ArticlePDF AvailableLiterature Review

Malignant hypertension: New aspects of an old clinical entity

Authors:

Abstract and Figures

Malignant hypertension, also known as accelerated-malignant hypertension or malignant phase hypertension, is the most severe form of hypertension; it is defined clinically as high blood pressure levels associated with bilateral retinal flame-shaped hemorrhages, exudates or cotton wool spots, with or without papilledema. Despite the vast range of antihypertensive agents, malignant hypertension continues to be an important clinical entity. Although its frequency is very low, the absolute number of new cases has not changed much over the past decades. While the role of the activation of the renin-angiotensin-aldosterone system and endothelial dysfunction in the pathogenesis of malignant hypertension has been well described, recent studies indicate that the immune system may play an important role in the development of malignant hypertension. Patients with malignant hypertension are characterized by pronounced target organ damage, including cardiac, structural and functional abnormalities. Malignant hypertension is frequently complicated by renal insufficiency and end-stage renal disease. The survival rates for malignant hypertension survival have considerably improved, with the introduction of antihypertensive therapy. However, in patients with malignant hypertension, renal insufficiency and end-stage renal disease still remain a significant cause of morbidity and mortality. In conclusion, malignant hypertension is not a 'vanishing disease', with a relatively stable number of cases per year, but it does have a significantly improved prognosis and survival rates due to earlier detection, stricter blood pressure control, lower BP targets, better choice of antihypertensive drugs and availability of hemodialysis and renal transplantation.
Content may be subject to copyright.
POLSKIE ARCHIWUM MEDYCYNY WEWNĘTRZNEJ
2016; 126 (1-2)
86
MHT is associated with failure in BP autoregu-
lation and develops when the mean arterial pres-
sure (MAP) reaches a critical level of 150 mmHg,
as reported in experimental animals. Fibrinoid ne-
crosis appears in the arterial walls, which may be
caused by vasoactive factor(s) or may be a non-
specific consequence of very high BP.1
In patients with MHT, very high systolic BP,
and in particular significantly elevated diastolic
BP, are surrogate markers of increased peripheral
vascular resistance associated with an increase in
MAP, an abnormal nonpulsatile component of BP.
4
Even though the introduction of modern mul-
tidrug antihypertensive therapy has resulted in
better BP control and markedly improved prog-
nosis, MHT is still a life-threating manifestation
of hypertension and represents a clinical enti-
ty characterized by high cardiovascular risk and
Introduction
Malignant hypertension (MHT),
also known as accelerated-malignant hyperten-
sion or malignant-phase hypertension, is the most
severe form of hypertension; it is defined clinically
as high blood pressure (BP) associated with bilat-
eral retinal flame-shaped hemorrhages, exudates,
or cotton wool spots, with or without papillede-
ma. In the 2010 revision of the Dutch guidelines
for the management of hypertensive crisis, van
den Born et al
1
replaced the term “malignant hy-
pertension” with “hypertensive crisis with reti-
nopathy”, which was followed by Kaplan and Vic-
tor in the 2015 edition of Kaplan’s Clinical Hyper-
tension. e authors concluded that the presence
of retinopathy may allow other target organs to
be included, making the description of this type
of emergency more accurate.1-3
Correspondence to:
Aleksander Prejbisz, MD, PhD,
Klinika Nadciśnienia Tętniczego,
Instytut Kardiologii, ul. Alpejska
42, 04-628 Warszawa, Poland,
phone: +48 22 343 43 39,
e-mail: a.prejbisz@ikard.pl
Received: November 12, 2015.
Accepted: November 16, 2015.
Published online:
December 10, 2015.
Conflict of interest: none declared.
Pol Arch Med Wewn. 2016;
126 (1-2): 86-93
Copyright by Medycyna Praktyczna,
Kraków 2016
KEY WORDS
malignant
hypertension,
pathogenesis,
prevalence, treatment
ABSTRACT
Malignant hypertension (MHT), also known as accelerated-malignant hypertension or malignant-phase
hypertension, is the most severe form of arterial hypertension. It is defined clinically as high blood pressure
(BP) levels associated with lesions of the retinal fundus (flame-shaped hemorrhages, exudates, or cotton
wool spots, with or without papilledema). Despite the availability of a vast range of antihypertensive
agents, MHT continues to be a significant clinical challenge. Although its prevalence is very low, the
absolute number of new cases has not changed over the past decades. While the role of the activation of
the renin–angiotensin–aldosterone system and endothelial dysfunction in the pathogenesis of MHT has
been well described, recent studies have indicated that the immune system may also play an important
role in the development of this condition. Patients with MHT are characterized by pronounced target organ
damage, including structural and functional cardiac abnormalities. MHT is frequently complicated by
renal insufficiency and end-stage renal disease. The survival rates for patients with MHT have improved
considerably with increased availability of antihypertensive treatment. However, renal insufficiency and
end-stage renal disease still remain a significant cause of morbidity and mortality in this patient group.
In conclusion, MHT is not a “vanishing disease” because there is a relatively stable number of new cases
per year. Nonetheless, prognosis and survival rates in these patients have improved significantly owing
to earlier detection, stricter BP control, lower BP targets, better choice of antihypertensive drugs, and
availability of hemodialysis and renal transplantation.
REVIEW ARTICLE
Malignant hypertension: new aspects of an old
clinical entity
AndrzejJanuszewicz
1
, TomaszGuzik
2
, AleksanderPrejbisz
1
,
TomaszMikołajczyk
2
, GrzegorzOsmenda
2
, WłodzimierzJanuszewicz
3
1 Department of Hypertension, Institute of Cardiology, Warsaw, Poland
2 Department of Internal and Agricultural Medicine, Jagiellonian University Medical College, Kraków, Poland
3 Warsaw, Poland
REVIEW ARTICLE Malignant hypertension: new aspects of an old clinical entity 87
admissions for MHT, identifying all hospitaliza-
tions between 2000 and 2011 during which a pri-
mary diagnosis of MHT was recorded.
7
e re-
sults clearly demonstrated a much higher rate of
increase in the number of hospitalizations with
a primary discharge diagnosis of MHT during
this time period. e rate of annual hospitaliza-
tions for MHT remained stable before 2007, but
then started to significantly increase at an esti-
mated rate of 2700 per year.
7
One explanation of
the results is that the reported data represented
a true change in the epidemiology of hyperten-
sive emergencies. Another cause of the change,
which is much more likely, is that this abrupt in-
crease resulted from a change in coding practic-
es in the United States.7
Pathophysiological mechanisms
Possible patho-
physiological mechanisms for the development
of MHT have been proposed, including rapid-
ly increasing BP, pressure diuresis and natriure-
sis, severe renal vasoconstriction, and ischemia
(
FIGURE 1
). In addition, activation of the renin–an-
giotensin–aldosterone system (
FIGURE 2
), microan-
giopathy, hemolytic anemia, and development of
retinopathy are observed in MHT. e vascular
lesions of MHT include myointimal proliferation
and fibrinoid necrosis.8-21
The immune system in malignant hypertension
While the role of endothelial dysfunction and kid-
ney damage in MHT has been addressed by nu-
merous other reviews, recent studies have indi-
cated that the immune system, and in particular
T cells, play a crucial role in the development of
elevated risk of developing end-stage renal dis-
ease (ESRD) in a long-term follow-up.1
Incidence of malignant hypertension
Despite the
vast range of antihypertensive agents and effec-
tive BP control, MHT remains an important clin-
ical entity. Although its frequency is very low, the
absolute number of new cases has not changed
much over the past decades, as documented by
recent studies.1
e study by Lane et al5 examined the chang-
ing demography and survival of 446 patients with
MHT attending the City Hospital in Birmingham,
United Kingdom, from 1964 to 2006, with a me-
dian follow-up of 103.8 months. In this largest
prospective analysis in the world literature, the
number of new cases of MHT has not changed
substantially over 40 years. e study supports
the concept that MHT is not a vanishing disease,
with a relatively stable number of 2 to 3 new cas-
es of MHT per 100 000 head of population per
year.5 e demography of the disease has not al-
tered during the follow-up, with no significant dif-
ferences in the mean age at diagnosis, but with
a slight predominance in men. It is of note that
during the follow-up, there was a significant in-
crease in the proportion of ethnic minorities, in-
cluding South-Asians and Afro-Caribbeans, and a
significant decrease in white Europeans.
5
Another
smaller study evaluating European cohorts dem-
onstrated a fairly stable incidence of MHT, with
2 to 3 new cases per 100 000 head of population
over the past 30 years.6
A recent retrospective cohort study investigat-
ed national trends in the United States in hospital
FIGURE 1 Postulated
pathophysiological
mechanisms of malignant
hypertension
excessive increase in
blood pressure
smoking
endothelial
dysfunction
key factors in the
pathogenesis of malignant
hypertension
oxidative stress
angiotensin II
endothelin 1
vasopressin
platelet activation
aldosterone
sympathetic nervous
system
inflammatory and
immune mechanisms
proinflammatory
cytokines
POLSKIE ARCHIWUM MEDYCYNY WEWNĘTRZNEJ
2016; 126 (1-2)
88
immunohistochemical analysis of biopsied re
-
nal tissue revealed a higher infiltration of T cells,
both CD4+ and CD8+, in patients with hyperten-
sive nephrosclerosis, compared with normoten-
sive controls. In patients with psoriasis treated
with mycophenolate mofetil targeting B and T
cells has shown to reduce BP. It is possible that
immunomodulating agents will emerge in the fu-
ture as a potential treatment option for patients
with MHT, especially in the context of concom-
itant immune disorders. However, at present,
there is no sufficient clinical evidence to support
such an approach.26,27
Clinical features
MHT may be accompanied by
various symptoms and complications, the most
characteristic being microangiopathic lesions or
renal dysfunction (FIGURE 3). Acute lesions in the
retinal fundus may include arteriolar spasm, reti-
nal edema, hemorrhages, exudates and papillede-
ma, and engorged retinal veins.1
In 28% of patients with MHT, van den Born et
al28 found thrombotic microangiopathy, charac-
terized by thromboses of small vessels, intravas-
cular hemolysis with fragmented red blood cells,
elevated lactic dehydrogenase, and consumption
of platelets. Less common clinical presentations
may include fibrinoid necrosis within abdominal
arteries producing major gastrointestinal tract in-
farction with an acute abdomen, necrotizing vas-
culitis as a feature of lupus, polyarteritis nodosa,
or Takayasu arteritis.29-33
Urine may contain protein and red cells. In a
few patients, acute oliguric renal failure may be
the presenting manifestation, and various fea-
tures of renal dysfunction (including protein-
uria) may be present. Approximately half of pa-
tients with MHT have hypokalemia, reflecting
experimental hypertension and vascular pathol-
ogy. In addition, our preliminary studies in hu-
mans have indicated that immune dysregulation
is particularly evident in hypertension.
22,23
Mice
lacking T and B lymphocytes (Rag1–/–) were re-
sistant to the development of severe hyperten-
sion in both angiotensin II and desoxycorticoste-
rone acetate-salt-induced hypertension models,
and this effect was greatly mediated by T-cell ac-
tivation.22-24 Subsequent studies have indicated
that a complex network of interactions between
T cells, dendritic cells, monocytes, and B cells
may be involved in hypertension. A mild form of
hypertension develops in the absence of T cells,
while severe hypertension is completely prevent-
ed. is could indicate a possible involvement of
an inflammatory and immune mechanism, par-
ticularly in MHT.22-24
Activated memory and effector T lympho-
cytes responding to the prohypertensive fac-
tors are able to infiltrate the adventitia and peri-
vascular fat, and possibly also the renal tissues
or even the central nervous system. In the peri-
vascular adipose tissue and in the kidney, these
cells release proinflammatory cytokines such
as interleukin (IL) 17, tumor necrosing factor α
(TNF-α), and interferon γ. us, the targeting
of TNF-α, IL-17, or chemotactic molecules such
as the RANTES/CCR5 axis might be considered,
and may become useful in the future manage-
ment of resistant hypertension.22,25
Evidence of T cell-derived inflammation in
hypertension is also slowly accumulating for
humans.26 ere is an increased fraction of im-
munosenescent, cytotoxic T cells in the circu-
lation of patients with hypertension compared
with normotensive subjects. Plasma TNF-α
and IL-6 levels are correlated with BP, and a
FIGURE 2 Urine
aldosterone excretion in
patients with essential
hypertension and
malignant hypertension.
Adapted from Laragh
et al.8
essential hypertension
3000
2500
2000
1500
1000
500
0
malignant hypertension
evaluated patients
aldosterone
µ
g/24 h
REVIEW ARTICLE Malignant hypertension: new aspects of an old clinical entity 89
patients suspected of fibromuscular dysplasia, or
in older patients with extensive atherosclerosis.
Idiopathic IgA nephropathy has also been report-
ed as a possible cause of MHT.1,34,35
Rarely, primary hyperaldosteronism may be
associated with MHT, as documented by a small
number of cases. A recent report described a
22-year-old patient with MHT, associated with
primary hyperaldosteronism and LVH mimick-
ing hypertrophic cardiomyopathy.
35,36
Primary al-
dosteronism is typically defined by an increased
plasma aldosterone level and suppressed plasma
renin activity through a negative feedback mech-
anism. Hence, primary aldosteronism and MHT
are at opposite ends of the renin spectrum, and
their coexistence is thought to be very rare.36,37
ere have also been sporadic reports of MHT
in patients with pheochromocytoma and renin-
-secreting tumors.38-40
It is of note that, in most cases, MHT is accom-
panied by various life-threatening symptoms,
signs, and associated complications. However, it
is not uncommon to see patients denying any pri
-
or symptoms in the end stages of the hyperten-
sive disease. In such cases, MHT is often unrec-
ognized until a later stage in the progress of the
disease, when clinical symptoms are becoming ap-
parent and patients are symptomatic.1
Impact of malignant hypertension on the heart
Afew
studies have confirmed that structural and func-
tional cardiac abnormalities are present in pa-
tients with MHT, indicating features of cardiac
damage and compensatory remodeling. However,
secondary aldosteronism from increased renin
secretion induced by intrarenal ischemia. Hypo-
natremia is common and may be severe.1
Multiple markers of inflammation, coagulation,
platelet activation, and fibrinolysis were found in
the blood of patients with various types of hyper-
tensive emergencies, compared with the levels ob-
served in normotensive controls.2,11
An electrocardiogram usually shows evidence
of left ventricular hypertrophy (LVH). Echocar-
diography may show impaired systolic and dia-
stolic function and delayed mitral valve opening.
Regression of these abnormalities usually occurs
after lowering of BP by antihypertensive therapy.
1
Identifiable causes of malignant hypertension
In pa-
tients with MHT, after an acute phase, an appro-
priate evaluation of identifiable causes of hyper-
tension should be performed as quickly as possi-
ble. It is preferable to obtain the necessary blood
and urine samples for required laboratory stud-
ies before instituting therapies that may mark-
edly complicate subsequent evaluation. Howev-
er, none of these procedures should delay effec-
tive therapy. MHT usually occurs in patients with
long-standing preexisting hypertension and may
be associated with a recent cessation of antihyper-
tensive therapy. However, any form of secondary
hypertension may progress to MHT.1-3
Renovascular hypertension is the most likely
secondary cause, and, in a large series of patients
with MHT, renal artery stenosis was the under-
lying cause of hypertension in 3.6% of the cas-
es. It should in particular be sought in younger
headache
visual disturbances
gastrointestinal symptoms
heart failure
neurological sequelae (encephalopathy)
left ventricular hypertrophy
mild-to-moderate renal impairment
severe renal impairment
microangiophatic hemolytic anemia
20
30
50
70
10
0
40
60
80
90
100
frequency, %
FIGURE 3 Presenting
symptoms and associated
complications in patients
with hypertensive crisis
and advanced
retinopathy;
gastrointestinal
symptoms: nausea,
vomiting, weight loss;
mild-to-moderate renal
impairment, serum
creatinine concentration
115–300 μmol/l; severe
renal impairment, serum
creatinine concentration
>300 μmol/l. Adapted
from van den Born.1
POLSKIE ARCHIWUM MEDYCYNY WEWNĘTRZNEJ
2016; 126 (1-2)
90
years of follow-up, respectively. e study con-
firmed that the severity of renal failure at presen-
tation, worse BP control during follow--up, and
severity of proteinuria were important prognos-
tic factors for long-term renal outcomes.46
In another study, van den Born et al28 reported
a relatively high prevalence of microangiopath-
ic hemolysis in patients with MHT and showed
that, in those subjects, microangiopathic hemo-
lysis was an important indicator of both renal in-
sufficiency and recovery.
It has been reported that, when intensive an-
tihypertensive therapy was started in patients
with MHT, renal function remained unchanged
or improved in nearly half of those with initial
renal insufficiency. In one series of 54 patients
with MHT requiring dialysis, 22% of the subjects
recovered sufficient renal function to allow with-
drawal of dialysis.47,4 8
Considering that the activation of the renin–
-angiotensin–aldosterone system is an impor-
tant pathogenic mechanism in MHT and that
drugs affecting the system have specific antipro-
teinuric properties, a long-term treatment of pa-
tients with MHT should be based mainly on this
type of agents.
Management MHT is a hypertensive emergen-
cy, and patients should receive immediate anti-
hypertensive treatment owing to a high risk of
renal failure, stroke, myocardial infarction, and
heart failure.
Following the 2013 guidelines on hyperten-
sion from the European Society of Hypertension
and European Society of Cardiology, the current
treatment is based on agents that can be admin-
istered by intravenous infusion and titrated ac-
cording to response, avoiding abrupt falls in BP
and excessive hypotension.3
e goal of the immediate therapy should be to
lower diastolic BP to approximately 110 mmHg.
ere is a consensus opinion to reduce MAP by
no more than 25% during the acute phase in or-
der to avoid cerebral hypoperfusion. Labetalol,
sodium nitroprusside, nicardipine, nitrates, and
furosemide are listed among the most frequent-
ly used intravenous drugs. However, due to the
low incidence of MHT, there are no data to doc-
ument which of these drugs is best, or whether
their use is followed by a decrease in morbidity
and mortality in this group of patients.1
e study by Immink et al
4
showed a significant
difference in the decrease of systemic vascular re-
sistance during the administration of sodium ni-
troprusside and labetalol (43% and 13%, respec-
tively) in patients with MHT. ese may be attrib-
uted to differential effects of these drug classes
on systemic vascular resistance and heart rate in
patients with MHT.
49
It has also been recognized
that antihypertensive drugs differ in their capac-
ity to reduce central BP, whereas their effects on
peripheral BP may be similar.4
Taken together, in the absence of definite ev-
idence, clinicians must continue to administer
these studies have had major limitations, such as
a short follow-up duration and the lack of use of
3-dimensional echocardiography.41-43
LVH, observed in patients with new-onset
MHT, may at least partly reflect an adaptive re-
sponse to persistently elevated BP, and its pres-
ence is associated with unfavorable prognosis.
Indeed, the duration of the follow-up and the
quality of BP control had a strong direct effect
on prognosis in patients with MHT presenting
with LVH.41-43
In contrast to previous observations based on
a cohort of MHT patients with less rigorous BP
control, the study by Shantsila et al41 showed that
patients with adequately controlled MHT have a
similar degree of LVH to that in control subjects
without MHT. e study documented that all pa-
tients had preserved left ventricular ejection frac-
tion, despite a median 12-year history of MHT,
which did not differ from the values observed in
the control group. In addition, the systolic and
diastolic function indices were out of range, but
did not differ between the MHT and non-MHT
groups.
41
ese findings appear to be similar to
the observations made by Gosse et al,
42
who dem-
onstrated a recovery of the left ventricular func-
tion with treatment during a short follow-up.
Taken together, the above observations prove
that, despite good long-term BP control, patients
with MHT have persistent structural and func-
tional changes in the left ventricle on echocar-
diography, comparable to those observed in con-
trol non-MHT patients.
Long-term renal outcome
Although the survival
rate of patients with MHT has considerably im-
proved with the introduction of antihyperten-
sive therapy, MHT is frequently complicated by
renal insufficiency, and ESRD still remains a sig-
nificant cause of morbidity and mortality in this
patient group. e STAT study
44
demonstrated
that, in patients hospitalized with acute severe
hypertension, the presence of acute kidney injury
was associated with a greater risk of serious out-
comes, including mortality and additional end-
-organ damage.44
A retrospective analysis of 120 patients with
MHT conducted in the Netherlands showed that
patients with MHT had a markedly increased risk
of ESRD after the acute phase.45 During a medi-
an follow-up period of 67 months, 31% of the pa-
tients reached the primary endpoint (ESRD), 15%
reached the secondary endpoint (all-cause mor-
tality), and 24% required kidney replacement. Af-
ter the acute phase, initial serum creatinine lev-
els and BP during follow-up were the main pre-
dictors of future ESRD.
In another retrospective study, Gonzalez etal
46
reported on long-term renal outcomes in a cohort
of 197 patients with MHT. Renal involvement was
a major component of MHT, with 63% of the pa-
tients presenting with acute renal function im-
pairment. e probability of renal survival in the
entire cohort was 84% and 72% after 5 and 10
REVIEW ARTICLE Malignant hypertension: new aspects of an old clinical entity 91
history of MHT remain at increased risk, and the
annual incidence of all-cause mortality was high-
er in MHT patients compared with that in nor-
motensive and hypertensive controls.
Taken together, despite satisfactory BP con-
trol with modern antihypertensive therapy, MHT
continues to be an important clinical entity and
may significantly contribute to an increase in to-
tal cardiovascular risk.1,14,50-55
Summary
In summary, MHT is not a “vanishing
disease”, having a relatively stable number of cas-
es per year, but its prognosis and survival rates
have significantly improved owing to earlier detec-
tion, tighter BP control, lower BP targets, better
choice of antihypertensive drugs, and availability
of hemodialysis and renal transplantation. How-
ever, despite the vast range of antihypertensive
agents and effective BP control, MHT remains a
significant clinical challenge.
Acknowledgments We would like to thank Mrs.
L. Kozłowska and Mrs. M. Ciechanowska for their
help during preparation of this manuscript. e
research of TG and TM is funded by the Na-
tional Science Centre, Poland (NCN 2011/03/B/
NZ4/02454).
REFERENCES
1 van den Born BJ, Beutler JJ, Gaillard CA, et al. Dutch guideline for the
management of hypertensive crisis - 2010 revision. Neth J Med. 2011; 69:
248-255.
2
Kaplan NM, Victor RG. Kaplan’s clinical hypertension. Wolters Kluwer,
Philadelphia, 2015.
3 Mancia G, Fagard R, Narkiewicz K, et al. 2013 ESH/ESC Guidelines for
the management of arterial hypertension: the Task Force for the manage-
ment of arterial hypertension of the European Society of Hypertension (ESH)
and of the European Society of Cardiology (ESC). J Hypertens. 2013; 31:
1281-1357.
parenteral drugs to markedly lower elevated BP
in patients with a hypertensive emergency. How-
ever, this should be done carefully, under close su-
pervision, and with the choice of drugs that allow
for a gradual reduction in BP, that have no inher-
ent toxicity, and that provide the ability to back
down if the target organ function deteriorates.
1-3
Prognosis
If left untreated, the 1-year surviv-
al rate has been reported to be only 10% to 20%,
with most patients dying within 6 months. Effec-
tive treatment has led to a substantial improve-
ment in the survival rate, with the rate at 5 years
reaching between 60% and 75% in developed
countries. However, in developing countries such
as Nigeria, the prognosis appears to be consid-
erably worse, with 1-year survival rates of only
37.5%.1,14,50-55
e results from the largest prospective regis-
try of 446 patients with MHT attending the City
Hospital in Birmingham, United Kingdom, clearly
showed a significant improvement in the 5-year
survival rates, from 32% before 1977 to as much
as 91% for patients diagnosed between 1997 and
2006 (FIGURE 4).5
e mortality rate was similar among white
Europeans and African-Caribbeans, but signifi-
cantly lower among South-Asians. e best inde-
pendent predictors of overall survival were age,
decade of MHT diagnosis, serum creatinine lev-
els at presentation, and follow-up systolic BP.5
In addition, a retrospective cohort study that
investigated national trends in the United States
in hospital admissions for MHT showed a de-
crease in the mortality rate by 36% among pa-
tients diagnosed with MHT after 2007.7
e study by Amraoui et al50 documented that,
despite a considerable improvement in the sur-
vival rate over the past decades, patients with a
FIGURE 4 5-year
survival of patients with
malignant hypertension
by the year of diagnosis
of malignant
hypertension in a
prospective registry of
446 patients with
malignant hypertension
attending the City
Hospital in Birmingham,
United Kingdom.
Adapted from Lane et al.5
20 30 50 7010040608090 100
year of diagnosis
5-year survival, %
<1967
1967–1976
1977–1986
1987–1996
1997–2006
POLSKIE ARCHIWUM MEDYCYNY WEWNĘTRZNEJ
2016; 126 (1-2)
92
30 Deguchi I, Uchino A, Suzuki H, et al. Malignant hypertension with re-
versible brainstem hypertensive encephalopathy and thrombotic microangi-
opathy. J Stroke Cerebrovasc Dis. 2012; 21: 915 e917-e920.
31
Akimoto T, Muto S, Ito C, et al. Clinical features of malignant hyperten-
sion with thrombotic microangiopathy. Clin Exp Hypertens. 2011; 33: 77-83.
32
Nzerue C, Oluwole K, Adejorin D, et al. Malignant hypertension with
thrombotic microangiopathy and persistent acute kidney injury (AKI). Clin
Kidney J. 2014; 7: 586-589.
33 Khanal N, Dahal S, Upadhyay S, et al. Differentiating malignant hyper-
tension-induced thrombotic microangiopathy from thrombotic thrombocyto-
penic purpura. Ther Adv Hematol. 2015; 6: 97-102.
34
Kumar P, Arora P, Kher V, et al. Malignant hypertension in children in In-
dia. Nephrol Dial Transplant. 1996; 11: 1261-1266.
35 Zotos P, Grivas D, Tsitsibis K, et al. Malignant hypertension associated
with IgA nephropathy. Intern Med J. 2010; 40: 668-669.
36 Prejbisz A, Klisiewicz A, Januszewicz A, et al. 22-Year-old patient with
malignant hypertension associated with primary aldosteronism. J Hum Hy-
pertens. 2013; 27: 138-140.
37
Kobayashi S, Hoshi A, Tanaka K, et al. Bilateral insular lesions related to
malignant hypertension. Intern Med. 2012; 51: 1805-1806.
38
Kumar UM, Pande P, Savita S, et al. An Extra-adrenal Pheochromocyto-
ma Presenting as Malignant Hypertension-A Report of two cases. J Clin Di-
agn Res. 2013; 7: 1177-1179.
39
Shera AH, Baba AA, Bakshi IH, et al. Recurrent malignant juxtaglomer-
ular cell tumor: A rare cause of malignant hypertension in a child. J Indian
Assoc Pediatr Surg. 2011; 16: 152-154.
40
Prejbisz A, Antoniewicz AA, Kabat M, et al. Reninsecreting juxtaglo-
merular cell tumor of the kidney causing severe hypertension and polyuria.
Pol Arch Med Wewn. 2014; 124: 207-208.
41
Shantsila A, Dwivedi G, Shantsila E, et al. A comprehensive assess-
ment of cardiac structure and function in patients with treated malignant
phase hypertension: the West Birmingham Malignant Hypertension project.
Int J Cardiol. 2013; 167: 67-72.
42
Gosse P, Coulon P, Papaioannou G, et al. Impact of malignant arterial hy-
pertension on the heart. J Hypertens. 2011; 29: 798-802.
43
Nadar S, Beevers DG, Lip GY. Echocardiographic changes in patients
with malignant phase hypertension: the West Birmingham Malignant Hyper-
tension Register. J Hum Hypertens. 2005; 19: 69-75.
44 Szczech LA, Granger CB, Dasta JF, et al. Acute kidney injury and car-
diovascular outcomes in acute severe hypertension. Circulation. 2010; 121:
2183-2191.
45
Amraoui F, Bos S, Vogt L , et al. Long-term renal outcome in patients
with malignant hypertension: a retrospective cohort study. BMC Nephrol.
2012; 13: 71.
46 Gonzalez R, Morales E, Segura J, et al. Long-term renal survival in ma-
lignant hypertension. Nephrol Dial Transplant. 2010; 25: 3266-3272.
47 James SH, Meyers AM, Milne FJ, et al. Partial recovery of renal func-
tion in black patients with apparent end-stage renal failure due to primary
malignant hypertension. Nephron. 1995; 71: 29-34.
48
Lip GY, Beevers M, Beevers DG. Does renal function improve after diag-
nosis of malignant phase hypertension? J Hypertens. 1997; 15: 1309-1315.
49 Immink RV, van den Born BJ, van Montfrans GA, et al. Cerebral hemo-
dynamics during treatment with sodium nitroprusside versus labetalol in ma-
lignant hypertension. Hypertension. 2008; 52: 236-240.
50 Amraoui F, Van Der Hoeven NV, Van Valkengoed IG, et al. Mortality and
cardiovascular risk in patients with a history of malignant hypertension: a
case-control study. J Clin Hypertens (Greenwich). 2014; 16: 122-126.
51 Lip GY, Beevers M, Beevers DG. Complications and survival of 315 pa-
tients with malignant-phase hypertension. J Hypertens. 1995; 13: 915-924.
52
Kadiri S, Olutade BO, Osobamiro O. Factors influencing the develop-
ment of malignant hypertension in Nigeria. J Hum Hypertens. 2000; 14:
171-174.
53
Ventura HO, Mehra MR, Messerli FH. Desperate diseases, desperate
measures: tackling malignant hypertension in the 1950s. Am Heart J. 2001;
142: 197-203.
54 Edmunds E, Beevers DG, Lip GY. What has happened to malignant hy-
pertension? A disease no longer vanishing. J Hum Hypertens. 2000; 14:
159-161.
55 Webster J, Petrie JC, Jeffers TA, et al. Accelerated hypertension-pat-
terns of mortality and clinical factors affecting outcome in treated patients.
Q J Med. 1993; 86: 485-493.
4 van den Bogaard B, Immink RV, Westerhof BE, et al. Central versus pe -
ripheral blood pressure in malignant hypertension; effects of antihyperten-
sive treatment. Am J Hypertens. 2013; 26: 574-579.
5 Lane DA, Lip GY, Beevers DG. Improving survival of malignant hyperten-
sion patients over 40 years. Am J Hypertens. 2009; 22: 1199-1204.
6
Scarpelli PT, Livi R, Caselli GM, et al. Accelerated (malignant) hyper-
tension: a study of 121 cases between 1974 and 1996. J Nephrol. 1997;
10: 207-215.
7
Polgreen LA, Suneja M, Tang F, et al. Increasing trend in admissions
for malignant hypertension and hypertensive encephalopathy in the United
States. Hypertension. 2015; 65: 1002-1007.
8 Laragh JH, Ulick S, Januszewicz V, et al. Aldosterone secretion and pri-
mary and malignant hypertension. J Clin Invest. 1960; 39: 1091-1106.
9 Howard CG, Mitchell KD. Renal functional responses to selective intra-
renal renin inhibition in Cyp1a1-Ren2 transgenic rats with ANG II-dependent
malignant hypertension. Am J Physiol Renal Physiol. 2012; 302: F52-F59.
10 Hilme E, Hansson L, Sandberg L, et al. Abnormal immune function in
malignant hypertension. J Hypertens. 1993; 11: 989-994.
11 van den Born BJ, Lowenberg EC, van der Hoeven NV, et al. Endothelial
dysfunction, platelet activation, thrombogenesis and fibrinolysis in patients
with hypertensive crisis. J Hypertens. 2011; 29: 922-927.
12
Shantsila A, Dwivedi G, Shantsila E, et al. Persistent macrovascular
and microvascular dysfunction in patients with malignant hypertension. Hy-
pertension. 2011; 57: 490-496.
13 Shantsila A, Lane DA, Beevers DG, et al. Lack of impact of pulse pres-
sure on outcomes in patients with malignant phase hypertension: the West
Birmingham Malignant Hypertension study. J Hypertens. 2012; 30: 974-979.
14
van den Born BJ, Koopmans RP, Groeneveld JO, et al. Ethnic disparities
in the incidence, presentation and complications of malignant hypertension.
J Hypertens. 2006; 24: 2299-2304.
15
Caro J, Morales E, Gutierrez E, et al. Malignant hypertension in patients
treated with vascular endothelial growth factor inhibitors. J Clin Hypertens
(Greenwich). 2013; 15: 215-216.
16
Cunningham MW, Jr., Sasser JM, West CA, et al. Renal nitric oxide
synthase and antioxidant preservation in Cyp1a1-Ren-2 transgenic rats with
inducible malignant hypertension. Am J Hypertens. 2013; 26: 1242-1249.
17
Efrati S, Berman S, Goldfinger N, et al. Enhanced angiotensin II pro-
duction by renal mesangium is responsible for apoptosis/proliferation of en-
dothelial and epithelial cells in a model of malignant hypertension. J Hyper-
tens. 2007; 25: 1041-1052.
18
van den Born BJ, van Montfrans GA, Uitterlinden AG, et al. The M235T
polymorphism in the angiotensinogen gene is associated with the risk of ma-
lignant hypertension in white patients. J Hypertens. 2007; 25: 2227-2233.
19 van den Born BJ, Koopmans RP, van Montfrans GA. The renin-angio-
tensin system in malignant hypertension revisited: plasma renin activity, mi-
croangiopathic hemolysis, and renal failure in malignant hypertension. Am J
Hypertens. 2007; 20: 900-906.
20 Milani CJ, Kobori H, Mullins JJ, et al. Enhanced urinary angiotensino-
gen excretion in Cyp1a1-Ren2 transgenic rats with inducible ANG II-depen-
dent malignant hypertension. Am J Med Sci. 2010; 340: 389-394.
21
Abrahamsen CT, Barone FC, Campbell WG, Jr, et al. The angiotensin
type 1 receptor antagonist, eprosartan, attenuates the progression of renal
disease in spontaneously hypertensive stroke-prone rats with accelerated
hypertension. J Pharmacol Exp Ther. 2002; 301: 21-28.
22 Guzik TJ, Hoch NE, Brown KA, et al. Role of the T cell in the genesis of
angiotensin II induced hypertension and vascular dysfunction. J Exp Med.
2007; 204: 2449-2460.
23
Crowley SD, Song YS, Lin EE, et al. Lymphocyte responses exacer-
bate angiotensin II-dependent hypertension. Am J Physiol Regul Integr Comp
Physiol. 2010; 298: R1089-R1097.
24
Mattson DL, Lund H, Guo C, et al. Genetic mutation of recombina-
tion activating gene 1 in Dahl salt-sensitive rats attenuates hypertension
and renal damage. Am J Physiol Regul Integr Comp Physiol. 2013; 304:
R407-R414.
25 Muller DN, Shagdarsuren E, Park JK, et al. Immunosuppressive treat-
ment protects against angiotensin II-induced renal damage. Am J Pathol.
2002; 161: 1679-1693.
26 Youn JC, Yu HT, Lim BJ, et al. Immunosenescent CD8+ T cells and C-
X-C chemokine receptor type 3 chemokines are increased in human hyper-
tension. Hypertension. 2013; 62: 126-133.
27
Herrera J, Ferrebuz A, MacGregor EG, et al. Mycophenolate mofetil
treatment improves hypertension in patients with psoriasis and rheumatoid
arthritis. J Am Soc Nephrol. 2006; 17: S218-S225.
28 van den Born BJ, Honnebier UP, Koopmans RP, et al. Microangiopathic
hemolysis and renal failure in malignant hypertension. Hypertension. 2005;
45: 246-251.
29
Eguchi K, Kasahara K, Nagashima A, et al. Two cases of malignant
hypertension with reversible diffuse leukoencephalopathy exhibiting a re-
versible nocturnal blood pressure “riser” pattern. Hypertens Res. 2002; 25:
467-473.
ARTYKUŁ POGLĄDOWY  Nadciśnienie tętnicze złośliwe – nowe aspekty starej jednostki klinicznej 93
Adres do korespondencji: 
dr n. med. Aleksander Prejbisz, Klinika
Nadciśnienia Tętniczego, Instytut
Kardiologii, Warszawa, ul. Alpejska 42,
04-628, Warszawa, tel.: 22 343 43
39, e-mail: a.prejbisz@ikard.pl
Praca wpłynęła: 12.11.2015.
Przyjęta do druku: 16.11.2015.
Publikacja online: 10.12.2015.
Nie zgłoszono sprzeczności 
interesów.
Pol Arch Med Wewn. 2016;
126 (1-2): 86-93
Copyright by Medycyna Praktyczna,
Kraków 2016
SŁOWA KLUczOWe
leczenie, nadciśnienie 
tętnicze złośliwe, 
patogeneza, 
występowanie
STReSzczenie
Nadciśnienie tętnicze  złośliwe  (malignant hypertension– MHT),  nazywane także przyspieszonym 
nadciśnieniem tętniczym złośliwym lub  fazą złośliwą  nadciśnienia tętniczego, jest najcięższą postacią 
nadciśnienia tętniczego. Definiuje się je jako istotnie podwyższone wartości ciśnienia tętniczego z towa-
rzyszącymi zmianami na dnie oka (płomykowate wybroczyny, wysięki, ogniska waty, z lub bez obrzęku 
tarczy nerwu wzrokowego). Pomimo dostępności licznych leków hipotensyjnych MHT pozostaje ważnym
problem klinicznym. Częstość jego występowania jest bardzo niska, jednak bezwzględna liczba nowych
przypadków nie  zmieniła się na  przestrzeni ostatnich dekad.  Chociaż rola aktywacji  układu renina-an-
giotensyna-aldosteron i  dysfunkcji śródbłonka w patogenezie  MHT została dobrze udokumentowana, 
wyniki ostatnich badań wskazują, że także układ odpornościowy może odgrywać istotną rolę w rozwoju 
tej choroby. Chorzy z MHT charakteryzują się nasilonymi powikłaniami narządowymi nadciśnienia tętni-
czego, w tym nieprawidłowościami strukturalnymi i funkcjonalnymi serca. Częstymi powikłaniami MHT 
są upośledzenie funkcji nerek oraz schyłkowa niewydolność nerek. Przeżycie chorych z MHT poprawiło 
się znacząco wraz ze zwiększeniem dostępności leczenia hipotensyjnego, niemniej jednak upośledzenie
funkcji nerek i rozwój schyłkowej niewydolności nerek wciąż stanowią istotną przyczynę chorobowo-
ści i śmiertelności w tej  grupie chorych. Podsumowując, MHT nie jest „znikającą chorobą”, ponieważ 
charakteryzuje się stałą  liczbą nowych przypadków rocznie.  Rokowanie i  przeżycie  pacjentów z MHT 
poprawiło się natomiast znacząco w wyniku wcześniejszego rozpoznania, ściślejszej kontroli ciśnienia 
tętniczego, niższych wartości docelowych ciśnienia tętniczego, większego wyboru leków hipotensyjnych 
oraz zwiększenia dostępności leczenia nerkozastępczego i transplantacji nerek.
ARTYKUŁ POGLĄDOWY
Nadciśnienie tętnicze złośliwe – nowe aspekty
starej jednostki klinicznej
AndrzejJanuszewicz
1
, TomaszGuzik
2
, AleksanderPrejbisz
1
,
TomaszMikołajczyk
2
, GrzegorzOsmenda
2
, WłodzimierzJanuszewicz
3
1 Klinika Nadciśnienia Tętniczego, Instytut Kardiologii, Warszawa
2 Katedra Chorób Wewnętrznych i Medycyny Wsi, Uniwersytet Jagielloński, Collegium Medicum, Kraków
3 Warszawa
... accelerated-malignant hypertension or malignant-phase hypertension, is the most serious form of hypertension and is a systemic disease characterized by extremely elevated systolic blood pressure and out-of-range office diastolic blood pressure [above 130 mmHg] at the time of diagnosis as well as acute ischemic organ damage (kidney, eye, etc.) [3][4][5]. Although the diagnosis of MHT is simple, clinically, it can only be diagnosed when the target organs are damaged [6], and the kidney is the organ most commonly affected by hypertension-mediated damage, and the damage to the kidney is irreversible. ...
... The clinical diagnosis of MHT was as follows: arterial blood pressure increased sharply in a short period, diastolic blood pressure ≥ 130 mmHg (or systolic blood pressure ≥ 180 mmHg), and fundus changes showed hypertensive retinal changes (retinal cotton exudation, hemorrhage, with or without optic papilla edema) [5]. The pathological diagnosis of renal damage in MHT was based on malignant hypertensive nephrosclerosis, characterized by proliferative arterioles (interlobular and arcuate arteries), endomyelitis (onion skin-like changes), and fibrinoid necrosis of the arteriole wall. ...
Article
Full-text available
Background Elabela, a recently discovered hormonal peptide containing 32 amino acids, is a ligand for the apelin receptor. It can lower blood pressure and attenuate renal fibrosis. However, the clinicopathological relationship between Elabela level and renal damage caused by benign hypertension (BHT) and malignant hypertension (MHT) has not been elucidated. Therefore, we investigated the clinicopathological correlation between serum Elabela level and renal damage caused by BHT and MHT. Methods The participants comprised 50 patients and 25 age-matched healthy adults. The 50 patients were separated into two groups: MHT (n = 25) and BHT groups (n = 25). We analyzed their medical histories, demographics, and clinical examinations, including physical and laboratory tests. Results The results showed that serum Elabela level decreased gradually with a continuous increase in blood pressure from the healthy control group, BHT, to MHT. Moreover, Elabela levels negatively correlated with BMI (R = − 0.27, P = 0.02), SBP (r = − 0.64, P < 0.01), DBP (r = − 0.58, P < 0.01), uric acid (r = − 0.39, P < 0.01), bun (r = − 0.53, P < 0.01), and Scr (r = − 0.53 P < 0.01) but positively correlated with eGFR (r = 0.54, P < 0.01). Stepwise multivariate linear regression analysis showed that SBP was the variable most related to Elabela (t = − 5.592, P < 0.01). Conclusions Serum Elabela levels decreased in patients with hypertension, especially malignant hypertension, and has the potential to be a marker of hypertension-related kidney damage.
... The results of the 2012-2015 survey using a strati ed multistage random sampling method to obtain a nationally representative sample of 451 755 residents ≥ 18 years of age from 31 provinces in Mainland China showed that the prevalence rate of hypertension among the Chinese adult population was approximately 23.2% [2]. Malignant hypertension (MHT), also known as accelerated-malignant hypertension or malignant-phase hypertension, is the most serious form of hypertension and is a systemic disease characterized by hypertension (extremely elevated systolic blood pressure and out-of-range o ce diastolic blood pressure [above 130 mmHg] at the time of diagnosis) and acute ischemic organ damage (kidney, eye, heart, etc.) [3][4][5]. Clinically, malignant hypertension can only be diagnosed when the target organs are damaged [6]. ...
... The clinical diagnosis of MHT was as follows: arterial blood pressure increased sharply in a short period, diastolic blood pressure ≥ 130 mmHg (or systolic blood pressure ≥ 180 mmHg), and fundus changes showed hypertensive retinal changes (retinal cotton exudation, hemorrhage, with or without optic papilla edema) [5]. The pathological diagnosis of renal damage in MHT was malignant hypertensive nephrosclerosis, characterized by proliferative arterioles (interlobular and arcuate arteries), endomyelitis (onion skin-like changes), and brinoid necrosis of the arteriole wall. ...
Preprint
Full-text available
Background Elabela, a recently discovered hormonal peptide containing 32 amino acids, is a ligand for the apelin receptor. It can lower blood pressure and attenuate renal fibrosis. However, the clinicopathological relationship between the Elabela level and renal damage caused by benign hypertension (BHT) and malignant hypertension (MHT) has not been elucidated. Therefore, we discussed the clinicopathological correlation between the serum Elabela level and renal damage caused by BHT and MHT in patients. Methods The participants comprised 50 patients and 25 age-matched healthy adults. The 50 patients were separated into two groups: the MHT (n = 25) and BHT groups (n = 25). We analyzed their medical histories, demographics, and clinical examinations, including physical and laboratory tests. Results The results showed that the serum Elabela level decreased gradually with a continuous increase in blood pressure from the healthy control group, BHT, to MHT. Moreover, the Elabela levels negatively correlated with BMI(R = − 0.27, P = 0.02), SBP (R = − 0.64, P < 0.01), DBP (R = − 0.58, P < 0.01), Uric acid(R = − 0.39, P < 0.01), BUN (R = − 0.53, P < 0.01), and Scr (R = − 0.53 P < 0.01) but positively correlated with eGFR (R = 0.54, P < 0.01). Stepwise multivariate linear regression analysis showed that SBP was the variable most related to Elabela (t = − 7.029, P < 0.01). Conclusions Serum Elabela levels decreased in patients with hypertension, especially malignant hypertension, and had a significant negative correlation with systolic blood pressure. Trial registration: retrospectively registered approval number:2020076.
... Activation of the renin-angiotensin-aldosterone system together with vasoconstriction and ischemia of the small renal arteries was reported to contribute to the development of malignant hypertension [2]. Van den Born et al. indicated that both plasma renin activity (PRA) and plasma aldosterone concentration (PAC) are elevated in patients with malignant hypertension and emphasized that in particular renin accelerates vascular damage and renal dysfunction [3]. ...
... Malignant hypertension (hypertensive emergency) is characterized by rapid blood pressure elevation, acute renal deterioration, retinopathy, encephalopathy and heart failure, and kidney histology is characterized by small renal arterial occlusive lesions including onion skin lesions. While high serum renin and aldosterone levels are another feature hyperreninemia [1][2][3][4]. There are many reports that increased renin secretion is based on the malignant hypertension, including the following. ...
Article
A 37-year-old Japanese man was admitted to our hospital for evaluation of severe hypertension and visual impairment. His serum creatinine was 4.16 mg/dL. Plasma renin activity was normal (2.7 ng/mL/h), but plasma aldosterone concentration was elevated (27.2 ng/dL). A kidney biopsy showed concentric subendothelial edematous thickening of the arterioles (onion skin pattern) with luminal narrowing or obstruction, and malignant nephrosclerosis was diagnosed. Antihypertensive therapies, including an angiotensin II receptor blocker and spironolactone, were administered and effectively preserved kidney function and normalized blood pressure. This case indicates that hyperaldosteronemia in the presence of normal renin levels might also cause malignant hypertension.
... Although being more worldwide prevalent, the prognosis of MHT was improved during the last decades especially in developed countries. The poor prognosis could occur if no treatment or management resulting in persistent target organ damage and other coronary complications (10,11) . ...
... aHUS is characterized by thrombocytopenia, microangiopathic hemolytic anemia, and acute kidney injury and can also present as progressive kidney damage, or as extrarenal manifestations resulting in damage to other organs [5][6][7]. Another condition that can result in TMA is malignant hypertension (MHT), a severe form of arterial hypertension traditionally diagnosed by high blood pressure (diastolic pressure > 120 mmHg) with papilledema/hypertensive retinopathy [8][9][10][11][12]. More recent experience has emphasized the role of multi-organ involvement/damage in the diagnosis and prognosis of MHT, and MHT with multi-organ involvement has also been referred to as hypertensive emergency [10,13]. ...
Article
Full-text available
Introduction Atypical hemolytic uremic syndrome (aHUS) is a rare form of thrombotic microangiopathy (TMA) often caused by alternative complement dysregulation. Patients with aHUS can present with malignant hypertension (MHT), which may also cause TMA. Methods This analysis of the Global aHUS Registry (NCT01522183) assessed demographics and clinical characteristics in eculizumab-treated and not-treated patients with aHUS, with ( n = 71) and without ( n = 1026) malignant hypertension, to further elucidate the potential relationship between aHUS and malignant hypertension. Results While demographics were similar, patients with aHUS + malignant hypertension had an increased need for renal replacement therapy, including kidney transplantation (47% vs 32%), and more pathogenic variants/anti-complement factor H antibodies (56% vs 37%) than those without malignant hypertension. Not-treated patients with malignant hypertension had the highest incidence of variants/antibodies (65%) and a greater need for kidney transplantation than treated patients with malignant hypertension (65% vs none). In a multivariate analysis, the risk of end-stage kidney disease or death was similar between not-treated patients irrespective of malignant hypertension and was significantly reduced in treated vs not-treated patients with aHUS + malignant hypertension (adjusted HR (95% CI), 0.11 [0.01–0.87], P = 0.036). Conclusions These results confirm the high severity and poor prognosis of untreated aHUS and suggest that eculizumab is effective in patients with aHUS ± malignant hypertension. Furthermore, these data highlight the importance of accurate, timely diagnosis and treatment in these populations and support consideration of aHUS in patients with malignant hypertension and TMA. Trial registration details Atypical Hemolytic-Uremic Syndrome (aHUS) Registry. Registry number: NCT01522183 (first listed 31st January, 2012; start date 30th April, 2012). Graphical abstract
... Malignant hypertension (MHT), a form of hypertensive emergency also known as accelerated-malignant hypertension or hypertensive crisis with retinopathy [1,2], is defined as an extreme elevation (above 120 to 130 mmHg) of the diastolic blood pressure (BP) with grade III or IV hypertensive retinopathy and has the potential to lead to progressive organ damage [3]. The prognosis of MHT is poor without treatment [3,4]; thus, it is important for physicians to diagnose the condition immediately and start treatment to control the BP. ...
Article
Full-text available
Patient: Male, 32-year-old Final Diagnosis: Malignant hypertension Symptoms: Epigastric pain Medication: — Clinical Procedure: — Specialty: Nephrology Objective Unusual clinical course Background Malignant hypertension (MHT), one of the severest forms of hypertension, can have deleterious effects on various organs, such as renal failure, retinopathy, and encephalopathy. These types of organ damage are common complications of MHT, but in several previous cases, damage to other organs, such as the gastrointestinal tract or pancreas, resulting from small vessel lesions, has also been reported, and these cases have had severe clinical outcomes and a poor prognosis. Case Report A 32-year-old male patient with untreated hypertension of a 5-year duration presented with breathlessness and edema. His blood pressure was 220/144 mmHg, and he had renal dysfunction, congestive heart failure, and hypertensive retinopathy. He immediately received treatment, including antihypertensive agents and intermittent hemodialysis, but experienced epigastric pain for several days. A cystic lesion appeared in the pancreatic head, and his serum pancreatic enzymes were elevated. Based on these findings, acute pancreatitis with a cystic lesion was diagnosed. He first received fluid management, pain control, and parenteral nutrition but experienced 2 relapses. Finally, he received transpapillary endoscopic drainage for the cystic lesion with suspected walled-off necrosis. Thereafter, his symptoms improved. Conclusions The present case of MHT is the first to demonstrate acute necrotizing pancreatitis and it illustrates the difficulty of treatment. Therefore, if a patient with MHT presents with abdominal pain, a thorough workup, including contrast-enhanced computed tomography, should be performed to rule out significant organ involvement.
... Its prevalence is estimated at 2-10/100,000/year [2,3]. MH most commonly appears in patients with essential hypertension, but it may also be caused by renal, endocrine, vascular or rheumatoid diseases, as well as juxtaglomerular area neoplasms [4][5][6]. The 5-year survival rate in patients who underwent MH ranges from <5% (if the episode was left untreated) to 90% [3,7]. ...
... Cannabis use disorder (dependence or abuse) has been linked to the risk of acute cardiocerebrovascular events [3]. Malignant hypertension is one of the hypertensive emergencies which could lead to acute end-organ damage like myocardial infarction, cardiac failure, and stroke, if not treated [4]. Furthermore, end-organ damage could lead to long-term comorbidities and increased health care burden. ...
Article
Full-text available
Reliable clinical signs associated with recent brain injuries during malignant hypertension crisis are lacking. In this single-center study we compared the prevalence of brain MRI injuries between fully asymptomatic patients and those with headache; we found no significant differences, suggesting that headache is not associated with recent brain injuries during malignant hypertension.
Article
Background: Acute and diffuse microvascular damage characterizes malignant hypertension (MHT), the deadliest form of hypertension (HTN). Although its ophthalmological, renal and cardiological repercussions are well known, brain involvement is considered rare with few descriptions, although it is one of the main causes of death. We hypothesized that brain MRI abnormalities are common in MHT, even in patients without objective neurological signs. Method: We analyzed retrospectively the brain MRI of patients admitted for acute MHT between 2008 and 2018 in Bordeaux University Hospital, regardless of their neurological status. A trained operator analyzed every brain MRI, looking for posterior reversible encephalopathy syndrome (PRES), ischemic stroke, intracerebral hematoma (ICH) and microangiopathy markers. We included 58 patients without neurological signs, 66% were men, and mean age was 45.6 ± 11.3 years. Results: Brain MRI were normal in 26% of patients but we found at least one acute abnormality on brain MRI in 29% and an Small Vessel Disease score (SVD score) of two or higher in 52%. In patients with neurological signs, these findings were 9, 53 and 70%, respectively. A PRES was found in 16% of asymptomatic patients and 31% had an ischemic stroke and/or a cerebral hematoma. Conclusion: PRES, recent hematoma, ischemic stroke and severe cerebral microangiopathy are common findings in MHT patients without neurological signs on admission. The impact of these findings on patient management, and their cerebrovascular and cognitive prognostic value, should be established. Brain MRI might need to become systematic in patients suffering from MHT episodes.
Article
Full-text available
Malignant hypertension can cause thrombotic microangiopathy (TMA) and the overall presentation may mimic thrombotic thrombocytopenic purpura (TTP). This presents a dilemma of whether or not to initiate plasma exchange. The objective of the study was to determine the clinical and laboratory manifestations of malignant hypertension-induced TMA, and its outcomes. Using several search terms, we reviewed English language articles on malignant hypertension-induced TMA, indexed in MEDLINE by 31 December 2013. We also report a new case. All these cases were analyzed using descriptive statistics. A total of 19 patients, with 10 males, had a median age of 38 years at diagnosis; 58% had a history of hypertension. Mean arterial pressure at presentation was 159 mmHg (range 123-190 mmHg). All had prominent renal dysfunction (mean creatinine of 5.2 mg/dl, range 1.7-13 mg/dl) but relatively modest thrombocytopenia (mean platelet count of 60 × 103/µl, range 12-131 × 10(3)/µl). Reported cases (n = 9) mostly had preserved ADAMTS-13 activity (mean 64%, range 18-96%). Following blood pressure control, the majority had improvement in presenting symptoms (100%) and platelet counts (84%); however, only 58% had significant improvement in creatinine. More than half (53%) needed hemodialysis. One patient died of cardiac arrest during pacemaker insertion. Prior history of hypertension, high mean arterial pressure, significant renal impairment but relatively modest thrombocytopenia and lack of severe ADAMTS-13 deficiency (activity <10%) at diagnosis are clues to diagnose malignant hypertension-induced TMA. Patients with malignant hypertension respond well to antihypertensive agents and have favorable nonrenal outcomes.
Article
Full-text available
Two cases of malignant hypertension presenting with acute kidney injury, thrombocytopenia and hemolytic anemia are presented. In both patients a prolonged duration of renal replacement therapy was required. The plasma levels of ADAMTS13 enzyme were not helpful in delineating the precise pathogenesis in both cases, as the decrements were not severe. We discuss the clinic-pathologic correlation of the biopsy findings and persistence of AKI.
Article
Full-text available
BACKGROUND Dietary administration of 0.30% indole-3-carbinol (I3C) to Cyp1a1-Ren2 transgenic rats (TGRs) generates angiotensin II (ANG II)-dependent malignant hypertension (HTN) and increased renal vascular resistance. However, TGRs with HTN maintain a normal or slightly reduced glomerular filtration rate. We tested the hypothesis that maintenance of renal function in hypertensive Cyp1a1-Ren2 TGRs is due to preservation of the intrarenal nitric oxide (NO) and antioxidant systems.METHODS Kidney cortex, kidney medulla, aortic endothelial (e) and neuronal (n) nitric oxide synthase (NOS), superoxide dismutases (SODs), and p22phox (nicotinamide adenine dinucleotide phosphate-oxidase subunit) protein abundances were measured along with kidney cortex total antioxidant capacity (TAC) and NOx. TGRs were fed a normal diet that contained 0.3% I3C or 0.3% I3C + candesartan (AT1 receptor antagonist; 25mg/L in drinking water) (n = 5-6 per group) for 10 days.RESULTSBlood pressure increased and body weight decreased in I3C-induced TGRs, while candesartan blunted these responses. Abundances of NOS, SOD, and p22phox as well as TAC were maintained in the kidney cortex of I3C-induced TGRs with and without candesartan, while kidney cortex NOx production increased in both groups. Kidney medulla eNOS and extracellular (EC) SOD decreased and nNOS were unchanged in both groups of I3C-induced TGRs. In addition, a compensatory increase occurred in kidney medulla Mn SOD in I3C-induced TGRs + candesartan. Aortic eNOS and nNOS∝ fell and p22phox and Mn SOD increased in hypertensive I3C-induced TGRs; all changes were reversed with candesartan.CONCLUSIONS The preservation of renal cortical NO and antioxidant capacity is associated with preserved renal function in Cyp1a1-Ren2 TGRs with ANG II-dependent malignant HTN. © American Journal of Hypertension, Ltd 2013. All rights reserved.
Article
Full-text available
Background: Sodium nitroprusside (SNP) and labetalol are recommended for the immediate treatment of malignant hypertension. Both are intravenous agents but have different effects on systemic hemodynamics, and may have differential effects on pulse-wave reflection and pulse-pressure amplification, with consequences for peripheral versus central blood pressures (BPs). Methods: We conducted a nonrandomized, open-label study of 8 patients treated with sodium nitroprusside (mean age (±SD), 44±14 years; 6 males; diastolic/systolic BP, 225±22/135±8mm Hg) and 6 patients treated with intravenous labetalol (mean age, 39±15 years; 4 males; systolic/diastolic BP, 232±22/138±17mm Hg) before and after treatment for malignant hypertension, aiming at a 25% reduction in mean arterial pressure. We measured peripheral pressures with an intra-arterial catheter in the radial artery and derived central pressures with a generalized transfer filter. Results: Mean arterial pressure was similarly reduced with sodium nitroprusside and labetalol (by 27% and 30%, respectively; P = 0.76). There was a nonsignificantly greater reduction in peripheral systolic blood pressure (SBP) with labetalol than with sodium nitroprusside (29±11% vs. 18±7%, P = 0.08). The decline in peripheral diastolic blood pressure (DBP) with the two agents was comparable, whereas the reduction in peripheral pulse pressure was 8±16% with SNP and 33±17% with labetalol (P = 0.01). The decline in reflection magnitude was greater with SNP than with labetalol. There were no significant differences in the reduction of central BP with SNP and labetalol. The amplification of PP increased with SNP but did not change with labetalol. Conclusions: We found no difference in central SBP or PP in subjects treated with SNP and labetalol, but labetalol produced a greater reduction in peripheral SBP and PP in the immediate treatment of malignant hypertension.
Article
The effects of the angiotensin type 1 (AT1) receptor antagonist, eprosartan, were studied in a model of severe, chronic hypertension. Treatment of male spontaneously hypertensive stroke prone rats (SHR-SP) fed a high-fat, high-salt diet with eprosartan (60 mg/kg/day i.p.) for 12 weeks resulted in a lowering of blood pressure (250 ± 9 versus 284 ± 8 mm Hg), renal expression of transforming growth factor-β mRNA (1.5 ± 0.2 versus 5.4 ± 1.4) and the matrix components: plasminogen activator inhibitor-1 (5.2 ± 1.4 versus 31.4 ± 10.7), fibronectin (2.2 ± 0.6 versus 8.2 ± 2.2), collagen I-α1 (5.6 ± 2.0 versus 23.8 ± 7.3), and collagen III (2.7 ± 0.9 versus 7.6 ± 2.1). Data were corrected for rpL32 mRNA expression and expressed relative to Wistar Kyoto (WKY) rats [=1.0]. Expression of fibronectin protein was also lowered by eprosartan (0.8 ± 0.1 versus 1.9 ± 0.5), relative to WKY rats. Eprosartan provided significant renoprotection to SHR-SP rats as measured by decreased proteinuria (22 ± 2 versus 127 ± 13 mg/day) and histological evidence of active renal damage (5 ± 2 versus 195 ± 6) and renal fibrosis (5.9 ± 0.7 versus 16.4 ± 1.9) in vehicle- versus eprosartan-treated rats, respectively. Our results demonstrated that AT1 receptor blockade with eprosartan can reduce blood pressure and preserve renal structure and function in this model of severe, chronic hypertension. These effects were accompanied by a decreased renal expression of transforming growth factor-β1, plasminogen activator inhibitor-1, and several other extracellular matrix proteins compared with vehicle-treated SHR-SP.
Article
Malignant hypertension and hypertensive encephalopathy are life-threating manifestations of hypertension. These syndromes primarily occur in patients with a history of poorly controlled hypertension. The purpose of this study was to investigate national trends in hospital admissions for malignant hypertension, hypertensive encephalopathy, and essential hypertension. This was a retrospective cohort study that used the Nationwide Inpatient Sample. We identified all hospitalizations between 2000 and 2011, during which a primary diagnosis of malignant hypertension (ICD 9 code: 401.0), hypertensive encephalopathy (ICD 9 code: 437.2), or essential hypertension (ICD 9 code: 401.9) was recorded. Time series models were estimated for malignant hypertension, hypertensive encephalopathy, essential hypertension and also for the combined series. A piecewise linear regression analyses was performed to investigate whether there were changes in the trends of these series. In addition, we also compared the characteristics of patients with these diagnoses. The estimated number of admissions for both malignant hypertension and hypertensive encephalopathy increased dramatically after 2007, whereas discharges for essential hypertension fell, and there was no change in trend for the combined series. Costs rose substantially for patients with these diagnoses after 2007, but mortality significantly fell for malignant hypertension and mortality for hypertensive encephalopathy did not change. The dramatic increase in the number of hospital admissions for hypertensive encephalopathy and malignant hypertension should have resulted in dramatic increases in morbidity, but it did not. The change is most likely related to changes in coding related to diagnostic-related groups that occurred in 2007. © 2015 American Heart Association, Inc.
Article
The survival of patients with malignant hypertension (MHT) has considerably improved over the past decades. Data regarding the excess risk of mortality and the contribution of conventional cardiovascular risk factors are lacking. The authors retrospectively assessed cardiovascular risk factors and all-cause mortality in 120 patients with a history of MHT and compared them with 120 normotensive and 120 hypertensive age-, sex-, and ethnicity-matched controls. Total cholesterol, low-density lipoprotein cholesterol, and body mass index were lower in MHT patients compared with hypertensive controls, whereas blood pressure, high-density lipoprotein cholesterol, and smoking habit were similar. Median estimated glomerular filtration rate was lower in MHT patients compared with normotensive and hypertensive controls (both P<.01). The annual incidence of all-cause mortality per 100 patient-years was higher in MHT patients (2.6) compared with normotensive (0.2) and hypertensive (0.5) controls (both P<.01). Mortality of patients with a history of MHT remains high compared with normotensive and hypertensive controls. Patients with MHT had a more favorable cardiovascular risk profile compared with hypertensive controls but a higher prevalence of renal insufficiency.
Article
The pathogenic role of T cells in hypertension has been documented well in recent animal studies. However, the existence of T-cell-driven inflammation in human hypertension has not been confirmed. Therefore, we undertook immunologic characterization of T cells from patients with hypertension and measured circulating levels of C-X-C chemokine receptor type 3 chemokines, which are well-known tissue-homing chemokines for T cells. We analyzed immunologic markers on T cells from patients with hypertension by multicolor flow cytometry. We then measured circulating levels of the C-X-C chemokine receptor type 3 chemokines, monokine induced by γ interferon (IFN), IFN γ-induced protein 10, and IFN-inducible T-cell α chemoattractant, in patients with hypertension and in age- and sex-matched control subjects by the cytometric bead array method. In addition, we examined histological features of IFN-inducible T-cell α chemoattractant expression from renal biopsy specimens of patients with hypertensive nephrosclerosis and control subjects. The total T-cell population from patients with hypertension showed an increased fraction of immunosenescent, proinflammatory, cytotoxic CD8(+) T cells. Circulating levels of C-X-C chemokine receptor type 3 chemokines were significantly higher in patients with hypertension than in control subjects. Furthermore, immunohistochemical staining revealed increased expression of the T-cell chemokine, IFN-inducible T-cell α chemoattractant, in the proximal and distal tubules of patients with hypertensive nephrosclerosis. Immunosenescent CD8(+) T cells and C-X-C chemokine receptor type 3 chemokines are increased in human hypertension, suggesting a role for T-cell-driven inflammation in hypertension. A more detailed characterization of CD8(+) T cells may offer new opportunities for the prevention and treatment of human hypertension.
Article
The development of drugs capable of modulatingangiogenesis has created a significant improvement inthe survival associated with metastatic tumors. Thisgroup of drugs includes anti-vascular endothelialgrowth factor (VEGF) monoclonal antibodies (bev-acizumab), intracellular VEGF receptor inhibitors (so-rafenib, sutinib, pazopanib), and the small VEGFinactivating molecules (VEGF trap).