkers W, Hierholzer K. Tubuloglomerular feedback in rat kidneys of different renin contents. Pflügers Arch 359: 33–56, 1975.
1057. Narayanan D, Bulley S, Leo MD, Burris SK, Gabrick KS, Boop FA, Jaggar JH. Smooth
muscle cell transient receptor potential polycystin-2 (TRPP2) channels contribute to
the myogenic response in cerebral arteries. J Physiol 591: 5031–5046, 2013.
1058. Narayanan D, Xi Q, Pfeffer LM, Jaggar JH. Mitochondria control functional CaV1.2
expression in smooth muscle cells of cerebral arteries. Circ Res 107: 631– 641, 2010.
1059. Narayanan J, Imig JD, Roman RJ, Harder DR. Pressurization of isolated renal arteries
increases inositol trisphosphate and diacylglycerol. Am J Physiol Heart Circ Physiol 266:
H1840 –H1845, 1994.
1060. Narishige T, Egashira K, Akatsuka Y, Katsuda Y, Numaguchi K, Sakata M, Takeshita
A. Glibenclamide, a putative ATP-sensitive K⫹ channel blocker, inhibits coronary
autoregulation in anesthetized dogs. Circ Res 73: 771–776, 1993.
1061. Narumiya S, Sugimoto Y, Ushikubi F. Prostanoid receptors: structures, properties,
functions. Physiol Rev 79: 1193–1226, 1999.
1062. Nath KA. Heme oxygenase-1: a provenance for cytoprotective pathways in the
kidney and other tissues. Kidney Int 70: 432– 443, 2006.
1041. Muller-Suur R, Norlen BJ, Persson AE. Resetting of tubuloglomerular feedback in rat
kidneys after unilateral nephrectomy. Kidney Int 18: 48 –57, 1980.
1063. Nath KA, Chmielewski DH, Hostetter TH. Regulatory role of prostanoids in glomerular microcirculation of remnant nephrons. Am J Physiol Renal Fluid Electrolyte
Physiol 252: F829 –F837, 1987.
1042. Müller-Suur R, Persson AE, Ulfendahl HR. Tubuloglomerular feedback in juxtamedullary nephrons. Kidney Int 22: S-104 –S-108, 1982.
1064. Navar LG. Renal autoregulation: perspectives from whole kidney and single nephron
studies. Am J Physiol Renal Fluid Electrolyte Physiol 234: F357–F370, 1978.
1043. Muller-Suur R, Ulfendahl HR, Persson AE. Evidence for tubuloglomerular feedback
in juxtamedullary nephrons of young rats. Am J Physiol Renal Fluid Electrolyte Physiol
244: F425–F431, 1983.
1065. Navar LG, Arendshorst WJ, Pallone TL, Inscho EW, Imig JD, Bell PD. The renal
microcirculation. In: Handbook of Physiology: Microcirculation, edited by Tuma R,
Duran WN, and Ley N. Boston: Academic, 2008.
1044. Mulvany MJ. Small artery remodelling in hypertension. Basic Clin Pharmacol Toxicol
110: 49 –55, 2012.
1066. Navar LG, Bell PD. Romancing the macula densa at UAB. Kidney Int Suppl: S34 –S40,
2004.
1045. Murphy TV, Spurrell BE, Hill MA. Tyrosine phosphorylation following alterations in
arteriolar intraluminal pressure and wall tension. Am J Physiol Heart Circ Physiol 281:
H1047–H1056, 2001.
1067. Navar LG, Bell PD, Burke TJ. Autoregulatory responses of superficial nephrons and
their association with sodium excretion during arterial pressure alterations in the
dog. Circ Res 41: 487– 496, 1977.
496
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
1068. Navar LG, Bell PD, White RW, Watts RL, Williams RH. Evaluation of the single
nephron glomerular filtration coefficient in the dog. Kidney Int 12: 137–149, 1977.
1069. Navar LG, Champion WJ, Thomas CE. Effects of calcium channel blockade on renal
vascular resistance responses to changes in perfusion pressure and angiotensinconverting enzyme inhibition in dogs. Circ Res 58: 874 – 881, 1986.
1070. Navar LG, Chomdej B, Bell PD. Absence of estimated glomerular pressure autoregulation during interrupted distal delivery. Am J Physiol 229: 1596 –1603, 1975.
1071. Navar LG, Inscho EW, Majid DSA, Imig JD, Harrison-Bernard LM, Mitchell KD.
Paracrine regulation of the renal microcirculation. Physiol Rev 76: 425–536, 1996.
1072. Navar LG, Jirakulsomchok D, Bell PD, Thomas CE, Huang WC. Influence of converting enzyme inhibition on renal hemodynamics and glomerular dynamics in sodium-restricted dogs. Hypertension 4: 58 – 68, 1982.
1073. Navar LG, LaGrange RA, Bell PD, Thomas CE, Ploth DW. Glomerular and renal
hemodynamics during converting enzyme inhibition (SQ20,881) in the dog. Hypertension 1: 371–377, 1979.
1074. Navar LG, Paul RV, Carmines PK, Chou CL, Marsh DJ. Intrarenal mechanisms mediating pressure natriuresis: role of angiotensin and prostaglandins. Federation Proc
45: 2885–2891, 1986.
1075. Navar LG, Ploth DW, Bell PD. Distal tubular feedback control of renal hemodynamics and autoregulation. Annu Rev Physiol 42: 557–571, 1980.
1091. Nishiyama A, Majid DS, Walker M, III, Miyatake A, Navar LG. Renal interstitial ATP
responses to changes in arterial pressure during alterations in tubuloglomerular
feedback activity. Hypertension 37: 753–759, 2001.
1092. Norman RA Jr, Dzielak DJ. Role of the renal nerves in onset and maintenance of
spontaneous hypertension. Am J Physiol Heart Circ Physiol 243: H284 –H288, 1982.
1093. Norman RA Jr, Enobakhare JA, DeClue JW, Douglas BH, Guyton AC. Arterial pressure-urinary output relationship in hypertensive rats. Am J Physiol Regul Integr Comp
Physiol 234: R98 –R103, 1978.
1094. Norrelund H, Christensen KL, Samani NJ, Kimber P, Mulvany MJ, Korsgaard N. Early
narrowed afferent arteriole is a contributor to the development of hypertension.
Hypertension 24: 301–308, 1994.
1095. Nouri P, Gill P, Li M, Wilcox CS, Welch WJ. p22phox in the macula densa regulates
single nephron GFR during angiotensin II infusion in rats. Am J Physiol Heart Circ
Physiol 292: H1685–H1689, 2007.
1096. Novak J, Ramirez RJ, Gandley RE, Sherwood OD, Conrad KP. Myogenic reactivity is
reduced in small renal arteries isolated from relaxin-treated rats. Am J Physiol Regul
Integr Comp Physiol 283: R349 –R355, 2002.
1097. Nowicki PT, Flavahan S, Hassanain H, Mitra S, Holland S, Goldschmidt-Clermont PJ,
Flavahan NA. Redox signaling of the arteriolar myogenic response. Circ Res 89:
114 –116, 2001.
1076. Navar LG, Prieto MC, Satou R, Kobori H. Intrarenal angiotensin II and its contribution to the genesis of chronic hypertension. Curr Opin Pharmacol 11: 180 –186, 2011.
1098. Numabe A, Komatsu K, Frohlich ED. Effects of ANG-converting enzyme and alpha
1-adrenoceptor inhibition on intrarenal hemodynamics in SHR. Am J Physiol Regul
Integr Comp Physiol 266: R1437–R1442, 1994.
1077. Navar LG, Saccomani G, Mitchell KD. Synergistic intrarenal actions of angiotensin on
tubular reabsorption and renal hemodynamics. Am J Hypertens 4: 90 –96, 1991.
1099. O’Connor PM, Cowley AW Jr. Modulation of pressure-natriuresis by renal medullary reactive oxygen species and nitric oxide. Curr Hypertens Rep 12: 86 –92, 2010.
1078. Nelson LD, Mashburn NA, Bell PD. Altered sodium-calcium exchange in afferent
arterioles of the spontaneously hypertensive rat. Kidney Int 50: 1889 –1896, 1996.
1100. O’Donnell MP, Crary GS, Oda H, Kasiske BL, Powell JR, Keane WF. Irbesartan
lowers blood pressure and ameliorates renal injury in experimental non-insulindependent diabetes mellitus. Kidney Int Suppl 63: S218 –S220, 1997.
1079. Nelson LD, Unlap MT, Lewis JL, Bell PD. Renal arteriolar Na⫹/Ca2⫹ exchange in
salt-sensitive hypertension. Am J Physiol Renal Physiol 276: F567–F573, 1999.
1080. Nelson MT, Conway MA, Knot HJ, Brayden JE. Chloride channel blockers inhibit
myogenic tone in rat cerebral arteries. J Physiol 502: 259 –264, 1997.
1101. O’Donnell MP, Kasiske BL, Cleary MP, Keane WF. Effects of genetic obesity on renal
structure and function in the Zucker rat. II. Micropuncture studies. J Lab Clin Med
106: 605– 610, 1985.
1081. Nelson MT, Quayle JM. Physiological roles and properties of potassium channels in
arterial smooth muscle. Am J Physiol Cell Physiol 268: C799 –C822, 1995.
1102. Ochodnicky P, Henning RH, Buikema HJ, de Zeeuw D, Provoost AP, Van Dokkum
RP. Renal vascular dysfunction precedes the development of renal damage in the
hypertensive Fawn-Hooded rat. Am J Physiol Renal Physiol 298: F625–F633, 2010.
1082. Nelson RG, Meyer TW, Myers BD, Bennett PH. Clinical and pathological course of
renal disease in non-insulin-dependent diabetes mellitus: the Pima Indian experience. Semin Nephrol 17: 124 –131, 1997.
1103. Oddie CJ, Dilley RJ, Bobik A. Long-term angiotensin II antagonism in spontaneously
hypertensive rats: effects on blood pressure and cardiovascular amplifiers. Clin Exp
Pharmacol Physiol 19: 392–395, 1992.
1083. Nelson RG, Meyer TW, Myers BD, Bennett PH. Course of renal disease in Pima
Indians with non-insulin-dependent diabetes mellitus. Kidney Int Suppl 63: S45–S48,
1997.
1104. Oeckler RA, Kaminski PM, Wolin MS. Stretch enhances contraction of bovine coronary arteries via an NAD (P) oxidase-medicated activation of the extracellular
signal- regulated kinase mitogen-activated protein kinase cascade. Circ Res 92: 24 –
31, 2003.
1084. New DI, Chesser AM, Thuraisingham RC, Yaqoob MM. Cerebral artery responses
to pressure and flow in uremic hypertensive and spontaneously hypertensive rats.
Am J Physiol Heart Circ Physiol 284: H1212–H1216, 2003.
1085. Newman JM, Dwyer RM, St-Pierre P, Richards SM, Clark MG, Rattigan S. Decreased
microvascular vasomotion and myogenic response in rat skeletal muscle in association with acute insulin resistance. J Physiol 587: 2579 –2588, 2009.
1086. Nguyen Dinh CA, Montezano AC, Burger D, Touyz RM. Angiotensin II, NADPH
oxidase, and redox signaling in the vasculature. Antioxid Redox Signal 19: 1110 –1120,
2013.
1087. Nilsson H, Aalkjaer C. Vasomotion: mechanisms and physiological importance. Mol
Interv 3: 5179 –5189, 2003.
1105. Ofstad J, Iversen BM, Morkrid L, Sekse I. Autoregulation of renal blood flow (RBF)
with and without participation of afferent arterioles. Acta Physiol Scand 130: 25–32,
1987.
1106. Ogawa N. Effect of nicardipine on the relationship of renal blood flow and of renal
vascular resistance to perfusion pressure in dog kidney. J Pharm Pharmacol 42: 138 –
140, 1990.
1107. Ogawa N, Ono H. No role for prostaglandins and bradykinin in the autoregulation of
renal blood flow. Jpn J Pharmacol 39: 349 –355, 1985.
1108. Ogawa N, Ono H. Different effects of noradrenaline, angiotensin II and BAY K 8644
on the abolition of autoregulation of renal blood flow by verapamil. Naunyn-Schmiedebergs Arch Pharmacol 333: 445– 449, 1986.
1088. Nishiyama A, Fukui T, Fujisawa Y, Rahman M, Tian RX, Kimura S, Youichi A. Systemic and regional hemodynamic responses to tempol in angiotensin II-infused hypertensive rats. Hypertension 37: 77– 83, 2001.
1109. Ogawa N, Ono H. Different effects of various vasodilators on autoregulation of renal
blood flow in anesthetized dogs. Jpn J Pharmacol 41: 299 –306, 1986.
1089. Nishiyama A, Jackson KE, Majid DS, Rahman M, Navar LG. Renal interstitial fluid ATP
responses to arterial pressure and tubuloglomerular feedback activation during calcium channel blockade. Am J Physiol Heart Circ Physiol 290: H772–H777, 2006.
1110. Ogawa N, Ono H. Role of Ca channel in the renal autoregulatory vascular response
analysed by the use of BAY K 8644. Naunyn-Schmiedebergs Arch Pharmacol 335:
189 –193, 1987.
1090. Nishiyama A, Majid DS, Taher KA, Miyatake A, Navar LG. Relation between renal
interstitial ATP concentrations and autoregulation-mediated changes in renal vascular resistance. Circ Res 86: 656 – 662, 2000.
1111. Ogawa N, Ono H. Effect of 8-(N,N-diethylamino)octyl-3,4,5-trimethoxybenzoate
(TMB-8), an inhibitor of intracellular Ca2⫹ release, on autoregulation of renal blood
flow in the dog. Naunyn-Schmiedebergs Arch Pharmacol 338: 293–296, 1988.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
497
RENAL AUTOREGULATORY MECHANISMS
1112. Ogawa N, Yokota S, Ono H. Different interaction of bepridil and diltiazem with BAY
K 8644 in the abolition of autoregulation of renal blood flow. J Cardiovasc Pharmacol
11: 147–150, 1988.
1113. Ohishi K, Carmines PK. Superoxide dismutase restores the influence of nitric oxide
on renal arterioles in diabetes mellitus. J Am Soc Nephrol 5: 1559 –1566, 1995.
1114. Ohmine T, Miwa Y, Takahashi-Yanaga F, Morimoto S, Maehara Y, Sasaguri T. The
involvement of aldosterone in cyclic stretch-mediated activation of NADPH oxidase
in vascular smooth muscle cells. Hypertension Res 32: 690 – 699, 2009.
1115. Okamoto K, Aoki K. Development of a strain of spontaneously hypertensive rats. Jpn
Circ J 27: 282–293, 1963.
1116. Okamoto K, Hazama F, Yamori Y, Haebara H, Nagaoka A. Pathogenesis and prevention of stroke in spontaneously hypertensive rats. Clin Sci Mol Med Suppl 2:
161s–163s, 1975.
1117. Okazaki K, Seki S, Kanaya N, Hattori J, Tohse N, Namiki A. Role of endotheliumderived hyperpolarizing factor in phenylephrine-induced oscillatory vasomotion in
rat small mesenteric artery. Anesthesiology 98: 1164 –1171, 2003.
⫺
1118. Okuda T, Kojima I, Ogata E, Kurokawa K. Ambient Cl ions modify rat mesangial cell
contraction by modulating cell inositol trisphosphate and Ca2⫹ via enhanced prostaglandin E2. J Clin Invest 84: 1866 –1872, 1989.
1119. Olson JL, Hostetter TH, Rennke HG, Brenner BM, Venkatachalam MA. Altered
glomerular permselectivity and progressive sclerosis following extreme ablation of
renal mass. Kidney Int 22: 112–126, 1982.
1120. Ono H, Kokubun H, Hashimoto K. Abolition by calcium antagonists of the autoregulation of renal blood flow. Arch Pharmacol 285: 201–207, 1974.
1121. Onozato ML, Tojo A, Goto A, Fujita T, Wilcox CS. Oxidative stress and nitric oxide
synthase in rat diabetic nephropathy: effects of ACEI or ARB. Kidney Int 61: 186 –194,
2002.
1122. Oppermann M, Carota I, Schiessl I, Eisner C, Castrop H, Schnermann J. Direct
assessment of tubulogomerular feedback responsiveness in connexin 40-deficient
mice. Am J Phyisol Renal Physiol 304: F1181–F1186, 2013.
1123. Oppermann M, Friedman DJ, Faulhaber-Walter R, Mizel D, Castrop H, Enjyoji K,
Robson SC, Schnermann J. Tubuloglomerular feedback and renin secretion in NTPDase1/CD39-deficient mice. Am J Physiol Renal Physiol 294: F965–F970, 2008.
1124. Oppermann M, Hansen PB, Castrop H, Schnermann J. Vasodilatation of afferent
arterioles and paradoxical increase of renal vascular resistance by furosemide in
mice. Am J Physiol Renal Physiol 293: F279 –F287, 2007.
1125. Oppermann M, Qin Y, Lai EY, Eisner C, Li L, Huang Y, Mizel D, Fryc J, Wilcox CS,
Briggs J, Schnermann J, Castrop H. Enhanced tubuloglomerular feedback in mice
with vascular overexpression of A1 adenosine receptors. Am J Physiol Renal Physiol
297: F1256 –F1264, 2009.
1126. Orlov SN, Koltsova SV, Tremblay J, Baskakov MB, Hamet P. NKCC1 and hypertension: role in the regulation of vascular smooth muscle contractions and myogenic
tone. Ann Med 44 Suppl 1: S111–S118, 2012.
1127. Ortiz MC, Atucha NM, Lahera V, Vargas F, Quesada T, Garcia-Estan J. Importance
of nitric oxide and prostaglandins in the control of rat renal papillary blood flow.
Hypertension 27: 377–381, 1996.
1128. Ortiz MC, Manriquez MC, Romero JC, Juncos LA. Antioxidants block angiotensin
II-induced increases in blood pressure and endothelin. Hypertension 38: 655– 659,
2001.
1133. Osol G, Halpern W. Myogenic properties of cerebral blood vessels from normotensive and hypertensive rats. Am J Physiol Heart Circ Physiol 249: H914 –H921, 1985.
1134. Osol G, Halpern W. Effect of antihypertensive treatment on myogenic properties of
brain arteries from the stroke-prone rat. J Hypertens Suppl 4: S517–S518, 1986.
1135. Osol G, Halpern W. Spontaneous vasomotion in pressurized cerebral arteries from
genetically hypertensive rats. Am J Physiol Heart Circ Physiol 254: H28 –H33, 1988.
1136. Osol G, Laher I, Cipolla M. Protein kinase C modulates basal myogenic tone in
resistance arteries from the cerebral circulation. Circ Res 68: 359 –367, 1991.
1137. Osol G, Laher I, Kelley M. Myogenic tone is coupled to phospholipase C and G
protein activation in small cerebral arteries. Am J Physiol Heart Circ Physiol 265:
H415–H420, 1993.
1138. Osswald H, Haas JA, Marchand GR, Knox FG. Glomerular dynamics in dogs at
reduced renal artery pressure. Am J Physiol Renal Fluid Electrolyte Physiol 236: F25–
F29, 1979.
1139. Ott CE, Marchand GR, Diaz-Buxo JA, Knox FG. Determinants of glomerular filtration rate in the dog. Am J Physiol 231: 235–239, 1976.
1140. Ott CE, Navar LG, Guyton AC. Pressures in static and dynamic states from capsules
implanted in the kidney. Am J Physiol 221: 394 – 400, 1971.
1141. Ott CE, Vari RC. Renal autoregulation of blood flow and filtration rate in the rabbit.
Am J Physiol Renal Fluid Electrolyte Physiol 237: F479 –F482, 1979.
1142. Owens GK, Kumar MS, Wamhoff BR. Molecular regulation of vascular smooth muscle cell differentiation in development and disease. Physiol Rev 84: 767– 801, 2004.
1143. Owsianik G, Talavera K, Voets T, Nilius B. Permeation and selectivity of TRP channels. Annu Rev Physiol 68: 685–717, 2006.
1144. Oyekan AO, McGiff JC. Functional response of the rat kidney to inhibition of nitric
oxide synthesis: role of cytochrome P450-derived arachidonic metabolites. Br J Pharm
125: 1065–1073, 1998.
1145. Ozawa Y, Hayashi K, Kanda T, Homma K, Takamatsu I, Tatematsu S, Yoshioka K,
Kumagai H, Wakino S, Saruta T. Impaired nitric oxide- and endothelium-derived
hyperpolarizing factor-dependent dilation of renal afferent arteriole in Dahl saltsensitive rats. Nephrology 9: 272–277, 2004.
1146. Ozawa Y, Hayashi K, Wakino S, Kanda T, Homma K, Takamatsu I, Tatematsu S,
Yoshioka K, Saruta T. Free radical activity depends on underlying vasoconstrictors in
renal microcirculation. Clin Exp Hypertens 26: 219 –229, 2004.
1147. Pabbidi MR, Juncos J, Juncos L, Renic M, Tullos HJ, Lazar J, Jacob HJ, Harder DR,
Roman RJ. Identification of a region of rat chromosome 1 that impairs the myogenic
response and autoregulation of cerebral blood flow in fawn-hooded hypertensive
rats. Am J Physiol Heart Circ Physiol 304: H311–H317, 2013.
1148. Pallone TL, Mattson DL. Role of nitric oxide in regulation of the renal medulla in
normal and hypertensive kidneys. Curr Opin Nephrol Hypertens 11: 93–98, 2002.
1149. Pallone TL, Robertson CR, Jamison RL. Renal medullary microcirculation. Physiol Rev
70: 885–920, 1990.
1150. Pandey D, Patel A, Patel V, Chen F, Qian J, Wang Y, Barman SA, Venema RC, Stepp
DW, Rudic RD, Fulton DJ. Expression and functional significance of NADPH oxidase
5 (Nox5) and its splice variants in human blood vessels. Am J Physiol Heart Circ Physiol
302: H1919 –H1928, 2012.
1129. Ortiz PA, Garvin JL. Role of nitric oxide in the regulation of nephron transport. Am J
Physiol Renal Physiol 282: F777–F784, 2002.
1151. Paravicini TM, Touyz RM. NADPH oxidases, reactive oxygen species, and hypertension: clinical implications and therapeutic possibilities. Diabetes Care 31 Suppl 2:
S170 –S180, 2008.
1130. Osborn JL, Francisco LL, DiBona GF. Effect of renal nerve stimulation on renal blood
flow autoregulation and antinatriuresis during reductions in renal perfusion pressure.
Proc Soc Exper Biol Med 168: 77– 81, 1981.
1152. Parekh N, Zou AP, Jungling I, Endlich K, Sadowski J, Steinhausen M. Sex differences
in control of renal outer medullary circulation in rats: role of prostaglandins. Am J
Physiol Renal Fluid Electrolyte Physiol 264: F629 –F636, 1993.
1131. Osmond DA, Inscho EW. P2X(1) receptor blockade inhibits whole kidney autoregulation of renal blood flow in vivo. Am J Physiol Renal Physiol 298: F1360 –F1368,
2010.
1153. Park J, Kemp BA, Howell NL, Gildea JJ, Keller SR, Carey RM. Intact microtubules are
required for natriuretic responses to nitric oxide and increased renal perfusion
pressure. Hypertension 51: 494 – 499, 2008.
1132. Osmond DA, Zhang S, Pollock JS, Yamamoto T, De Miguel C, Inscho EW. Clopidogrel preserves whole kidney autoregulatory behavior in ANG II-induced hypertension. Am J Physiol Renal Physiol 306: F619 –F628, 2014.
1154. Park KS, Kim Y, Lee YH, Earm YE, Ho WK. Mechanosensitive cation channels in
arterial smooth muscle cells are activated by diacylglycerol and inhibited by phospholipase C inhibitor. Circ Res 93: 557–564, 2003.
498
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
1155. Park S, Bivona BJ, Feng Y, Lazartigues E, Harrison-Bernard LM. Intact renal afferent
arteriolar autoregulatory responsiveness in db/db mice. Am J Physiol Renal Physiol 295:
F1504 –F1511, 2008.
1175. Pesic A, Madden JA, Pesic M, Rusch NJ. High blood pressure upregulates arterial
L-type Ca2⫹ channels: is membrane depolarization the signal? Circ Res 94: e97–104,
2004.
1156. Parsa A, Kao WH, Xie D, Astor BC, Li M, Hsu CY, Feldman HI, Parekh RS, Kusek JW,
Greene TH, Fink JC, Anderson AH, Choi MJ, Wright JT Jr, Lash JP, Freedman BI, Ojo
A, Winkler CA, Raj DS, Kopp JB, He J, Jensvold NG, Tao K, Lipkowitz MS, Appel LJ.
APOL1 risk variants, race, and progression of chronic kidney disease. N Engl J Med
369: 2183–2196, 2013.
1176. Peti-Peterdi J. Calcium wave of tubuloglomerular feedback. Am J Physiol Renal Physiol
291: F473–F480, 2006.
1157. Parving HH, Lehnert H, Brochner-Mortensen J, Gomis R, Andersen S, Arner P. The
effect of irbesartan on the development of diabetic nephropathy in patients with type
2 diabetes. N Engl J Med 345: 870 – 878, 2001.
1178. Peti-Peterdi J, Chambrey R, Bebok Z, Biemesderfer D, St John PL, Abrahamson DR,
Warnock DG, Bell PD. Macula densa Na⫹/H⫹ exchange activities mediated by apical
NHE2 and basolateral NHE4 isoforms. Am J Physiol Renal Physiol 278: F452–F463,
2000.
1158. Patschan O, Kuttler B, Heemann U, Uber A, Rettig R. Kidneys from normotensive
donors lower blood pressure in young transplanted spontaneously hypertensive
rats. Am J Physiol Regul Integr Comp Physiol 273: R175–R180, 1997.
1159. Patzak A, Lai EY, Fahling M, Sendeski M, Martinka P, Persson PB, Persson AE.
Adenosine enhances long term the contractile response to angiotensin II in afferent
arterioles. Am J Physiol Regul Integr Comp Physiol 293: R2232–R2242, 2007.
1160. Patzak A, Mrowka R, Storch E, Hocher B, Persson PB. Interaction of angiotensin II
and nitric oxide in isolated perfused afferent arterioles of mice. J Am Soc Nephrol 12:
1122–1127, 2001.
1161. Patzak A, Petzhold D, Wronski T, Martinka P, Babu GJ, Periasamy M, Haase H,
Morano I. Constriction velocities of renal afferent and efferent arterioles of mice are
not related to SMB expression. Kidney Int 68: 2726 –2734, 2005.
1162. Pavenstadt H, Kriz W, Kretzler M. Cell biology of the glomerular podocyte. Physiol
Rev 83: 253–307, 2003.
1163. Pavlov AN, Sosnovtseva OV, Pavlova ON, Mosekilde E, Holstein-Rathlou NH. Characterizing multimode interaction in renal autoregulation. Physiol Meas 29: 945–958,
2008.
1164. Payne GW, Smeda JS. Cerebrovascular alterations in pressure and protein kinase
C-mediated constriction in Dahl salt-sensitive rats. J Hypertens 20: 1355–1363, 2002.
1165. Pelayo JC, Quan AH, Shanley PF. Angiotensin II control of the renal microcirculation
in rats with reduced renal mass. Am J Physiol Renal Fluid Electrolyte Physiol 258:
F414 –F422, 1990.
1166. Pelayo JC, Shanley PF. Glomerular and tubular adaptive responses to acute nephron
loss in the rat. Effect of prostaglandin synthesis inhibition. J Clin Invest 85: 1761–1769,
1990.
1167. Pelayo JC, Westcott JY. Impaired autoregulation of glomerular capillary hydrostatic
pressure in the rat remnant nephron. J Clin Invest 88: 101–105, 1991.
1168. Peled N, Bendayan D, Shitrit D, Fox B, Yehoshua L, Kramer MR. Peripheral endothelial dysfunction in patients with pulmonary arterial hypertension. Respir Med 102:
1791–1796, 2008.
1177. Peti-Peterdi J, Bell PD. Regulation of macula densa Na:H exchange by angiotensin II.
Kidney Int 54: 2021–2028, 1998.
1179. Pfaltzgraff ER, Shelton EL, Galindo CL, Nelms BL, Hooper CW, Poole SD, Labosky
PA, Bader DM, Reese J. Embryonic domains of the aorta derived from diverse origins
exhibit distinct properties that converge into a common phenotype in the adult. J Mol
Cell Cardiol 69: 88 –96, 2014.
1180. Phillips SA, Sylvester FA, Frisbee JC. Oxidant stress and constrictor reactivity impair
cerebral artery dilation in obese Zucker rats. Am J Physiol Regul Integr Comp Physiol
288: R522–R530, 2005.
1181. Pickering TG, Sos TA, Saddekni S, Rozenblit G, James GD, Orenstein A, Helseth G,
Laragh JH. Renal angioplasty in patients with azotaemia and renovascular hypertension. J Hypertens 4 Suppl 6: 667– 669, 1986.
1182. Pires SL, Barres C, Sassard J, Julien C. Renal blood flow dynamics and arterial pressure lability in the conscious rat. Hypertension 38: 147–152, 2001.
1183. Pires SL, Julien C, Chapuis B, Sassard J, Barres C. Spontaneous renal blood flow
autoregulation curves in conscious sinoaortic baroreceptor-denervated rats. Am J
Physiol Renal Physiol 282: F51–F58, 2002.
1184. Pittner J, Wolgast M, Casellas D, Persson AE. Increased shear stress-released NO
and decreased endothelial calcium in rat isolated perfused juxtamedullary nephrons.
Kidney Int 67: 227–236, 2005.
1185. Plato CF, Stoos BA, Wang D, Garvin JL. Endogenous nitric oxide inhibits chloride
transport in the thick ascending limb. Am J Physiol Renal Physiol 276: F159 –F163,
1999.
1186. Ploth DW, Kleeman K, Morrill L, Rademacher R, Jackson CA. Effects of verapamil
and converting enzyme inhibition on bilateral renal function of two-kidney, one-clip
hypertensive rats. Clin Sci 72: 657– 667, 1987.
1187. Ploth DW, Mackenzie HS. Tubuloglomerular feedback in one-kidney, one-clip hypertensive rats. Kidney Int Suppl 32: S115–S118, 1991.
1188. Ploth DW, Roy RN, Huang WC, Navar LG. Impaired renal blood flow and cortical
pressure autoregulation in contralateral kidneys of Goldblatt hypertensive rats. Hypertension 3: 67–74, 1981.
1169. Peng H, Matchkov V, Ivarsen A, Aalkjaer C, Nilsson H. Hypothesis for the initiation
of vasomotion. Circ Res 88: 810 – 815, 2001.
1189. Ploth DW, Rudulph J, LaGrange R, Navar LG. Tubuloglomerular feedback and single
nephron function after converting enzyme inhibition in the rat. J Clin Invest 64:
1325–1335, 1979.
1170. Persson AE, Hahne B, Selen G. The effect of tubular perfusion with PGE2, PGF2 alpha,
and PGI2 on the tubuloglomerular feedback control in the rat. Can J Physiol Pharmacol
61: 1317–1323, 1983.
1190. Ploth DW, Rudulph J, Thomas C, Navar LG. Renal and tubuloglomerular feedback
responses to plasma expansion in the rat. Am J Physiol Renal Fluid Electrolyte Physiol
235: F156 –F162, 1978.
1171. Persson PB, Ehmke H, Kirchheim H. Influence of the renin-angiotensin system on
the autoregulation of renal blood flow and glomerular filtration rate in conscious
dogs. Acta Physiol Scand 134: 1–7, 1988.
1191. Ploth DW, Schnermann J, Dahlheim H, Hermle M, Schmidmeier E. Autoregulation
and tubuloglomerular feedback in normotensive and hypertensive rats. Kidney Int 12:
253–267, 1977.
1172. Persson PB, Ehmke H, Kirchheim HR, Janssen B, Baumann JE, Just A, Nafz B. Autoregulation and non-homeostatic behaviour of renal blood flow in conscious dogs. J
Physiol 462: 261–273, 1993.
1192. Pogoda K, Fuller M, Pohl U, Kameritsch P. NO, via its target Cx37, modulates
calcium signal propagation selectively at myoendothelial gap junctions. Cell Commun
Signal 12: 33–12, 2014.
1173. Persson PB, Ehmke H, Nafz B, Kirchheim HR. Resetting of renal autoregulation in
conscious dogs: angiotensin II and alpha1-adrenoceptors. Pflügers Arch 417: 42– 47,
1990.
1193. Polichnowski AJ, Griffin KA, Long J, Williamson GA, Bidani AK. Blood pressure-renal
blood flow relationships in conscious angiotensin II- and phenylephrine-infused rats.
Am J Physiol Renal Physiol 305: F1074 –F1084, 2013.
1174. Persson PB, Ehmke H, Nafz B, Kirchheim HR. Sympathetic modulation of renal
autoregulation by carotid occlusion in conscious dogs. Am J Physiol Renal Fluid Electrolyte Physiol 258: F364 –F370, 1990.
1194. Pollock CA, Lawrence JR, Field MJ. Tubular sodium handling and tubuloglomerular
feedback in experimental diabetes mellitus. Am J Physiol Renal Fluid Electrolyte Physiol
260: F946 –F952, 1991.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
499
RENAL AUTOREGULATORY MECHANISMS
1195. Pollock DM, Arendshorst WJ. Tubuloglomerular feedback and blood flow autoregulation during DA1-induced renal vasodilation. Am J Physiol Renal Fluid Electrolyte
Physiol 258: F627–F635, 1990.
1196. Polzin DJ, Leininger JR, Osborne CA, Jeraj K. Development of renal lesions in dogs
after 11/12 reduction of renal mass. Influences of dietary protein intake. Lab Invest
58: 172–183, 1988.
1216. Ren Y, Arima S, Carretero OA, Itoh S. Possible role of adenosine in macula densa
control of glomerular hemodynamics. Kidney Int 61: 169 –176, 2002.
1217. Ren Y, Carretero OA, Garvin JL. Mechanism by which superoxide potentiates tubuloglomerular feedback. Hypertension 39: 624 – 628, 2002.
1218. Ren Y, Carretero OA, Garvin JL. Role of mesangial cells and gap junctions in tubuloglomerular feedback. Kidney Int 62: 525–531, 2002.
1197. Povstyan OV, Harhun MI, Gordienko DV. Ca2⫹ entry following P2X receptor activation induces IP3 receptor-mediated Ca2⫹ release in myocytes from small renal
arteries. Br J Pharmacol 162: 1618 –1638, 2011.
1219. Ren Y, D’Ambrosio MA, Garvin JL, Carretero OA. Angiotensin II enhances connecting tubule glomerular feedback. Hypertension 56: 636 – 642, 2010.
1198. Premen AJ, Hall JE, Mizelle HL, Cornell JE. Maintenance of renal autoregulation
during infusion of aminophylline or adenosine. Am J Physiol Renal Fluid Electrolyte
Physiol 248: F366 –F373, 1985.
1220. Ren Y, D’Ambrosio MA, Garvin JL, Peterson EL, Carretero OA. Mechanism of
impaired afferent arteriole myogenic response in Dahl salt sensitive rats: role of
20-HETE. Am J Physiol Renal Physiol 307: F533–F538, 2014.
1199. Pretus HA, Williams A, Navar LG. Impaired autoregulation of glomerular capillary
hydrostatic pressure in the rat remnant nephron (Abstract). J Am Soc Nephrol 4: 586,
1994.
1200. Prior HM, Yates MS, Beech DJ. Functions of large conductance Ca2⫹-activated
(BKCa), delayed rectifier (KV) and background K⫹ channels in the control of membrane potential in rabbit renal arcuate artery. J Physiol 511: 159 –169, 1998.
1201. Prosser BL, Ward CW, Lederer WJ. X-ROS signaling: rapid mechano-chemo transduction in heart. Science 333: 1440 –1445, 2011.
1202. Provoost AP. Spontaneous glomerulosclerosis: insights from the fawn-hooded rat.
Kidney Int Suppl 45: S2–S5, 1994.
1203. Racasan S, Joles JA, Boer P, Koomans HA, Braam B. NO dependency of RBF and
autoregulation in the spontaneously hypertensive rat. Am J Physiol Renal Physiol 285:
F105–F112, 2003.
1204. Racasan S, Turkstra E, Joles JA, Koomans HA, Braam B. Hypoxanthine plus xanthine
oxidase causes profound natriuresis without affecting renal blood flow autoregulation. Kidney Int 64: 226 –231, 2003.
1205. Raghavan R, Chen X, Yip KP, Marsh DJ, Chon KH. Interactions between TGFdependent and myogenic oscillations in tubular pressure and whole kidney blood
flow in both SDR and SHR. Am J Physiol Renal Physiol 290: F720 –F732, 2006.
⫹
1206. Raina H, Ella S, Hill MA. Decreased activity of the smooth muscle Na /Ca
changer impairs arteriolar reactivity. J Physiol 586: 1669 –1681, 2008.
2⫹
ex-
1207. Ramsey CR, Berndt TJ, Knox FG. Indomethacin blocks enhanced paracellular backflux in proximal tubules. J Am Soc Nephrol 13: 1449 –1454, 2002.
1208. Rapp JP, Dene H. Development and characteristics of inbred strains of Dahl saltsensitive and salt-resistant rats. Hypertension 7: 340 –349, 1985.
1209. Rath G, Dessy C, Feron O. Caveolae, caveolin and control of vascular tone: nitric
oxide (NO) and endothelium derived hyperpolarizing factor (EDHF) regulation. J
Physiol Pharmacol 60 Suppl 4: 105–109, 2009.
1210. Raychowdhury MK, Yukawa M, Collins LJ, McGrail SH, Kent KC, Ware JA. Alternative splicing produces a divergent cytoplasmic tail in the human endothelial thromboxane A2 receptor. J Biol Chem 269: 19256 –19261, 1994.
1211. Rayshubskiy A, Wojtasiewicz TJ, Mikell CB, Bouchard MB, Timerman D, Youngerman BE, McGovern RA, Otten ML, Canoll P, McKhann GM, Hillman EM. Direct,
intraoperative observation of ⬃0.1 Hz hemodynamic oscillations in awake human
cortex: implications for fMRI. Neuroimage 87: 323–331, 2014.
1212. Reading SA, Earley S, Waldron BJ, Welsh DG, Brayden JE. TRPC3 mediates pyrimidine receptor-induced depolarization of cerebral arteries. Am J Physiol Heart Circ
Physiol 288: H2055–H2061, 2005.
1213. Reho JJ, Zheng X, Fisher SA. Smooth muscle contractile diversity in the control of
regional circulations. Am J Physiol Heart Circ Physiol 306: H163–H172, 2014.
1221. Ren Y, D’Ambrosio MA, Garvin JL, Wang H, Carretero OA. Possible mediators of
connecting tubule glomerular feedback. Hypertension 53: 319 –323, 2009.
1222. Ren Y, D’Ambrosio MA, Garvin JL, Wang H, Carretero OA. Mechanism of inhibition
of tubuloglomerular feedback by CO and cGMP. Hypertension 62: 99 –104, 2013.
1223. Ren Y, D’Ambrosio MA, Garvin JL, Wang H, Carretero OA. Prostaglandin E2 mediated connecting tubule glomerular feedback. Hypertension 62: 1123–1128, 2013.
1224. Ren Y, D’Ambrosio MA, Liu R, Pagano PJ, Garvin JL, Carretero OA. Enhanced
myogenic response in the afferent arteriole of spontaneously hypertensive rats. Am
J Physiol Heart Circ Physiol 298: H1769 –H1775, 2010.
1225. Ren Y, D’Ambrosio MA, Wang H, Falck JR, Peterson EL, Garvin JL, Carretero OA.
Mechanisms of carbon monoxide attenuation of tubuloglomerular feedback. Hypertension 59: 1139 –1144, 2012.
1226. Ren Y, D’Ambrosio MA, Wang H, Liu R, Garvin JL, Carretero OA. Heme oxygenase
metabolites inhibit tubuloglomerular feedback (TGF). Am J Physiol Renal Physiol 295:
F1207–F1212, 2008.
1227. Ren Y, D’Ambrosio MA, Wang H, Peterson EL, Garvin JL, Carretero OA. Mechanisms of angiotensin II-enhanced connecting tubule glomerular feedback. Am J Physiol
Renal Physiol 303: F259 –F265, 2012.
1228. Ren Y, Garvin JL, Carretero OA. Role of macula densa nitric oxide and cGMP in the
regulation of tubuloglomerular feedback. Kidney Int 58: 2053–2060, 2000.
1229. Ren Y, Garvin JL, Liu R, Carretero OA. Role of macula densa adenosine triphosphate
(ATP) in tubuloglomerular feedback. Kidney Int 66: 1479 –1485, 2004.
1230. Ren Y, Garvin JL, Liu R, Carretero OA. Crosstalk between the connecting tubule and
the afferent arteriole regulates renal microcirculation. Kidney Int 71: 1116 –1121,
2007.
1231. Reslerova M, Loutzenhiser R. Divergent mechanisms of ATP-sensitive K⫹ channelinduced vasodilation in renal afferent and efferent arterioles. Evidence of L-type
Ca2⫹ channel-dependent and -independent actions of pinacidil. Circ Res 77: 1114 –
1120, 1995.
1232. Reslerova M, Loutzenhiser R. Renal microvascular actions of calcitonin gene-related
peptide. Am J Physiol Renal Physiol 274: F1078 –F1085, 1998.
1233. Resta TC, Broughton BR, Jernigan NL. Reactive oxygen species and RhoA signaling in
vascular smooth muscle: role in chronic hypoxia-induced pulmonary hypertension.
Adv Exp Med Biol 661: 355–373, 2010.
1234. Retailleau K, Toutain B, Galmiche G, Fassot C, Sharif-Naeini R, Kauffenstein G,
Mericskay M, Duprat F, Grimaud L, Merot J, Lardeux A, Pizard A, Baudrie V, Jeunemaitre X, Feil R, Gothert JR, Lacolley P, Henrion D, Li Z, Loufrani L. Selective
involvement of serum response factor in pressure-induced myogenic tone in resistance arteries. Arterioscler Thromb Vasc Biol 33: 339 –346, 2013.
1235. Rettig R, Bandelow N, Patschan O, Kuttler B, Frey B, Uber A. The importance of the
kidney in primary hypertension: insights from cross-transplantation. J Hum Hypertens 10: 641– 644, 1996.
1214. Reimann K, Krishnamoorthy G, Wangemann P. NOS inhibition enhances myogenic
tone by increasing rho-kinase mediated Ca2⫹ sensitivity in the male but not the
female gerbil spiral modiolar artery. PLoS One 8: e53655, 2013.
1236. Rettig R, Grisk O. The kidney as a determinant of genetic hypertension: evidence
from renal transplantation studies. Hypertension 46: 463– 468, 2005.
1215. Ren H, Liu NY, Andreasen A, Thomsen JS, Cao L, Christensen EI, Zhai XY. Direct
physical contact between intercalated cells in the distal convoluted tubule and the
afferent arteriole in mouse kidneys. PLoS One 8: e70898, 2013.
1237. Rhaleb NE, Pokharel S, Sharma U, Carretero OA. Renal protective effects of Nacetyl-Ser-Asp-Lys-Pro in deoxycorticosterone acetate-salt hypertensive mice. J Hypertens 29: 330 –338, 2011.
500
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
1238. Ritter ER. Pressure/flow relations in the kidney: Alleged effects of pulse pressure. Am
J Physiol 168: 480 – 489, 1952.
1239. Rivadulla C, de LC, Grieve KL, Cudeiro J. Vasomotion and neurovascular coupling in
the visual thalamus in vivo. PLoS One 6: e28746, 2011.
1240. Roald AB, Ofstad J, Iversen BM. Attenuated buffering of renal perfusion pressure
variation in juxtamedullary cortex in SHR. Am J Physiol Renal Physiol 282: F506 –F511,
2002.
1241. Robertson CR, Deen WM, Troy JL, Brenner BM. Dynamics of glomerular ultrafiltration in the rat. 3. Hemodynamics and autoregulation. Am J Physiol 223: 1191–1200,
1972.
1242. Rodriguez F, Zhang F, Dinocca S, Nasjletti A. Nitric oxide synthesis influences the
renal vascular response to heme oxygenase inhibition. Am J Physiol Renal Physiol 284:
F1255–F1262, 2003.
1243. Roman RJ. Alterations in renal medullary hemodynamics and the pressure-natriuretic response in genetic hypertension. Am J Hypertens 3: 893–900, 1990.
1244. Roman RJ. Abnormal renal hemodynamics and pressure-natriuresis relationship in
Dahl salt-sensitive rats. Am J Physiol Renal Fluid Electrolyte Physiol 251: F57–F65,
1986.
1245. Roman RJ. Pressure-diuresis in volume-expanded rats. Tubular reabsorption in superficial and deep nephrons. Hypertension 12: 177–183, 1988.
1246. Roman RJ, Cowley AW Jr. Abnormal pressure-diuresis-natriuresis response in spontaneously hypertensive rats. Am J Physiol Renal Fluid Electrolyte Physiol 248: F199 –
F205, 1985.
1261. Rosivall L, Youngblood P, Navar LG. Renal autoregulatory efficiency during angiotensin-converting enzyme inhibition in dogs on a low sodium diet. Renal Physiol 9:
18 –28, 1986.
1262. Rossi M, Bertuglia S, Varanini M, Giusti A, Santoro G, Carpi A. Generalised wavelet
analysis of cutaneous flowmotion during post-occlusive reactive hyperaemia in patients with peripheral arterial obstructive disease. Biomed Pharmacother 59: 233–
239, 2005.
1263. Rothe CF, Nash FD, Thompson DE. Patterns in autoregulation of renal blood flow in
the dog. Am J Physiol 220: 1621–1626, 1971.
1264. Rothermund L, Traupe T, Dieterich M, Kossmehl P, Yagil C, Yagil Y, Kreutz R.
Nephroprotective effects of the endothelin ET(A) receptor antagonist darusentan in
salt-sensitive genetic hypertension. Eur J Pharmacol 468: 209 –216, 2003.
1265. Ruan X, Arendshorst WJ. Role of protein kinase C in angiotensin II-induced renal
vasoconstriction in genetically hypertensive rats. Am J Physiol Renal Fluid Electrolyte
Physiol 270: F945–F952, 1996.
1266. Ruan X, Wagner C, Chatziantoniou C, Kurtz A, Arendshorst WJ. Regulation of
angiotensin II receptor AT1 subtypes in renal afferent arterioles during chronic
changes in sodium diet. J Clin Invest 99: 1072–1081, 1997.
1267. Rubin MJ, Bohlen HG. Cerebral vascular autoregulation of blood flow and tissue PO2
in diabetic rats. Am J Physiol Heart Circ Physiol 249: H540 –H546, 1985.
1268. Rudd MA, Grippo RS, Arendshorst WJ. Acute renal denervation produces a diuresis
and natriuresis in young SHR but not WKY rats. Am J Physiol Renal Fluid Electrolyte
Physiol 251: F655–F661, 1986.
1247. Roman RJ, Cowley AW Jr. Characterization of a new model for the study of pressure-natriuresis in the rat. Am J Physiol Renal Fluid Electrolyte Physiol 248: F190 –F198,
1985.
1269. Sachidanandam K, Hutchinson JR, Elgebaly MM, Mezzetti EM, Dorrance AM, Motamed K, Ergul A. Glycemic control prevents microvascular remodeling and increased
tone in type 2 diabetes: link to endothelin-1. Am J Physiol Regul Integr Comp Physiol
296: R952–R959, 2009.
1248. Roman RJ, Cowley AW Jr, Garcia-Estan J, Lombard JH. Pressure-diuresis in volumeexpanded rats Cortical and medullary hemodynamics. Hypertension 12: 168 –176,
1988.
1270. Sadowski J, Kulczykowska E, Kulczykowski M, Badzynska B. Renal vein outflow
recording in rats and rabbits: alternative method of RBF measurement. Acta Physiol
Hung 72: 335–342, 1988.
1249. Roman RJ, Harder DR. Cellular and ionic signal transduction mechanisms for the
mechanical activation of renal arterial vascular smooth muscle. J Am Soc Nephrol 4:
986 –996, 1993.
1271. Saeed A, DiBona GF, Grimberg E, Nguy L, Mikkelsen ML, Marcussen N, Guron G.
High-NaCl diet impairs dynamic renal blood flow autoregulation in rats with adenine-induced chronic renal failure. Am J Physiol Regul Integr Comp Physiol 306: R411–
R419, 2014.
1250. Roman RJ, Hoagland KM, Lopez B, Kwitek AE, Garrett MR, Rapp JP, Lazar J, Jacob
HJ, Sarkis A. Characterization of blood pressure and renal function in chromosome
5 congenic strains of Dahl S rats. Am J Physiol Renal Physiol 290: F1463–F1471, 2006.
1251. Roman RJ, Kaldunski M. Pressure natriuresis and cortical and papillary blood flow in
inbred Dahl rats. Am J Physiol Regul Integr Comp Physiol 261: R595–R602, 1991.
1272. Saeed A, DiBona GF, Guron G. Effects of endothelin receptor antagonists on renal
hemodynamics in angiotensin II-infused rats on high NaCl intake. Kidney Blood Press
Res 36: 258 –267, 2012.
1252. Roman RJ, Kaldunski ML. Renal cortical and papillary blood flow in spontaneously
hypertensive rats. Hypertension 11: 657– 663, 1988.
1273. Saeed A, DiBona GF, Marcussen N, Guron G. High-NaCl intake impairs dynamic
autoregulation of renal blood flow in ANG II-infused rats. Am J Physiol Regul Integr
Comp Physiol 299: R1142–R1149, 2010.
1253. Roman RJ, Kaldunski ML, Mattson DL, Mistry M, Nasjletti A. Influence of eicosanoids
on renal function of DOCA-salt hypertensive rats. Hypertension 12: 287–294, 1988.
1274. Sakai T, Hallman E, Marsh DJ. Frequency domain analysis of renal autoregulation in
the rat. Am J Physiol Renal Fluid Electrolyte Physiol 250: F364 –F373, 1986.
1254. Roman RJ, Lianos E. Influence of prostaglandins on papillary blood flow and pressurenatriuretic response. Hypertension 15: 29 –35, 1990.
1275. Salamanca DA, Khalil RA. Protein kinase C isoforms as specific targets for modulation
of vascular smooth muscle function in hypertension. Biochem Pharmacol 70: 1537–
1547, 2005.
1255. Roman RJ, Osborn JL. Renal function and sodium balance in conscious Dahl S and R
rats. Am J Physiol Regul Integr Comp Physiol 252: R833–R841, 1987.
1256. Roman RJ, Smits C. Laser-Doppler determination of papillary blood flow in young
and adult rats. Am J Physiol Renal Fluid Electrolyte Physiol 251: F115–F124, 1986.
1276. Sallstrom J, Carlsson PO, Fredholm BB, Larsson E, Persson AE, Palm F. Diabetesinduced hyperfiltration in adenosine A(1)-receptor deficient mice lacking the tubuloglomerular feedback mechanism. Acta Physiol 190: 253–259, 2007.
1257. Roman RJ, Zou AP. Influence of the renal medullary circulation on the control of
sodium excretion. Am J Physiol Regul Integr Comp Physiol 265: R963–R973, 1993.
1277. Salom MG, Lahera V, Miranda-Guardiola F, Romero JC. Blockade of pressure natriuresis induced by inhibition of renal synthesis of nitric oxide in dogs. Am J Physiol Renal
Fluid Electrolyte Physiol 262: F718 –F722, 1992.
1258. Romero JC, Knox FG. Mechanisms underlying pressure-related natriuresis: the role
of the renin-angiotensin and prostaglandin systems. Hypertension 11: 724 –738,
1988.
1278. Salomonsson M, Braunstein TH, Holstein-Rathlou NH, Jensen LJ. Na⫹-independent,
nifedipine-resistant rat afferent arteriolar Ca2⫹ responses to noradrenaline: possible
role of TRPC channels. Acta Physiol 200: 265–278, 2010.
1259. Roos MH, van Rodijnen WF, Van Lambalgen AA, Ter Wee PM, Tangelder GJ. Renal
microvascular constriction to membrane depolarization and other stimuli: pivotal
role for rho-kinase. Pflügers Arch 452: 471– 477, 2006.
1279. Salomonsson M, Gustafsson F, Andreasen D, Jensen BL, Holstein-Rathlou NH. Local
electric stimulation causes conducted calcium response in rat interlobular arteries.
Am J Physiol Renal Physiol 283: F473–F480, 2002.
1260. Rosen S, Epstein FH, Brezis M. Determinants of intrarenal oxygenation: factors in
acute renal failure. Renal Failure 14: 321–325, 1992.
1280. Salomonsson M, Sorensen CM, Arendshorst WJ, Steendahl J, Holstein-Rathlou NH.
Calcium handling in afferent arterioles. Acta Physiol Scand 181: 421– 429, 2004.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
501
RENAL AUTOREGULATORY MECHANISMS
1281. Sanchez-Ferrer CF, Roman RJ, Harder DR. Pressure-dependent contraction of rat
juxtamedullary afferent arterioles. Circ Res 64: 790 –798, 1989.
1282. Sanchez-Lozada LG, Tapia E, Soto V, Avila-Casado C, Franco M, Wessale JL, Zhao L,
Johnson RJ. Effect of febuxostat on the progression of renal disease in 5/6 nephrectomy rats with and without hyperuricemia. Nephron Physiol 108: 69 –78, 2008.
1283. Sandner P, Kornfeld M, Ruan X, Arendshorst WJ, Kurtz A. Nitric oxide/cAMP interactions in the control of rat renal vascular resistance. Circ Res 84: 186 –192, 1999.
1284. Sanowski J, Wocial B. Renin release and autoregulation of blood flow in a new model
of non-filtering non-transporting kidney. J Physiol 266: 219 –233, 1977.
1285. Sarubbi D, Quilley J. Evidence against a role of arachidonic acid metabolites in
autoregulatory responses of the isolated perfused rat kidney. Eur J Pharmacol 197:
27–31, 1991.
1286. Satoh M, Haruna Y, Fujimoto S, Sasaki T, Kashihara N. Telmisartan improves endothelial dysfunction and renal autoregulation in Dahl salt-sensitive rats. Hypertension
Res 33: 135–142, 2010.
1287. Satriano J, Wead L, Cardus A, Deng A, Boss GR, Thomson SC, Blantz RC. Regulation
of ecto-5’-nucleotidase by NaCl and nitric oxide: potential roles in tubuloglomerular
feedback and adaptation. Am J Physiol Renal Physiol 291: F1078 –F1082, 2006.
1288. Savage T, McMahon AC, Mullen A, Tribe RM, Yaqoob MM. Ramipril prevents basal
arterial constriction and enhanced myogenic tone in the femoral artery in mildly
uraemic normotensive rats. Clin Sci 97: 233–237, 1999.
1289. Savage T, McMahon AC, Mullen AM, Nott CA, Dodd SM, Tribe RM, Yaqoob MM.
Increased myogenic tone precedes structural changes in mild experimental uraemia
in the absence of hypertension in rats. Clin Sci 95: 681– 686, 1998.
1290. Schaldecker T, Kim S, Tarabanis C, Tian D, Hakroush S, Castonguay P, Ahn W,
Wallentin H, Heid H, Hopkins CR, Lindsley CW, Riccio A, Buvall L, Weins A, Greka
A. Inhibition of the TRPC5 ion channel protects the kidney filter. J Clin Invest 123:
5298 –5309, 2013.
1291. Schiffrin EL. Endothelin:role in experimental hypertension. J Cardiovasc Pharmacol
35: S33–S35, 2000.
1292. Schiffrin EL, Lariviere R, Li JS, Sventek P, Touyz RM. Deoxycorticosterone acetate
plus salt induces overexpression of vascular endothelin-1 and severe vascular hypertrophy in spontaneously hypertensive rats. Hypertension 25: 769 –773, 1995.
1293. Schjoedt KJ, Christensen PK, Jorsal A, Boomsma F, Rossing P, Parving HH. Autoregulation of glomerular filtration rate during spironolactone treatment in hypertensive
patients with type 1 diabetes: a randomized crossover trial. Nephrol Dial Transplant
24: 3343–3349, 2009.
1294. Schleifenbaum J, Kassmann M, Szijarto IA, Hercule HC, Tano JY, Weinert S, Heidenreich M, Pathan AR, Anistan YM, Alenina N, Rusch NJ, Bader M, Jentsch TJ, Gollasch
M. Stretch-activation of angiotensin II type 1a receptors contributes to the myogenic
response of mouse mesenteric and renal arteries. Circ Res 115: 263–272, 2014.
1301. Schnackenberg CG, Welch WJ, Wilcox CS. TP receptor-mediated vasoconstriction
in microperfused afferent arterioles: role O2⫺ and NO. Am J Physiol Renal Physiol 279:
F302–F308, 2000.
1302. Schnackenberg CG, Wilcox CS. Two-week administration of tempol attenuates
both hypertension and renal excretion of 8-Iso prostaglandin F2alpha. Hypertension
33: 424 – 428, 1999.
1303. Schnackenberg CG, Wilcox CS. The SOD mimetic tempol restores vasodilation in
afferent arterioles of experimental diabetes. Kidney Int 59: 1859 –1864, 2001.
1304. Schneider MP, Ge Y, Pollock DM, Pollock JS, Kohan DE. Collecting duct-derived
endothelin regulates arterial pressure and Na excretion via nitric oxide. Hypertension
51: 1605–1610, 2008.
1305. Schnermann J. Micropuncture analysis of tubuloglomerular feedback regulation in
transgenic mice. J Am Soc Nephrol 10: 2614 –2619, 1999.
1306. Schnermann J. Maintained tubuloglomerular feedback responses during acute inhibition of P2 purinergic receptors in mice. Am J Physiol Renal Physiol 300: F339 –F344,
2011.
1307. Schnermann J, Briggs JP. Participation of renal cortical prostaglandins in the regulation of glomerular filtration rate. Kidney Int 19: 802– 815, 1981.
1308. Schnermann J, Briggs JP. Interaction between loop of Henle flow and arterial pressure as determinants of glomerular pressure. Am J Physiol Renal Fluid Electrolyte
Physiol 256: F421–F429, 1989.
1309. Schnermann J, Briggs JP. Restoration of tubuloglomerular feedback in volume-expanded rats by angiotensin II. Am J Physiol Renal Fluid Electrolyte Physiol 259: F565–
F572, 1990.
1310. Schnermann J, Briggs JP. Role of NO in the function of the juxtaglomerular apparatus.
In: Nitric Oxide and the Kidney: Physiology and Pathophysiology, edited by Goligorsky
MS and Gross SS. New York: Chapman & Hall, 1997.
1311. Schnermann J, Briggs JP. Function of the juxtaglomerular apparatus: control of glomerular hemodynamics and renin secretion. In: Seldin and Giebisch’s The Kidney:
Physiology and Pathophysiology, edited by Alpern RJ and Hebert SC. New York:
Academic, 2007.
1312. Schnermann J, Briggs JP. Tubuloglomerular feedback: mechanistic insights from
gene-manipulated mice. Kidney Int 74: 418 – 426, 2008.
1313. Schnermann J, Briggs JP, Weber PC. Tubuloglomerular feedback, prostaglandins,
and angiotensin in the autoregulation of glomerular filtration rate. Kidney Int 25:
53– 64, 1984.
1314. Schnermann J, Gokel M, Weber PC, Schubert G, Briggs JP. Tubuloglomerular feedback and glomerular morphology in Goldblatt hypertensive rats on varying protein
diets. Kidney Int 29: 520 –529, 1986.
1315. Schnermann J, Hermle M, Schmidmeier E, Dahlheim H. Impaired potency for feedback regulation of gloemrular filtration rate in DOCA escaped rats. Pflügers Arch 358:
325–338, 1975.
1295. Schmid HE Jr, Spencer MP. Characteristics of pressure-flow regulation by the kidney. J Appl Physiol 17: 201–204, 1962.
1316. Schnermann J, Levine DZ. Paracrine factors in tubuloglomerular feedback: adenosine, ATP, and nitric oxide. Annu Rev Physiol 65: 501–529, 2003.
1296. Schmidt VJ, Jobs A, von MJ, Worsdorfer P, Willecke K, de Wit C. Connexin45 is
expressed in vascular smooth muscle but its function remains elusive. PLoS One 7:
e42287, 2012.
1317. Schnermann J, Ploth DW, Dahlheim H. Inter-relationship between autoregulation of
glomerular filtration rate, tubuloglomerular feedback and juxtaglomerular renin activity in normotensive and hypertensive rats. Clin Sci Mol Med 51: 105s–107s, 1976.
1297. Schmieder RE, Messerli FH, Garavaglia G, Nunez B. Glomerular hyperfiltration
indicates early target organ damage in essential hypertension. JAMA 264: 2775–2780,
1990.
1318. Schnermann J, Traynor TR, Yang T, Huang YG, Oliverio MI, Coffman T, Briggs JP.
Absence of tubuloglomerular feedback responses in AT1A receptor-deficient mice.
Am J Physiol Renal Physiol 273: F315–F320, 1997.
1298. Schmieder RE, Veelken R, Gatzka CD, Ruddel H, Schachinger H. Predictors for
hypertensive nephropathy: results of a 6-year follow-up study in essential hypertension. J Hypertens 13: 357–365, 1995.
1299. Schmitz PG, O’Donnell MP, Kasiske BL, Katz SA, Keane WF. Renal injury in obese
Zucker rats: glomerular hemodynamic alterations and effects of enalapril. Am J
Physiol Renal Fluid Electrolyte Physiol 263: F496 –F502, 1992.
1300. Schnackenberg CG, Welch WJ, Wilcox CS. Normalization of blood pressure and
renal vascular resistance in SHR with a membrane-permeable superoxide dismutase
mimetic: role of nitric oxide. Hypertension 32: 59 – 64, 1998.
502
1319. Schnermann J, Weber PC. Reversal of indomethacin-induced inhibition of tubuloglomerular feedback by protaglandin infusion. Prostaglandins 24: 351–361, 1982.
1320. Schnermann J, Weihprecht H, Briggs JP. Inhibition of tubuloglomerular feedback
during adenosine1 receptor blockade. Am J Physiol Renal Fluid Electrolyte Physiol 258:
F553–F561, 1990.
1321. Scholey JW, Meyer TW. Control of glomerular hypertension by insulin administration in diabetic rats. J Clin Invest 83: 1384 –1389, 1989.
1322. Scholz H, Kurtz A. Role of protein kinase C in renal vasoconstriction caused by
angiotensin II. Am J Physiol Cell Physiol 259: C421–C426, 1990.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
1323. Scholz H, Kurtz A. Disparate effects of calcium channel blockers on pressure dependence of renin secretion and flow in the isolated perfused rat kidney. Pflügers Arch
421: 155–162, 1992.
1345. Selkurt EE, Hall PW, Spencer MP. Influence of graded arterial pressure decrement
on renal clearance of creatinine, p-aminohippurate and sodium. Am J Physiol 159:
369 –378, 1949.
1324. Scholz H, Kurtz A. Differential regulation of cytosolic calcium between afferent
arteriol ar smooth muscle cells from mouse kidney. Pflügers Arch 431: 46 –51, 1995.
1346. Sell M, Boldt W, Markwardt F. Desynchronising effect of the endothelium on intracellular Ca2⫹ concentration dynamics in vascular smooth muscle cells of rat mesenteric arteries. Cell Calcium 32: 105–120, 2002.
1325. Schoonmaker GC, Fallet RW, Carmines PK. Superoxide anion curbs nitric oxide
modulation of afferent arteriolar Ang II responsiveness in diabetes mellitus. Am J
Physiol Renal Physiol 278: F302–F309, 2000.
1347. Semple SJ, de Wardener HE. Effect of increased renal venous pressure on circulatory
autoregulation of isolated dog kidneys. Circ Res 7: 643– 648, 1959.
1326. Schrier RW, Estacio RO, Esler A, Mehler P. Effects of aggressive blood pressure
control in normotensive type 2 diabetic patients on albuminuria, retinopathy and
strokes. Kidney Int 61: 1086 –1097, 2002.
1348. Sendeski MM, Liu ZZ, Perlewitz A, Busch JF, Ikromov O, Weikert S, Persson PB,
Patzak A. Functional characterization of isolated, perfused outermedullary descending human vasa recta. Acta Physiol 208: 50 –56, 2013.
1327. Schrijvers BF, De Vriese AS, Van de Voorde J, Rasch R, Lameire NH, Flyvbjerg A.
Long-term renal changes in the Goto-Kakizaki rat, a model of lean type 2 diabetes.
Nephrol Dial Transplant 19: 1092–1097, 2004.
1349. Seppey D, Sauser R, Koenigsberger M, Beny JL, Meister JJ. Intercellular calcium
waves are associated with the propagation of vasomotion along arterial strips. Am J
Physiol Heart Circ Physiol 298: H488 –H496, 2010.
1328. Schubert R, Lidington D, Bolz SS. The emerging role of Ca2⫹ sensitivity regulation in
promoting myogenic vasoconstriction. Cardiovasc Res 77: 8 –18, 2008.
1350. Sgouralis I, Layton AT. Autoregulation and conduction of vasomotor responses in a
mathematical model of the rat afferent arteriole. Am J Physiol Renal Physiol 303:
F229 –F239, 2012.
1329. Schubert R, Mulvany MJ. The myogenic response: established facts and attractive
hypotheses. Clin Sci 96: 313–326, 1999.
1330. Schulman IH, Zhou MS, Raij L. Nitric oxide, angiotensin II, and reactive oxygen
species in hypertension and atherogenesis. Curr Hypertens Rep 7: 61– 67, 2005.
1331. Schurek HJ, Jost U, Baumgartl H, Bertram H, Heckmann U. Evidence for a preglomerular oxygen diffusion shunt in rat renal cortex. Am J Physiol Renal Fluid Electrolyte
Physiol 259: F910 –F915, 1990.
1332. Schurek HJ, Kriz W. Morphologic and functional evidence for oxygen deficiency in
the isolated perfused rat kidney. Lab Invest 53: 145–155, 1985.
1333. Schuster A, Lamboley M, Grange C, Oishi H, Beny JL, Stergiopulos N, Meister JJ.
Calcium dynamics and vasomotion in rat mesenteric arteries. J Cardiovasc Pharmacol
43: 539 –548, 2004.
1334. Schweda F, Klar J, Narumiya S, Nusing RM, Kurtz A. Stimulation of renin release by
prostaglandin E2 is mediated by EP2 and EP4 receptors in mouse kidneys. Am J
Physiol Renal Physiol 287: F427–F433, 2004.
1335. Schweda F, Seebauer H, Kramer BK, Kurtz A. Functional role of sodium-calcium
exchange in the regulation of renal vascular resistance. Am J Physiol Renal Physiol 280:
F155–F161, 2001.
1336. Schweitzer G, Gertz KH. Single-nephron hemodynamics in one-kidney one-clip
hypertension in the rat. Renal Physiol 7: 46 –53, 1984.
1337. Schwietzer G, Gertz KH. Changes of hemodynamics and glomerular ultrafiltration in
renal hypertension of rats. Kidney Int 15: 134 –143, 1979.
1338. Schwietzer G, Gertz KH. Goldblatt hypertension in rats: a model for unilateral renal
artery stenosis in man. Contrib Nephrol 19: 134 –138, 1980.
1339. Scotland RS, Chauhan S, Vallance PJ, Ahluwalia A. An endothelium-derived hyperpolarizing factor-like factor moderates myogenic constriction of mesenteric resistance arteries in the absence of endothelial nitric oxide synthase-derived nitric oxide.
Hypertension 38: 833– 839, 2001.
1340. Scully CG, Mitrou N, Braam B, Cupples WA, Chon KH. Detecting physiological
systems with laser speckle perfusion imaging of the renal cortex. Am J Physiol Regul
Integr Comp Physiol 304: R929 –R939, 2013.
1341. Scully CG, Siu KL, Cupples WA, Braam B, Chon KH. Time-frequency approaches for
the detection of interactions and temporal properties in renal autoregulation. Ann
Biomed Eng 41: 172–184, 2013.
1342. Seeliger E, Wronski T, Ladwig M, Dobrowolski L, Vogel T, Godes M, Persson PB,
Flemming B. The renin-angiotensin system and the third mechanism of renal blood
flow autoregulation. Am J Physiol Renal Physiol 296: F1334 –F1345, 2009.
1343. Selen G, Persson AE. Effects of reduced renal artery pressure on feedback control of
glomerular filtration. Am J Physiol Renal Fluid Electrolyte Physiol 244: F342–F348,
1983.
1344. Selkurt EE. Effect of pulse pressure and mean arterial pressure modification on renal
hemodynamics and electrolyte and water excretion. Circ 4: 541–551, 1951.
1351. Sgouralis I, Layton AT. Theoretical assessment of renal autoregulatory mechanisms.
Am J Physiol Renal Physiol 306: F1357–F1371, 2014.
1352. Sharif-Naeini R, Dedman A, Folgering JH, Duprat F, Patel A, Nilius B, Honore E. TRP
channels and mechanosensory transduction: insights into the arterial myogenic response. Pflügers Arch 456: 529 –540, 2008.
1353. Sharif-Naeini R, Folgering JH, Bichet D, Duprat F, Lauritzen I, Arhatte M, Jodar M,
Dedman A, Chatelain FC, Schulte U, Retailleau K, Loufrani L, Patel A, Sachs F,
Delmas P, Peters DJ, Honore E. Polycystin-1 and -2 dosage regulates pressure
sensing. Cell 139: 587–596, 2009.
1354. Sharma K, Cook A, Smith M, Valancius C, Inscho EW. TGF-beta impairs renal
autoregulation via generation of ROS. Am J Physiol Renal Physiol 288: F1069 –F1077,
2005.
1355. Sharma K, McCue P, Dunn SR. Diabetic kidney disease in the db/db mouse. Am J
Physiol Renal Physiol 284: F1138 –F1144, 2003.
1356. Shi S, Carattino MD, Hughey RP, Kleyman TR. ENaC regulation by proteases and
shear stress. Curr Mol Pharmacol 6: 28 –34, 2013.
1357. Shi Y, Lau C, Cupples WA. Interactive modulation of renal myogenic autoregulation
by nitric oxide and endothelin acting through ET-B receptors. Am J Physiol Regul
Integr Comp Physiol 292: R354 –R361, 2007.
1358. Shi Y, Wang X, Chon KH, Cupples WA. Tubuloglomerular feedback-dependent
modulation of renal myogenic autoregulation by nitric oxide. Am J Physiol Regul Integr
Comp Physiol 290: R982–R991, 2006.
1359. Shibuya J, Ohyanagi M, Iwasaki T. Enhanced myogenic response in resistance small
arteries from spontaneously hypertensive rats: relationship to the voltage-dependent calcium channel. Am J Hypertens 11: 767–773, 1998.
1360. Shipley RE, Study RS. Changes in renal blood flow, extraction of inulin, glomerular
filtation rate, tissue pressure and urine flow with acute alterations of renla artery
blood pressure. Am J Physiol 167: 676 – 688, 1951.
1361. Shiraishi M, Wang X, Walsh MP, Kargacin G, Loutzenhiser K, Loutzenhiser R. Myosin
heavy chain expression in renal afferent and efferent arterioles: relationship to contractile kinetics and function. FASEB J 17: 2284 –2286, 2003.
1362. Siegl D, Koeppen M, Wolfle SE, Pohl U, de Wit C. Myoendothelial coupling is not
prominent in arterioles within the mouse cremaster microcirculation in vivo. Circ Res
97: 781–788, 2005.
1363. Sigmon DH, Beierwaltes WH. Influence of nitric oxide in the chronic phase of
two-kidney, one clip renovascular hypertension. Hypertension 31: 649 – 656, 1998.
1364. Silbernagl S. The renal handling of amino acids and oligopeptides. Physiol Rev 68:
911–1007, 1988.
1365. Sima CA, Koeners MP, Joles JA, Braam B, Magil AB, Cupples WA. Increased susceptibility to hypertensive renal disease in streptozotocin-treated diabetic rats is not
modulated by salt intake. Diabetologia 55: 2246 –2255, 2012.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
503
RENAL AUTOREGULATORY MECHANISMS
1366. Simchon S, Manger WM, Shi GS, Brensilver J. Impaired renal vascular reactivity in
prehypertensive Dahl salt-sensitive rats. Hypertension 20: 524 –532, 1992.
1367. Simon Bulley SB, Jaggar JH. Cl⫺ channels in smooth muscle cells. Pflügers Arch 466:
861– 872, 2013.
1388. Smirnov SV, Loutzenhiser K, Loutzenhiser RD. Voltage-activated Ca2⫹ channels in
rat renal afferent and efferent myocytes: no evidence for the T-type Ca2⫹ current.
Cardiovasc Res 97: 293–301, 2013.
1389. Smith TL, Hutchins PM. Central hemodynamics in the developmental stage of spontaneous hypertension in the unanesthetized rat. Hypertension 1: 508 –517, 1979.
1368. Simon AM, McWhorter AR. Decreased intercellular dye-transfer and downregulation of non-ablated connexins in aortic endothelium deficient in connexin37 or
connexin40. J Cell Sci 116: 2223–2236, 2003.
1390. Soboloff J, Rothberg BS, Madesh M, Gill DL. STIM proteins: dynamic calcium signal
transducers. Nat Rev Mol Cell Biol 13: 549 –565, 2012.
1369. Simon AM, McWhorter AR. Role of connexin37 and connexin40 in vascular development. Cell Commun Adhes 10: 379 –385, 2003.
1391. Somers MJ, Mavromatis K, Galis ZS, Harrison DG. Vascular superoxide production
and vasomotor function in hypertension induced by deoxycorticosterone acetatesalt. Circulation 101: 1722–1728, 2000.
1370. Simonet S, Isabelle M, Bousquenaud M, Clavreul N, Feletou M, Vayssettes-Courchay
C, Verbeuren TJ. KCa 3.1 channels maintain endothelium-dependent vasodilatation
in isolated perfused kidneys of spontaneously hypertensive rats after chronic inhibition of NOS. Br J Pharmacol 167: 854 – 867, 2012.
1392. Somlyo AP, Somlyo AV. Ca2⫹ sensitivity of smooth muscle and nonmuscle myosin II:
modulated by G proteins, kinases, and myosin phosphatase. Physiol Rev 83: 1325–
1358, 2003.
1371. Simons JL, Provoost AP, Anderson S, Troy JL, Rennke HG, Sandstrom DJ, Brenner
BM. Pathogenesis of glomerular injury in the fawn-hooded rat: early glomerular
capillary hypertension predicts glomerular sclerosis. J Am Soc Nephrol 3: 1775–1782,
1993.
1372. Singh P, Blantz RC, Rosenberger C, Gabbai FB, Schoeb TR, Thomson SC. Aberrant
tubuloglomerular feedback and HIF-1 alpha confer resistance to ischemia after subtotal nephrectomy. J Am Soc Nephrol 23: 483– 493, 2012.
1373. Singh P, Deng A, Blantz RC, Thomson SC. Unexpected effect of angiotensin AT1
receptor blockade on tubuloglomerular feedback in early subtotal nephrectomy. Am
J Physiol Renal Physiol 296: F1158 –F1165, 2009.
1374. Sipos A, Vargas S, Peti-Peterdi J. Direct demonstration of tubular fluid flow sensing
by macula densa cells. Am J Physiol Renal Physiol 299: F1087–F1093, 2010.
1375. Siu KL, Sung B, Cupples WA, Moore LC, Chon KH. Detection of low-frequency
oscillations in renal blood flow. Am J Physiol Renal Physiol 297: F155–F162,
2009.
1376. Skarlatos S, Metting PJ, Britton SL. Spontaneous pressure-flow patterns in the kidney
of conscious rats. Am J Physiol Heart Circ Physiol 265: H2151–H2159, 1993.
1377. Skott O, Baumbach L. Effects of adenosine on renin release from isolated rat glomeruli and kidney slices. Pflügers Arch 404: 232–237, 1985.
1378. Skott O, Briggs JP. Direct demonstration of macula densa-mediated renin secretion.
Science 237: 1618 –1620, 1987.
1379. Skov K, Fenger-Gron J, Mulvany MJ. Effects of an angiotensin-converting enzyme
inhibitor, a calcium antagonist, and an endothelin receptor antagonist on renal afferent arteriolar structure. Hypertension 28: 464 – 471, 1996.
1380. Skov K, Mulvany MJ. Structure of renal afferent arterioles in the pathogenesis of
hypertension. Acta Physiol Scand 181: 397– 405, 2004.
1381. Skov K, Mulvany MJ, Korsgaard N. Morphology of renal afferent arterioles in spontaneously hypertensive rats. Hypertension 20: 821– 827, 1992.
1382. Skov K, Nyengaard JR, Korsgaard N, Mulvany MJ. Number and size of renal glomeruli
in spontaneously hypertensive rats. J Hypertens 12: 1373–1376, 1994.
1393. Song J, Lu Y, Lai EY, Wei J, Wang L, Chandrashekar K, Wang S, Shen C, Juncos LA, Liu
R. Oxidative status in the macula densa modulates tubuloglomerular feedback responsiveness in Ang II-induced hypertension. Acta Physiol Oxf. In press.
1394. Sonnenberg H, Honrath U, Wilson DR. Effects of increased perfusion pressure on
medullary collecting duct function. Can J Physiol Pharmacol 68: 402– 407, 1990.
1395. Sorensen CM, Giese I, Braunstein TH, Brasen JC, Salomonsson M, Holstein-Rathlou
NH. Role of connexin 40 in the autoregulatory response of the afferent arteriole. Am
J Phyisol Renal Physiol 303: F855–F863, 2012.
1396. Sorensen CM, Giese I, Braunstein TH, Holstein-Rathlou NH, Salomonsson M. Closure of multiple types of K⫹ channels is necessary to induce changes in renal vascular
resistance in vivo in rats. Pflügers Arch 462: 655– 667, 2011.
1397. Sorensen CM, Leyssac PP, Skott O, Holstein-Rathlou NH. Role of the renin-angiotensin system in regulation and autoregulation of renal blood flow. Am J Physiol Regul
Integr Comp Physiol 279: R1017–R1024, 2000.
1398. Sorensen CM, Leyssac PP, Skott O, Holstein-Rathlou NH. NO mediates downregulation of RBF after a prolonged reduction of renal perfusion pressure in SHR. Am J
Physiol Regul Integr Comp Physiol 285: R329 –R338, 2003.
1399. Sorensen CM, Salomonsson M, Braunstein TH, Nielsen MS, Holstein-Rathlou NH.
Connexin mimetic peptides fail to inhibit vascular conducted calcium responses in
renal arterioles. Am J Physiol Regul Integr Comp Physiol 295: R840 –R847, 2008.
1400. Sosnovtseva OV, Pavlov AN, Mosekilde E, Holstein-Rathlou NH, Marsh DJ. Doublewavelet approach to study frequency and amplitude modulation in renal autoregulation. Phys Rev E Stat Nonlin Soft Matter Phys 70: 031915, 2004.
1401. Sosnovtseva OV, Pavlov AN, Mosekilde E, Yip KP, Holstein-Rathlou NH, Marsh DJ.
Synchronization among mechanisms of renal autoregulation is reduced in hypertensive rats. Am J Physiol Renal Physiol 293: F1545–F1555, 2007.
1402. Sosnovtseva OV, Pavlov AN, Pavlova ON, Mosekilde E, Holstein-Rathlou NH. The
effect of L-NAME on intra- and inter-nephron synchronization. Eur J Pharm Sci 36:
39 –50, 2009.
1403. Spassova MA, Hewavitharana T, Xu W, Soboloff J, Gill DL. A common mechanism
underlies stretch activation and receptor activation of TRPC6 channels. Proc Natl
Acad Sci USA 103: 16586 –16591, 2006.
1383. Slish DF, Welsh DG, Brayden JE. Diacylglycerol and protein kinase C activate cation
channels involved in myogenic tone. Am J Physiol Heart Circ Physiol 283: H2196 –
H2201, 2002.
1404. Sporkova A, Kopkan L, Varcabova S, Huskova Z, Hwang SH, Hammock BD, Imig JD,
Kramer HJ, Cervenka L. Role of cytochrome P-450 metabolites in the regulation of
renal function and blood pressure in 2-kidney 1-clip hypertensive rats. Am J Physiol
Regul Integr Comp Physiol 300: R1468 –R1475, 2011.
1384. Smeda JS, Daneshtalab N. The effects of poststroke captopril and losartan treatment
on cerebral blood flow autoregulation in SHRsp with hemorrhagic stroke. J Cereb
Blood Flow Metab 31: 476 – 485, 2011.
1405. St Lezin E, Liu W, Wang JM, Yang Y, Qi N, Kren V, Zidek V, Kurtz TW, Pravenec M.
Genetic analysis of rat chromosome 1 and the Sa gene in spontaneous hypertension.
Hypertension 35: 225–230, 2000.
1385. Smeda JS, Lee RM, Forrest JB. Structural and reactivity alterations of the renal
vasculature of spontaneously hypertensive rats prior to and during established hypertension. Circ Res 63: 518 –533, 1988.
1406. Stec DE, Trolliet MR, Krieger JE, Jacob HJ, Roman RJ. Renal cytochrome P4504A
activity and salt sensitivity in spontaneously hypertensive rats. Hypertension 27:
1329 –1336, 1996.
1386. Smeda JS, Payne GW. Alterations in autoregulatory and myogenic function in the
cerebrovasculature of Dahl salt-sensitive rats. Stroke 34: 1484 –1490, 2003.
1407. Steendahl J, Holstein-Rathlou NH, Sorensen CM, Salomonsson M. Effects of chloride
channel blockers on rat renal vascular responses to angiotensin II and norepinephrine. Am J Physiol Renal Physiol 286: F323–F330, 2004.
1387. Smeda JS, van Vliet BN, King SR. Stroke-prone spontaneously hypertensive rats lose
their ability to auto-regulate cerebral blood flow prior to stroke. J Hypertens 17:
1697–1705, 1999.
504
1408. Stefanovska A. Coupled oscillators. Complex but not complicated cardiovascular
and brain interactions. IEEE Eng Med Biol Mag 26: 25–29, 2007.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
1409. Stegbauer J, Vonend O, Oberhauser V, Sellin L, Rump LC. Angiotensin II receptor
modulation of renal vascular resistance and neurotransmission in young and adult
spontaneously hypertensive rats. Kidney Blood Press Res 28: 20 –26, 2005.
1410. Steiner RW, Tucker BJ, Leslie MS, Gushwa LC, Gifford J, Wilson CB, Blantz RC.
Glomerular hemodynamics in moderate Goldblatt hypertension in the rat. Hypertension 4: 51–57, 1982.
1411. Steinhausen M, Baehr M, Dussel R. Vasomotion and vasoconstriction induced by a
Ca-agonist in the split hydronephrotic kidney. Prog Appl Microvasc 14: 25–39, 1989.
1412. Steinhausen M, Ballantyne D, Fretschner M, Parekh N. Sex differences in autoregulation of juxtamedullary glomerular blood flow in hydronephrotic rats. Am J Physiol
Renal Fluid Electrolyte Physiol 258: F863–F869, 1990.
1413. Steinhausen M, Blum M, Fleming JT, Holz FG, Parekh N, Wiegman DL. Visualization
of renal autoregulation in the split hydronephrotic kidney of rats. Kidney Int 35:
1151–1160, 1989.
1414. Steinhausen M, Endlich K, Nobiling R, Parekh N, Schutt F. Electrically induced vasomotor responses and their propagation in rat renal vessels in vivo. J Physiol 505:
493–501, 1997.
1415. Steinhausen M, Snoei H, Parekh N, Baker R, Johnson PC. Hydronephrosis: a new
method to visualize vas afferens, efferens, and glomerular network. Kidney Int 23:
794 – 806, 1983.
1416. Stepp DW, Merkus D, Nishikawa Y, Chilian WM. Nitric oxide limits coronary vasoconstriction by a shear stress-dependent mechanism. Am J Physiol Heart Circ Physiol
281: H796 –H803, 2001.
1430. Sward K, Albinsson S, Rippe C. Arterial dysfunction but maintained systemic blood
pressure in cavin-1-deficient mice. PLoS One 9: e92428, 2014.
1431. Syntichaki P, Tavernarakis N. Genetic models of mechanotransduction: the nematode Caenorhabditis elegans. Physiol Rev 84: 1097–1153, 2004.
1432. Szentivanyi M Jr, Maeda CY, Cowley AW Jr. Local renal medullary L-NAME infusion
enhances the effect of long-term angiotensin II treatment. Hypertension 33: 440 –
445, 1999.
1433. Takahashi N, Mori Y. TRP channels as sensors and signal integrators of redox status
changes. Front Pharmacol 2: 1–11, 2011.
1434. Takata Y, Koga T, Kobayashi K, Takishita S, Fujishima M. Decrease in endothelium
dependent hypotension in spontaneously hypertensive rats. Jpn Circ J 54: 183–191,
1990.
1435. Takenaka T, Epstein M, Forster H, Landry DW, Iijima K, Goligorsky MS. Attenuation
of endothelin effects by a chloride channel inhibitor, indanyloxyacetic acid. Am J
Physiol Renal Fluid Electrolyte Physiol 262: F799 –F806, 1992.
1436. Takenaka T, Forster H, De Micheli AG, Epstein M. Impaired myogenic responsiveness of renal microvessels in Dahl salt-sensitive rats. Circ Res 71: 471– 480, 1992.
1437. Takenaka T, Harrison-Bernard LM, Inscho EW, Carmines PK, Navar LG. Autoregulation of afferent arteriolar blood flow in juxtamedullary nephrons. Am J Physiol
Renal Fluid Electrolyte Physiol 267: F879 –F887, 1994.
1438. Takenaka T, Inoue T, Kanno Y, Okada H, Hill CE, Suzuki H. Connexins 37 and 40
transduce purinergic signals mediating renal autoregulation. Am J Physiol Regul Integr
Comp Physiol 294: R1–R11, 2008.
1417. Stern MD, Bowen PD, Parma R, Osgood RW, Bowman RL, Stein JH. Measurement
of renal cortical and medullary blood flow by laser-Doppler spectroscopy in the rat.
Am J Physiol Renal Fluid Electrolyte Physiol 236: F80 –F87, 1979.
1439. Takenaka T, Inoue T, Kanno Y, Okada H, Meaney KR, Hill CE, Suzuki H. Expression
and role of connexins in the rat renal vasculature. Kidney Int 73: 415– 422, 2008.
1418. Sterzel RB, Luft FC, Gao Y, Schnermann J, Briggs JP, Ganten D, Waldherr R, Schnabel
E, Kriz W. Renal disease and the development of hypertension in salt-sensitive Dahl
rats. Kidney Int 33: 1119 –1129, 1988.
1440. Takenaka T, Inoue T, Ohno Y, Miyazaki T, Nishiyama A, Ishii N, Suzuki H. Elucidating
mechanisms underlying altered renal autoregulation in diabetes. Am J Physiol Regul
Integr Comp Physiol 303: R495–R504, 2012.
1419. Stoos BA, Garvin JL. Actions of nitric oxide on renal epithelial transport. Clin Exp
Pharmacol Physiol 24: 591–594, 1997.
1441. Takenaka T, Inoue T, Okada H, Ohno Y, Miyazaki T, Chaston DJ, Hill CE, Suzuki H.
Altered gap junctional communication and renal haemodynamics in Zucker fatty rat
model of type 2 diabetes. Diabetologia 54: 2192–2201, 2011.
1420. Stoos BA, Metting PJ, Britton SL. Autoregulatory capacity in renal and mesenteric
vasculatures of mineralocorticoid hypertensive dogs. Am J Physiol Heart Circ Physiol
261: H1205–H1213, 1991.
1421. Storch U, Mederos y Schnitzler M, Gudermann T. G protein-mediated stretch reception. Am J Physiol Heart Circ Physiol 302: H1241–H1249, 2012.
1422. Stowe N, Schnermann J, Hermle M. Feedback regulation of nephron filtration rate
during pharmacologic interference with the renin-angiotensin and adrenergic systems in rats. Kidney Int 15: 473– 486, 1979.
1423. Strick DM, Fiksen-Olsen MJ, Lockhart JC, Roman RJ, Romero JC. Direct measurement of renal medullary blood flow in the dog. Am J Physiol Regul Integr Comp Physiol
267: R253–R259, 1994.
1424. Sun D, Cuevas AJ, Goltlinger K, Hwang SH, Hammock BD, Laniado SM, Huang A.
Soluble epoxide hydrolase-dependent regulation of myogenic response and blood
pressure. Am J Physiol Heart Circ Physiol 306: H1146 –H1153, 2014.
1425. Sun D, Huang A, Yan EH, Wu Z, Yan C, Kaminski PM, Oury TD, Wolin MS, Kaley G.
Reduced release of nitric oxide to shear stress in mesenteric arteries of aged rats. Am
J Physiol Heart Circ Physiol 286: H2249 –H2256, 2004.
1426. Sun D, Samuelson LC, Yang T, Huang Y, Paliege A, Saunders T, Briggs J, Schnermann
J. Mediation of tubuloglomerular feedback by adenosine: evidence from mice lacking
adenosin 1 receptors. Proc Natl Acad Sci USA 98: 9983–9988, 2001.
1427. Sun L, McArdle S, Chun M, Wolff DW, Pettinger WA. Cosegregation of the renin
gene with an increase in mean arterial blood pressure in the F2 rats of SHR-WKY
cross. Clin Exp Hypertens 15: 797– 805, 1993.
1428. Susic D, Frohlich ED, Kobori H, Shao W, Seth D, Navar LG. Salt-induced renal injury
in SHRs is mediated by AT1 receptor activation. J Hypertens 29: 716 –723, 2011.
1429. Suzaki Y, Yoshizumi M, Kagami S, Nishiyama A, Ozawa Y, Kyaw M, Izawa Y, Kanematsu Y, Tsuchiya K, Tamaki T. BMK1 is activated in glomeruli of diabetic rats and in
mesangial cells by high glucose conditions. Kidney Int 65: 1749 –1760, 2004.
1442. Takenaka T, Kanno Y, Kitamura Y, Hayashi K, Suzuki H, Saruta T. Role of chloride
channels in afferent arteriolar constriction. Kidney Int 50: 864 – 872, 1996.
1443. Takenaka T, Ohno Y, Hayashi K, Saruta T, Suzuki H. Governance of arteriolar
oscillation by ryanodine receptors. Am J Physiol Regul Integr Comp Physiol 285: R125–
R131, 2003.
1444. Takenaka T, Okada H, Kanno Y, Inoue T, Ryuzaki M, Nakamoto H, Kawachi H,
Shimizu F, Suzuki H. Exogenous 5’-nucleotidase improves glomerular autoregulation
in Thy-1 nephritic rats. Am J Physiol Renal Physiol 290: F844 –F853, 2006.
1445. Takenaka T, Suzuki H, Ikenaga H, Itaya Y, Yamakawa H, Sakamaki Y, Saruta T.
Effects of a calcium channel blocker, nicardipine, on pressure-natriuresis in Dahl
salt-sensitive rats. Clin Exp Hypertens 16: 77– 88, 1994.
1446. Takenaka T, Suzuki H, Okada H, Hayashi K, Kanno Y, Saruta T. Mechanosensitive
cation channels mediate afferent arteriolar myogenic constriction in the isolated rat
kidney. J Physiol 511: 245–253, 1998.
1447. Takenaka T, Suzuki H, Okada H, Hayashi K, Ozawa Y, Saruta T. Biophysical signals
underlying myogenic responses in rat interlobular artery. Hypertension 32: 1060 –
1065, 1998.
1448. Takenaka T, Suzuki H, Okada H, Inoue T, Kanno Y, Ozawa Y, Hayashi K, Saruta T.
Transient receptor potential channels in rat renal microcirculation: actions of angiotensin II. Kidney Int 62: 558 –565, 2002.
1449. Tang G, Wu L, Wang R. Interaction of hydrogen sulfide with ion channels. Clin Exp
Pharmacol Physiol 37: 753–763, 2009.
1450. Tang L, Parker M, Fei Q, Loutzenhiser R. Afferent arteriolar adenosine A2a receptors are coupled to KATP in in vitro perfused hydronephrotic rat kidney. Am J Physiol
Renal Physiol 277: F926 –F933, 1999.
1451. Tapia E, Franco M, Sanchez-Lozada LG, Soto V, Avila-Casado C, Santamaria J,
Quiroz Y, Rodriguez-Iturbe B, Herrera-Acosta J. Mycophenolate mofetil prevents
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
505
RENAL AUTOREGULATORY MECHANISMS
arteriolopathy and renal injury in subtotal ablation despite persistent hypertension.
Kidney Int 63: 994 –1002, 2003.
1452. Tapia E, Gabbai FB, Calleja C, Franco M, Cermeno JL, Bobadilla NA, Perez JM,
Alvarado JA, Herrera-Acosta J. Determinants of renal damage in rats with systemic
hypertension and partial renal ablation. Kidney Int 38: 642– 648, 1990.
1453. Taylor NE, Glocka P, Liang M, Cowley AW Jr. NADPH oxidase in the renal medulla
causes oxidative stress and contributes to salt-sensitive hypertension in Dahl S rats.
Hypertension 47: 692– 698, 2006.
1454. Terashvili M, Pratt PF, Gebremedhin D, Narayanan J, Harder DR. Reactive oxygen
species cerebral autoregulation in health and disease. Pediatr Clin North Am 53:
1029 –1037, 2006.
1455. Textor SC, Wilcox CS. Renal artery stenosis: a common, treatable cause of renal
failure? Annu Rev Med 52: 421– 442, 2001.
1456. Thai TL, Churchill GC, Arendshorst WJ. NAADP receptors mediate calcium signaling stimulated by endothelin-1 and norepinephrine in renal afferent arterioles. Am J
Physiol Renal Physiol 297: F510 –F516, 2009.
1457. Thai TL, Fellner SK, Arendshorst WJ. ADP-ribosyl cyclase and ryanodine receptor
activity contribute to basal renal vasomotor tone and agonist-induced renal vasoconstriction in vivo. Am J Physiol Renal Physiol 293: F1107–F1114, 2007.
1471. Tolbert EM, Weisstuch J, Feiner HD, Dworkin LD. Onset of glomerular hypertension with aging precedes injury in the spontaneously hypertensive rat. Am J Physiol
Renal Physiol 278: F839 –F846, 2000.
1472. Tolins JP, Shultz PJ, Raij L, Brown DM, Mauer SM. Abnormal renal hemodynamic
response to reduced renal perfusion pressure in diabetic rats: role of NO. Am J
Physiol Renal Fluid Electrolyte Physiol 265: F886 –F895, 1993.
1473. Toma I, Bansal E, Meer EJ, Kang JJ, Vargas SL, Peti-Peterdi J. Connexin 40 and
ATP-dependent intercellular calcium wave in renal glomerular endothelial cells. Am
J Physiol Regul Integr Comp Physiol 294: R1769 –R1776, 2008.
1474. Tomoda F, Bergstrom G, Evans RG, Anderson WP. Evidence for decreased structurally determined preglomerular resistance in the young spontaneously hypertensive rat after 4 weeks of renal denervation. J Hypertens 15: 1187–1195, 1997.
1475. Toney GM, Vallon V, Stockand JD. Intrinsic control of sodium excretion in the distal
nephron by inhibitory purinergic regulation of the epithelial Na⫹ channel. Curr Opin
Nephrol Hypertens 21: 52– 60, 2012.
1476. Tornel J, Madrid MI, Garcia-Salom M, Wirth KJ, Fenoy FJ. Role of kinins in the control
of renal papillary blood flow, pressure natriuresis, and arterial pressure. Circ Res 86:
589 –595, 2000.
1477. Touyz RM, Schiffrin EL. Reactive oxygen species in vascular biology: implications in
hypertension. Histochem Cell Biol 122: 339 –352, 2004.
1458. Thomsen AB, Kim S, Aalbaek F, Aalkjaer C, Boedtkjer E. Intracellular acidification
alters myogenic responsiveness and vasomotion of mouse middle cerebral arteries.
J Cereb Blood Flow Metab 34: 161–168, 2014.
1478. Touyz RM, Turgeon A, Schiffrin EL. Endothelin-A-receptor blockade improves renal
function and doubles the lifespan of stroke-prone spontaneously hypertensive rats. J
Cardiovasc Pharmacol 36: S300 –S304, 2000.
1459. Thomson S, Bao D, Deng A, Vallon V. Adenosine formed by 5’-nucleotidase mediates tubuloglomerular feedback. J Clin Invest 106: 289 –298, 2000.
1479. Traynor TR, Schnermann J. Renin-angiotensin system dependence of renal hemodynamics in mice. J Am Soc Nephrol 10 Suppl 11: S184 –S188, 1999.
1460. Thomson SC, Deng A. Cyclic GMP mediates influence of macula densa nitric oxide
over tubuloglomerular feedback. Kidney Blood Press Res 26: 10 –18, 2003.
1480. Traynor TR, Yang T, Huang YG, Arend L, Oliverio MI, Coffman T, Briggs JP, Schnermann J. Inhibition of adenosine-1 receptor-mediated preglomerular vasoconstriction in AT1A receptor-deficient mice. Am J Physiol Renal Physiol 275: F922–F927,
1998.
1461. Thomson SC, Deng A, Komine N, Hammes JS, Blantz RC, Gabbai FB. Early diabetes
as a model for testing the regulation of juxtaglomerular NOS I. Am J Physiol Renal
Physiol 287: F732–F738, 2004.
1462. Thorup C, Ollerstam A, Persson AE, Torffvit O. Increased tubuloglomerular feedback reactivity is associated with increased NO production in the streptozotocindiabetic rat. J Diabetes Complications 14: 46 –52, 2000.
1481. Traynor TR, Yang T, Huang YG, Krege JH, Briggs JP, Smithies O, Schnermann J.
Tubuloglomerular feedback in ACE-deficient mice. Am J Physiol Renal Physiol 276:
F751–F757, 1999.
1482. Trebak M. STIM/Orai signalling complexes in vascular smooth muscle. J Physiol 590:
4201– 4208, 2012.
1463. Thorup C, Persson AE. Inhibition of locally produced nitric oxide resets tubuloglomerular feedback mechanism. Am J Physiol Renal Fluid Electrolyte Physiol 267: F606 –
F611, 1994.
1483. Trippodo NC, Frohlich ED. Similarities of genetic (spontaneous) hypertension. Man
and rat. Circ Res 48: 309 –319, 1981.
1464. Thorup C, Persson AE. Impaired effect of nitric oxide synthesis inhibition on tubuloglomerular feedback in hypertensive rats. Am J Physiol Renal Fluid Electrolyte Physiol
271: F246 –F252, 1996.
1484. Troncoso Brindeiro CM, Fallet RW, Lane PH, Carmines PK. Potassium channel
contributions to afferent arteriolar tone in normal and diabetic rat kidney. Am J
Physiol Renal Physiol 295: F171–F178, 2008.
1465. Thurau K. Autoregulation of renal blood flow and glomerular filtration rate, including
data on tubular and peritubular capillary pressures and vessel wall tension. Circ Res
15 Suppl: 132–141, 1964.
1485. Troncoso Brindeiro CM, Lane PH, Carmines PK. Tempol prevents altered K⫹ channel regulation of afferent arteriolar tone in diabetic rat kidney. Hypertension 59:
657– 664, 2012.
1466. Thurau K, Schnermann J. Die Natriumkonzentration an den Macula densa Zellen als
regulierender Faktor fuer das Glomerulumfiltrat (Mikropunktionsversuche). Klin
Wochenschr 43: 410 – 413, 1965.
1486. Trottier G, Triggle CR, O’Neill SK, Loutzenhiser R. Cyclic GMP-dependent and
cyclic GMP-independent actions of nitric oxide on the renal afferent arteriole. Br J
Pharmacol 125: 563–569, 1998.
1467. Tobia AJ, Walsh GM, Tadepalli S, Lee JY. Unaltered distribution of cardiac output in
the conscious young spontaneously hypertensive rat: evidence for uniform elevation
of regional vascular resistances. Blood Vessels 11: 287–294, 1974.
1487. Tucker BJ, Anderson CM, Thies RS, Collins RC, Blantz RC. Glomerular hemodynamic alterations during acute hyperinsulinemia in normal and diabetic rats. Kidney
Int 42: 1160 –1168, 1992.
1468. Tobian L, Johnson MA, Lange J, Magraw S. Effect of varying perfusion pressures on
the output of sodium and renin and the vascular resistance in kidneys of rats with
“post-salt” hypertension and Kyoto spontaneous hypertension. Circ Res 36: 162–
170, 1975.
1469. Tobin AA, Joseph BK, Al-Kindi HN, Albarwani S, Madden JA, Nemetz LT, Rusch NJ,
Rhee SW. Loss of cerebrovascular Shaker-type K⫹ channels: a shared vasodilator
defect of genetic and renal hypertensive rats. Am J Physiol Heart Circ Physiol 297:
H293–H303, 2009.
1470. Tojo A, Onozato ML, Fukuda S, Asaba K, Kimura K, Fujita T. Nitric oxide generated
by nNOS in the macula densa regulates the afferent arteriolar diameter in rat kidney.
Med Electron Microsc 37: 236 –241, 2004.
506
1488. Tucker BJ, Blantz RC. An analysis of the determinants of nephron filtration rate. Am
J Physiol Renal Fluid Electrolyte Physiol 232: F477–F483, 1977.
1489. Tucker BJ, Blantz RC. Effect of furosemide administration on glomerular and tubular
dynamics in the rat. Kidney Int 26: 112–121, 1984.
1490. Tucker BJ, Mundy CA, Blantz RC. Can causality be determined from proximal tubular reabsorption and peritubular physical factors? Am J Physiol Renal Fluid Electrolyte
Physiol 250: F169 –F175, 1986.
1491. Turkstra E, Boer P, Braam B, Koomans HA. Increased availability of nitric oxide leads
to enhanced nitric oxide dependency of tubuloglomerular feedback in the contralateral kidney of rats with 2-kidney, 1-clip Goldblatt hypertension. Hypertension 34:
679 – 684, 1999.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
1492. Turkstra E, Braam B, Koomans HA. Losartan attenuates modest but not strong renal
vasoconstriction induced by nitric oxide inhibition. J Cardiovasc Pharmacol 32: 593–
600, 1998.
1512. Vallon V, Richter K, Huang DY, Rieg T, Schnermann J. Functional consequences at
the single-nephron level of the lack of adenosine A1 receptors and tubuloglomerular
feedback in mice. Pflügers Arch 448: 214 –221, 2004.
1493. Turkstra E, Braam B, Koomans HA. Nitric oxide release as an essential mitigating
step in tubuloglomerular feedback: observations during intrarenal nitric oxide clamp.
J Am Soc Nephrol 9: 1596 –1603, 1998.
1513. Vallon V, Schnermann J. Tubuloglomerular feedback. Methods Mol Med 86: 429 –
441, 2003.
1494. Turkstra E, Braam B, Koomans HA. Impaired renal blood flow autoregulation in
two-kidney, one-clip hypertensive rats is caused by enhanced activity of nitric oxide.
J Am Soc Nephrol 11: 847– 855, 2000.
1495. Turkstra E, Braam B, Koomans HA. Normal TGF responsiveness during chronic
treatment with angiotensin-converting enzyme inhibition: role of AT1 receptors.
Hypertension 36: 818 – 823, 2000.
1496. Turner SR, Macdonald JA. Novel contributions of the smoothelin-like 1 protein in
vascular smooth muscle contraction and its potential involvement in myogenic tone.
Microcirculation 21: 249 –258, 2014.
1497. Turoni CM, Reynoso HA, Maranon RO, Coviello A, Peral De BM. Structural changes
in the renal vasculature in streptozotocin-induced diabetic rats without hypertension. Nephron Physiol 99: 50 –57, 2005.
1498. Uchino K, Frohlich ED, Nishikimi T, Isshiki T, Kardon MB. Spontaneously hypertensive rats demonstrate increased renal vascular alpha 1-adrenergic receptor responsiveness. Am J Physiol Regul Integr Comp Physiol 260: R889 –R893, 1991.
1499. Uhrenholt TR, Schjerning J, Vanhoutte PM, Jensen BL, Skott O. Intercellular calcium
signaling and nitric oxide feedback during constriction of rabbit renal afferent arterioles. Am J Physiol Renal Physiol 292: F1124 –F1131, 2007.
1500. Umesh A, Thompson MA, Chini EN, Yip KP, Sham JS. Integrin ligands mobilize Ca2⫹
from ryanodine receptor-gated stores and lysosome-related acidic organelles in
pulmonary arterial smooth muscle cells. J Biol Chem 281: 34312–34323, 2006.
1501. Ungvari Z, Csiszar A, Huang A, Kaminski PM, Wolin MS, Koller A. High pressure
induces superoxide production in isolated arteries via protein kinase C-dependent
activation of NAD(P)H oxidase. Circulation 108: 1253–1258, 2003.
1502. Ungvari Z, Koller A. Endothelin and prostaglandin H2/thromboxane A2 enhance
myogenic constriction in hypertension by increasing Ca2⫹ sensitivity of arteriolar
smooth muscle. Hypertension 36: 856 – 861, 2000.
1503. Ungvari Z, Koller A. Selected contribution: NO released to flow reduces myogenic
tone of skeletal muscle arterioles by decreasing smooth muscle Ca2⫹ sensitivity. J
Appl Physiol 91: 522–527, 2001.
1504. Ungvari Z, Pacher P, Kecskemeti V, Papp G, Szollar L, Koller A. Increased myogenic
tone in skeletal muscle arterioles of diabetic rats. Possible role of increased activity of
smooth muscle Ca2⫹ channels and protein kinase C. Cardiovasc Res 43: 1018 –1028,
1999.
1505. Unlap MT, Bates E, Williams C, Komlosi P, Williams I, Kovacs G, Siroky B, Bell PD.
Na⫹/Ca2⫹ exchanger: target for oxidative stress in salt-sensitive hypertension. Hypertension 42: 363–368, 2003.
1506. Uriu K, Kaizu K, Qie YL, Kai K, Ito A, Ikeda M, Hashimoto O, Sun XF, Morita E, Eto
S. Renal hemodynamics in Otsuka Long-Evans Tokushima fatty rat, a model rat of
spontaneous non-insulin-dependent diabetes mellitus with obesity. J Lab Clin Med
134: 483– 491, 1999.
1507. Uyehara CF, Gellai M. Impairment of renal function precedes establishment of hypertension in spontaneously hypertensive rats. Am J Physiol Regul Integr Comp Physiol
265: R943–R950, 1993.
1508. Vallon V. Tubuloglomerular feedback in the kidney: insights from gene-targeted
mice. Pflügers Arch 445: 470 – 476, 2003.
1509. Vallon V, Blantz RC, Thomson S. Homeostatic efficiency of tubuloglomerular feedback is reduced in established diabetes mellitus in rats. Am J Physiol Renal Fluid
Electrolyte Physiol 269: F876 –F883, 1995.
1510. Vallon V, Muhlbauer B, Osswald H. Adenosine and kidney function. Physiol Rev 86:
901–940, 2006.
1511. Vallon V, Osswald H. Adenosine receptors and the kidney. Handb Exp Pharmacol
443– 470, 2009.
1514. Vallon V, Schroth J, Satriano J, Blantz RC, Thomson SC, Rieg T. Adenosine A(1)
receptors determine glomerular hyperfiltration and the salt paradox in early streptozotocin diabetes mellitus. Nephron Physiol 111: 30 –38, 2009.
1515. Vallon V, Traynor TR, Barajas L, Huang YG, Briggs JP, Schnermann J. Feedback
control of glomerular vascular tone in neuronal nitric oxide synthase knockout mice.
Am J Soc Nephrol 12: 1599 –1606, 2001.
1516. Van der Mark J, Kline RL. Altered pressure natriuresis in chronic angiotensin II
hypertension in rats. Am J Physiol Regul Integr Comp Physiol 266: R739 –R748, 1994.
1517. Van Dijk PC, Jager KJ, Stengel B, Gronhagen-Riska C, Feest TG, Briggs JD. Renal
replacement therapy for diabetic end-stage renal disease: data from 10 registries in
Europe (1991–2000). Kidney Int 67: 1489 –1499, 2005.
1518. Van Dijk SJ, Specht PA, Lazar J, Jacob HJ, Provoost AP. Renal damage susceptibility
and autoregulation in RF-1 and RF-5 congenic rats. Nephron Exp Nephrol 101: e59 –
e66, 2005.
1519. Van Dijk SJ, Specht PA, Lazar J, Jacob HJ, Provoost AP. Absence of an interaction
between the Rf-1 and Rf-5 QTLs influencing susceptibility to renal damage in rats.
Nephron Exp Nephrol 104: e96 – e102, 2006.
1520. Van Dijk SJ, Specht PA, Lazar J, Jacob HJ, Provoost AP. Synergistic QTL interactions
between Rf-1 and Rf-3 increase renal damage susceptibility in double congenic rats.
Kidney Int 69: 1369 –1376, 2006.
1521. Van Dijk SJ, Specht PA, Lutz MM, Lazar J, Jacob HJ, Provoost AP. Interaction between Rf-1 and Rf-4 quantitative trait loci increases susceptibility to renal damage in
double congenic rats. Kidney Int 68: 2462–2472, 2005.
1522. Van Dokkum RP, Sun CW, Provoost AP, Jacob HJ, Roman RJ. Altered renal hemodynamics and impaired myogenic responses in the fawn-hooded rat. Am J Physiol
Regul Integr Comp Physiol 276: R855–R863, 1999.
1523. Van Rodijnen WF, van Lambalgen TA, Tangelder G, Van Dokkum RP, Provoost AP,
Ter Wee PM. Reduced reactivity of renal microvessels to pressure and angiotensin II
in Fawn-Hooded rats. Hypertension 39: 111–115, 2002.
1524. Van Rodijnen WF, van Lambalgen TA, van Wijhe MH, Tangelder GJ, Ter Wee PM.
Renal microvascular actions of angiotensin II fragments. Am J Physiol Renal Physiol 283:
F86 –F92, 2002.
1525. VanBavel E, Giezeman MJ, Mooij T, Spaan JA. Influence of pressure alterations on
tone and vasomotion of isolated mesenteric small arteries of the rat. J Physiol 436:
371–383, 1991.
1526. Vaneckova I. Function of the isolated perfused kidney in young and adult spontaneously hypertensive and dahl salt-sensitive rats. Kidney Blood Press Res 25: 315–321,
2002.
1527. Vavrinec P, Henning RH, Goris M, Landheer SW, Buikema H, Van Dokkum RP. Renal
myogenic constriction protects the kidney from age-related hypertensive renal damage in the Fawn-Hooded rat. J Hypertens 31: 1637–1645, 2013.
1528. Vavrinec P, Van Dokkum RP, Goris M, Buikema H, Henning RH. Losartan protects
mesenteric arteries from ROS-associated decrease in myogenic constriction following 5/6 nephrectomy. J Renin Angiotensin Aldosterone Syst 12: 184 –199, 2011.
1529. Vaziri ND. Roles of oxidative stress and antioxidant therapy in chronic kidney disease
and hypertension. Curr Opin Nephrol Hypertens 13: 93–99, 2004.
1530. Veerareddy S, Cooke CL, Baker PN, Davidge ST. Gender differences in myogenic
tone in superoxide dismutase knockout mouse: animal model of oxidative stress. Am
J Physiol Heart Circ Physiol 287: H40 –H45, 2004.
1531. Venuto RC, O’Dorisio T, Ferris TF, Stein JH. Prostaglandins and renal function. II.
The effect of prostaglandin inhibition on autoregulation of blood flow in the intact
kidney of the dog. Prostaglandins 9: 817– 828, 1975.
1532. Verbeuren TJ, Vallez MO, Lavielle G, Bouskela E. Activation of thromboxane receptors and the induction of vasomotion in the hamster cheek pouch microcirculation.
Br J Pharmacol 122: 859 – 866, 1997.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
507
RENAL AUTOREGULATORY MECHANISMS
1533. Vermeij A, Meel-van den Abeelen AS, Kessels RP, van Beek AH, Claassen JA. Verylow-frequency oscillations of cerebral hemodynamics and blood pressure are affected by aging and cognitive load. Neuroimage 85: 608 – 615, 2014.
1534. Verseput GH, Braam B, Provoost AP, Koomans HA. Tubuloglomerular feedback
and prolonged ACE-inhibitor treatment in the hypertensive fawn-hooded rat. Nephrol Dial Transplant 13: 893– 899, 1998.
1535. Verseput GH, Provoost AP, Braam BB, Weening JJ, Koomans HA. Angiotensinconverting enzyme inhibition in the prevention and treatment of chronic renal damage in the hypertensive fawn-hooded rat. J Am Soc Nephrol 8: 249 –259, 1997.
1555. Wang H, D’Ambrosio MA, Garvin JL, Ren Y, Carretero OA. Connecting tubule
glomerular feedback in hypertension. Hypertension 62: 738 –745, 2013.
1556. Wang H, Garvin JL, Carretero OA. Angiotensin II enhances tubuloglomerular feedback via luminal AT(1) receptors on the macula densa. Kidney Int 60: 1851–1857,
2001.
1557. Wang H, Garvin JL, D’Ambrosio MA, Falck JR, Leung P, Liu R, Ren Y, Carretero OA.
Heme oxygenase metabolites inhibit tubuloglomerular feedback in vivo. Am J Physiol
Heart Circ Physiol 300: H1320 –H1326, 2011.
1536. Vetri F, Menicucci D, Lapi D, Gemignani A, Colantuoni A. Pial arteriolar vasomotion
changes during cortical activation in rats. Neuroimage 38: 25–33, 2007.
1558. Wang H, Garvin JL, D’Ambrosio MA, Ren Y, Carretero OA. Connecting tubule
glomerular feedback antagonizes tubuloglomerular feedback in vivo. Am J Physiol
Renal Physiol 299: F1374 –F1378, 2010.
1537. Vettoretti S, Ochodnicky P, Buikema H, Henning RH, Kluppel CA, de Zeeuw D, Van
Dokkum RP. Altered myogenic constriction and endothelium-derived hyperpolarizing factor-mediated relaxation in small mesenteric arteries of hypertensive subtotally nephrectomized rats. J Hypertens 24: 2215–2223, 2006.
1559. Wang H, Smeda JS, Lee RM. Prevention of stroke and preservation of the functions
of cerebral arteries by treatment with perindopril in stroke-prone spontaneously
hypertensive rats. Can J Physiol Pharmacol 76: 26 –34, 1998.
1538. Vio CP, Figueroa CD, Caorsi I. Anatomical relationship between kallikrein-containing tubules and the juxtaglomerular apparatus in the human kidney. Am J Hypertens 1:
269 –271, 1988.
1539. Vitzhum H, Abt I, Einhellig S, Kurtz A. Gene expression of prostanoid forming
enzymes along the rat nephron. Kidney Int 62: 1570 –1581, 2002.
1560. Wang T. Role of iNOS and eNOS in modulating proximal tubule transport and
acid-base balance. Am J Physiol Renal Physiol 283: F658 –F662, 2002.
1561. Wang T, Inglis FM, Kalb RG. Defective fluid and HCO3⫺ absorption in proximal tubule
of neuronal nitric oxide synthase-knockout mice. Am J Physiol Renal Physiol 279:
F518 –F524, 2000.
1540. Vitzthum H, Weiss B, Bachleitner W, Kramer BK, Kurtz A. Gene expression of
adenosine receptors along the nephron. Kidney Int 65: 1180 –1190, 2004.
1562. Wang T, Kobayashi Y, Nabika T, Takabatake T. Enhanced sympathetic control of
renal function in rats congenic for the hypertension-related region on chromosome
1. Clin Exp Pharmacol Physiol 32: 1055–1060, 2005.
1541. Volhard F. Clinical aspects of Bright s disease. In: The Kidney in Health and Disease,
edited by Berglund H, Medes G, and Huber GL. Philadelphia, PA: Lea & Febiger,
1935, p. 665– 688.
1563. Wang X, Ajikobi DO, Salevsky FC, Cupples WA. Impaired myogenic autoregulation
in kidneys of Brown Norway rats. Am J Physiol Renal Physiol 278: F962–F969, 2000.
1542. Vyas SJ, Jackson EK. Angiotensin II: enhanced renal responsiveness in young genetically hypertensive rats. J Pharmacol Exp Ther 273: 768 –777, 1995.
1564. Wang X, Aukland K, Bostad L, Iversen BM. Autoregulation of total and zonal glomerular filtration rate in spontaneously hypertensive rats with mesangiolysis. Kidney
Blood Press Res 20: 11–17, 1997.
1543. Wagenfeld L, von Domarus F, Weiss S, Klemm M, Richard G, Zeitz O. The effect of
reactive oxygen species on the myogenic tone of rat ophthalmic arteries with and
without endothelium. Graefes Arch Clin Exp Ophthalmol 251: 2339 –2344, 2013.
1544. Wagner AJ, Holstein-Rathlou NH, Marsh DJ. Internephron coupling by conducted
vasomotor responses in normotensive and spontaneously hypertensive rats. Am J
Physiol Renal Physiol 272: F372–F379, 1997.
1565. Wang X, Aukland K, Iversen BM. Autoregulation of total and zonal glomerular filtration rate in spontaneously hypertensive rats during antihypertensive therapy. J Cardiovasc Pharmacol 28: 833– 841, 1996.
1566. Wang X, Aukland K, Iversen BM. Acute effects of angiotensin II receptor antagonist
on autoregulation of zonal glomerular filtration rate in renovascular hypertensive
rats. Kidney Blood Press Res 20: 225–232, 1997.
1545. Wagner C. Function of connexins in the renal circulation. Kidney Int 73: 547–555, 2008.
1546. Wagner C, de Wit C, Kurtz L, Grunberger C, Kurtz A, Schweda F. Connexin 40 is
essential for the pressure control of renin synthesis and secretion. Circ Res 100:
556 –563, 2007.
1547. Walker M, Harrison-Bernard LM, Cook AK, Navar LG. Dynamic interaction between myogenic and TGF mechanisms in afferent arteriolar blood flow autoregulation. Am J Physiol Renal Physiol 279: F858 –F865, 2000.
1567. Wang X, Aukland K, Ofstad J, Iversen BM. Autoregulation of zonal glomerular filtration rate and renal blood flow in spontaneously hypertensive rats. Am J Physiol Renal
Fluid Electrolyte Physiol 269: F515–F521, 1995.
1568. Wang X, Breaks J, Loutzenhiser K, Loutzenhiser R. Effects of inhibition of the Na⫹/
K⫹/2Cl⫺ cotransporter on myogenic and angiotensin II responses of the rat afferent
arteriole. Am J Physiol Renal Physiol 292: F999 –F1006, 2007.
1548. Walsh MP, Cole WC. The role of actin filament dynamics in the myogenic response
of cerebral resistance arteries. J Cereb Blood Flow Metab 33: 1–12, 2013.
1569. Wang X, Cupples WA. Interaction between nitric oxide and renal myogenic autoregulation in normotensive and hypertensive rats. Can J Physiol Pharmacol 79: 238 –
245, 2001.
1549. Wang CT, Chin SY, Navar LG. Impairment of pressure-natriuresis and renal autoregulation in ANG II-infused hypertensive rats. Am J Physiol Renal Physiol 279: F319 –
F325, 2000.
1570. Wang X, Cupples WA. Brown Norway rats show impaired nNOS-mediated information transfer in renal autoregulation. Can J Physiol Pharmacol 87: 29 –36, 2009.
1550. Wang CT, Zou LX, Navar LG. Renal responses to AT1 blockade in angiotensin
II-induced hypertensive rats. J Am Soc Nephrol 8: 535–542, 1997.
1571. Wang X, Loutzenhiser R. Determinants of renal microvascular repsonse to ACh:
afferent and efferent arteriolar actions of EDHF. Am J Physiol Renal Physiol 282:
F124 –F132, 2002.
1551. Wang D, Borrego-Conde L, Falck JR, Sharma KK, Wilcox CS, Umans JG. Contributions of NO, EDHF and EETs to endothelium-dependent relaxation in rabbit renal
afferent arterioles. Kidney Int 63: 2187–2193, 2003.
1552. Wang D, Chen Y, Chabrashvili T, Aslam S, Borrego L, Umans J, Wilcox CS. Role of
oxidative stress in endothelial dysfunction and enhanced responses to Ang II of
afferent arterioles from rabbits infused with Ang II. J Am Soc Nephrol 14: 2783–2789,
2003.
1572. Wang X, Loutzenhiser RD, Cupples WA. Frequency modulation of renal myogenic
autoregulation by perfusion pressure. Am J Physiol Regul Integr Comp Physiol 293:
R1199 –R1204, 2007.
1573. Wang X, Salevsky FC, Cupples WA. Nitric oxide, atrial natriuretic factor, and dynamic renal autoregulation. Can J Physiol Pharmacol 77: 777–786, 1999.
1553. Wang H, Carretero OA, Garvin JL. Nitric oxide produced by THAL nitrc oxide
synthase inhibitis TGF. Hypertension 39: 662– 666, 2002.
1574. Wang X, Takeya K, Aaronson PI, Loutzenhiser K, Loutzenhiser R. Effects of
amiloride, benzamil, and alterations in extracellular Na⫹ on the rat afferent arteriole
and its myogenic response. Am J Physiol Renal Physiol 295: F272–F282, 2008.
1554. Wang H, D’Ambrosio MA, Garvin JL, Ren Y, Carretero OA. Connecting tubule
glomerular feedback mediates acute tubuloglomerular feedback resetting. Am J
Physiol Renal Physiol 302: F1300 –F1304, 2012.
1575. Wang YX, Brooks DP, Edwards RM. Attenuated glomerular cGMP production and
renal vasodilation in streptozotocin-induced diabetic rats. Am J Physiol Regul Integr
Comp Physiol 264: R952–R956, 1993.
508
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
1576. Watanabe J, Horiguchi S, Karibe A, Keitoku M, Takeuchi M, Satoh S, Takishima T,
Shirato K. Effects of ryanodine on development of myogenic response in rat small
skeletal muscle arteries. Cardiovasc Res 28: 480 – 484, 1994.
1598. White JV, Finco DR, Crowell WA, Brown SA, Hirakawa DA. Effect of dietary protein
on functional, morphologic, and histologic changes of the kidney during compensatory renal growth in dogs. Am J Vet Res 52: 1357–1365, 1991.
1577. Watanabe J, Karibe A, Horiguchi S, Keitoku M, Satoh S, Takishima T, Shirato K.
Modification of myogenic intrinsic tone and [Ca2⫹]i of rat isolated arterioles by
ryanodine and cyclopiazonic acid. Circ Res 73: 465– 472, 1993.
1599. White TW, Paul DL, Goodenough DA, Bruzzone R. Functional analysis of selective
interactions among rodent connexins. Mol Biol Cell 6: 459 – 470, 1995.
1578. Waugh WH. Circulatory autoregulation in the fully isolated kidney and in the humorally supported, isolated kidney. Circ Res 15 Suppl: 156 –169, 1964.
1600. Wickramasekera NT, Gebremedhin D, Carver KA, Vakeel P, Ramchandran R,
Schuett A, Harder DR. Role of dual-specificity protein phosphatase-5 in modulating
the myogenic response in rat cerebral arteries. J Appl Physiol 114: 252–261, 2013.
1579. Weichert W, Paliege A, Provoost AP, Bachmann S. Upregulation of juxtaglomerular
NOS1 and COX-2 precedes glomerulosclerosis in fawn-hooded hypertensive rats.
Am J Physiol Renal Physiol 280: F706 –F714, 2001.
1601. Wiederhielm CA, Bouskela E, Heald R, Black L. A method for varying arterial and
venous pressures in intact, unanesthetized mammals. Microvasc Res 18: 124 –128,
1979.
1580. Weihprecht H, Lorenz JN, Briggs JP, Schnermann J. Synergistic effects of angiotensin
and adenosine in the renal microvasculature. Am J Physiol Renal Fluid Electrolyte
Physiol 266: F227–F239, 1994.
1602. Wilcox CS. Role of macula densa NOS in tubuloglomerular feedback. Nephrol Hypertens 7: 443– 449, 1998.
1581. Welch WJ. Effects of isoprostane on tubuloglomerular feedback: roles of TP receptors, NOS, and salt intake. Am J Physiol Renal Physiol 288: F757–F762, 2005.
1582. Welch WJ, Baumgärtl H, Lübbers D, Wilcox CS. Nephron pO2 and renal oxygen
usage in the hypertensive rat kidney. Kidney Int 59: 230 –237, 2001.
1583. Welch WJ, Baumgärtl H, Lübbers D, Wilcox CS. Renal oxygenation defects in the
spontaneously hypertensive rat: role of AT1 receptors. Kidney Int 63: 202–208, 2003.
1584. Welch WJ, Chabrashvili T, Solis G, Chen Y, Gill P, Aslam S, Wang X, Ji H, Sandberg
K, Jose P, Wilcox CS. Role of extracellular superoxide dismutase in the mouse
angiotensin slow pressor response. Hypertension 48: 934 –941, 2006.
1585. Welch WJ, Mendonca M, Blau J, Karber A, Dennehy K, Lao Y, José P, Wilcox CS.
Antihypertensive response to prolonged tempol in the spontaneously hypertensive
rat. Kidney Int 68: 179 –187, 2005.
1586. Welch WJ, Schmidt HHHW, Murad F, Gross SS, Taylor G, Levi R, Wilcox CS. Nitric
oxide synthase in macula densa: localization and functional studies. In: The Biology of
Nitric Oxide. 1. Physiologic and Clinical Aspects, edited by Moncada S, Marletta MA,
Hibbs JB, and Higgs EA. London: Portland, 1992.
1603. Wilcox CS. Redox regulation of the afferent arteriole and tubuloglomerular feedback. Acta Physiol Scand 179: 217–223, 2003.
1604. Wilcox CS. Oxidative stress and nitric oxide deficiency in the kidney: a critical link to
hypertension? Am J Physiol Regul Integr Comp Physiol 289: R913–R935, 2005.
1605. Wilcox CS, Palm F, Welch WJ. Renal oxygenation and function of the rat kidney:
effects of inspired oxygen and preglomerular oxygen shunting. Adv Exp Med Biol 765:
329 –334, 2013.
1606. Wilcox CS, Pearlman A. Chemistry and antihypertensive effects of tempol and other
nitroxides. Pharm Rev 60: 418 – 469, 2008.
1607. Wilcox CS, Sterzel RB, Dunckel PT, Mohrmann M, Perfetto M. Renal interstitial
pressure and sodium excretion during hilar lymphatic ligation. Am J Physiol Renal Fluid
Electrolyte Physiol 247: F344 –F351, 1984.
1608. Wilcox CS, Welch WJ. Interaction between nitric oxide and oxygen radicals in regulation of tubuloglomerular feedback. Acta Physiol Scand 168: 119 –124, 2000.
1609. Wilcox CS, Welch WJ, Murad F, Gross SS, Taylor G, Levi R, Schmidt HH. Nitric oxide
synthase in macula densa regulates glomerular capillary pressure. Proc Natl Acad Sci
USA 89: 11993–11997, 1992.
1587. Welch WJ, Tojo A, Lee JU, Kang DG, Schnackenberg CG, Wilcox CS. Nitric oxide
synthase in the JGA of the SHR: expression and role in tubuloglomerular feedback.
Am J Physiol Renal Physiol 277: F130 –F138, 1999.
1610. Williams JM, Burke M, Lazar J, Jacob HJ, Roman RJ. Temporal characterization of the
development of renal injury in FHH rats and FHH.1BN congenic strains. Am J Physiol
Renal Physiol 300: F330 –F338, 2011.
1588. Welch WJ, Tojo A, Wilcox CS. Roles of NO and oxygen radicals in tubuloglomerular
feedback in SHR. Am J Physiol Renal Physiol 278: F769 –F776, 2000.
1611. Williams JM, Sarkis A, Hoagland KM, Fredrich K, Ryan RP, Moreno C, Lopez B, Lazar
J, Fenoy FJ, Sharma M, Garrett MR, Jacob HJ, Roman RJ. Transfer of the CYP4A
region of chromosome 5 from Lewis to Dahl S rats attenuates renal injury. Am J
Physiol Renal Physiol 295: F1764 –F1777, 2008.
1589. Welch WJ, Wilcox CS. Modulating role for thromboxane in the tubuloglomerular
feedback response in the rat. J Clin Invest 81: 1843–1849, 1988.
1590. Welch WJ, Wilcox CS. Feedback responses during sequential inhibition of angiotensin and thromboxane. Am J Physiol Renal Fluid Electrolyte Physiol 258: F457–F466,
1990.
1591. Welch WJ, Wilcox CS. Potentiation of tubuloglomerular feedback in the rat by
thromboxane mimetic. Role of macula densa. J Clin Invest 89: 1857–1865, 1992.
1592. Welch WJ, Wilcox CS. AT1 receptor antagonist combats oxidative stress and restores nitric oxide signaling in the SHR. Kidney Int 59: 1257–1263, 2001.
1593. Welch WJ, Wilcox CS, Folger WH. Mechanism of renal vasoconstriction with thromboxane mimetic: engagement of tubuloglomerular feedback. Adv Prostaglandin
Thromboxane Leukotriene Res 21B: 693– 696, 1991.
1594. Welsh DG, Morielli AD, Nelson MT, Brayden JE. Transient receptor potential channels regulate myogenic tone of resistance arteries. Circ Res 90: 248 –250, 2002.
1595. Wesselman JP, Schubert R, VanBavel ED, Nilsson H, Mulvany MJ. KCa-channel
blockade prevents sustained pressure-induced depolarization in rat mesenteric
small arteries. Am J Physiol Heart Circ Physiol 272: H2241–H2249, 1997.
1596. Westcott EB, Goodwin EL, Segal SS, Jackson WF. Function and expression of ryanodine receptors and inositol 1,4,5-trisphosphate receptors in smooth muscle cells of
murine feed arteries and arterioles. J Physiol 590: 1849 –1869, 2012.
1597. Westcott EB, Jackson WF. Heterogeneous function of ryanodine receptors, but not
IP3 receptors, in hamster cremaster muscle feed arteries and arterioles. Am J Physiol
Heart Circ Physiol 300: H1616 –H1630, 2011.
1612. Williams JM, Sarkis A, Lopez B, Ryan RP, Flasch AK, Roman RJ. Elevations in renal
interstitial hydrostatic pressure and 20-hydroxyeicosatetraenoic acid contribute to
pressure natriuresis. Hypertension 49: 687– 694, 2007.
1613. Williams RH, Thomas C, Bell PD, Navar LG. Autoregulation of nephron filtration
rate in the dog assessed by indicator-dilution technique. Am J Physiol Renal Fluid
Electrolyte Physiol 233: F282–F289, 1977.
1614. Williamson GA, Loutzenhiser RD, Wang X, Griffin KA, Bidani AK. Systolic and mean
blood pressures and afferent arteriolar myogenic response dynamics: a modeling
approach. Am J Physiol Regul Integr Comp Physiol 295: R1502–R1511, 2008.
1615. Wilson HM, Stewart KN. Glomerular epithelial and mesangial cell culture and characterization. Methods Mol Biol 806: 187–201, 2012.
1616. Wilson SM, Mason HS, Smith GD, Nicholson N, Johnston L, Janiak R, Hume JR.
Comparative capacitative calcium entry mechanisms in canine pulmonary and renal
arterial smooth muscle cells. J Physiol 543: 917–931, 2002.
1617. Winternitz SR, Katholi RE, Oparil S. Role of the renal sympathetic nerves in the
development and maintenance of hypertension in the spontaneously hypertensive
rat. J Clin Invest 66: 971–978, 1980.
1618. Wittmann U, Nafz B, Ehmke H, Kirchheim HR, Persson PB. Frequency domain of
renal autoregulation in the conscious dog. Am J Physiol Renal Fluid Electrolyte Physiol
269: F317–F322, 1995.
1619. Wolf G, Ritz E. Diabetic nephropathy in type 2 diabetes prevention and patient
management. J Am Soc Nephrol 14: 1396 –1405, 2003.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
509
RENAL AUTOREGULATORY MECHANISMS
1620. Wolfle SE, Navarro-Gonzalez MF, Grayson TH, Stricker C, Hill CE. Involvement of
nonselective cation channels in the depolarisation initiating vasomotion. Clin Exp
Pharmacol Physiol 37: 536 –543, 2010.
1641. Yu H, Harrap SB, Di NR. Cosegregation of spontaneously hypertensive rat renin
gene with elevated blood pressure in an F2 generation. J Hypertens 16: 1141–1147,
1998.
1621. Wolin MS, Gupte SA, Oeckler RA. Superoxide in the vascular system. J Vasc Res 39:
191–207, 2002.
1642. Yu JZ, Zhang DX, Zou AP, Campbell WB, Li PL. Nitric oxide inhibits Ca2⫹ mobilization through cADP-ribose signaling in coronary arterial smooth muscle cells. Am J
Physiol Heart Circ Physiol 279: H873–H881, 2000.
1622. Woods LL, DeYoung DR, Smith BE. Regulation of renal hemodynamics after protein
feeding: effects of loop diuretics. Am J Physiol Renal Fluid Electrolyte Physiol 261:
F815–F823, 1991.
1643. Yuan J, Pannabecker TL. Architecture of inner medullary descending and ascending
vasa recta: pathways for countercurrent exchange. Am J Physiol Renal Physiol 299:
F265–F272, 2010.
1623. Woods LL, Mizelle HL, Hall JE. Autoregulation of renal blood flow and glomerular
filtration rate in the pregnant rabbit. Am J Physiol Regul Integr Comp Physiol 252:
R69 –R72, 1987.
1644. Yuen PST, Garbers DL. Guanylyl cyclase-linked receptors. Annu Rev Neurosci 15:
193–225, 1992.
1624. Woods LL, Mizelle HL, Hall JE. Control of renal hemodynamics in hyperglycemia:
possible role of tubuloglomerular feedback. Am J Physiol Renal Fluid Electrolyte Physiol
252: F65–F73, 1987.
1645. Yuill KH, McNeish AJ, Kansui Y, Garland CJ, Dora KA. Nitric oxide suppresses
cerebral vasomotion by sGC-independent effects on ryanodine receptors and voltage-gated calcium channels. J Vasc Res 47: 93–107, 2010.
1625. Wray S, Burdyga T. Sarcoplasmic reticulum function in smooth muscle. Physiol Rev
90: 113–178, 2010.
1646. Zagorac D, Yamaura K, Zhang C, Roman RJ, Harder DR. The effect of superoxide
anion on autoregulation of cerebral blood flow. Stroke 36: 2589 –2594, 2005.
1626. Wronski T, Seeliger E, Persson PB, Forner C, Fichtner C, Scheller J, Flemming B. The
step response: a method to characterize mechanisms of renal blood flow autoregulation. Am J Physiol Renal Physiol 285: F758 –F764, 2003.
1647. Zambraski EJ, Ciccone CD. Effects of aortic constriction and renal denervation in
DOCA-hypertensive swine. Am J Physiol Renal Fluid Electrolyte Physiol 253: F1223–
F1231, 1987.
1627. Wu F, Park F, Cowley AW Jr, Mattson DL. Quantification of nitric oxide synthase
activity in microdissected segments of the rat kidney. Am J Physiol Renal Physiol 276:
F874 –F881, 1999.
1648. Zatz R, Meyer TW, Rennke HG, Brenner BM. Predominance of hemodynamic
rather than metabolic factors in the pathogenesis of diabetic glomerulopathy. Proc
Natl Acad Sci USA 82: 5963–5967, 1985.
1628. Wu X, Davis MJ. Characterization of stretch-activated cation current in coronary
smooth muscle cells. Am J Physiol Heart Circ Physiol 280: H1751–H1761, 2001.
1649. Zhan CD, Sindhu RK, Vaziri ND. Up-regulation of kidney NAD(P)H oxidase and
calcineurin in SHR: reversal by lifelong antioxidant supplementation. Kidney Int 65:
219 –227, 2004.
1629. Xi Q, Adebiyi A, Zhao G, Chapman KE, Waters CM, Hassid A, Jaggar JH. IP3 constricts cerebral arteries via IP3 receptor-mediated TRPC3 channel activation and
independently of sarcoplasmic reticulum Ca2⫹ release. Circ Res 102: 1118 –1126,
2008.
1630. Xiong Z, Sperelakis N. Regulation of L-type calcium channels of vascular smooth
muscle cells. J Mol Cell Cardiol 27: 75–91, 1995.
1631. Yang T, Hassan SA, Singh I, Smart A, Brosius FC, Holzman LB, Schnermann J, Briggs
JP. SA gene expression in the proximal tubule of normotensive and hypertensive rats.
Hypertension 27: 541–551, 1996.
1632. Yang Y, Sohma Y, Nourian Z, Ella SR, Li M, Stupica A, Korthuis RJ, Davis MJ, Braun
AP, Hill MA. Mechanisms underlying regional differences in the Ca2⫹ sensitivity of
BKCa current in arteriolar smooth muscle. J Physiol 591: 1277–1293, 2013.
1633. Yeh HI, Rothery S, Dupont E, Coppen SR, Severs NJ. Individual gap junction plaques
contain multiple connexins in arterial endothelium. Circ Res 83: 1248 –1263, 1998.
1634. Yeon DS, Kim JS, Ahn DS, Kwon SC, Kang BS, Morgan KG, Lee YH. Role of protein
kinase C- or RhoA-induced Ca2⫹ sensitization in stretch-induced myogenic tone.
Cardiovasc Res 53: 431– 438, 2002.
1635. Yip KP, Holstein-Rathlou NH, Marsh DJ. Chaos in blood flow control in genetic and
renovascular hypertensive rats. Am J Physiol Renal Fluid Electrolyte Physiol 261: F400 –
F408, 1991.
1636. Yip KP, Holstein-Rathlou NH, Marsh DJ. Mechanisms of temporal variation in singlenephron blood flow in rats. Am J Physiol Renal Fluid Electrolyte Physiol 264: F427–
F434, 1993.
1637. Yip KP, Marsh DJ. [Ca2⫹]i in rat afferent arteriole during constriction measured with
confocal fluorescence microscopy. Am J Physiol Renal Fluid Electrolyte Physiol 271:
F1004 –F1011, 1996.
1650. Zhang A, Carella A, Altura BT, Altura BM. Interactions of magnesium and chloride
ions on tone and contractility of vascular muscle. Eur J Pharmacol 203: 223–235,
1991.
1651. Zhang AY, Yi F, Teggatz EG, Zou AP, Li PL. Enhanced production and action of cyclic
ADP-ribose during oxidative stress in small bovine coronary arterial smooth muscle.
Microvasc Res 67: 159 –167, 2004.
1652. Zhang DX, Borbouse L, Gebremedhin D, Mendoza SA, Zinkevich NS, Li R, Gutterman DD. H2O2-induced dilation in human coronary arterioles: role of protein kinase
G dimerization and large-conductance Ca2⫹-activated K⫹ channel activation. Circ
Res 110: 471– 480, 2012.
1653. Zhang F, Kaide J, Wei Y, Jiang H, Yu C, Balazy M, Abraham NG, Wang W, Nasjletti
A. Carbon monoxide produced by isolated arterioles attenuates pressure-induced
vasoconstriction. Am J Physiol Heart Circ Physiol 281: H350 –H358, 2001.
1654. Zhang J, Lee MY, Cavalli M, Chen L, Berra-Romani R, Balke CW, Bianchi G, Ferrari
P, Hamlyn JM, Iwamoto T, Lingrel JB, Matteson DR, Wier WG, Blaustein MP. Sodium
pump alpha2 subunits control myogenic tone and blood pressure in mice. J Physiol
569: 243–256, 2005.
1655. Zhang J, Ren C, Chen L, Navedo MF, Antos LK, Kinsey SP, Iwamoto T, Philipson KD,
Kotlikoff MI, Santana LF, Wier WG, Matteson DR, Blaustein MP. Knockout of Na⫹/
Ca2⫹ exchanger in smooth muscle attenuates vasoconstriction and L-type Ca2⫹
channel current and lowers blood pressure. Am J Physiol Heart Circ Physiol 298:
H1472–H1483, 2010.
1656. Zhang Q, Cao C, Zhang Z, Wier WG, Edwards A, Pallone TL. Membrane current
oscillations in descending vasa recta pericytes. Am J Physiol Renal Physiol 294: F656 –
F666, 2008.
1638. Yip KP, Marsh DJ. An Arg-Gly-Asp peptide stimulates constriction in rat afferent
arteriole. Am J Physiol Renal Physiol 273: F768 –F776, 1997.
1657. Zhang R, Harding P, Garvin JL, Juncos R, Peterson E, Juncos LA, Liu R. Isoforms and
functions of NAD(P)H oxidase at the macula densa. Hypertension 53: 556 –563,
2009.
1639. Yoshida T, Inoue R, Morii T, Takahashi N, Yamamoto S, Hara Y, Tominaga M,
Shimizu S, Sato Y, Mori Y. Nitric oxide activates TRP channels by cysteine S-nitrosylation. Nat Chem Biol 2: 596 – 607, 2006.
1658. Zhang SL, Yu Y, Roos J, Kozak JA, Deerinck TJ, Ellisman MH, Stauderman KA,
Cahalan MD. STIM1 is a Ca2⫹ sensor that activates CRAC channels and migrates
from the Ca2⫹ store to the plasma membrane. Nature 437: 902–905, 2005.
1640. Young DK, Marsh DJ. Pulse wave propagation in rat renal tubules: implications for
GFR autoregulation. Am J Physiol Renal Fluid Electrolyte Physiol 240: F446 –F458,
1981.
1659. Zhang Z, Rhinehart K, Kwon W, Weinman E, Pallone TL. ANG II signaling in vasa
recta pericytes by PKC and reactive oxygen species. Am J Physiol Heart Circ Physiol
287: H773–H781, 2004.
510
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
CARLSTRÖM ET AL.
1660. Zhao D, Zhang J, Blaustein MP, Navar LG. Attenuated renal vascular responses to
acute angiotensin II infusion in smooth muscle-specific Na⫹/Ca2⫹ exchanger knockout mice. Am J Physiol Renal Physiol 301: F574 –F579, 2011.
1661. Zhao X, Cook AK, Field M, Edwards B, Zhang S, Zhang Z, Pollock JS, Imig JD, Inscho
EW. Impaired Ca2⫹ signaling attenuates P2X receptor-mediated vasoconstriction of
afferent arterioles in angiotensin II hypertension. Hypertension 46: 562–568, 2005.
1662. Zhao X, Yamamoto T, Newman JW, Kim IH, Watanabe T, Hammock BD, Stewart
J, Pollock JS, Pollock DM, Imig JD. Soluble epoxide hydrolase inhibition protects the
kidney from hypertension-induced damage. J Am Soc Nephrol 15: 1244 –1253, 2004.
1663. Zhong XZ, Harhun MI, Olesen SP, Ohya S, Moffatt JD, Cole WC, Greenwood IA.
Participation of KCNQ (Kv7) potassium channels in myogenic control of cerebral
arterial diameter. J Physiol 588: 3277–3293, 2010.
1664. Zimmermann PA, Knot HJ, Stevenson AS, Nelson MT. Increased myogenic tone and
diminished responsiveness to ATP-sensitive K⫹ channel openers in cerebral arteries
from diabetic rats. Circ Res 81: 996 –1004, 1997.
1665. Zou AP, Billington H, Su N, Cowley AW Jr. Expression and actions of heme oxygenase in the renal medulla of rats. Hypertension 35: 342–347, 2000.
1666. Zou AP, Drummond HA, Roman RJ. Role of 20-HETE in elevating loop chloride
reabsorption in Dahl SS/Jr rats. Hypertension 27: 631– 635, 1996.
1667. Zou AP, Fleming JT, Falck JR, Jacobs ER, Gebremedhin D, Harder DR, Roman RJ.
Stereospecific effects of epoxyeicosatrienoic acids on renal vascular tone and
K⫹-channel activity. Am J Physiol Renal Fluid Electrolyte Physiol 270: F822–F832,
1996.
1668. Zou AP, Imig JD, Kaldunski ML, Ortiz de Montellano PR, Sui Z, Roman RJ. Inhibition
of renal vascular 20-HETE production impairs autoregulation of renal blood flow. Am
J Physiol Renal Fluid Electrolyte Physiol 266: F275–F282, 1994.
1669. Zou AP, Imig JD, Ortiz de Montellano PR, Sui Z, Falck JR, Roman RJ. Effect of P-450
omega-hydroxylase metabolites of arachidonic acid on tubuloglomerular feedback.
Am J Phyisol 266: F934 –F941, 1994.
1670. Zou AP, Li N, Cowley AW Jr. Production and actions of superoxide in the renal
medulla. Hypertension 37: 547–553, 2001.
1671. Zou H, Ratz PH, Hill MA. Role of myosin phosphorylation and [Ca2⫹]i in myogenic
reactivity and arteriolar tone. Am J Physiol Heart Circ Physiol 269: H1590 –H1596,
1995.
1672. Zou R, Cupples WA, Yip KP, Holstein-Rathlou NH, Chon KH. Time-varying properties of renal autoregulatory mechanisms. IEEE Trans Biomed Eng 49: 1112–1120,
2002.
1673. Zou Y, Akazawa H, Qin Y, Sano M, Takano H, Minamino T, Makita N, Iwanaga K,
Zhu W, Kudoh S, Toko H, Tamura K, Kihara M, Nagai T, Fukamizu A, Umemura
S, Iiri T, Fujita T, Komuro I. Mechanical stress activates angiotensin II type 1
receptor without the involvement of angiotensin II. Nat Cell Biol 6: 499 –506,
2004.
Physiol Rev • VOL 95 • APRIL 2015 • www.prv.org
511
Download