eISSN: 1897-4252
ISSN: 1731-5530
Kardiochirurgia i Torakochirurgia Polska/Polish Journal of Thoracic and Cardiovascular Surgery
Current issue Archive Manuscripts accepted About the journal Supplements Editorial board Reviewers Abstracting and indexing Contact Instructions for authors Ethical standards and procedures
Editorial System
Submit your Manuscript
SCImago Journal & Country Rank
4/2008
vol. 5
 
Share:
Share:

FORUM EKSPERTÓW
Immunomodulation of cardiac allograft vasculopathy: beyond immunosuppression

Miguel A Ortiz
,
Gonzalo Campana
,
John R Woods
,
Carlos A. Labarrere

Kardiochirurgia i Torakochirurgia Polska 2008; 5 (4): 369–373
Online publish date: 2008/12/30
Article file
- Immunomodulation.pdf  [0.08 MB]
Get citation
 
 
Introduction
Following heart transplantation a rapid form of atherosclerosis develops known as cardiac allograft vasculopathy (CAV). As a leading cause of graft failure, CAV remains the dominant limiting factor for the long-term survival of human heart transplant recipients [1, 2]. The angiographically determined incidence of CAV is around 30% at 3 years post-transplantation [3] and by 10 years more than half of all surviving recipients have angiographic CAV [4].
Even though CAV appears to be related to the native form of atherosclerosis found in coronary artery disease (CAD), no direct association has been found between standard risk factors for CAD in the general population and the development of CAV in transplanted hearts [1, 5] and its etiology remains unknown. However, several immunological and nonimmunological mechanisms have been proposed, including mechanisms that result in the establishment of a prothrombotic microvasculature within the allografts.

Immunosuppression and CAV
The introduction of effective immunosuppressive regimens, along with a variety of therapies designed to target the immune response, have improved heart transplant outcome. Examples include the use of calcineurin inhibitors, which has been pivotal in reducing the frequency of acute rejection and improving early survival. The use of steroids, cyclosporine, tacrolimus and MMF has also had a significant beneficial impact, although these drugs have strong adverse side effects. Antiproliferative immunosuppressants have been used to inhibit T-cell and B-cell proliferation and proliferation-signal inhibitors have been used to block activation of T cells after autocrine stimulation by IL-2. The best results have been achieved with combined immunotherapy and, of the current commonly used immunosuppressive regimens, tacrolimus in combination with MMF appears to produce the greatest benefit [6].
Although the introduction of new immunosuppressive agents has shown beneficial effects on graft rejection and early survival, the development of CAV continues to be the principal limiting factor for long-term survival of heart transplant recipients, and the incidence of angiographically detected CAV has not changed appreciably over the past two decades [7]. Thus, current research is focused on identifying factors that stimulate or limit CAV in an effort to better understand the mechanisms of this disease and to develop new therapeutic strategies that might inhibit its development and progression.

Microvascular changes and CAV
During the past several years we have shown that the development of CAV in heart transplant patients is associated with several abnormalities within the cardiac microvasculature that are detectable immediately following transplantation [8-16]. These abnormalities are characterized by a heightened activation of the coagulation system, leading to a build-up of fibrin deposits in the arterial microvasculature, coupled with a loss of factors, such as antithrombin and tissue plasminogen activator (tPA), that would normally impede and remove fibrin deposits.
Patients whose cardiac allografts exhibit myocardial fibrin deposits during the first month post-transplant develop significantly more CAV and have a higher incidence of graft failure in the ensuing years than patients whose allografts are without fibrin deposits [10, 11]. Increasing levels of fibrin deposition are also associated with myocardial cell damage as evidenced by detectable levels of cardiac troponin I in the circulation of these patients [17].
To explain why some patients have early fibrin deposition within the cardiac allograft microvasculature, we have analyzed the status of the heparan sulfate proteoglycan-antithrombin anticoagulant pathway and the fibrinolytic pathway in the grafts [12-15]. This research has demonstrated that vascular antithrombin, which is found on arterial and arteriolar smooth muscle cells, arterial intima, and venous endothelium of normal hearts, is associated with the absence of microvascular fibrin deposition when present in transplanted hearts [18-21]. Conversely, we have also observed that the early loss of vascular antithrombin following heart transplantation is associated with fibrin deposition in the microvasculature and myocardial cells, particularly in areas of infarction, and with the subsequent development of CAV and cardiac graft failure [10, 11, 14, 15, 17, 21-23]. Importantly, the recovery of vascular antithrombin binding following the development of microinfarction is always associated with new capillary antithrombin binding in areas of vascular remodeling or new vessel growth, and is correlated with improved survival [23].
We have demonstrated further that when allografts lose vascular antithrombin in the first three months post-transplant and, consequently, exhibit fibrin deposits, these allografts also show a significant depletion of tPA from arteriolar smooth muscle cells [14, 15] and tPA depletion is also predictive of CAV and graft failure [13]. It is not clear what causes this depletion of tPA, although it is known that proinflammatory molecules such as C-reactive protein (CRP) can downregulate tPA [24]. It is also possible that the depletion of tPA might be related to the presence of uninhibited thrombin [13, 14]. In any case, all the above changes are known to promote the establishment of a prothrombotic microenvironment and to weaken thromboresistance [21].
Factors that activate the vascular endothelium are also thrombogenic and, therefore, increase the risk of CAV, because activated endothelial cells express tissue factor, the principal activator of the coagulation cascade. We have studied allografts with microvascular abnormalities to determine whether they also show signs of endothelial activation following transplantation, and whether such activation is associated with outcome [9, 15]. What we have observed is that when allografts show signs of fibrin deposits and loss of vascular antithrombin and tPA, they also tend to express arterial endothelial ICAM-1 and HLA-DR [15, 25], and the early expression of these markers on arterial and arteriolar endothelium correlates with the subsequent development of CAV and graft failure following cardiac transplantation [9]. There is also a correlation between the blood concentration of soluble ICAM-1 early after transplantation and the upregulation of ICAM-1 in endomyocardial biopsies of transplanted hearts, with both of these markers predicting subsequent CAV and graft failure [25].
The idea that adhesion molecules contribute to the development of CAV is further supported by studies in animal models [26-30]. It has been shown, for example, that CAV can be prevented by the short-term blockade of ICAM-1 and lymphocyte function-associated antigen-1 adhesion with monoclonal antibodies in a heterotopic heart transplantation mouse model [30]. It has also been shown that ICAM-1 participates in the neutrophil-mediated damage associated with ischemia-reperfusion injury in the heart [31].
Adhesion molecules have been found on arterial endothelium taken from CAV lesions in both animal models and human patients [26-30, 32], and in humans the expression of adhesion molecules on endothelium is considered a risk factor for the development of both native atherosclerosis [33] and CAV [8, 9].
The plasma concentration of soluble ICAM-1 is also considered a risk factor for future coronary occlusion and myocardial infarction in the general population [33]. Further studies are needed to investigate the mechanism(s) responsible for the upregulation of ICAM-1 in cardiac allografts and to explain why these allografts concomitantly show increased prothrombotic activity.

Proinflammatory molecules and CAV
Another line of investigation that we have pursued involves the association between various inflammatory factors and the development of CAV following heart transplantation. One such factor, C-reactive protein (CRP), is known to be an important risk factor for native atherosclerosis and CAD in the general population [34]. Elevated plasma levels of CRP have been shown to be predictive of subsequent cardiovascular events among apparently healthy men [35-38] and women [39], patients with stable and unstable angina [40-44], and patients with a previous history of myocardial infarction [45]. Elevated CRP serum levels may promote atherosclerosis through its effect on adhesion molecule expression, since it has been shown that CRP induces ICAM-1 expression in coronary artery endothelial cells [46, 47]. Proinflammatory molecules, such as CRP, also downregulate tPA [48]. Possible links between CRP and adhesion molecule expression, and between CRP and atherosclerosis, have been reported in the clinical research literature and we and others have demonstrated an association between elevated CRP levels and the subsequent development of CAV [49] and graft failure [49, 50] in cardiac transplant recipients.
Despite the wealth of epidemiological and in vitro evidence suggesting that CRP is a key contributor to the development of atherosclerosis [35, 47, 51-53], mechanistic evidence from experimental mouse models on the role of CRP has been controversial. Some investigators have observed accelerated progression of atherosclerosis in CRP-transgenic apoE–/– mice compared with CRP-nontransgenic controls [54]. Others, however, have reported that CRP is not proatherogenic in either apoE–/– [55-57] or LDL receptor-deficient mice [58]. Schwedler et al. [59] have shown that pentameric native CRP (nCRP) increases atherosclerosis, while modified monomeric CRP reduces atherosclerosis in apoE–/– mice, and Tennent et al. [60] reported that transgenic human CRP is neither proatherogenic, proatherothrombotic, nor proinflammatory in apoE–/– mice. A recent study by Kovacs et al. [61] showed that CRP is antiatherogenic in CRP-transgenic LDL receptor-deficient apoB100 mice. Our own research has shown that pentameric CRP is not proatherosclerotic or proinflammatory in apoE–/– mice [62]. Thus, the role of CRP in atherosclerosis, whether spontaneous or transplant-associated, needs to be elucidated further before serious consideration can be given to the development of new therapeutic strategies directed to antagonize CRP.

Immunomodulation:
Beyond immunosuppression

The role of innate immunity in the prevention or inhibition of transplant-associated CAV and graft failure represents another promising avenue for investigation. We have been especially interested in exploring the possible beneficial role that the early presence of IgM antibodies may play following heart transplantation. Preliminary data from our laboratory reported 15 years ago were the first, and the only data to date, showing a protective effect for IgM antibodies in transplanted human heart patients [63]. Those data indicated that the early presence of these antibodies is associated with reduced morbidity and mortality when compared with patients lacking IgM reactivity. Furthermore, the absence of IgM was significantly related to the deposition of fibrin within the heart, which, as discussed above, has a detrimental effect on allograft and patient survival.
Our initial report was supported by other studies that showed an atheroprotective effect of natural IgM antibodies in mouse models of atherosclerosis [64]. These innate antibodies are present even in naive germ-free mice in the absence of any exogenous antigen exposure [65]. Binder et al. have proposed that these antibodies confer their atheroprotective effect by responding to oxidation-specific epitopes that constitute a “self-altered” danger signal, which occurs as a result of oxidative stress [66]. Oxidative stress may arise from events that occur when cells undergo apoptosis or in association with endothelial accumulation of oxidized low-density lipoproteins (ox-LDL). Either of these events could occur within the allografts of heart transplant recipients. Based on the experience gained from recent animal studies, we hypothesize that the presence of IgM antibodies in human cardiac allografts is associated with reduced CAV and reduced morbidity and mortality because of their capacity to remove the source of oxidative stress related to either the accumulation of cells undergoing apoptosis, the accumulation of endothelial ox-LDL, or both.

Summary and conclusions
Despite the success of immunosuppressive agents in the management of graft rejection in heart transplant recipients, the current leading cause of long-term graft failure, CAV, continues to be a significant problem and the principal factor limiting long-term survival in these patients. Research conducted during the past two decades by us and others suggests several promising avenues for developing new therapeutic approaches to prevent or control the progression of CAV. Some of these avenues are familiar, such as: (1) approaches that seek to block the coagulation system, either directly by preventing activation of the coagulation pathway, or indirectly by promoting factors, such as antithrombin and tPA, that hold it in check; (2) approaches that seek to prevent the activation of arterial endothelium and block the coagulation cascade by suppressing the expression of cell adhesion molecules and tissue factor; and (3) approaches that seek to prevent the chronic inflammation that seems to induce a prothrombotic state by blocking the effects of proinflammatory proteins such as CRP, or its principal activator, interleukin-6. Although all of these factors are familiar and considered to be important disease markers in heart transplantation, the specific biochemical mechanisms responsible for fibrin deposition in the microvessels, activation of the arterial endothelium, promotion of a procoagulant and proinflammatory microenvironment, and, ultimately, the establishment of an atherothrombotic vasculature in heart transplant patients are not completely understood.
Other avenues suggested by the research literature are, perhaps, less familiar, involving approaches that would seek to stimulate the innate immune response in order to counter the adverse effects of oxidative stress arising from cell apoptosis or the accumulation of ox-LDL, thereby establishing an atheroprotective microenvironment. The possible beneficial role of IgM antibodies is especially interesting in this regard. We are currently studying the possibility of reinforcing the immunosuppressive approaches currently used in heart transplantation with an immunomodulation approach designed to favor the production of IgM natural antibodies directed against oxidization-specific epitopes.
The enhancement of atheroprotective factors early after transplantation, when most oxidative and inflammatory processes occur, may be extremely relevant to counteract the effects of proinflammatory and proatherogenic molecules.

References
1. Gao SZ, Hunt SA, Schroeder JS, Alderman EL, Hill IR, Stinson EB. Early development of accelerated graft coronary artery disease: risk factors and course. J Am Coll Cardiol 1996; 28: 673-679.
2. Miller LW, Wolford TL, Donohue TJ, Drury JH. Cardiac allograft vasculopathy: new insights from intravascular ultrasound and coronary flow measurements. Transplant Rev 1995; 9: 77-96.
3. Mehra MR. Contemporary concepts in prevention and treatment of cardiac allograft vasculopathy. Am J Transplant 2006; 6: 1248-1256.
4. Taylor DO, Edwards LB, Aurora P, Christie JD, Dobbels F, Kirk R, Rahmel AO, Kucheryavaya AY, Hertz MI. Registry of the International Society for Heart and Lung Transplantation: Twenty-fifth Official Adult Heart Transplant Report – 2008. J Heart Lung Transplant 2008; 27: 943-956.
5. Gao SZ, Hunt SA, Schroeder JS, Alderman E, Hill IR, Stinson EB. Does rapidity of development of transplant coronary artery disease portend a worse prognosis? J Heart Lung Transplant 1994; 13: 1119-1124.
6. Kobashigawa JA, Patel J K. Immunosuppression for heart transplantation: where are we now? Nat Clin Pract Cardiovasc Med 2006; 3: 203-212.
7. Dhaliwal A, Thohan V. Cardiac allograft vasculopathy: the Achilles’ heel of long-term survival after cardiac transplantation. Curr Atheroscler Rep 2006; 8: 119-130.
8. Labarrere CA, Pitts D, Nelson DR, Faulk WP. Coronary artery disease in cardiac allografts: association with arteriolar endothelial HLA-DR and ICAM-1 antigens. Transplant Proc 1995; 27: 1939-1940.
9. Labarrere CA, Nelson DR, Faulk WP. Endothelial activation and development of coronary artery disease in transplanted human hearts. JAMA 1997; 278: 1169-1175.
10. Labarrere CA, Pitts D, Halbrook H, Faulk WP. Tissue plasminogen activator, plasminogen activator inhibitor-1, and fibrin as indexes of clinical course in cardiac allograft recipients. An immunocytochemical study. Circulation 1994; 89: 1599-1608.
11. Labarrere CA, Nelson DR, Faulk WP. Myocardial fibrin deposits in the first month after transplantation predict subsequent coronary artery disease and graft failure in cardiac allograft recipients. Am J Med 1998; 105: 207-213.
12. Labarrere CA, Pitts D, Nelson DR, Faulk WP. Coronary artery disease in cardiac allografts: association with depleted arteriolar tissue plasminogen activator. Transplant Proc 1995; 27: 1941-1943.
13. 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: 1111-1116.
14. Faulk WP, Labarrere CA, Nelson DR, Pitts D. Hemostasis, fibrinolysis, and natural anticoagulation in transplant vascular sclerosis. J Heart Lung Transplant 1995; 14: S158-164.
15. Labarrere CA, Faulk WP. Antithrombin determinants of coronary artery disease in transplanted human hearts. Semin Hematol 1995; 32: 61-66.
16. Faulk WP, Labarrere CA, Torry RJ, Nelson DR. Serum cardiac troponin-T concentrations predict development of coronary artery disease in heart transplant patients. Transplantation 1998; 66: 1335-1339.
17. Labarrere CA, Nelson DR, Cox CJ, Pitts D, Kirlin P, Halbrook H. Cardiac-specific troponin I levels and risk of coronary artery disease and graft failure following heart transplantation. JAMA 2000; 284: 457-464.
18. Labarrere CA, Pitts D, Halbrook H, Faulk WP. Natural anticoagulant pathways in normal and transplanted human hearts. J Heart Lung Transplant 1992;
11: 342-347.
19. Faulk WP, Labarrere CA, Pitts D, Halbrook H. Vascular lesions in biopsy specimens devoid of cellular infiltrates: qualitative and quantitative immunocytochemical studies of human cardiac allografts. J Heart Lung Transplant 1993;
12: 219-229.
20. Faulk WP, Labarrere CA. Fibrinolytic and anticoagulant control of hemostasis in human cardiac and renal allografts. Major Probl Pathol 1994; 30: 49-65.
21. Faulk WP, Labarrere CA. Antithrombin III in normal and transplanted human hearts: indications of vascular disease. Semin Hematol 1994; 31: 26-33.
22. Faulk WP, Labarrere CA. Modulation of vascular antithrombin III in human cardiac allografts. Haemostasis 1993; 23 (Suppl 1): 194-201.
23. Labarrere CA, Torry RJ, Nelson DR, Miller SJ, Pitts DE, Kirlin PC, Halbrook HG. Vascular antithrombin and clinical outcome in heart transplant patients. Am J Cardiol 2001; 87: 425-431.
24. Labarrere CA, Jaeger BR. Role of C-reactive protein as a cardiovascular risk predictor. In: Subclinical Atherosclerosis. Assessing the Risks. Bianco JA (ed.). Taylor and Francis, 2006; 117-135.
25. Labarrere CA, Nelson DR, Miller SJ, Nieto JM, Conner JA, Pitts DE, Kirlin PC, Halbrook HG. 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: 1549-1555.
26. Fuster V, Poon M, Willerson JT. Learning from the transgenic mouse: endothelium, adhesive molecules, and neointimal formation. Circulation 1998; 97: 16-18.
27. Adams DH, Russell ME, Hancock WW, Sayegh MH, Wyner LR, Karnovsky MJ. Chronic rejection in experimental cardiac transplantation: studies in the Lewis-F344 model. Immunol Rev 1993; 134: 5-19.
28. Karnovsky MJ, Russell ME, Hancock W, Sayegh MH, Adams DH. Chronic rejection in experimental cardiac transplantation in a rat model. Clin Transplant 1994; 8: 308-312.
29. Koskinen PK, Lemstrom KB. Adhesion molecule P-selectin and vascular cell adhesion molecule-1 in enhanced heart allograft arteriosclerosis in the rat. Circulation 1997; 95: 191-196.
30. Suzuki, J, Isobe, M, Yamazaki, S, Horie, S, Okubo, Y, Sekiguchi, M, Inhibition of accelerated coronary atherosclerosis with short-term blockade of intercellular adhesion molecule-1 and lymphocyte function- associated antigen-1 in
a heterotopic murine model of heart transplantation. J Heart Lung Transplant 1997; 16: 1141-1148.
31. Poston RS, Jr, Billingham ME, Pollard J, Hoyt EG, Robbins RC. Effects of increased ICAM-1 on reperfusion injury and chronic graft vascular disease. Ann Thorac Surg 1997; 64: 1004-1012.
32. Salomon RN, Hughes CC, Schoen FJ, Payne DD, Pober JS, Libby P. Human coronary transplantation-associated arteriosclerosis. Evidence for a chronic immune reaction to activated graft endothelial cells. Am J Pathol 1991; 138: 791-798.
33. Ridker PM, Hennekens CH, Roitman-Johnson B, Stampfer MJ, Allen J. Plasma concentration of soluble intercellular adhesion molecule 1 and risks of future myocardial infarction in apparently healthy men. Lancet 1998; 351: 88-92.
34. Shah PK. Circulating markers of inflammation for vascular risk prediction: are they ready for prime time. Circulation 2000; 101: 1758-1759.
35. Ridker PM, Cushman M, Stampfer MJ, Tracy RP, Hennekens CH. Inflammation, aspirin, and the risk of cardiovascular disease in apparently healthy men.
N Engl J Med 1997; 336: 973-979.
36. Ridker PM, Cushman M, Stampfer MJ, Tracy RP, Hennekens CH. Plasma concentration of C-reactive protein and risk of developing peripheral vascular disease. Circulation 1998; 97: 425-428.
37. Ridker PM, Glynn RJ, Hennekens CH. C-reactive protein adds to the predictive value of total and HDL cholesterol in determining risk of first myocardial infarction. Circulation 1998; 97: 2007-2011.
38. Koenig W, Sund M, Frohlich M, Fischer HG, Lowel H, Doring A, Hutchinson WL, Pepys MB. C-Reactive protein, a sensitive marker of inflammation, predicts future risk of coronary heart disease in initially healthy middle-aged men: results from the MONICA (Monitoring Trends and Determinants in Cardiovascular Disease) Augsburg Cohort Study, 1984 to 1992. Circulation 1999; 99: 237-242.
39. Ridker PM, Buring JE, Shih J, Matias M, Hennekens CH. Prospective study of C-reactive protein and the risk of future cardiovascular events among apparently healthy women. Circulation 1998; 98: 731-733.
40. Thompson SG, Kienast J, Pyke SD, Haverkate F, van de Loo JC. Hemostatic factors and the risk of myocardial infarction or sudden death in patients with angina pectoris. European Concerted Action on Thrombosis and Disabilities Angina Pectoris Study Group. N Engl J Med 1995; 332: 635-641.
41. Liuzzo G, Biasucci LM, Gallimore JR, Grillo RL, Rebuzzi AG. Pepys MB, Maseri A. The prognostic value of C-reactive protein and serum amyloid A protein in severe unstable angina. N Engl J Med 1994; 331: 417-424.
42. Haverkate F, Thompson SG, Pyke SD, Gallimore JR, Pepys MB. Production of C-reactive protein and risk of coronary events in stable and unstable angina. European Concerted Action on Thrombosis and Disabilities Angina Pectoris Study Group. Lancet 1997; 349: 462-466.
43. Rebuzzi AG, Quaranta G, Liuzzo G, Caligiuri G, Lanza GA, Gallimore JR, Grillo RL, Cianflone D, Biasucci LM, Maseri A. Incremental prognostic value of serum levels of troponin T and C-reactive protein on admission in patients with unstable angina pectoris. Am J Cardiol 1998; 82: 715-719.
44. Biasucci LM, Liuzzo G, Grillo RL, Caligiuri G, Rebuzzi AG, Buffon A, Summaria F, Ginnetti F, Fadda G, Maseri A. Elevated levels of C-reactive protein at discharge in patients with unstable angina predict recurrent instability. Circulation 1999; 99: 855-860.
45. Ridker PM, Rifai N, Pfeffer MA, Sacks FM, Moye LA, Goldman S, Flaker GC, Braunwald E. Inflammation, pravastatin, and the risk of coronary events after myocardial infarction in patients with average cholesterol levels. Cholesterol and Recurrent Events (CARE) Investigators. Circulation 1998; 98: 839-844.
46. Pasceri V, Willerson JT, Yeh ET. Direct proinflammatory effect of C-reactive protein on human endothelial cells. Circulation 2000; 102: 2165-2168.
47. Labarrere CA, Zaloga GP. C-reactive protein: from innocent bystander to pivotal mediator of atherosclerosis. Am J Med 2004; 117: 499-507.
48. Singh U, Devaraj S, Jialal I. C-reactive protein decreases tissue plasminogen activator activity in human aortic endothelial cells: evidence that C-reactive protein is a procoagulant. Arterioscler Thromb Vasc Biol 2005; 25: 2216-2221.
49. Labarrere CA, Lee JB, Nelson DR, Al-Hassani M, Miller SJ, Pitts DE. C-reactive protein, arterial endothelial activation, and development of transplant coronary artery disease: a prospective study. Lancet 2002; 360: 1462-1467.
50. Eisenberg MS, Chen HJ, Warshofsky MK, Sciacca RR, Wasserman HS, Schwartz A, Rabbani LE. Elevated levels of plasma C-reactive protein are associated with decreased graft survival in cardiac transplant recipients. Circulation 2000; 102: 2100-2104.
51. Verma S, Szmitko PE, Ridker PM. C-reactive protein comes of age. Nat Clin Pract Cardiovasc Med 2005; 2: 29-36.
52. Ridker PM, Hennekens CH, Buring JE, Rifai N. C-reactive protein and other markers of inflammation in the prediction of cardiovascular disease in women. N Engl J Med 2000; 342: 836-843.
53. Torzewski M, Carsten R, Mortensen RF, Zwaka TP, Bienek M, Waltenberger J, Koenig W, Schmitz G, Hombach V, Torzewski J. C-reactive protein in the arterial intima: Role of C-Reactive Protein Receptor-Dependent Monocyte Recruitment in Atherogenesis. Arterioscler Thromb Vasc Biol 2000; 20: 2094-2099.
54. Paul A, Ko KW, Li L, Yechoor V, McCrory MA, Szalai AJ, Chan L. C-reactive protein accelerates the progression of atherosclerosis in apolipoprotein E-deficient mice. Circulation 2004; 109: 647-655.
55. Hirschfield GM, Gallimore JR, Kahan MC, Hutchinson WL, Sabin CA, Benson GM, Dhillon AP, Tennent GA, Pepys MB. Transgenic human C-reactive protein is not proatherogenic in apolipoprotein E-deficient mice. Proc Nat Acad Sci USA 2005; 102: 8309-8314.
56. Reifenberg K, Lehr HA, Baskal D, Wiese E, Schaefer SC, Black S, Samols D, Torzewski M, Lackner KJ, Husmann M, Blettner M, Bhakdi S. Role of C-reactive protein in atherogenesis: can the apolipoprotein E knockout mouse provide the answer? Arterioscler Thromb Vasc Biol 2005; 25: 1641-1646.
57. Trion A, de Maat MP, Jukema JW, van der Laarse A, Maas MC, Offerman EH, Havekes LM, Szalai AJ, Princen HM, Emeis JJ. No effect of C-reactive protein on early atherosclerosis development in apolipoprotein E*3-leiden/human
C-reactive protein transgenic mice. Arterioscler Thromb Vasc Biol 2005;
25: 1635-1640.
58. Torzewski M, Reifenberg K, Cheng F, Wiese E, Kupper I, Crain J, Lackner KJ, Bhakdi S. No effect of C-reactive protein on early atherosclerosis in LDLR(-/-) / human C-reactive protein transgenic mice. Thromb Haemost 2008; 99: 196-201.
59. Schwedler SB, Amann K, Wernicke K, Krebs A, Nauck M, Wanner C, Potempa LA, Galle J. Native C-reactive protein increases whereas modified C-reactive protein reduces atherosclerosis in apolipoprotein E-knockout mice. Circulation 2005; 112: 1016-1023.
60. Tennent GA, Hutchinson WL, Kahan MC, Hirschfield GM, Gallimore JR, Lewin J, Sabin CA, Dhillon AP, Pepys MB. Transgenic human CRP is not pro-atherogenic, pro-atherothrombotic or pro-inflammatory in apoE-/- mice. Atherosclerosis 2008; 196: 248-255.
61. Kovacs A, Tornvall P, Nilsson R, Tegnér J, Hamsten A, Björkegren J. Human
C-reactive protein slows atherosclerosis development in a mouse model with human-like hypercholesterolemia. Proc Nat Acad Sci USA 2007; 104: 13768-13773.
62. Ortiz MA, Campana GL, Woods JR, Boguslawski G, Sosa MJ, Walker CL, Labarrere CA. Continuously-infused human C-reactive protein is not proatherosclerotic or proinflammatory in apolipoprotein E-deficient mice. 2008; submitted.
63. Labarrere CA, Pitts D, Halbrook H, Faulk WP. Immunoglobulin M antibodies in transplanted human hearts. J Heart Lung Transplant 1993; 12: 394-402.
64. Chou MY, Hartvigsen K, Hansen LF, Fogelstrand L, Shaw PX, Boullier A, Binder CJ, Witztum JL. Oxidation-specific epitopes are important targets of innate immunity. J Intern Med 2008; 263: 479-488.
65. Binder CJ, Hartvigsen K, Witztum JL. Promise of immune modulation to inhibit atherogenesis. J Am Coll Cardiol 2007; 50: 547-550.
66. Binder CJ, Shaw PX, Chang MK, Boullier A, Hartvigsen K, Horkko S, Miller YI, Woelkers DA, Corr M, Witztum JL. The role of natural antibodies in atherogenesis. J Lipid Res 2005; 46: 1353-1363.
Copyright: © 2008 Polish Society of Cardiothoracic Surgeons (Polskie Towarzystwo KardioTorakochirurgów) and the editors of the Polish Journal of Cardio-Thoracic Surgery (Kardiochirurgia i Torakochirurgia Polska). This is an Open Access article distributed under the terms of the Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International (CC BY-NC-SA 4.0) License (http://creativecommons.org/licenses/by-nc-sa/4.0/), allowing third parties to copy and redistribute the material in any medium or format and to remix, transform, and build upon the material, provided the original work is properly cited and states its license.
Quick links
© 2024 Termedia Sp. z o.o.
Developed by Bentus.