ﺑﺎﺯﮔﺸﺖ ﺑﻪ ﺻﻔﺤﻪ ﻗﺒﻠﯽ
خرید پکیج
تعداد آیتم قابل مشاهده باقیمانده : 3 مورد
نسخه الکترونیک
medimedia.ir

Heart transplantation in adults: Cardiac allograft vasculopathy pathogenesis and risk factors

Heart transplantation in adults: Cardiac allograft vasculopathy pathogenesis and risk factors
Literature review current through: Jan 2024.
This topic last updated: Jan 27, 2022.

INTRODUCTION — Cardiac transplantation is the definitive therapy for eligible patients with end-stage heart failure. The major limitations to survival in the early post-transplant period (first year) are nonspecific graft failure, multiorgan failure, acute rejection, and infection [1]. Beyond the first year, cardiac allograft vasculopathy (CAV, also called transplant coronary artery disease or cardiac transplant vasculopathy) is among the top three causes of death [1]. Approximately 30 percent of patients have angiographic coronary artery disease at five years and 50 percent at 10 years, with the incidence increasing progressively with time [1]. (See "Heart transplantation in adults: Prognosis", section on 'Causes of death' and "Heart transplantation: Clinical manifestations, diagnosis, and prognosis of cardiac allograft vasculopathy".)

The pathogenesis of CAV will be reviewed here. The diagnosis, prevention, and treatment of this disease are discussed separately. (See "Heart Transplantation: Prevention and treatment of cardiac allograft vasculopathy" and "Heart transplantation: Clinical manifestations, diagnosis, and prognosis of cardiac allograft vasculopathy".)

PATHOLOGY — Cardiac allograft vasculopathy (CAV) is a panarterial disease confined to the allograft and is characterized by diffuse concentric longitudinal intimal hyperplasia in the epicardial coronary arteries (figure 1) [2-4] and concentric medial disease in the microvasculature [5-7]. In contrast, traditional atherosclerosis is focal, noncircumferential, and most often observed proximally in the epicardial vessels.

Transplant recipients also frequently develop proximal coronary artery disease. However, these lesions more closely resemble traditional atherosclerosis pathologically and probably evolve from pre-existing disease in the donor heart that is accelerated by the plethora of cardiac risk factors after transplantation [4].

Serial intravascular (intracoronary) ultrasound testing has shown that most of the intimal thickening occurs during the first year after transplantation [8]. Lumen loss is also due to arterial remodeling, with early expansion and late constriction of the external elastic membrane area. (See "Heart transplantation: Clinical manifestations, diagnosis, and prognosis of cardiac allograft vasculopathy", section on 'Epidemiology'.)

Coronary angioscopy has demonstrated heterogeneous intimal lesions with varied clinical correlates (picture 1) [9].

CAV is also associated with intracoronary mural and occlusive thrombi, which can cause acute myocardial infarction [10,11]. Thrombi typically develop more than two months after transplantation [10]. Early, thin mural thrombi primarily contain platelets, while later organized thrombi, which are often occlusive, primarily consist of fibrin [11]. The important role of platelets is further supported by a strong association between platelet activation and the development and progression of CAV [12].

PATHOGENESIS — Cardiac allograft vasculopathy (CAV) appears to be multifactorial in origin with both immunologic and nonimmunologic factors implicated. Among the factors associated with vasculopathy are cellular and antibody-mediated rejection, donor-specific anti-HLA antibodies, cytomegalovirus infection, and hypercholesterolemia.

Immunologic events appear to be most important, since CAV develops in the donor's but not the recipient's arteries. It has been suggested that CAV reflects an accelerated "normal" healing process following allograft-induced immunologic injury [13].

Consistent with the importance of inflammation is the observation that elevations in serum C-reactive protein, an acute phase reactant, predict a greater likelihood of CAV and allograft failure [14-16]. (See "Heart transplantation in adults: Prognosis".)

Immunologic factors — The development of CAV involves both the cellular and humoral arms of the immune system. Immunologic mechanisms that result in the initiation and maintenance of cellular rejection correlate with the development of vascular injury. These factors include histocompatibility antigen HLA mismatching, T cell activation, endothelial cell activation, and altered cytokine expression. Increasing evidence suggests that donor specific antibodies and antibody mediated rejection are also major contributors to CAV [17]. The basic aspects of transplantation immunobiology are reviewed in detail separately. (See "Transplantation immunobiology".)

HLA mismatch — Several studies have shown a greater incidence of HLA mismatching in recipients who develop CAV than in those who do not [2,3]. HLA-DR mismatching has generally been the most important of the HLA subtypes [18], but a positive correlation with HLA-A mismatching has also been observed [19,20]. Development of de novo donor specific anti HLA Class I or II antibodies after transplant have been associated with CAV [21].

Acute cellular rejection — The number of episodes of moderate to severe cellular rejection appears to correlate with the development of CAV [2,22-25]. In addition a summary of overall rejection, the total rejection score, has also been associated with CAV [26]. This observation is consistent with animal data that suggest that the immunogenicity of the graft is probably the most important stimulus to the development of vasculopathy [27]. However, some studies have not supported a relationship between rejection ≥2R (3A) and vasculopathy [28].

Antibodies and antibody-mediated rejection — Humoral activation, resulting in the production of anti-HLA and antiendothelial antibodies [29-33], enhances the development of CAV, and is associated with acute antibody-mediated rejection (also called humoral or vascular rejection) [29,32,34]. (See "Heart transplantation in adults: Diagnosis of allograft rejection", section on 'Acute antibody-mediated (humoral) rejection'.)

Episodes of antibody-mediated rejection are associated with the development of CAV and poorer long-term outcomes [33-35].

Anti-HLA and antiendothelial antibodies also appear to increase the risk of CAV independent of antibody-mediated rejection. In one study, for example, cardiac transplant recipients who developed anti-HLA antibodies had a lower four-year survival rate than those who did not develop such antibodies (90 versus 38 percent) [30]. CAV was responsible for many of the late (more than one year post-transplant) deaths in these patients. At two years post-transplant in another report, a higher incidence of CAV was found in patients who continued to make anti-donor HLA antibodies than in those who did not (18 versus 3 percent) [31].

A higher rate of accelerated coronary artery disease has also been associated with the presence of antiendothelial cell antibodies (72 versus 5 percent in those without such antibodies) [33].

T cell activation — A complex interplay between activated T lymphocytes and endothelial cells may explain the sequential and temporal changes observed in the infiltrating T cell subtypes found in the coronary arteries of transplanted hearts (figure 2) [36-41]. T cells, HLA-DR+ endothelial cells, and HLA-DR+ macrophages are frequently seen in vasculopathic coronary artery lesions [41].

Endothelial cells, in particular, show significantly increased expression of MHC class I alloantigens [36]. These antigens are thought to be recognized by CD8+ cells, resulting in the secretion of CD8+ derived cytokines that further activate coronary endothelial cells. The activated endothelial cells, in turn, express increased levels of MHC class II antigens, which subsequently activate CD4+ T cells. As a result, CD8+ cells are predominant in the early vasculopathic lesion, while CD4+ cells are found in increasing numbers in later pathologic stages.

Endothelial cell activation — Activation of the endothelial cell provides the proper milieu for the recruitment of proinflammatory cells from the vasculature, the initiation of the immune response, and the subsequent development of CAV. Cytokines released from T cells, macrophages, and other cells (see 'Cytokines' below) dramatically increase the expression of intercellular cell adhesion molecule-1 (ICAM-1), vascular cell adhesion molecule-1 (VCAM-1), and P-selectin on the surface of allograft endothelial cells at the time of acute cellular rejection [42-44].

The prognostic significance of endothelial activation was suggested in a study in which serial endomyocardial biopsy specimens were obtained in 121 donor hearts just prior to and at three months following transplantation [45]. The patients with evidence of endothelial activation during the first three months following transplant were at significantly greater risk for developing vasculopathy (64 versus 28 percent in patients without activation).

Serum levels of soluble ICAM-1 (sICAM-1) also may predict the risk of CAV. Elevated levels correlate with the biopsy finding of ICAM-1 on arterial and arteriolar endothelium and are associated with a greater risk of vasculopathy and graft failure [46].

Thus, evidence of either endothelial activation on endomyocardial biopsy or elevated serum sICAM-I concentration can predict the development of CAV.

In a cohort of 76 heart transplant recipients biopsied at 12 months post-transplantation, antibody-mediated rejection (AMR) positive biopsies showed significantly greater endothelial localization of VEGF than time-matched AMR negative biopsies [47]. The presence of diffuse endothelial expression of VEGF was associated with a 2.5-fold risk of developing CAV while biopsy evidence of AMR was associated with a fivefold risk.

Endothelial dysfunction (assessed via Doppler flow wire arterial responsiveness to acetylcholine, adenosine, and nifedipine) was found to be an independent predictor of cardiovascular events and death [48].

Cytokines — Cytokine expression by infiltrating lymphocytes and macrophages has been detected in transplant vasculopathic lesions. Among the cytokines involved are interleukin (IL)-1, IL-6, tumor necrosis factor (TNF)-alpha, platelet derived growth factor (PDGF), insulin-like growth factor-I (IGF-1), macrophage chemoattractant protein-1 (MCP-1), fibroblastic growth factor, vascular endothelial growth factor, transforming growth factor-alpha (TGF-alpha), and TGF-beta-1 [49-56]. These cytokines have trophic and proliferative effects on coronary artery smooth muscle cells.

Interferon-gamma is produced by certain T cell subsets and may play a central role in CAV by activating macrophages and enhancing the production of MHC products, other components of the antigen presentation pathway, and the adhesion molecules ICAM-1 and VCAM-1 [57]. Interferon-gamma deficient mice appear to be protected against transplant arterial disease but not graft rejection [57].

An increasing number of genetic polymorphisms have been associated with the development of vasculopathy. These include the genes for TNF-alpha and IL-6 [58,59].

Nonimmunologic factors — Nonimmunologic factors contribute to the development of CAV. These factors may act indirectly through immunologic pathways.

Age and sex — Age and sex in both the donor and recipient are predictors of CAV:

Among donor characteristics, older age and male sex are associated with higher risk [25,60,61]. In a study of 489 one-year heart transplant survivors who underwent 1435 coronary angiograms, the relative risk for CAV was 1.26 for every 10 years of donor age [25].

Among recipient characteristics, younger age is associated with higher risk [18,25,62].

Hyperlipidemia — Elevations in serum total cholesterol, low-density lipoprotein (LDL)-cholesterol, oxidized LDL-cholesterol, and triglycerides are common after heart transplantation. These changes are probably related to immunosuppressive therapy with corticosteroids and cyclosporine and to obesity [63]. In comparison, lipoprotein(a) levels decrease during the first six months after transplantation and are not associated with vasculopathy [64,65]. (See "Heart transplantation: Hyperlipidemia after transplantation".)

The lipid abnormalities appear to correlate with the severity of the vasculopathy [66-68]. Support for a direct role of hypercholesterolemia comes from a study in mice in which hypercholesterolemia increased the rates of neointima formation and vascular occlusion by a mechanism that depends upon smooth muscle cell accumulation [69]. Further support comes from the observation in heart transplant recipients that lipid-lowering with statins is associated with a significant reduction in the incidence and severity of CAV [70]. (See "Heart Transplantation: Prevention and treatment of cardiac allograft vasculopathy", section on 'Statins'.)

Cytomegalovirus infection — Cytomegalovirus infection has been associated with a higher incidence of CAV [71,72]. One or more of the following mechanisms may contribute:

A direct endothelial assault, which results in the enhancement of vascular adhesiveness, activation of the coagulation cascade, and elaboration of cytokines.

Adverse vascular remodeling with greater net lumen loss [73].

Stimulation of cellular immune responses in the vasculature [63], perhaps via induced expression of MHC antigens on the endothelial cells [74].

Glycemic control and insulin resistance — Both impaired glycemic control and insulin resistance may play a role in the pathogenesis of CAV. As an example, an association has been noted between an elevated HbA1c concentration, a marker of chronic glucose intolerance, and the incidence and severity of CAV [75].

The metabolic syndrome, also called the insulin resistance syndrome or syndrome X, is characterized by abdominal obesity, hypertension, diabetes, and an atherogenic lipid profile (hypertriglyceridemia and low high density lipoprotein [HDL] cholesterol) and is seen frequently after heart transplantation. In a series of 66 patients, high serum insulin or glucose concentrations were associated with a significantly lower likelihood of freedom from CAV (57 versus 82 percent in patients with normal values) and with reduced survival [76]. (See "Metabolic syndrome (insulin resistance syndrome or syndrome X)".)

In 98 consecutive heart transplant recipients, high C-reactive protein (CRP) and insulin resistance (defined as triglyceride to HDL ratio >3.0), were significantly associated with CAV identified by angiography. The combination of increased CRP and insulin resistance was synergistic and associated with a fourfold increased risk in developing CAV [77].

Coronary disease history — Coronary artery disease in either the donor [60] or recipient [18,22] appears to contribute to CAV. In one report, CAV at three years after transplantation was more common in hearts from donors with angiographic coronary disease and from older donors (figure 3) [76]. Although older donor age is one of the strongest predictors for CAV, donor age did not affect recipient survival or freedom from ischemic events at a mean follow-up of 3.8 years.

Fibrinolysis — Animal models suggest that the plasminogen system may contribute to the development of CAV by mediating elastin degradation, macrophage infiltration, media remodeling, medial smooth muscle migration, and the formation of a neointima; as an example, plasminogen deficient mice have a reduced risk of vasculopathy [78].

In contrast, deficient fibrinolysis may be a contributing factor to CAV in humans. Grafts with persistent depletion of tissue-type plasminogen activator (tPA) and expression of its inhibitor, plasminogen activator inhibitor-1 (PAI-1) are much more likely to develop CAV (78 versus 24 percent) and patients with such grafts are much more likely to receive a second transplant or die (30 versus 2.5 percent) [79].

The cause of t-PA depletion is unknown, but may have a genetic basis. It is possible that genotype-specific overexpression of PAI-1 leads to local depletion of t-PA. Consistent with this hypothesis is the observation that PAI-1 levels and impairment of fibrinolysis are linked to the presence of PAI-1 gene polymorphisms in the recipient, ie, 1/1, 1/2, or 2/2 genotypes [80]. In one series of 48 transplant recipients, a donor 2/2 PAI-1 genotype was associated with a significant risk of CAV [81]. The actuarial freedom from any coronary artery disease at 12 and 24 months for the 1/1, 1/2, and 2/2 genotypes was 100 and 100 percent, 92 and 92 percent, and 75 and 45 percent, respectively. There was no association with donor t-PA genotype.

Grafts with a persistent loss of vascular antithrombin also have a higher incidence of vasculopathy. In one report, the vascular disease was more severe and progressed more rapidly in such grafts compared with those that lost and later recovered antithrombin [82].

Endothelial dysfunction — The development of vasculopathy may correlate with the presence of endothelial dysfunction and cardiac death, as established by paradoxical coronary artery constriction in response to acetylcholine or to the cold pressor test [83-86]. (See "Coronary endothelial dysfunction: Clinical aspects".)

The following observations illustrate this relationship:

In a report of 20 patients, endothelial dysfunction in the early post-transplant period was associated with a significant increase in angiographic vasculopathy at one year (58 versus 13 percent in those without endothelial dysfunction) [84].

In a series of 45 patients followed with annual angiography, IVUS, and Doppler assessment of endothelial function, the development of endothelial dysfunction was significantly associated with both CAV and a higher event rate [86].

Inducible nitric oxide synthase is upregulated in grafts with CAV and may play a protective role. Consistent with this hypothesis are the observations that CAV is exacerbated in animals deficient in inducible nitric oxide synthase and reduced in animals given gene therapy with endothelial nitric oxide synthase [87,88]. A similar protective effect can be achieved in animals by lowering the levels of asymmetric dimethylarginine, an endogenous inhibitor of nitric oxide synthase [89].

On the other hand, an excessive nitric oxide response may be deleterious, promoting CAV by a process that may involve reaction with the free radical superoxide to form peroxynitrite, a strong oxidant that damages cellular proteins [90].

There are a number of different responses to vessel injury, some of which are cytoprotective and/or inflammatory. Using a global proteomic approach examining protein expression in endomyocardial biopsies from patients with and without angiographic vasculopathy, an increase in heat shock protein 27 (HSP27) was found in patients without vasculopathy [91]. This observation suggests that normal vascular expression of HSP27 may protect against vessel injury, possibly by reducing apoptosis.

ACE gene polymorphism — Angiotensin converting enzyme (ACE) polymorphism may be associated with the development of CAV [92,93]. In a series of 80 heart transplant recipients, the DD genotype in the donors but not the recipients was associated with the development CAV [92]. This observation suggests the importance of tissue rather than circulating ACE. (See "Pathophysiology of heart failure: Neurohumoral adaptations", section on 'ACE gene polymorphism'.)

Endothelin — The endothelial cells release endothelin-1, a potent direct vasoconstrictor that also may stimulate the proliferation of vascular smooth muscle cells and fibroblasts [94]. Intense endothelin-1 immunoreactivity has been demonstrated in arteries with CAV, but not in normal coronary arteries [95].

The potential pathogenetic importance of endothelin-1 was shown in an animal model in which there was local upregulation of endothelin-1 in the thickened neointima and media of the coronary arteries [96]. The administration of bosentan, a nonselective endothelin receptor antagonist, suppressed the development of graft atherosclerosis. (See "Pathophysiology of heart failure: Neurohumoral adaptations", section on 'Endothelin'.)

Vascular remodeling — Adverse vascular remodeling can occur. When an atheromatous plaque develops within an artery, the artery can respond with compensatory dilatation, resulting in preserved net lumen area, or with shrinkage and a further net reduction in lumen area [97-99]. Intravascular ultrasound (IVUS) has allowed an increased understanding of this process. In one report, 78 percent of patients had overcompensation or partial compensatory dilatation, but 22 percent had no compensatory dilatation or shrinkage in the coronary segment [97].

IVUS has also shown that 95 percent of heart transplant recipients had an increase in intimal area over time. Among these patients, only 37 percent had some compensatory dilation, suggesting that abnormal vascular remodeling contributes to the decrease in luminal area [97,98].

Early left ventricular dysfunction — In a review of 117 patients, multivariate analysis showed that mean lumen diameter loss at one year was inversely related to early left ventricular function as defined by fractional shortening on echocardiography performed within the first week after transplantation [61]. It was speculated that myocardial injury associated with brain death in the donor, myocardial preservation, and ischemia-reperfusion injury might lead to injury to coronary endothelium as well as the myocardium.

Etiology of brain death in donor — Explosive brain death (eg, gunshot wound, head trauma, or intracerebral hemorrhage, as opposed to ischemic stroke) in the donor has been associated with an increased risk of late CAV and reduced survival in the recipient in some [100,101] but not all studies [61]. Systemic activation of matrix metalloproteinase (MMP)-2 and MMP-2 is associated with intracerebral hemorrhage and may contribute to progression of hemorrhagic stroke [101]. Although activation occurs before the heart is removed from such donors, endomyocardial biopsies obtained in the recipient at one week after transplantation show a marked increase in expression of both MMP-2 and MMP-9 [101].

Other — A number of other factors may importantly contribute to an increased risk of CAV [18,62]. These include:

Recipient obesity [62].

Pretransplantation diagnosis of ischemic compared with nonischemic cardiomyopathy (see 'Coronary disease history' above).

Longer graft ischemia time (the time between heart explant from the donor and implantation in the recipient) and the development of fibrosis [102].

Increases in elastase activity [103], thrombospondin-1 (a matrix glycoprotein that inhibits angiogenesis and facilitates the smooth muscle proliferation that is characteristic of CAV) [104], and the expression of tissue factor and the vitronectin receptor [105].

Hepatitis C virus seropositivity in the donor [106].

Possibly hyperhomocysteinemia [107,108]. (See "Heart Transplantation: Prevention and treatment of cardiac allograft vasculopathy", section on 'Homocysteine'.)

Angiogenesis within the intima [109].

Higher von Willebrand factor levels [110].

SUMMARY

Cardiac allograft vasculopathy (CAV) appears to be multifactorial in origin with both immunologic and nonimmunologic factors implicated. Among the factors associated with vasculopathy are cellular and antibody-mediated rejection, donor-specific anti-HLA antibodies, cytomegalovirus infection, and hypercholesterolemia. (See 'Pathogenesis' above.)

Beyond the first year after cardiac transplantation, CAV, also called transplant coronary artery disease or cardiac transplant vasculopathy) is among the top three causes of death. (See 'Introduction' above.)

CAV is a panarterial disease confined to the allograft and is characterized by diffuse concentric longitudinal intimal hyperplasia in the epicardial coronary arteries (figure 1) and concentric medial disease in the microvasculature. In contrast, traditional atherosclerosis is focal, noncircumferential, and most often observed proximally in the epicardial vessels. (See 'Pathology' above.)

CAV appears to be multifactorial in origin with both immunologic and nonimmunologic factors implicated. Among the factors associated with vasculopathy are cellular and antibody mediated rejection, donor specific anti-HLA antibodies, cytomegalovirus infection, and hypercholesterolemia. Immunologic events appear to be most important, since CAV develops in the donor's but not the recipient's arteries. (See 'Pathogenesis' above.)

  1. Lund LH, Edwards LB, Kucheryavaya AY, et al. The registry of the International Society for Heart and Lung Transplantation: thirty-first official adult heart transplant report--2014; focus theme: retransplantation. J Heart Lung Transplant 2014; 33:996.
  2. Uretsky BF, Murali S, Reddy PS, et al. Development of coronary artery disease in cardiac transplant patients receiving immunosuppressive therapy with cyclosporine and prednisone. Circulation 1987; 76:827.
  3. Pollack MS, Ballantyne CM. HLA match and other immunologic parameters in relation to survival, rejection severity and accelerated coronary artery disease after heart transplant. Clin Transplant 1990; 4:269.
  4. Tuzcu EM, De Franco AC, Goormastic M, et al. Dichotomous pattern of coronary atherosclerosis 1 to 9 years after transplantation: insights from systematic intravascular ultrasound imaging. J Am Coll Cardiol 1996; 27:839.
  5. Billingham ME. Graft coronary disease: the lesions and the patients. Transplant Proc 1989; 21:3665.
  6. Tanaka H, Swanson SJ, Sukhova G, et al. Early proliferation of medial smooth muscle cells in coronary arteries of rabbit cardiac allografts during immunosuppression with cyclosporine A. Transplant Proc 1995; 27:2062.
  7. Hiemann NE, Wellnhofer E, Knosalla C, et al. Prognostic impact of microvasculopathy on survival after heart transplantation: evidence from 9713 endomyocardial biopsies. Circulation 2007; 116:1274.
  8. Tsutsui H, Ziada KM, Schoenhagen P, et al. Lumen loss in transplant coronary artery disease is a biphasic process involving early intimal thickening and late constrictive remodeling: results from a 5-year serial intravascular ultrasound study. Circulation 2001; 104:653.
  9. Mehra MR, Ventura HO, Jain SP, et al. Heterogeneity of cardiac allograft vasculopathy: clinical insights from coronary angioscopy. J Am Coll Cardiol 1997; 29:1339.
  10. Arbustini E, Dal Bello B, Morbini P, et al. Frequency and characteristics of coronary thrombosis in the epicardial coronary arteries after cardiac transplantation. Am J Cardiol 1996; 78:795.
  11. Arbustini E, Dal Bello B, Morbini P, et al. Immunohistochemical characterization of coronary thrombi in allograft vascular disease. Transplantation 2000; 69:1095.
  12. Fateh-Moghadam S, Bocksch W, Ruf A, et al. Changes in surface expression of platelet membrane glycoproteins and progression of heart transplant vasculopathy. Circulation 2000; 102:890.
  13. Hillebrands JL, Klatter FA, van den Hurk BM, et al. Origin of neointimal endothelium and alpha-actin-positive smooth muscle cells in transplant arteriosclerosis. J Clin Invest 2001; 107:1411.
  14. Eisenberg MS, Chen HJ, Warshofsky MK, et al. Elevated levels of plasma C-reactive protein are associated with decreased graft survival in cardiac transplant recipients. Circulation 2000; 102:2100.
  15. Labarrere CA, Lee JB, Nelson DR, et al. C-reactive protein, arterial endothelial activation, and development of transplant coronary artery disease: a prospective study. Lancet 2002; 360:1462.
  16. Hognestad A, Endresen K, Wergeland R, et al. Plasma C-reactive protein as a marker of cardiac allograft vasculopathy in heart transplant recipients. J Am Coll Cardiol 2003; 42:477.
  17. Kaczmarek I, Deutsch MA, Kauke T, et al. Donor-specific HLA alloantibodies: long-term impact on cardiac allograft vasculopathy and mortality after heart transplant. Exp Clin Transplant 2008; 6:229.
  18. Taylor DO, Edwards LB, Boucek MM, et al. Registry of the International Society for Heart and Lung Transplantation: twenty-fourth official adult heart transplant report--2007. J Heart Lung Transplant 2007; 26:769.
  19. Narrod J, Kormos R, Armitage J, et al. Acute rejection and coronary artery disease in long-term survivors of heart transplantation. J Heart Transplant 1989; 8:418.
  20. Radovancevic B, Birovljev S, Vega JD, et al. Inverse relationship between human leukocyte antigen match and development of coronary artery disease. Transplant Proc 1991; 23:1144.
  21. Tambur AR, Pamboukian SV, Costanzo MR, et al. The presence of HLA-directed antibodies after heart transplantation is associated with poor allograft outcome. Transplantation 2005; 80:1019.
  22. Radovancevic B, Poindexter S, Birovljev S, et al. Risk factors for development of accelerated coronary artery disease in cardiac transplant recipients. Eur J Cardiothorac Surg 1990; 4:309.
  23. Zerbe T, Uretsky B, Kormos R, et al. Graft atherosclerosis: effects of cellular rejection and human lymphocyte antigen. J Heart Lung Transplant 1992; 11:S104.
  24. Lindelöw B, Bergh C, Lamm C, et al. Graft coronary artery disease is strongly related to the aetiology of heart failure and cellular rejections. Eur Heart J 1999; 20:1326.
  25. Stoica SC, Cafferty F, Pauriah M, et al. The cumulative effect of acute rejection on development of cardiac allograft vasculopathy. J Heart Lung Transplant 2006; 25:420.
  26. Raichlin E, Edwards BS, Kremers WK, et al. Acute cellular rejection and the subsequent development of allograft vasculopathy after cardiac transplantation. J Heart Lung Transplant 2009; 28:320.
  27. Schmid C, Heemann U, Tilney NL. Factors contributing to the development of chronic rejection in heterotopic rat heart transplantation. Transplantation 1997; 64:222.
  28. Tuzcu EM, Kapadia SR, Sachar R, et al. Intravascular ultrasound evidence of angiographically silent progression in coronary atherosclerosis predicts long-term morbidity and mortality after cardiac transplantation. J Am Coll Cardiol 2005; 45:1538.
  29. Reed EF, Demetris AJ, Hammond E, et al. Acute antibody-mediated rejection of cardiac transplants. J Heart Lung Transplant 2006; 25:153.
  30. Suciu-Foca N, Reed E, Marboe C, et al. The role of anti-HLA antibodies in heart transplantation. Transplantation 1991; 51:716.
  31. Petrossian GA, Nichols AB, Marboe CC, et al. Relation between survival and development of coronary artery disease and anti-HLA antibodies after cardiac transplantation. Circulation 1989; 80:III122.
  32. Cherry R, Nielsen H, Reed E, et al. Vascular (humoral) rejection in human cardiac allograft biopsies: relation to circulating anti-HLA antibodies. J Heart Lung Transplant 1992; 11:24.
  33. Dunn MJ, Crisp SJ, Rose ML, et al. Anti-endothelial antibodies and coronary artery disease after cardiac transplantation. Lancet 1992; 339:1566.
  34. Fredrich R, Toyoda M, Czer LS, et al. The clinical significance of antibodies to human vascular endothelial cells after cardiac transplantation. Transplantation 1999; 67:385.
  35. Taylor DO, Yowell RL, Kfoury AG, et al. Allograft coronary artery disease: clinical correlations with circulating anti-HLA antibodies and the immunohistopathologic pattern of vascular rejection. J Heart Lung Transplant 2000; 19:518.
  36. Hruban RH, Beschorner WE, Baumgartner WA, et al. Accelerated arteriosclerosis in heart transplant recipients is associated with a T-lymphocyte-mediated endothelialitis. Am J Pathol 1990; 137:871.
  37. Salomon RN, Hughes CC, Schoen FJ, et al. Human coronary transplantation-associated arteriosclerosis. Evidence for a chronic immune reaction to activated graft endothelial cells. Am J Pathol 1991; 138:791.
  38. Nielsen H, Symmans F, Marboe CC, et al.. Arteriopathy in human cardiac allografts: An immunopathologic study: Cell phenotyping and expression of adhesions molecules. Am J Pathol (in press).
  39. Marboe CC, Schierman SW, Rose E, et al. Characterization of mononuclear cell infiltrates in human cardiac allografts. Transplant Proc 1984; 16:1598.
  40. Doherty PC, Allan JE, Lynch F, Ceredig R. Dissection of an inflammatory process induced by CD8+ T cells. Immunol Today 1990; 11:55.
  41. Chomette G, Auriol M, Delcourt A, et al. Human cardiac transplants. Diagnosis of rejection by endomyocardial biopsy. Causes of death (about 30 autopsies). Virchows Arch A Pathol Anat Histopathol 1985; 407:295.
  42. Briscoe DM, Schoen FJ, Rice GE, et al. Induced expression of endothelial-leukocyte adhesion molecules in human cardiac allografts. Transplantation 1991; 51:537.
  43. Tanio JW, Basu CB, Albelda SM, Eisen HJ. Differential expression of the cell adhesion molecules ICAM-1, VCAM-1, and E-selectin in normal and posttransplantation myocardium. Cell adhesion molecule expression in human cardiac allografts. Circulation 1994; 89:1760.
  44. Koskinen PK, Lemström KB. Adhesion molecule P-selectin and vascular cell adhesion molecule-1 in enhanced heart allograft arteriosclerosis in the rat. Circulation 1997; 95:191.
  45. Labarrere CA, Nelson DR, Faulk WP. Endothelial activation and development of coronary artery disease in transplanted human hearts. JAMA 1997; 278:1169.
  46. Labarrere CA, Nelson DR, Miller SJ, et al. Value of serum-soluble intercellular adhesion molecule-1 for the noninvasive risk assessment of transplant coronary artery disease, posttransplant ischemic events, and cardiac graft failure. Circulation 2000; 102:1549.
  47. Bayliss J, Bailey M, Leet A, et al. Late onset antibody-mediated rejection and endothelial localization of vascular endothelial growth factor are associated with development of cardiac allograft vasculopathy. Transplantation 2008; 86:991.
  48. Kübrich M, Petrakopoulou P, Kofler S, et al. Impact of coronary endothelial dysfunction on adverse long-term outcome after heart transplantation. Transplantation 2008; 85:1580.
  49. Libby P, Warner SJ, Friedman GB. Interleukin 1: a mitogen for human vascular smooth muscle cells that induces the release of growth-inhibitory prostanoids. J Clin Invest 1988; 81:487.
  50. Raines EW, Dower SK, Ross R. Interleukin-1 mitogenic activity for fibroblasts and smooth muscle cells is due to PDGF-AA. Science 1989; 243:393.
  51. Fellström B, Dimeny E, Larsson E, et al. Importance of PDGF receptor expression in accelerated atherosclerosis-chronic rejection. Transplant Proc 1989; 21:3689.
  52. Lou H, Zhao Y, Delafontaine P, et al. Estrogen effects on insulin-like growth factor-I (IGF-I)-induced cell proliferation and IGF-I expression in native and allograft vessels. Circulation 1997; 96:927.
  53. Gullestad L, Simonsen S, Ueland T, et al. Possible role of proinflammatory cytokines in heart allograft coronary artery disease. Am J Cardiol 1999; 84:999.
  54. Miller GG, Davis SF, Atkinson JB, et al. Longitudinal analysis of fibroblast growth factor expression after transplantation and association with severity of cardiac allograft vasculopathy. Circulation 1999; 100:2396.
  55. Border WA, Noble NA. Transforming growth factor beta in tissue fibrosis. N Engl J Med 1994; 331:1286.
  56. Densem CG, Hutchinson IV, Cooper A, et al. Polymorphism of the transforming growth factor-beta 1 gene correlates with the development of coronary vasculopathy following cardiac transplantation. J Heart Lung Transplant 2000; 19:551.
  57. Nagano H, Mitchell RN, Taylor MK, et al. Interferon-gamma deficiency prevents coronary arteriosclerosis but not myocardial rejection in transplanted mouse hearts. J Clin Invest 1997; 100:550.
  58. Ternstrom L, Jeppsson A, Ricksten A, Nilsson F. Tumor necrosis factor gene polymorphism and cardiac allograft vasculopathy. J Heart Lung Transplant 2005; 24:433.
  59. Densem CG, Ray M, Hutchinson IV, et al. Interleukin-6 polymorphism: a genetic risk factor for cardiac transplant related coronary vasculopathy? J Heart Lung Transplant 2005; 24:559.
  60. Gao HZ, Hunt SA, Alderman EL, et al. Relation of donor age and preexisting coronary artery disease on angiography and intracoronary ultrasound to later development of accelerated allograft coronary artery disease. J Am Coll Cardiol 1997; 29:623.
  61. Bolad IA, Robinson DR, Webb C, et al. Impaired left ventricular systolic function early after heart transplantation is associated with cardiac allograft vasculopathy. Am J Transplant 2006; 6:161.
  62. Johnson MR. Transplant coronary disease: nonimmunologic risk factors. J Heart Lung Transplant 1992; 11:S124.
  63. Hosenpud JD, Shipley GD, Wagner CR. Cardiac allograft vasculopathy: current concepts, recent developments, and future directions. J Heart Lung Transplant 1992; 11:9.
  64. Chang G, DeNofrio D, Desai S, et al. Lipoprotein(a) levels and heart transplantation atherosclerosis. Am Heart J 1998; 136:329.
  65. DeNofrio D, Desai S, Rader DJ, et al. Changes in lipoprotein(a) concentration after orthotopic heart transplantation. Am Heart J 2000; 139:729.
  66. Escobar A, Ventura HO, Stapleton DD, et al. Cardiac allograft vasculopathy assessed by intravascular ultrasonography and nonimmunologic risk factors. Am J Cardiol 1994; 74:1042.
  67. Kapadia SR, Nissen SE, Ziada KM, et al. Impact of lipid abnormalities in development and progression of transplant coronary disease: a serial intravascular ultrasound study. J Am Coll Cardiol 2001; 38:206.
  68. Holvoet P, Van Cleemput J, Collen D, Vanhaecke J. Oxidized low density lipoprotein is a prognostic marker of transplant-associated coronary artery disease. Arterioscler Thromb Vasc Biol 2000; 20:698.
  69. Shi C, Lee WS, Russell ME, et al. Hypercholesterolemia exacerbates transplant arteriosclerosis via increased neointimal smooth muscle cell accumulation: studies in apolipoprotein E knockout mice. Circulation 1997; 96:2722.
  70. Kobashigawa JA, Katznelson S, Laks H, et al. Effect of pravastatin on outcomes after cardiac transplantation. N Engl J Med 1995; 333:621.
  71. Fateh-Moghadam S, Bocksch W, Wessely R, et al. Cytomegalovirus infection status predicts progression of heart-transplant vasculopathy. Transplantation 2003; 76:1470.
  72. Bonaros NE, Kocher A, Dunkler D, et al. Comparison of combined prophylaxis of cytomegalovirus hyperimmune globulin plus ganciclovir versus cytomegalovirus hyperimmune globulin alone in high-risk heart transplant recipients. Transplantation 2004; 77:890.
  73. Potena L, Grigioni F, Ortolani P, et al. Relevance of cytomegalovirus infection and coronary-artery remodeling in the first year after heart transplantation: a prospective three-dimensional intravascular ultrasound study. Transplantation 2003; 75:839.
  74. Sedmak DD, Roberts WH, Stephens RE, et al. Inability of cytomegalovirus infection of cultured endothelial cells to induce HLA class II antigen expression. Transplantation 1990; 49:458.
  75. Kato T, Chan MC, Gao SZ, et al. Glucose intolerance, as reflected by hemoglobin A1c level, is associated with the incidence and severity of transplant coronary artery disease. J Am Coll Cardiol 2004; 43:1034.
  76. Valantine H, Rickenbacker P, Kemna M, et al. Metabolic abnormalities characteristic of dysmetabolic syndrome predict the development of transplant coronary artery disease: a prospective study. Circulation 2001; 103:2144.
  77. Biadi O, Potena L, Fearon WF, et al. Interplay between systemic inflammation and markers of insulin resistance in cardiovascular prognosis after heart transplantation. J Heart Lung Transplant 2007; 26:324.
  78. Moons L, Shi C, Ploplis V, et al. Reduced transplant arteriosclerosis in plasminogen-deficient mice. J Clin Invest 1998; 102:1788.
  79. Labarrere CA, Pitts D, Nelson DR, Faulk WP. Vascular tissue plasminogen activator and the development of coronary artery disease in heart-transplant recipients. N Engl J Med 1995; 333:1111.
  80. Dawson S, Hamsten A, Wiman B, et al. Genetic variation at the plasminogen activator inhibitor-1 locus is associated with altered levels of plasma plasminogen activator inhibitor-1 activity. Arterioscler Thromb 1991; 11:183.
  81. Benza RL, Grenett HE, Bourge RC, et al. Gene polymorphisms for plasminogen activator inhibitor-1/tissue plasminogen activator and development of allograft coronary artery disease. Circulation 1998; 98:2248.
  82. Labarrere CA, Torry RJ, Nelson DR, et al. Vascular antithrombin and clinical outcome in heart transplant patients. Am J Cardiol 2001; 87:425.
  83. Fish RD, Nabel EG, Selwyn AP, et al. Responses of coronary arteries of cardiac transplant patients to acetylcholine. J Clin Invest 1988; 81:21.
  84. Davis SF, Yeung AC, Meredith IT, et al. Early endothelial dysfunction predicts the development of transplant coronary artery disease at 1 year posttransplant. Circulation 1996; 93:457.
  85. Hollenberg SM, Klein LW, Parrillo JE, et al. Coronary endothelial dysfunction after heart transplantation predicts allograft vasculopathy and cardiac death. Circulation 2001; 104:3091.
  86. Hollenberg SM, Klein LW, Parrillo JE, et al. Changes in coronary endothelial function predict progression of allograft vasculopathy after heart transplantation. J Heart Lung Transplant 2004; 23:265.
  87. Koglin J, Glysing-Jensen T, Mudgett JS, Russell ME. Exacerbated transplant arteriosclerosis in inducible nitric oxide-deficient mice. Circulation 1998; 97:2059.
  88. Iwata A, Sai S, Moore M, et al. Gene therapy of transplant arteriopathy by liposome-mediated transfection of endothelial nitric oxide synthase. J Heart Lung Transplant 2000; 19:1017.
  89. Tanaka M, Sydow K, Gunawan F, et al. Dimethylarginine dimethylaminohydrolase overexpression suppresses graft coronary artery disease. Circulation 2005; 112:1549.
  90. Ravalli S, Albala A, Ming M, et al. Inducible nitric oxide synthase expression in smooth muscle cells and macrophages of human transplant coronary artery disease. Circulation 1998; 97:2338.
  91. De Souza AI, Wait R, Mitchell AG, et al. Heat shock protein 27 is associated with freedom from graft vasculopathy after human cardiac transplantation. Circ Res 2005; 97:192.
  92. Cunningham DA, Crisp SJ, Barbir M, et al. Donor ACE gene polymorphism: a genetic risk factor for accelerated coronary sclerosis following cardiac transplantation. Eur Heart J 1998; 19:319.
  93. Richter M, Skupin M, Grabs R, et al. New approach in the therapy of chronic rejection? ACE- and AT1-blocker reduce the development of chronic rejection after cardiac transplantation in a rat model. J Heart Lung Transplant 2000; 19:1047.
  94. Bobik A, Grooms A, Millar JA, et al. Growth factor activity of endothelin on vascular smooth muscle. Am J Physiol 1990; 258:C408.
  95. Ravalli S, Szabolcs M, Albala A, et al. Increased immunoreactive endothelin-1 in human transplant coronary artery disease. Circulation 1996; 94:2096.
  96. Okada K, Nishida Y, Murakami H, et al. Role of endogenous endothelin in the development of graft arteriosclerosis in rat cardiac allografts: antiproliferative effects of bosentan, a nonselective endothelin receptor antagonist. Circulation 1998; 97:2346.
  97. Lim TT, Liang DH, Botas J, et al. Role of compensatory enlargement and shrinkage in transplant coronary artery disease. Serial intravascular ultrasound study. Circulation 1997; 95:855.
  98. Kobashigawa J, Wener L, Johnson J, et al. Longitudinal study of vascular remodeling in coronary arteries after heart transplantation. J Heart Lung Transplant 2000; 19:546.
  99. Wong C, Ganz P, Miller L, et al. Role of vascular remodeling in the pathogenesis of early transplant coronary artery disease: a multicenter prospective intravascular ultrasound study. J Heart Lung Transplant 2001; 20:385.
  100. Mehra MR, Uber PA, Ventura HO, et al. The impact of mode of donor brain death on cardiac allograft vasculopathy: an intravascular ultrasound study. J Am Coll Cardiol 2004; 43:806.
  101. Yamani MH, Starling RC, Cook DJ, et al. Donor spontaneous intracerebral hemorrhage is associated with systemic activation of matrix metalloproteinase-2 and matrix metalloproteinase-9 and subsequent development of coronary vasculopathy in the heart transplant recipient. Circulation 2003; 108:1724.
  102. Yamani MH, Haji SA, Starling RC, et al. Myocardial ischemic-fibrotic injury after human heart transplantation is associated with increased progression of vasculopathy, decreased cellular rejection and poor long-term outcome. J Am Coll Cardiol 2002; 39:970.
  103. Cowan B, Baron O, Crack J, et al. Elafin, a serine elastase inhibitor, attenuates post-cardiac transplant coronary arteriopathy and reduces myocardial necrosis in rabbits afer heterotopic cardiac transplantation. J Clin Invest 1996; 97:2452.
  104. Zhao XM, Hu Y, Miller GG, et al. Association of thrombospondin-1 and cardiac allograft vasculopathy in human cardiac allografts. Circulation 2001; 103:525.
  105. Yamani MH, Masri CS, Ratliff NB, et al. The role of vitronectin receptor (alphavbeta3) and tissue factor in the pathogenesis of transplant coronary vasculopathy. J Am Coll Cardiol 2002; 39:804.
  106. Haji SA, Starling RC, Avery RK, et al. Donor hepatitis-C seropositivity is an independent risk factor for the development of accelerated coronary vasculopathy and predicts outcome after cardiac transplantation. J Heart Lung Transplant 2004; 23:277.
  107. Miriuka SG, Delgado DH, Cole DE, et al. Hyperhomocysteinemia in heart transplantation: from bench to bedside. J Heart Lung Transplant 2003; 22:1069.
  108. Kutschka I, Pethig K, Harringer W, et al. Increased plasma homocysteine concentrations accelerate cardiac allograft vasculopathy. J Heart Lung Transplant 2004; 23:1260.
  109. Atkinson C, Southwood M, Pitman R, et al. Angiogenesis occurs within the intimal proliferation that characterizes transplant coronary artery vasculopathy. J Heart Lung Transplant 2005; 24:551.
  110. Martínez-Dolz L, Almenar L, Reganon E, et al. Follow-up study on the utility of von Willebrand factor levels in the diagnosis of cardiac allograft vasculopathy. J Heart Lung Transplant 2008; 27:760.
Topic 3517 Version 21.0

References

آیا می خواهید مدیلیب را به صفحه اصلی خود اضافه کنید؟