Recent Advances in Pulmonary Hypertension Pathophysiology and Treatment of High-Altitude Pulmonary Vascular Disease Martin R. Wilkins, MD; Hossein-Ardeschir Ghofrani, MD; Norbert Weissmann, PhD; Almaz Aldashev, PhD; Lan Zhao, MD, PhD

I

with ventilation. A fall in alveolar Po2 is the main stimulus for HPV, but a reduction in mixed venous and bronchial arterial Po2 may also contribute.9 Ventilation of intact lungs with a hypoxic gaseous mixture (eg, fraction of inspired oxygen=0.10) leads to acute pulmonary vasoconstriction throughout the pulmonary vascular bed, including nonmuscular arterioles, capillaries, and veins, but is most pronounced in small pulmonary arterioles.10–13 That said, HPV is not distributed evenly throughout the lung and lung perfusion is inhomogeneous during hypoxia.14 HPV has at least 2 phases (Figure 1A). An initial constrictor response that starts within seconds and reaches a maximum within minutes is followed by a sustained phase, which develops after 30 to 120 minutes.9 A transient phase of vasodilation may be observed linking the two, and a third phase of even more pronounced vasoconstriction can occur after 120 minutes. The phases are regulated, at least in part, by different signaling pathways.15 In vivo, factors such as neurohumoral mediators, red blood cells, and lung innervation may influence the response.16 Alveolar capillaries have been proposed as a site for oxygen sensing with propagation of the hypoxic signal by endothelial membrane depolarization to upstream arterioles in a connexin 40-dependent manner.17 However, the recapitulation of contraction to hypoxia in cultured pulmonary artery smooth muscle cells, the effector cells, indicates that an oxygen-sensing mechanism is intrinsic to these cells.18 Both mitochondria and nicotinamide adenine dinucleotide (phosphate) oxidases have been suggested as oxygen sensors. A change in the levels of reactive oxygen species is thought to be important, but there is a lack of agreement regarding whether the signal is an increase or decrease in reactive oxygen species (Figure 2).19– 21 Differences in techniques used contribute to the different observations, but the spatial distribution of reactive oxygen species signaling may also be significant.22 The second phase of HPV is influenced by endothelial cell function. The endothelium releases a variety of vasoactive mediators, such as endothelin 1, prostacyclin, and nitric oxide (NO; Figure 3),9,16 and their production is perturbed by hypoxia. For example, oxygen is a substrate for NO

t is estimated that >140 million people live above 2500 m in various regions of the world.1 There are many challenges to living at high altitude, but chronic exposure to alveolar hypoxia is prominent among them. Inspired Po2 falls from ≈150 mm Hg at sea level to ≈100 mm Hg at 3000 m and 43 mm Hg on the summit of Everest (8400 m).2,3 The body responds by hyperventilating, increasing resting heart rate, and stimulating red cell production in an attempt to maintain the oxygen content of arterial blood at or above sea level values.2 However, hypoxic pulmonary vasoconstriction (HPV) and vascular remodeling, together with increased erythropoiesis, place an increased pressure load on the right ventricle (RV). How well healthy humans adapt to hypoxia depends on their rate of ascent to altitude, the severity and duration of their exposure, and their genetic background.

Pathophysiology of Acclimatization to Hypoxia Pulmonary Vascular Response to Hypoxia For most mammals, including humans, a rise in pulmonary artery pressure (PAP) is an early and inevitable consequence of ascent to high altitude. Resting mean PAP increases along a parabolic curve from 15 mm Hg at 2000 m to ≈30 mm Hg at 4500 m.4 The exceptions and interindividual variation in the magnitude of response offer a natural experiment that might provide insight into fundamental underlying mechanisms (vide infra). The initial rise in PAP on exposure to hypoxia is attributed to HPV. With chronic hypoxia, other mechanisms that likely drive vascular remodeling soon contribute to the elevated pressure (Figure 1A). After 2 or 3 weeks of hypoxia, there is little response to rebreathing 100% oxygen, indicating a structural resistance to pulmonary blood flow rather than one based solely on increased vasomotor tone.6 A fall in PAP on re-exposure to a normal oxygen environment is evident in rats monitored by telemetry over days after removal from a hypoxic chamber7 (Figure 1B) and is also documented in humans.4,8

Pulmonary Arteriolar Vasoconstriction Dominates the Acute Pulmonary Vascular Response to Hypoxia Oxygen tension is a major regulator of pulmonary vascular tone and a physiological mechanism for matching perfusion

From Experimental Medicine, Imperial College London, Hammersmith Hospital, United Kingdom (M.R.W., H.-A.G., L.Z.); Excellence Cluster Cardio-Pulmonary System, Universities of Giessen, Germany (M.R.W., H.-A.G., N.W., L.Z.); University of Giessen Marburg Lung Center, Justus-LiebigUniversity, Germany (M.R.W., H.-A.G., N.W., L.Z.); Kerckhoff Clinic, Bad Nauheim, Germany (H.-A.G.); Institute of Molecular Biology and Medicine, Bishkek, Kyrgyzstan (A.A.). Correspondence to Martin R. Wilkins, MD, NIHR Imperial Clinical Research Facility, Imperial College London, Hammersmith Hospital, Du Cane Road, London W12 0NN, United Kingdom. E-mail [email protected] (Circulation. 2015;131:582-590. DOI: 10.1161/CIRCULATIONAHA.114.006977.) © 2015 American Heart Association, Inc. Circulation is available at http://circ.ahajournals.org

DOI: 10.1161/CIRCULATIONAHA.114.006977

582

Wilkins et al   High-Altitude Pulmonary Vascular Response   583

ΔPAP (mmHg)

A

5 4

exchange of perfusate

3 2 1

0 -3

B

normoxic ventilation (21% O2) hypoxic ventilation (3% O2)

normoxic control

-10 0

30

60

90

p < 0.05

120

Time (min)

150

180

hypoxia

70

PAP (mmHg)

60 50

normoxia

normoxia

40

30 20 10 0

0

5

10

15

20

25

30

Time (days)

35

40

45

Figure 1. Contribution of hypoxic pulmonary vasoconstriction (HPV) and vascular remodeling to the rise in pulmonary artery pressure (PAP) in chronic hypoxia. A, The initial rise in PAP in hypoxia is driven by HPV. The pressor response to hypoxia does not return to baseline on return to normoxia in isolated perfused rabbit lungs, even if the perfusate is replaced to remove hypoxia-stimulated circulating vasoactive factors. Modified from Weissmann et al.5 B, Elevated PAP in a telemetered rat takes days to return to baseline after removal from 2 weeks in a hypoxic chamber.

synthases, and NO bioavailability is reduced by hypoxia.23 In addition, sustained HPV has been shown to depend on glucose level.24 In addition to changes in intracellular Ca2+ levels, changes in pulmonary artery smooth muscle cell myofilament sensitivity to Ca2+ arising from inhibition of myosin light chain phosphatase via RhoA/Rho kinase or protein kinase C or a decreased activation of myosin light chain phosphatase by decreased NO signaling9 also contribute to sustained HPV. A recent observation that still has to be reconciled with the effects of global hypoxia-inducible factor (HIF)-1α haploinsufficiency on the pulmonary vasculature is that HIF-1α levels in pulmonary artery smooth muscle cells maintain low pulmonary vascular tone by decreasing myosin light chain phosphorylation.25 Rho kinase inhibitors are very effective at preventing and attenuating HPV, underlying the importance of the RhoA/Rho kinase signaling pathway in regulating pulmonary vascular tone. Acute administration of a Rho kinase inhibitor significantly reduces pulmonary vascular resistance in chronically hypoxic rats,26 advancing the argument that vasoconstriction is an important pathophysiological mechanism in high-altitude pulmonary hypertension (HAPH), perhaps as important or more important than vascular remodeling.27

Vascular Remodeling Contributes to and Sustains the Chronic Pulmonary Vascular Response to Hypoxia Chronic global alveolar hypoxia is accompanied by structural remodeling of pulmonary vessels. This has been described in a number of species, including rat,28 cow,29 and humans,30

although some species seem resistant.31 All of the layers of the vascular wall, including fibroblasts, are involved in the remodeling (Figure 4),32,33 but the hallmark of the vascular response to chronic hypoxia is increased muscularization of distal vessels with extension of muscle into previously unmuscularized arterioles.28,30 Hypoxia leads to changes in endothelial cell membrane properties that compromise barrier function, resulting in an influx of plasma proteins that may activate vascular wall proteases.34 In addition, mechanical stress, blood-borne and locally produced factors, and the recruitment of circulating cells act collectively to drive the vascular remodeling of small and large pulmonary vessels, with increasing recognition of the role of inflammation (Figure 4).35–37 Rapid expansion of the adventitial vasa vasorum serves to facilitate the arrival of these cells.33 HIFs and nuclear factor-κB are key transcriptional regulators of the proliferative and inflammatory responses to hypoxia. Pulmonary hypertension in hypoxic mice haploinsufficient for HIF-1α (Hif1α+/-) or HIF-2α (Hif2α+/-) is attenuated.38,39 Conversely, gain-of-function mutations in HIF2α are associated with the development of pulmonary hypertension in mice and humans.40–42 Although in part caused by and adaptive to the increase in hemodynamic stress, the vascular remodeling contributes to and sustains the elevated PAP. Indeed, the structural changes take a considerable time to resolve on return to a sea-level oxygen environment and may persist in some form.43 The extent to which the structural changes in pulmonary resistance vessels infringe on the lumen and contribute a physical obstruction to blood flow (ie, the argument being that vascular growth is outward rather than inward) and the extent of vascular rarefaction in response to chronic hypoxia are unclear.44 These may vary between species. Mechanisms other than narrowing of the vessel lumen are relevant to this discussion, specifically the contribution of changes in vascular compliance.45 Changes in the stiffness of proximal vessels leads to changes in the propagation of highenergy pulsatile waves. These are transmitted to the microcirculation, where they might perpetuate or potentially even cause the microcirculatory changes, as well as contribute to the burden on the RV.46

Cardiac Response to Chronic Hypoxia Studies of healthy subjects exposed to hypoxia report an increase in resting heart rate and an initial increase in cardiac output in an attempt to maintain oxygen delivery to tissues.47 After 2 to 3 days of hypoxia, stroke volume falls. Heart rate remains elevated, and so cardiac output remains at or just below sea level. Oxygen delivery to tissues is maintained by increased oxygen extraction. The roles of activation of the sympathetic nervous system, hypovolemia (from hyperventilation and increased diuresis), hypocapnia (from hyperventilation), and myocardial contractility in this response are difficult to discern.47 On the whole, myocardial contractility is preserved, although reversible reductions in cardiac mass and myocardial phosphocreatine/ ATP have been documented in healthy volunteers after a 17-day trek to 5300 m.48 Right heart failure is a risk in some previously healthy individuals, precipitated by more extreme pulmonary

584  Circulation  February 10, 2015

Figure 2. Signaling mechanisms underlying acute hypoxic pulmonary vasoconstriction (HPV). Pathways activated by hypoxia are depicted in blue; those inhibited by hypoxia are depicted in red. Both mitochondria and nicotinamide adenine dinucleotide (phosphate) oxidases have been suggested as oxygen sensors. A reduction in the cytosolic redox state could inhibit voltage-dependent potassium channels, subsequent membrane depolarization of PASMCs, opening of l-type calcium channels and Ca2+ influx.20 By contrast, increased cytosolic ROS levels can result in Ca2+ release from the SR, possibly through the oxidation of cysteine residues in RyRs and the opening of IP3-gated calcium stores.19 Increased ROS could also provoke an influx of extracellular Ca2+ or Na+ through transient receptor potential channels (TRPC6).21 In this scenario, the increase of acute hypoxia-induced ROS triggers an accumulation of DAG, resulting from the activation of phospholipase C or phospholipase D or inhibition of DAG-degrading DAG kinases. Another proposal assumes that acute hypoxia leads to inhibition of the respiratory chain and a subtle decrease in ATP production, which does not affect energy state, but rather acts as a mediator and alters the cellular AMP/ATP ratio. An increase in the AMP/ATP ratio activates AMPK, followed by an increase in cADPR that triggers the release of [Ca2+]i through RyR of SR.9 The level of ROS could be relevant through ROS-dependent alteration of function of AMPK and cADPR. AMPK indicates AMP–activated protein kinase; Ca2+, calcium; cADPR, cyclic ADP–ribose; DAG, diacylglycerol; DGK, DAG kinases; Em, membrane potential; GSH, glutathione; GSSG, glutathione disulfide; IP3, inositol trisphosphate; IP3R, inositol trisphosphate receptor; K+, potassium; Kv, voltage-dependent potassium channels; Na+, sodium; NAD+, nicotinamide adenine dinucleotide; NADH, reduced form of NAD; NAD(P)H, nicotinamide adenine dinucleotide (phosphate) oxidase; O2, oxygen; PASMC, pulmonary arterial smooth muscle cell; ROS, reactive oxygen species; RyR, ryanodine receptors; SR, sarcoplasmic reticulum; TRPC6, transient receptor potential cation channel 6; and VOCC, voltage–dependent calcium channels (includes l-type calcium channels).

vascular responses to hypoxia, and also in the context of chronic mountain sickness (CMS), from pronounced erythrocytosis and fluid retention (from relatively higher Pco2).

Exercise Exercise capacity is reduced at altitude, even after acclimatization, but the contribution of pulmonary hypertension is controversial.49,50 PAP increases more sharply with the increase in cardiac output on exercise at altitude than at sea level.51 This augmented rise in PAP with exercise can persist for some time in acclimatized highlanders on descent to sea level, most likely reflecting structural remodeling of the pulmonary vasculature with chronic exposure.5,8 The increase in PAP may impair gas exchange from interstitial and alveolar edema and reduce maximal cardiac output, leading to a reduction in oxygen transport to exercising muscles.49 However, definitive data

from direct measurements of RV function at altitude are few, and not all are convinced that the improvement in exercise capacity at altitude reported with some pulmonary vasodilators is attributed to a reduction in RV afterload.50

Erythropoietic Response to Hypoxia Exposure to hypoxia leads to changes in blood-O2 affinity and stimulates red cell production in an attempt to improve tissue oxygenation. Increased red cell 2,3 diphosphoglycerate levels have an allosteric effect on hemoglobin, reducing its affinity for O2 and facilitating its release to tissues, although this is at the expense of impairing O2 capture as blood passes through the lungs. Increasing red cell mass also has its downside, as it increases blood viscosity. Little attention has been paid to the contribution of increased blood viscosity to PAP, because the increase in PAP precedes the rise in hemoglobin, and patients

Wilkins et al   High-Altitude Pulmonary Vascular Response   585 at altitude with high-altitude pulmonary edema (HAPE) or over weeks, months, and years with right heart failure secondary to severe pulmonary hypertension or excessive erythrocytosis.

High-Altitude Pulmonary Edema

Figure 3. Signaling mechanisms underlying sustained hypoxic pulmonary vasoconstriction (HPV). The endothelium releases a variety of vasoactive mediators, such as endothelin 1, prostacyclin, and nitric oxide (NO),9,16 and their production is perturbed by hypoxia. In addition to changes in intracellular Ca2+ levels, changes in PASMC myofilament sensitivity to Ca2+, arising from inhibition of MLCP via RhoA/Rho kinase or protein kinase C (PKC), or a decreased activation of MLCP by decreased NO signaling,9 also contributes to sustained HVP. ADMA indicates asymmetrical dimethylarginine; cGMP, cyclic guanosine monophosphate; DDAH, dimethylarginine dimethylaminohydrolase; DMA, dimethylamine; ET-1, endothelin 1; EC, endothelial cell; MLC20, regulatory myosin light chain; MLCP, myosin light chain phosphatase; NO, nitric oxide; NOS, nitric oxide synthase; O2, oxygen; PASMC, pulmonary arterial smooth muscle cell; PGI2, prostacyclin; Rho, Ras homolog gene family; ROS, reactive oxygen species; and sGC, soluble guanylyl cyclase. Phosphorylated proteins are indicated by a white “P” in a blue circle.

with polycythemia at sea level do not have pulmonary hypertension. A recent re-evaluation of the effect of increased blood viscosity on PAP measurements at altitude suggests that some correction for hematocrit is important.52,53

Clinical Presentations From Maladaptation to Hypoxia Variation in Susceptibility to High-Altitude Pulmonary Vascular Disease The rise in PAP in chronic hypoxia is generally modest, certainly compared with that seen in idiopathic pulmonary arterial hypertension, and is compatible with a normal life at high altitude. However, variation in the pulmonary vascular response to hypoxia is well recognized, both between and within species,16,31,54,55 and in humans the magnitude of HPV can vary ≈5-fold among individuals.16,55 Extreme responders are at highest risk of presenting acutely on arrival

The pathognomonic clinical feature is breathlessness accompanied by cough, initially dry but later productive of white and then pink frothy sputum.3,56 Tachycardia, mild pyrexia, and sometimes cyanosis are also evident. The chest radiograph shows pulmonary edema. The condition develops in susceptible individuals within the first 2 to 4 days of arrival at altitudes above 2500 m. It is precipitated by rapid ascent and exercise and is potentially lethal if not treated. The incidence is variously recorded, depending on the subject population, rapidity of ascent, and final altitude; everyone is at risk of HAPE if they ascend fast and high enough. It is caused by exaggerated and uneven HPV leading to high capillary pressures in regions of the lung and an exudation that might invoke a secondary inflammatory response.57–59 Susceptible individuals exhibit a marked rise in PAP on exposure to hypoxia mediated by pulmonary arteriolar vasoconstriction and a greater rise in PAP on exercise in a normal oxygen environment, indicating a hyperactive pulmonary circulation.56 The emphasis in management is on prevention. Travelers should manage their rate of ascent to 300 to 500 m per day along with a day of rest every 3 to 4 days when traveling above 3000 m.56 Pharmacologic measures that have been evaluated and demonstrated efficacy in reducing the incidence of HAPE include slow-release nifedipine (30 mg BID),60 the phosphodiesterase type 5 inhibitor tadalafil (10 mg BID), and dexamethasone (8 mg BID).61 The β2-agonist salmeterol (125 μg BID) by inhalation appears to be less effective.62 When treatment is required, consideration should be given to descent to a lower altitude coupled with supplemental oxygen (2–4 L/min) where possible.56,63 Nifedine is the standard treatment. A phosphodiesterase type 5 inhibitor may be helpful but has not been formally trialed. There is no role for diuretics. A fully conscious person with mild-to-moderate HAPE may be managed at altitude if the appropriate expertise and facilities are available.

HAPH and Heart Failure By convention, the definition of HAPH is a resting mean PAP >30 mm Hg.64 This needs revisiting and reconciling with international guidelines for the definition of pulmonary hypertension, which is a resting mean PAP >25 mm Hg.53,65 Data on the prevalence of HAPH are sparse. Prevalence will vary according to altitude and ethnic background, but some 14% of Kyrgyz highlanders have been found to have ECG evidence of right ventricular hypertrophy.66 A much smaller percentage progress to and present with heart failure. The RV generally copes well with a pressure load, and there is doubt as to whether pressure load itself is sufficient to cause heart failure, suggesting that other factors, such as myocardial hypoxia and neurohumoral factors, are important.67 Nonetheless, pulmonary hypertension progressing to fatal right heart failure, recognized as infantile subacute mountain

586  Circulation  February 10, 2015

Figure 4. Mechanisms of vascular remodeling in chronic hypoxia. The vascular remodeling in response to chronic hypoxia involves all 3 layers of the vascular wall–intima, media, and adventitia.32 Endothelial cell dysfunction and intimal proliferation are evident and prominent in some species, such as the rat, exposed to moderately severe hypoxia.34 However, the hallmark of HAPH is increased muscularization of distal vessels with extension of muscle into previously unmuscularized arterioles.28,30 There is thickening of the media of large and medium pulmonary arteries, which, in large mammals (including humans), is attributable mainly to the proliferation of smooth muscle cells, along with increased collagen and elastin deposition.32 The medial layer of proximal vessels contains a heterogeneous population of smooth muscle cells, which includes a reservoir of cells that can divide when exposed to hypoxia and could contribute to vascular hyperplasia.32 Medial smooth muscle cells from distal resistance pulmonary arterioles have a lower capacity for proliferation. Remodeling in the distal vasculature may depend on the presence of cells from other vascular compartments, including circulating inflammatory cells and peripheral blood hematopoietic cells, that might stimulate proliferation or transdifferentiate into smooth muscle-like cells. Increasing recognition is given to an adventitial reaction mediated by the fibroblast in response to hypoxia and vascular wall stress.33 In addition, an inflammatory cell infiltrate, composed of monocytes, dendritic cells, and T cells, is evident.35 Evidence for epigenetic regulation of pulmonary vascular remodeling comes from experiments with HDACs. Class I HDAC inhibitors markedly decreased cytokine/chemokine mRNA expression levels in these cells proliferating adventitial fibroblasts and their ability to induce monocyte migration and inflammation.36 This in part may be attributed to modulation of microRNA-124 expression.37 ECAM indicates endothelial cell adhesion molecule; FGF, fibroblast growth factor; HDAC, histone deacetylase; GM-CSF, granulocyte macrophage colony stimulating factor; HIF, hypoxia-inducible factor; ICAM, intercellular adhesion molecule; IL-1β, interleukin 6; IL-6, interleukin 6; NO-sGC-cGMP, nitric oxide-soluble guanylate cyclase-cyclic GMP; PDGF, plateletderived growth factor; PGI2, prostacyclin; ROS, reactive oxygen species; TRPC6, transient receptor potential cation channel 6; and VCAM, vascular cell adhesion molecule.

sickness, has been described in infants of Chinese Han origin who are born at low altitude and taken to high altitudes.68 They develop heart failure within a few weeks or months and the pathology reveals extreme medial hypertrophy of the small pulmonary arteries accompanied by hypertrophy and dilatation of the RV. Heart failure has also been described in Indian soldiers posted at the high-altitude borders in China69 and occasionally in previously healthy travelers,70 and HAPH is thought to be the major factor.71 Descent to lower altitude is life saving for severe cases of heart failure. There are a number of potential pharmacologic treatments for managing less severe disease, but few have been formally trialed in HAPH. Phosphodiesterase type 5 inhibitors appear effective at reducing pulmonary vascular resistance,72 and acetazolamide73 and the Rho-kinase inhibitor, fasudil74, are promising. The benefits of endothelin receptor antagonists are less clear.75,76

Chronic Mountain Sickness The defining feature of chronic mountain sickness (CMS) is excessive erythrocytosis accompanied by neurologic symptoms, such as headache, dizziness, and fatigue.64 By consensus, the hemoglobin should exceed ≥21g/dL in men and ≥19 g/dL in women. Hypoventilation leading to hypoxemia may stimulate red cell production,4 but an alternative possibility is that polycythemia is the primary abnormality, which, by reducing Pco2 drive, leads to hypoventilation. Pulmonary hypertension may accompany the polycythemia but is not a prerequisite. Descent to low altitudes is the best treatment but may not be acceptable to patients for personal reasons. An alternative to phlebotomy is acetazolamide. Acetazolamide (250 mg daily) has been shown to reduce hematocrit, increase Pao2 and oxygen saturation, and decrease in Paco2 in CMS, most likely via metabolic acidosis stimulating ventilation.73 Pulmonary

Wilkins et al   High-Altitude Pulmonary Vascular Response   587 vascular resistance was also reduced and cardiac output increased without a change in pulmonary pressure.

Insights From Humans Adapted to Hypoxia Variation in Susceptibility to High-Altitude Pulmonary Vascular Disease The variation in pulmonary vascular response has a strong genetic basis, which provides the substrate for environmental selection pressures and adaptation to high-altitude living. Tibetans appear less susceptible than recent migrants to HAPH77,78 and CMS,79 most likely the result of living above 3000 m for thousands of years. Data are few, but PAP measurements in ethnic Tibetans living over 3600 m are in the range typical of healthy adults at sea level,77,78 and postmortem studies show little vascular remodeling.78,80 A blunted pulmonary vascular pressor response to acute and sustained hypoxia is retained by Tibetans at sea level.81

Recent Genomic Studies in Humans A number of attempts have been made to understand adaptation to high-altitude life based on differences in candidate pathways, such as the ability of Tibetans to preserve NO production at altitude82 and candidate genes,66,83 but this approach is selective and the data come from small subject numbers. The advent of high-throughput genome sequencing has enabled a less-biased strategy for investigating gene associations. The priority to date has been to compare the genetic architecture of people living at high altitude with that of lowlanders or recent migrants (genome-wide selection scans) rather than to compare well-phenotyped populations with and without pulmonary vascular disease at altitude (genome-wide association studies). Seven genome-wide selection scans of Tibetan DNA have been reported. All have shown evidence of natural selection for noncoding variants in and around 2 HIF pathway genes, EPAS1 (HIF-2α) and EGNL1 (HIF prolyl 4-hydroxylase 2).84–90 Key to the interpretation of genetic data is robust phenotyping. Tibetans average ≥1 g/dL and as much as 3.5 g/ dL (ie, ≈10% to 20%) lower hemoglobin concentration compared with acclimatized lowlanders. At first sight, a paradox, a lower red cell mass by reducing blood viscosity may be an important factor enabling the cardiopulmonary circulation to adapt to high-altitude life. Significantly, the polymorphisms in EPAS1 and EGLN1 in Tibetans correlate with hemoglobin concentration.84,86–88,90 A high-frequency missense mutation has recently been identified in EGLN1 that encodes a variant prolyl 4-hydroxylase 2 with increased hydroxylase activity under hypoxic conditions that would contribute to this adaptive response.91 A genome study in Andeans has found evidence of positive selection for EGLN1 but not EPAS1.92 Neither were candidates in reported studies in Ethiopian highlanders.93–95 Moreover, Andeans exhibit a robust erythropoietic response to altitude and polymorphisms identified in EGLN1 in Andeans, albeit different from those in Tibetans, did not associate with hemoglobin level. Analysis of other quantitative traits, such as resting ventilation, hypoxic ventilator response, and oxygen saturation, also show differences between Tibetans and

other Asian and European populations studied at the same altitude.96 It is likely that the Andean and Tibetan populations have developed different genetic adaptations to high-altitude hypoxia, although pathways may converge. In the case of Tibetans, one source of adaptation is likely to be attributed to the introduction of genetic variants from archaic Denisovanlike individuals into the ancestral Tibetan gene pool.97 Aside from HIF, genes encoding downstream components of the HIF pathway remain a priori candidates for natural selection to hypoxia. A recent study of single nucleotide polymorphisms in 5 Ethiopian populations at altitude suggest positive selection for BHLHE41, a gene transcriptionally regulated by HIF-1α and with a major regulatory role in the same hypoxia-sensing pathway described in Tibetans, indicative of convergent evolution.95 Other pathways may emerge from unbiased genome-wide studies in larger population cohorts. Evidence for adaptation outside the HIF family comes from a study of Eurasians exposed to mild-to-moderate hypoxia, where the strongest adaptive signal came from the μ-opioid receptor-encoding gene (OPRM1, 2.54_10_9).98 Whole-genome sequencing of Andean highlanders, 10 with and 10 without CMS, followed by expression studies in fibroblasts identified 2 genes, SENP1 and ANP32D, that exhibit a higher transcription response to hypoxia in CMS subjects.99 Downregulation of the orthologs of these genes in flies enhanced their survival rates in a hypoxic environment. SENP1 is known to regulate erythropoiesis and SENP1-/- mice die early because of anemia, lending biological plausibility to this gene as a candidate for a role in CMS. There is widely believed to be a genetic predisposition to HAPE, but to date only candidate genes have been examined with no consensus observations.100

Conclusions Humans can live a normal life at high altitudes given sufficient time to acclimatize. People who exhibit a marked pulmonary vascular or erythropoietic response to hypoxia identify themselves as at risk of heart failure. Studies in Tibetan people adapted genetically to high altitude highlight the importance of the HIF pathway in determining the magnitude of response, but other pathways may emerge from studies of Tibetan cohorts better phenotyped for pulmonary hemodynamics, as well as studies of other ethnic groups. Considerable progress has been made in understanding the pathology of HAPH, but few drugs studied in animal models have been formally trialed in humans.

Acknowledgments We thank Dr Oleg Pak for his assistance with the figures.

Sources of Funding Drs Wilkins and Zhao are funded by the British Heart Foundation. Drs Weissmann and Ghofrani are funded by the German Research Foundation, Excellence Cluster Cardio-Pulmonary System (EXC 147).

Disclosures None.

588  Circulation  February 10, 2015

References 1. Moore LG, Niermeyer S, Zamudio S. Human adaptation to high altitude: regional and life-cycle perspectives. Am J Phys Anthropol. 1998;Suppl 27:25–64. 2. Grocott MP, Martin DS, Levett DZ, McMorrow R, Windsor J, Montgomery HE; Caudwell Xtreme Everest Research Group. Arterial blood gases and oxygen content in climbers on Mount Everest. N Engl J Med. 2009;360:140–149. doi: 10.1056/NEJMoa0801581. 3. West JB. High-altitude medicine. Am J Respir Crit Care Med. 2012;186:1229–1237. doi: 10.1164/rccm.201207-1323CI. 4. Penaloza D, Arias-Stella J. The heart and pulmonary circulation at high altitudes: healthy highlanders and chronic mountain sickness. Circulation. 2007;115:1132–1146. doi: 10.1161/CIRCULATIONAHA.106.624544. 5. Weissmann N, Winterhalder S, Nollen M, Voswinckel R, Quanz K, Ghofrani HA, Schermuly RT, Seeger W, Grimminger F. NO and reactive oxygen species are involved in biphasic hypoxic vasoconstriction of isolated rabbit lungs. Cell Mol Physiol. 2001;280:L638–L645. 6. Groves BM, Reeves JT, Sutton JR, Wagner PD, Cymerman A, Malconian MK, Rock PB, Young PM, Houston CS. Operation Everest II: elevated high-altitude pulmonary resistance unresponsive to oxygen. J Appl Physiol (1985). 1987;63:521–530. 7. Sebkhi A, Strange JW, Phillips SC, Wharton J, Wilkins MR. Phosphodiesterase type 5 as a target for the treatment of hypoxiainduced pulmonary hypertension. Circulation. 2003;107:3230–3235. doi: 10.1161/01.CIR.0000074226.20466.B1. 8. Sime F, Peñaloza D, Ruiz L. Bradycardia, increased cardiac output, and reversal of pulmonary hypertension in altitude natives living at sea level. Br Heart J. 1971;33:647–657. 9. Sylvester JT, Shimoda LA, Aaronson PI, Ward JP. Hypoxic pulmonary vasoconstriction. Physiol Rev. 2012;92:367–520. doi: 10.1152/ physrev.00041.2010. 10. Euler Von US, Liljestrand G. Observations on the pulmonary arterial blood pressure of the cat. Acta Physiologica Scandinavica. 1946;12:301–320. 11. Motley HL, Cournand A. The influence of short periods of induced acute anoxia upon pulmonary artery pressures in man. Am J Physiol. 1947;150:315–320. 12. Nagasaka Y, Bhattacharya J, Nanjo S, Gropper MA, Staub NC. Micropuncture measurement of lung microvascular pressure profile during hypoxia in cats. Circ Res. 1984;54:90–95. 13. Schwenke DO, Pearson JT, Umetani K, Kangawa K, Shirai M. Imaging of the pulmonary circulation in the closed-chest rat using synchrotron radiation microangiography. J Appl Physiol (1985). 2007;102:787–793. doi: 10.1152/japplphysiol.00596.2006. 14. Dehnert C, Risse F, Ley S, Kuder TA, Buhmann R, Puderbach M, Menold E, Mereles D, Kauczor HU, Bärtsch P, Fink C. Magnetic resonance imaging of uneven pulmonary perfusion in hypoxia in humans. Am J Respir Crit Care Med. 2006;174:1132–1138. doi: 10.1164/rccm.200606-780OC. 15. Weissmann N, Dietrich A, Fuchs B, Kalwa H, Ay M, Dumitrascu R, Olschewski A, Storch U, Mederos y Schnitzler M, Ghofrani HA, Schermuly RT, Pinkenburg O, Seeger W, Grimminger F, Gudermann T. Classical transient receptor potential channel 6 (TRPC6) is essential for hypoxic pulmonary vasoconstriction and alveolar gas exchange. Proc Natl Acad Sci U S A. 2006;103:19093–19098. doi: 10.1073/pnas.0606728103. 16. Swenson ER. Hypoxic pulmonary vasoconstriction. High Alt Med Biol. 2013;14:101–110. doi: 10.1089/ham.2013.1010. 17. Wang L, Yin J, Nickles HT, Ranke H, Tabuchi A, Hoffmann J, Tabeling C, Barbosa-Sicard E, Chanson M, Kwak BR, Shin HS, Wu S, Isakson BE, Witzenrath M, de Wit C, Fleming I, Kuppe H, Kuebler WM. Hypoxic pulmonary vasoconstriction requires connexin 40-mediated endothelial signal conduction. J Clin Invest. 2012;122:4218–4230. doi: 10.1172/JCI59176. 18. Murray TR, Chen L, Marshall BE, Macarak EJ. Hypoxic contraction of cultured pulmonary vascular smooth muscle cells. Am J Respir Cell Mol Biol. 1990;3:457–465. doi: 10.1165/ajrcmb/3.5.457. 19. Sommer N, Dietrich A, Schermuly RT, Ghofrani HA, Gudermann T, Schulz R, Seeger W, Grimminger F, Weissmann N. Regulation of hypoxic pulmonary vasoconstriction: basic mechanisms. Eur Respir J. 2008;32:1639–1651. doi: 10.1183/09031936.00013908. 20. Weir EK, Archer SL. The role of redox changes in oxygen sensing. Respir Physiol Neurobiol. 2010;174:182–191. doi: 10.1016/j.resp.2010.08.015. 21. Dietrich A, Kalwa H, Fuchs B, Grimminger F, Weissmann N, Gudermann T. In vivo TRPC functions in the cardiopulmonary vasculature. Cell Calcium. 2007;42:233–244. doi: 10.1016/j.ceca.2007.02.009. 22. Waypa GB, Marks JD, Guzy RD, Mungai PT, Schriewer JM, Dokic D, Ball MK, Schumacker PT. Superoxide generated at mitochondrial complex III triggers acute responses to hypoxia in the pulmonary

circulation. Am J Respir Crit Care Med. 2013;187:424–432. doi: 10.1164/ rccm.201207-1294OC. 23. Cremona G, Higenbottam T, Takao M, Hall L, Bower EA. Exhaled nitric oxide in isolated pig lungs. J Appl Physiol (1985). 1995;78:59–63. 24. Wiener CM, Sylvester JT. Effects of glucose on hypoxic vasoconstriction in isolated ferret lungs. J Appl Physiol (1985). 1991;70:439–446. 25. Kim YM, Barnes EA, Alvira CM, Ying L, Reddy S, Cornfield DN. Hypoxia-inducible factor-1 in pulmonary artery smooth muscle cells lowers vascular tone by decreasing myosin light chain phosphorylation. Circ Res. 2013;112:1230–1233. doi: 10.1161/CIRCRESAHA.112.300646 26. Oka M, Fagan KA, Jones PL, McMurtry IF. Therapeutic potential of RhoA/Rho kinase inhibitors in pulmonary hypertension. Br J Pharmacol. 2008;155:444–454. doi: 10.1038/bjp.2008.239. 27. Stenmark KR, McMurtry IF. Vascular remodeling versus vasoconstriction in chronic hypoxic pulmonary hypertension: a time for reappraisal? Circ Res. 2005;97:95–98. doi: 10.1161/01.RES.00000175934.68087.29. 28. Hislop A, Reid L. New findings in pulmonary arteries of rats with hypoxiainduced pulmonary hypertension. Br J Exp Pathol. 1976;57:542–554. 29. Alexander AF, Jensen R. Pulmonary vascular pathology of high altitude-induced pulmonary hypertension in cattle. Am J Vet Res. 1963;24:1112–1122. 30. Arias-Stella J, Saldana M. The terminal portion of the pulmonary arterial tree in people native to high altitudes. Circulation. 1963;28:915–925. 31. Rhodes J. Comparative physiology of hypoxic pulmonary hyperten sion: historical clues from brisket disease. J Appl Physiol (1985). 2005;98:1092–1100. doi: 10.1152/japplphysiol.01017.2004. 32. Stenmark KR, Fagan KA, Frid MG. Hypoxia-induced pulmonary vascular remodeling: cellular and molecular mechanisms. Circ Res. 2006;99: 675–691. doi: 10.1161/01.RES.0000243584.45145.3f. 33. Stenmark KR, Yeager ME, El Kasmi KC, Nozik-Grayck E, Gerasimovskaya EV, Li M, Riddle SR, Frid MG. The adventitia: essential regulator of vascular wall structure and function. Annu Rev Physiol. 2013;75:23–47. doi: 10.1146/annurev-physiol-030212-183802. 34. Meyrick B, Reid L. Endothelial and subintimal changes in rat hilar pulmonary artery during recovery from hypoxia: a quantitative ultrastructural study. Lab Invest. 1980;42:603–615. 35. Burke DL, Frid MG, Kunrath CL, Karoor V, Anwar A, Wagner BD, Strassheim D, Stenmark KR. Sustained hypoxia promotes the development of a pulmonary artery-specific chronic inflammatory microenvironment. Am J Physiol Lung Cell Mol Physiol. 2009;297:L238–L250. doi: 10.1152/ajplung.90591.2008. 36. Li M, Riddle SR, Frid MG, El Kasmi KC, McKinsey TA, Sokol RJ, Strassheim D, Meyrick B, Yeager ME, Flockton AR, McKeon BA, Lemon DD, Horn TR, Anwar A, Barajas C, Stenmark KR. Emergence of fibroblasts with a proinflammatory epigenetically altered phenotype in severe hypoxic pulmonary hypertension. J Immunol. 2011;187:2711–2722. doi: 10.4049/jimmunol.1100479. 37. Wang D, Zhang H, Li M, Frid MG, Flockton AR, McKeon BA, Yeager ME, Fini MA, Morrell NW, Pullamsetti SS, Velegala S, Seeger W, McKinsey TA, Sucharov CC, Stenmark KR. MicroRNA-124 controls the proliferative, migratory, and inflammatory phenotype of pulmonary vascular fibroblasts. Circ Res. 2014;114:67–78. doi: 10.1161/ CIRCRESAHA.114.301633. 38. Yu AY, Shimoda LA, Iyer NV, Huso DL, Sun X, McWilliams R, Beaty T, Sham JSK, Wiener CM, Sylvester JT, Semenza GL. Impaired physiological responses to chronic hypoxia in mice partially deficient for hypoxiainducible factor 1α. J Clin Invest. 1999;103:691–696. 39. Brusselmans K, Compernolle V, Tjwa M, Wiesener MS, Maxwell PH, Collen D, Carmeliet P. Heterozygous deficiency of hypoxia-inducible factor–2α protects mice against pulmonary hypertension and right ventricular dysfunction during prolonged hypoxia. J Clin Invest. 2003;111:1519–1527. 40. Gale DP, Harten SK, Reid CDL, Tuddenham EGD, Maxwell PH. Autosomal dominant erythrocytosis and pulmonary arterial hypertension associated with an activating HIF2 mutation. Blood. 2008;112:919–921. doi: 10.1182/blood-2008-04-153718. 41. Formenti F, Beer PA, Croft QPP, Dorrington KL, Gale DP, Lappin TRJ, Lucas GS, Maher ER, Maxwell PH, McMullin MF, O’Connor DF, Percy MJ, Pugh CW, Ratcliffe PJ, Smith TG, Talbot NP, Robbins PA. Cardiopulmonary function in two human disorders of the hypoxia-inducible factor (HIF) pathway: von Hippel-Lindau disease and HIF-2 gain-of-function mutation. FASEB J. 2011;25:2001–2011. doi: 10.1096/fj.10-177378. 42. Tan Q, Kerestes H, Percy MJ, Pietrofesa R, Chen L, Khurana TS, Christofidou-Solomidou M, Lappin TR, Lee FS. Erythrocytosis and pulmonary hypertension in a mouse model of human HIF2A gain of

Wilkins et al   High-Altitude Pulmonary Vascular Response   589 function mutation. J Biol Chem. 2013;288:17134–17144. doi: 10.1074/ jbc.M112.444059. 43. Hislop A, Reid L. Changes in the pulmonary arteries of the rat during recovery from hypoxia-induced pulmonary hypertension. Br J Exp Pathol. 1977;58:653–662. 44. Cahill E, Rowan SC, Sands M, Banahan M, Ryan D, Howell K, McLoughlin P. The pathophysiological basis of chronic hypoxic pulmonary hypertension in the mouse: vasoconstrictor and structural mechanisms contribute equally. Exp Physiol. 2012;97:796–806. doi: 10.1113/ expphysiol.2012.065474. 45. Zuckerman BD, Orton EC, Latham LP, Barbiere CC, Stenmark KR, Reeves JT. Pulmonary vascular impedance and wave reflections in the hypoxic calf. J Appl Physiol (1985). 1992;72:2118–2127. 46. Lammers S, Scott D, Hunter K, Tan W, Shandas R, Stenmark KR. Mechanics and function of the pulmonary vasculature: implications for pulmonary vascular disease and right ventricular function. Compr Physiol. 2012;2:295–319. doi: 10.1002/cphy.c100070. 47. Naeije R. Physiological adaptation of the cardiovascular system to high altitude. Prog Cardiovasc Dis. 2010;52:456–466. doi: 10.1016/j. pcad.2010.03.004. 48. Holloway CJ, Montgomery HE, Murray AJ, Cochlin LE, Codreanu I, Hopwood N, Johnson AW, Rider OJ, Levett DZ, Tyler DJ, Francis JM, Neubauer S, Grocott MP, Clarke K; Caudwell Xtreme Everest Research Group. Cardiac response to hypobaric hypoxia: persistent changes in cardiac mass, function, and energy metabolism after a trek to Mt. Everest Base Camp. FASEB J. 2011;25:792–796. doi: 10.1096/fj.10-172999. 49. Naeije R. Pro: hypoxic pulmonary vasoconstriction is a limiting factor of exercise at high altitude. High Alt Med Biol. 2011;12:309–312. doi: 10.1089/ham.2011.1060. 50. Anholm JD, Foster GP. Con: hypoxic pulmonary vasoconstriction is not a limiting factor of exercise at high altitude. High Alt Med Biol. 2011;12:313–317. doi: 10.1089/ham.2011.1059. 51. Banchero N, Sime F, Penaloza D, Cruz J, Gamboa R, Marticorena E. Pulmonary pressure, cardiac output, and arterial oxygen saturation during exercise at high altitude and at sea level. Circulation. 1966;33:249–262. 52. Wang Z, Chesler NC. Pulmonary vascular mechanics: important contributors to the increased right ventricular afterload of pulmonary hypertension. Exp Physiol. 2013;98:1267–1273. doi: 10.1113/expphysiol.2012.069096. 53. Naeije R, Vanderpool R. Pulmonary hypertension and chronic mountain sickness. High Alt Med Biol. 2013;14:117–125. doi: 10.1089/ ham.2012.1124. 54. Zhao L, Sebkhi A, Nunez DJ, Long L, Haley CS, Szpirer J, Szpirer C, Williams AJ, Wilkins MR. Right ventricular hypertrophy secondary to pulmonary hypertension is linked to rat chromosome 17: evaluation of cardiac ryanodine Ryr2 receptor as a candidate. Circulation. 2001;103:442–447. 55. Grünig E, Mereles D, Hildebrandt W, Swenson ER, Kübler W, Kuecherer H, Bärtsch P. Stress Doppler echocardiography for identification of susceptibility to high altitude pulmonary edema. J Am Coll Cardiol. 2000;35:980–987. 56. Bärtsch P, Swenson ER. Acute high-altitude illnesses. N Engl J Med. 2013;368:2294–2302. doi: 10.1056/NEJMc1309747. 57. Maggiorini M, Mélot C, Pierre S, Pfeiffer F, Greve I, Sartori C, Lepori M, Hauser M, Scherrer U, Naeije R. High-altitude pulmonary edema is initially caused by an increase in capillary pressure. Circulation. 2001;103:2078–2083. 58. Swenson ER, Maggiorini M, Mongovin S, Gibbs JS, Greve I, Mairbäurl H, Bärtsch P. Pathogenesis of high-altitude pulmonary edema: inflammation is not an etiologic factor. JAMA. 2002;287:2228–2235. 59. Hopkins SR, Garg J, Bolar DS, Balouch J, Levin DL. Pulmonary blood flow heterogeneity during hypoxia and high-altitude pulmonary edema. Am J Respir Crit Care Med. 2005;171:83–87. doi: 10.1164/ rccm.200406-707OC. 60. Bärtsch P, Maggiorini M, Ritter M, Noti C, Vock P, Oelz O. Prevention of high-altitude pulmonary edema by nifedipine. N Engl J Med. 1991;325:1284–1289. doi: 10.1056/NEJM199110313251805. 61. Maggiorini M, Brunner-La Rocca H-P, Peth S, Fischler M, Böhm T, Bernheim A, Kiencke S, Bloch KE, Dehnert C, Naeije R. Both tadalafil and dexamethasone may reduce the incidence of high-altitude pulmonary edemaa randomized trial. Ann Intern Med. 2006;145:497–506. 62. Sartori C, Allemann Y, Duplain H, Lepori M, Egli M, Lipp E, Hutter D, Turini P, Hugli O, Cook S, Nicod P, Scherrer U. Salmeterol for the prevention of high-altitude pulmonary edema. N Engl J Med. 2002;346: 1631–1636. doi: 10.1056/NEJMoa013183. 63. Netzer N, Strohl K, Faulhaber M, Gatterer H, Burtscher M. Hypoxia-related altitude illnesses. J Travel Med. 2013;20:247–255. doi: 10.1111/jtm.12017.

64. León-Velarde F, Maggiorini M, Reeves JT, Aldashev A, Asmus I, Bernardi L, Ge RL, Hackett P, Kobayashi T, Moore LG, Penaloza D, Richalet JP, Roach R, Wu T, Vargas E, Zubieta-Castillo G, Zubieta-Calleja G. Consensus statement on chronic and subacute high altitude diseases. High Alt Med Biol. 2005;6:147–157. doi: 10.1089/ham.2005.6.147. 65. Hoeper MM, Bogaard HJ, Condliffe R, Frantz R, Khanna D, Kurzyna M, Langleben D, Manes A, Satoh T, Torres F, Wilkins MR, Badesch DB. Definitions and diagnosis of pulmonary hypertension. J Am Coll Cardiol. 2013;62(25 suppl):D42–D50. doi: 10.1016/j.jacc.2013.10.032. 66. Aldashev AA, Sarybaev AS, Sydykov AS, Kalmyrzaev BB, Kim EV, Mamanova LB, Maripov R, Kojonazarov BK, Mirrakhimov MM, Wilkins MR, Morrell NW. Characterization of high-altitude pulmonary hypertension in the Kyrgyz: association with angiotensin-converting enzyme genotype. Am J Respir Crit Care Med. 2002;166:1396–1402. doi: 10.1164/ rccm.200204-345OC. 67. Bogaard HJ, Natarajan R, Henderson SC, Long CS, Kraskauskas D, Smithson L, Ockaili R, McCord JM, Voelkel NF. Chronic pulmonary artery pressure elevation is insufficient to explain right heart failure. Circulation. 2009;120:1951–1960. doi: 10.1161/CIRCULATIONAHA.109.883843. 68. Sui GJ, Liu YH, Cheng XS, Anand IS, Harris E, Harris P, Heath D. Subacute infantile mountain sickness. J Pathol. 1988;155:161–170. doi: 10.1002/path.1711550213. 69. Anand IS, Malhotra RM, Chandrashekhar Y, Bali HK, Chauhan SS, Jindal SK, Bhandari RK, Wahi PL. Adult subacute mountain sickness–a syndrome of congestive heart failure in man at very high altitude. Lancet. 1990;335:561–565. 70. Huez S, Faoro V, Vachiéry JL, Unger P, Martinot JB, Naeije R. Images in cardiovascular medicine: high-altitude-induced rightheart failure. Circulation. 2007;115:e308–e309. doi: 10.1161/ CIRCULATIONAHA.106.650994. 71. Anand IS, Wu T. Syndromes of subacute mountain sickness. High Alt Med Biol. 2004;5:156–170. doi: 10.1089/1527029041352135. 72. Aldashev AA, Kojonazarov BK, Amatov TA, Sooronbaev TM, Mirrakhimov MM, Morrell NW, Wharton J, Wilkins MR. Phosphodiesterase type 5 and high altitude pulmonary hypertension. Thorax. 2005;60:683–687. doi: 10.1136/thx.2005.041954. 73. Richalet JP, Rivera-Ch M, Maignan M, Privat C, Pham I, Macarlupu JL, Petitjean O, León-Velarde F. Acetazolamide for Monge’s disease: efficiency and tolerance of 6-month treatment. Am J Respir Crit Care Med. 2008;177:1370–1376. doi: 10.1164/rccm.200802-196OC. 74. Kojonazarov B, Myrzaakhmatova A, Sooronbaev T, Ishizaki T, Aldashev A. Effects of fasudil in patients with high-altitude pulmonary hypertension. Eur Respir J. 2012;39:496–498. doi: 10.1183/09031936.00095211. 75. Seheult RD, Ruh K, Foster GP, Anholm JD. Prophylactic bosentan does not improve exercise capacity or lower pulmonary artery systolic pressure at high altitude. Respir Physiol Neurobiol. 2009;165:123–130. doi: 10.1016/j.resp.2008.10.005. 76. Kojonazarov B, Isakova J, Imanov B, Sovkhozova N, Sooronbaev T, Ishizaki T, Aldashev AA. Bosentan reduces pulmonary artery pressure in high altitude residents. High Alt Med Biol. 2012;13:217–223. doi: 10.1089/ham.2011.1107. 77. Groves BM, Droma T, Sutton JR, McCullough RG, McCullough RE, Zhuang J, Rapmund G, Sun S, Janes C, Moore LG. Minimal hypoxic pulmonary hypertension in normal Tibetans at 3,658 m. J Appl Physiol (1985). 1993;74:312–318. 78. Wu T, Kayser B. High altitude adaptation in Tibetans. High Alt Med Biol. 2006;7:193–208. doi: 10.1089/ham.2006.7.193. 79. Pei SX, Chen XJ, Si Ren BZ, Liu YH, Cheng XS, Harris EM, Anand IS, Harris PC. Chronic mountain sickness in Tibet. Q J Med. 1989;71:555–574. 80. Gupta ML, Rao KS, Anand IS, Banerjee AK, Boparai MS. Lack of smooth muscle in the small pulmonary arteries of the native Ladakhi: is the Himalayan highlander adapted? Am Rev Respir Dis. 1992;145: 1201–1204. doi: 10.1164/ajrccm/145.5.1201. 81. Petousi N, Croft QPP, Cavalleri GL, Cheng H-Y, Formenti F, Ishida K, Lunn D, McCormack M, Shianna KV, Talbot NP, Ratcliffe PJ, Robbins PA. Tibetans living at sea level have a hyporesponsive hypoxia-inducible factor (HIF) system and blunted physiological responses to hypoxia. J Appl Physiol. 2013;116:893–904. doi: 10.1152/japplphysiol.00535.2013. 82. Beall CM, Laskowski D, Strohl KP, Soria R, Villena M, Vargas E, Alarcon AM, Gonzales C, Erzurum SC. Pulmonary nitric oxide in mountain dwellers. Nature. 2001;414:411–412. doi: 10.1038/35106641. 83. León-Velarde F, Mejía O. Gene expression in chronic high altitude diseases. High Alt Med Biol. 2008;9:130–139. doi: 10.1089/ham.2007.1077. 84. Beall CM, Cavalleri GL, Deng L, Elston RC, Gao Y, Knight J, Li C, Li JC, Liang Y, McCormack M, Montgomery HE, Pan H, Robbins PA, Shianna

590  Circulation  February 10, 2015 KV, Tam SC, Tsering N, Veeramah KR, Wang W, Wangdui P, Weale ME, Xu Y, Xu Z, Yang L, Zaman MJ, Zeng C, Zhang L, Zhang X, Zhaxi P, Zheng YT. Natural selection on EPAS1 (HIF2alpha) associated with low hemoglobin concentration in Tibetan highlanders. Proc Natl Acad Sci U S A. 2010;107:11459–11464. doi: 10.1073/pnas.1002443107. 85. Simonson TS, Yang Y, Huff CD, Yun H, Qin G, Witherspoon DJ, Bai Z, Lorenzo FR, Xing J, Jorde LB, Prchal JT, Ge R. Genetic evidence for high-altitude adaptation in Tibet. Science. 2010;329:72–75. doi: 10.1126/ science.1189406. 86. Yi X, Liang Y, Huerta-Sanchez E, Jin X, Cuo ZX, Pool JE, Xu X, Jiang H, Vinckenbosch N, Korneliussen TS, Zheng H, Liu T, He W, Li K, Luo R, Nie X, Wu H, Zhao M, Cao H, Zou J, Shan Y, Li S, Yang Q, Asan, Ni P, Tian G, Xu J, Liu X, Jiang T, Wu R, Zhou G, Tang M, Qin J, Wang T, Feng S, Li G, Huasang, Luosang J, Wang W, Chen F, Wang Y, Zheng X, Li Z, Bianba Z, Yang G, Wang X, Tang S, Gao G, Chen Y, Luo Z, Gusang L, Cao Z, Zhang Q, Ouyang W, Ren X, Liang H, Zheng H, Huang Y, Li J, Bolund L, Kristiansen K, Li Y, Zhang Y, Zhang X, Li R, Li S, Yang H, Nielsen R, Wang J, Wang J. Sequencing of 50 human exomes reveals adaptation to high altitude. Science. 2010;329:75–78. doi: 10.1126/science.1190371. 87. Bigham A, Bauchet M, Pinto D, Mao X, Akey JM, Mei R, Scherer SW, Julian CG, Wilson MJ, López Herráez D, Brutsaert T, Parra EJ, Moore LG, Shriver MD. Identifying signatures of natural selection in Tibetan and Andean populations using dense genome scan data. PLoS Genet. 2010;6:e1001116. doi: 10.1371/journal.pgen.1001116. 88. Peng Y, Yang Z, Zhang H, Cui C, Qi X, Luo X, Tao X, Wu T, Ouzhuluobu, Basang, Ciwangsangbu, Danzengduojie, Chen H, Shi H, Su B. Genetic variations in Tibetan populations and high-altitude adaptation at the Himalayas. Mol Biol Evol. 2011;28:1075–1081. doi: 10.1093/molbev/ msq290. 89. Wang B, Zhang YB, Zhang F, Lin H, Wang X, Wan N, Ye Z, Weng H, Zhang L, Li X, Yan J, Wang P, Wu T, Cheng L, Wang J, Wang DM, Ma X, Yu J. On the origin of Tibetans and their genetic basis in adapting high-altitude environments. PLoS One. 2011;6:e17002. doi: 10.1371/journal.pone.0017002. 90. Xu S, Li S, Yang Y, Tan J, Lou H, Jin W, Yang L, Pan X, Wang J, Shen Y, Wu B, Wang H, Jin L. A genome-wide search for signals of high-altitude adaptation in Tibetans. Mol Biol Evol. 2011;28:1003–1011. doi: 10.1093/ molbev/msq277. 91. Lorenzo FR, Huff C, Myllymäki M, Olenchock B, Swierczek S, Tashi T, Gordeuk V, Wuren T, Ri-Li G, McClain DA, Khan TM, Koul PA, Guchhait P, Salama ME, Xing J, Semenza GL, Liberzon E, Wilson A, Simonson TS, Jorde LB, Kaelin WG Jr, Koivunen P, Prchal JT. A genetic mechanism for Tibetan high-altitude adaptation. Nat Genet. 2014;46:951–956. doi: 10.1038/ng.3067.

92. Bigham AW, Wilson MJ, Julian CG, Kiyamu M, Vargas E, Leon-Velarde F, Rivera-Chira M, Rodriquez C, Browne VA, Parra E, Brutsaert TD, Moore LG, Shriver MD. Andean and Tibetan patterns of adaptation to high altitude. Am J Hum Biol. 2013;25:190–197. doi: 10.1002/ajhb.22358. 93. Alkorta-Aranburu G, Beall CM, Witonsky DB, Gebremedhin A, Pritchard JK, Di Rienzo A. The genetic architecture of adaptations to high altitude in Ethiopia. PLoS Genet. 2012;8:e1003110. doi: 10.1371/journal. pgen.1003110. 94. Scheinfeldt LB, Soi S, Thompson S, Ranciaro A, Woldemeskel D, Beggs W, Lambert C, Jarvis JP, Abate D, Belay G, Tishkoff SA. Genetic adaptation to high altitude in the Ethiopian highlands. Genome Biol. 2012;13:R1. doi: 10.1186/gb-2012-13-1-r1. 95. Huerta-Sánchez E, Degiorgio M, Pagani L, Tarekegn A, Ekong R, Antao T, Cardona A, Montgomery HE, Cavalleri GL, Robbins PA, Weale ME, Bradman N, Bekele E, Kivisild T, Tyler-Smith C, Nielsen R. Genetic signatures reveal high-altitude adaptation in a set of ethiopian populations. Mol Biol Evol. 2013;30:1877–1888. doi: 10.1093/molbev/mst089. 96. Beall CM. Two routes to functional adaptation: Tibetan and Andean highaltitude natives. Proc Natl Acad Sci U S A. 2007;104(suppl 1):8655–8660. doi: 10.1073/pnas.0701985104. 97. Huerta-Sánchez E, Jin X, Asan, Bianba Z, Peter BM, Vinckenbosch N, Liang Y, Yi X, He M, Somel M, Ni P, Wang B, Ou X, Huasang, Luosang J, Cuo ZX, Li K, Gao G, Yin Y, Wang W, Zhang X, Xu X, Yang H, Li Y, Wang J, Wang J, Nielsen R. Altitude adaptation in Tibetans caused by introgression of Denisovan-like DNA. Nature. 2014;512:194–197. doi: 10.1038/nature13408. 98. Ji LD, Qiu YQ, Xu J, Irwin DM, Tam SC, Tang NL, Zhang YP. Genetic adaptation of the hypoxia-inducible factor pathway to oxygen pressure among eurasian human populations. Mol Biol Evol. 2012;29:3359–3370. doi: 10.1093/molbev/mss144. 99. Zhou D, Udpa N, Ronen R, Stobdan T, Liang J, Appenzeller O, Zhao HW, Yin Y, Du Y, Guo L, Cao R, Wang Y, Jin X, Huang C, Jia W, Cao D, Guo G, Gamboa JL, Villafuerte F, Callacondo D, Xue J, Liu S, Frazer KA, Li Y, Bafna V, Haddad GG. Whole-genome sequencing uncovers the genetic basis of chronic mountain sickness in Andean highlanders. Am J Hum Genet. 2013;93:452–462. doi: 10.1016/j.ajhg.2013.07.011. 100. Pasha MAQ, Newman JH. High-altitude disorders: pulmonary hypertension. Chest. 2010;137:13S–19S. Key Words: anoxia ◼ genetics ◼ hypertension, pulmonary ◼ vascular remodeling ◼ vasoconstriction

Pathophysiology and treatment of high-altitude pulmonary vascular disease.

Pathophysiology and treatment of high-altitude pulmonary vascular disease. - PDF Download Free
1MB Sizes 0 Downloads 11 Views