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

Mechanisms of benefit of lipid-lowering drugs in patients with coronary heart disease

Mechanisms of benefit of lipid-lowering drugs in patients with coronary heart disease
Literature review current through: Jan 2024.
This topic last updated: May 10, 2022.

INTRODUCTION — Lipid lowering with statins in patients with hypercholesterolemia has a proven survival benefit for both primary prevention (ie, in patients without clinical evidence of coronary disease) and secondary prevention (ie, in patients with established coronary disease), even when serum cholesterol concentrations are "normal" for the population or borderline high. Statins may also be beneficial in patients with heart failure. (See "Management of low density lipoprotein cholesterol (LDL-C) in the secondary prevention of cardiovascular disease" and "Statin therapy in patients with heart failure", section on 'Potential benefits and harms of statin therapy in HF'.)

TIMING AND MECHANISMS OF BENEFIT — The mechanisms by which lipid-lowering therapy (particularly with statins) is beneficial are incompletely explained by the serum low density lipoprotein (LDL) concentration at baseline or after treatment [1-6]. Although statins probably cause regression of atherosclerosis, an improvement in outcome can be demonstrated as early as six months (figure 1) [7,8], a time considered too early for significant regression. In addition, the amount of lesion regression is small compared with the magnitude of the observed clinical benefit.

A case-control study of adults with a first clinical presentation of coronary heart disease found that use of statins was associated with a decreased likelihood of that presentation being an acute myocardial infarction (odds ratio 0.45) and thus an increased likelihood of presenting with stable angina [9]. These observations suggest a mechanism of benefit with statins that might involve unstable coronary lesions.

Among the nonlipid mechanisms that may be involved are plaque stabilization, reduced inflammation, reversal of endothelial dysfunction, and decreased thrombogenicity [10,11].

Regression of atherosclerosis — Regression of the atherosclerotic lesions can occur after lipid lowering, without change in vessel wall thickness or vessel wall area, and may be clinically important [12]. One limitation to arteriographic studies is the relative lack of sensitivity to morphologic changes in atheroma, which can be better characterized by intracoronary ultrasonography (ICUS) [13], or to small changes, which might be detectable by other techniques such as B-mode ultrasonography for measurement of carotid intima thickness [14], electron beam or multidetector row computed tomography to detect coronary artery calcification [15], or high-resolution magnetic resonance imaging (MRI) [12]. (See "Coronary artery calcium scoring (CAC): Overview and clinical utilization" and "Clinical utility of cardiovascular magnetic resonance imaging".)

The benefit of lipid lowering is illustrated by the observation that in different studies, coronary angiography has shown increases in lumen diameter at two and four years or a lesser degree of progression of stenosis at three years after the onset of statin therapy [16,17].

There are limited data on the exact timing of regression of atherosclerosis, particularly coronary atherosclerosis, after statin therapy. This issue was addressed using high-resolution MRI to assess aortic and carotid artery plaques. No changes were seen at six months [12], but progressive regression was noted at 12 and 24 months [18]. The earliest change, seen at 12 months, was a reduction in plaque size followed by an increase in luminal area due to arterial remodeling [18].

Plaque stabilization — Coronary artery plaque rupture is a central component in most patients with an acute coronary syndrome. Furthermore, there is increasing evidence that many of these patients have multiple unstable plaques in different coronary arteries, suggesting widespread inflammation in the coronary circulation. Thus, intervention aimed only at the culprit lesion is not likely to be optimal and the ability of statins to induce plaque stabilization may be an important mechanism of benefit. (See "Mechanisms of acute coronary syndromes related to atherosclerosis".)

Animal models and human studies have shown that statin therapy can reduce the rate of progression of plaque development and stabilize atherosclerotic plaques that have ruptured as well as those that are vulnerable to rupture [13,19-21]. One report of 131 patients evaluated the effect of 12 months of atorvastatin therapy using ICUS [13]. Compared with placebo, atorvastatin reduced the progression of mean plaque volume or thickness (1.2 versus 9.6 mm3 for placebo) and increased the hyperechogenicity of the plaque, indicating a change in plaque composition (from lipid-rich to fibrotic and calcified) that corresponds to increased plaque stability and a reduced tendency for rupture.

A variety of factors may contribute to the pathogenesis of plaque rupture and the induction of acute coronary syndromes [22,23]. (See "Mechanisms of acute coronary syndromes related to atherosclerosis".) One important effect of statin therapy may be maintenance of integrity of the fibrous cap of the plaque, thereby protecting against plaque rupture. This effect appears to be mediated by inhibition of macrophage proliferation, reduced expression of matrix metalloproteinases (MMPs) and tissue factor (which promotes thrombus formation) by macrophages, and an increase in tissue inhibitor of metalloproteinase-1 [20,24,25]:

The inhibitory effect on MMPs may be mediated in part by statin-induced inhibition of prostaglandin synthesis [26]. In cell cultures of mouse and human macrophages, lipophilic statins (fluvastatin, simvastatin) inhibit MMP-9 activity by 20 to 40 percent in a dose-dependent manner [25]. This effect was fully reversed by mevalonate.

In animal models, certain statins induce smooth muscle cell apoptosis. This may diminish collagen biosynthesis and thus the formation of the protective fibrous cap [27]. There are differences among the statins in their effects on smooth muscle cell apoptosis and collagen biosynthesis [28,29].

The clinical relevance of the differential tissue effects of statins is uncertain; a comparison study of statins in a high-risk population at equivalent LDL-cholesterol-lowering doses would be required to sort this out. It is likely that these effects are in part due to lipid lowering. In animal models, reductions in macrophage content, MMP, and tissue factor expression can be induced by dietary modification [30,31]. Additional support for the importance of lipid lowering comes from examination of the coronary arteries and plaque morphology of 113 men who died suddenly [32]. By multivariate analysis, the serum concentrations of total cholesterol, high density lipoprotein (HDL)-cholesterol, and the ratio of total cholesterol to HDL-cholesterol were independently associated with plaque rupture.

Reduced inflammation — Inflammation appears to be another important contributor to atherosclerosis and plaque rupture. (See "Mechanisms of acute coronary syndromes related to atherosclerosis".) Elevated serum markers of inflammation, particularly C-reactive protein (CRP), are associated with progression of atherosclerosis, predict the risk of a first myocardial infarction among apparently healthy men, and are associated with a worse prognosis among patients with stable and unstable angina and those who undergo coronary stenting. (See "C-reactive protein in cardiovascular disease".)

Statin therapy, given as primary or secondary prevention, reduces the serum CRP concentration, an effect that is mostly unrelated to lipid levels at baseline or during therapy [33-36]. The fall in serum CRP begins within 14 days [36].

Additional evidence that statins may have an anti-inflammatory effect is provided by a randomized trial that found that patients with rheumatoid arthritis experienced modest clinical improvement with atorvastatin; atorvastatin also reduced CRP levels and the erythrocyte sedimentation rate compared with placebo [37]. (See "Coronary artery disease in rheumatoid arthritis: Implications for prevention and management", section on 'Lipid lowering with statins'.)

Such an antiinflammatory effect could contribute to the benefit from the early institution of statin therapy in patients with an acute coronary syndrome [38]. (See "Low-density lipoprotein-cholesterol (LDL-C) lowering after an acute coronary syndrome".) Results from PROVE IT, MIRACL, and Phase Z of the A to Z trial raise the possibility that the antiinflammatory effects of statins may differ among statins [39]. (See "Management of low density lipoprotein cholesterol (LDL-C) in the secondary prevention of cardiovascular disease".) The first two trials showed early benefit with atorvastatin [40,41], while in the last trial, despite achieving an LDL-C concentration of 68 mg/dL (1.76 mmol/L) after one month of simvastatin 40 mg daily, there was no evidence of clinical benefit and no reduction in CRP [42].

Baseline inflammatory markers — In clinical trials, statins appear to have greater effects in patients with evidence of inflammation at baseline.

The potential importance of statin-induced reduction in serum markers of inflammation was illustrated by an analysis from the secondary prevention CARE trial [43]. Patients with baseline serum concentrations of CRP and serum amyloid A in the highest quintile had a relative risk for a recurrent event 75 percent higher than those with levels in the lowest quintile (figure 2). However, in patients who were treated with pravastatin, the association between inflammation and risk was attenuated and was no longer statistically significant (figure 3).

Similar findings were noted in a prospective study of patients with angiographically severe coronary disease; the improvement in survival with statin therapy occurred primarily in those with elevated serum CRP [44]. In addition, in a prospective observational study of patients undergoing percutaneous coronary interventions, pretreatment with statins was associated with a marked improvement in survival in those patients in the highest quartile of CRP levels [45].

An analysis from the AFCAPS/TexCAPS trial of primary prevention enhanced our understanding of the interactions among statin therapy, serum lipids, serum CRP, and patient outcomes [46]. Lovastatin reduced serum CRP by almost 15 percent. The following results were noted:

Among patients with a total cholesterol-to-HDL cholesterol ratio above the median, the beneficial effect of lovastatin on coronary events was independent of serum CRP; the number of subjects needed to treat for five years to prevent one event was 47.

Lovastatin was also effective in those with a ratio that was lower than the median in whom the serum CRP level was above the median; the number of subjects needed to treat for five years to prevent one event was 43.

Lovastatin was ineffective in patients with a low ratio and a CRP level below the median. (See 'Reduction in CRP' below.)

An effect of statins unrelated to lipid lowering has also been shown in a prospective study including patients with antibodies to cytomegalovirus (CMV). Statin usage, CMV seropositivity, and serum CRP level were measured in 2315 patients with angiographically significant coronary disease [47]. After 2.4 years, patients who were CMV positive and had high serum CRP levels had a lower mortality rate with statin use (6 versus 17 percent without statins). Those who had both a low serum CRP and a negative CMV titer had no improvement in mortality rate with statins (5 versus 4 percent).

Reduction in CRP — Analyses from randomized trials that compared intensity of statin therapy suggest that beyond reduction in LDL-cholesterol, the differential effects of statins are explained, at least in part, by reduction in CRP. This issue is discussed in detail elsewhere. (See "C-reactive protein in cardiovascular disease" and "Management of low density lipoprotein cholesterol (LDL-C) in the secondary prevention of cardiovascular disease".)

Mechanism — How statins might interfere with the inflammatory response is not well understood. One possible mechanism is impairment of inflammatory cell adhesion by inhibition of the main beta-2 integrin, LFA-1 [48,49]. However, pravastatin has antiinflammatory activity but does not interfere with this integrin [48]. Other contributing factors may include reduced lipidation of intracellular proteins and reduced expression of major histocompatibility complex class II molecules on antigen-presenting cells in response to interferon, decreasing subsequent T-lymphocyte activation [49,50]. (See "Transplantation immunobiology".)

These effects could also contribute to the reduction in transplant vasculopathy and the improved outcome in heart transplant recipients treated with pravastatin. (See "Heart Transplantation: Prevention and treatment of cardiac allograft vasculopathy".)

The fall in CRP induced by statins may be mediated in part by reduced monocyte expression of proinflammatory cytokines that stimulate the release of acute phase proteins [51]. (See "Acute phase reactants".)

Reversal of endothelial dysfunction — Endothelial dysfunction is a frequent finding in atherosclerotic coronary arteries, one characteristic of which is the induction of vasoconstriction by acetylcholine rather than the expected, nitric oxide-mediated vasodilation [52-54]. Normal arterial smooth muscle tone results from a balance of vasodilating (eg, nitric oxide) and vasoconstricting (eg, endothelin) influences, with the former predominating in the resting state. (See "Coronary artery endothelial dysfunction: Basic concepts", section on 'Nitric oxide function' and "Coronary artery endothelial dysfunction: Basic concepts", section on 'Endothelial dysfunction'.)

Most [52-55], but not all [56], studies have shown that vasoconstriction associated with endothelial dysfunction can be attenuated or abolished with statin therapy, an effect that can improve overall vasodilator capacity and myocardial blood flow reserve within six weeks [57-61].

Statins improve endothelial function through one or more of the following mechanisms:

Increase in endothelial nitric oxide (NO) activity due in large part to an increase in NO synthesis [61-65].

Reduction in serum oxidized LDL as oxidized LDL, but not native LDL, downregulates endothelial NO synthase activity [66].

Inhibition of endothelin synthesis through inhibition of pre-pro-endothelin mRNA [66-68]

Improvement in endothelial integrity with a reduction in permeability to LDL cholesterol [69]

Inactivation of superoxide, which is known to impair the function of NO [62,70]

Many of the above effects appear to be independent of LDL-cholesterol lowering. This was illustrated in two four-week clinical trials comparing a statin to ezetimibe or a statin-ezetimibe combination [60,61]. In the larger trial, 60 patients with dyslipidemia, but no cardiovascular disease, were randomly assigned to simvastatin 40 mg daily, simvastatin/ezetimibe 10/10 mg daily, or placebo [61]. Compared to placebo, both treatment groups had similar 35 to 40 percent reductions in LDL-cholesterol and serum CRP [61]. In contrast, a significant improvement in flow-mediated dilatation was only seen in the high-dose simvastatin group (32.1 percent improvement versus 0.5 percent worsening with combination therapy).

Different mechanisms may be involved in the improvement in endothelial function with fibrate therapy in patients with type 2 diabetes [71]. This effect may be mediated by an elevation in serum HDL and an attenuation of postprandial lipemia and the associated oxidative stress. (See "HDL cholesterol: Clinical aspects of abnormal values", section on 'Effect of increasing HDL cholesterol on clinical outcome'.)

Decreased thrombogenicity — Thrombus formation at the site of plaque rupture appears to account for most acute coronary syndromes. (See "Mechanisms of acute coronary syndromes related to atherosclerosis".)

Lipid lowering, particularly with statin therapy, has a variety of effects that may reduce thrombus formation [72]. These include:

Reduced expression of tissue factor in endothelial cells and by macrophages in the atherosclerotic plaque [24,31,73].

Decreased prothrombin activation and thrombin generation [74,75]

Improved fibrinolytic profile [76]

Decreased platelet activation [77-79], perhaps mediated in part by an antioxidant effect [79]

These changes are either independent of [74] or only partially explained by cholesterol lowering [76,79].

The molecular mechanisms by which these events occur are becoming clearer. Both membrane bound and soluble CD40 ligand (sCD40L) interact with CD40 expressed on vascular cells, resulting in inflammatory and prothrombotic responses [80,81]. Elevated levels of sCD40L have been implicated in acute coronary syndromes and predict an increased risk of future cardiovascular events in healthy subjects [82,83].

A study of 80 patients with hypercholesterolemia found that compared with matched controls, hypercholesterolemic patients had higher levels of sCD40L, factor VIIa, and prothrombin fragment [84]. The level of sCD40L correlated with serum total and LDL-cholesterol, and it was positively associated with both in vivo platelet activation and with a procoagulant state. Inhibition of cholesterol biosynthesis with either pravastatin or cerivastatin was associated with comparable, significant reductions in sCD40L, factor VIIa, and prothrombin fragment. Atorvastatin has also been found to inhibit platelet CD40L and CD40L-mediated thrombin generation [85].

The timing and mechanism of the antithrombotic effect of statins were evaluated in an open label trial in which 30 hypercholesterolemic patients were randomly assigned to either atorvastatin 40 mg daily or a Mediterranean diet with low cholesterol intake [79]. Multiple laboratory tests of platelet activation and oxidative stress were performed during 168 hours of follow-up. The following findings were noted:

Patients assigned to the diet group showed no differences from baseline in parameters of platelet function or oxidative stress.

Measures of oxidative stress and platelet activation were significantly and progressively reduced in the atorvastatin group, with a measurable effect as early as two hours after drug initiation. These early effects appeared to be independent of cholesterol lowering.

In a separate in-vitro study, atorvastatin, in a dose-dependent manner, inhibited markers of oxidative stress and platelet activation. This study suggested that the antiplatelet effect of atorvastatin may occur by inhibition of platelet Nox2, a marker of NADPH oxidase activation.

Reduction in ventricular arrhythmias — A major cause of cardiac mortality in patients with coronary heart disease (CHD) is sudden death, which is primarily due to a life-threatening ventricular tachyarrhythmia.

Lipid lowering in patients with CHD reduces the incidence of cardiac death and, in studies of patients with an implantable cardioverter-defibrillator, reduced the rate of life-threatening ventricular arrhythmia (unstable ventricular tachycardia or ventricular fibrillation) [86,87].

Other — Statin therapy has a variety of other effects that may contribute to improved outcomes in patients with CHD:

Reduced monocyte adhesion to the endothelium [88]. Monocyte recruitment into the vascular wall is important for the initiation and progression of an atherosclerotic lesion.

Reduced oxidative modification of LDL, which is thought to play an important role in the pathogenesis of atherosclerosis [89,90].

Increased mobilization and differentiation of endothelial progenitor cells, suggesting a possible role in new vessel formation [91,92].

There is some variability among the statins and fibrates on plasma fibrinogen and viscosity, which may be risk factors for coronary disease. Several small uncontrolled studies found that plasma fibrinogen levels rise with atorvastatin and lovastatin, may rise or fall with pravastatin, and are generally unchanged with simvastatin or fluvastatin [76,93,94]. Similar variability occurs with fibrates. Plasma fibrinogen tends to rise with gemfibrozil and fall with bezafibrate and fenofibrate [95].

KIF6 gene — Some [96,97] but not all studies [98,99] have suggested that carriers of the kinesin family member 6 (KIF6) 719Arg allele might derive greater benefit from statin therapy than non-carriers [100]. In an analysis of nearly 8800 White participants from the randomized JUPITER trial, there was no significant difference in the rate of the primary outcome (first major cardiovascular event) between carriers and non-carriers of the allele in either the rosuvastatin or placebo groups [100]. We do not recommend testing for the KIF6 polymorphisms to guide statin therapy.

SUMMARY

The mechanisms by which lipid-lowering therapy (particularly with statins) is beneficial are incompletely understood. While lowering of low density lipoprotein (LDL) concentration is associated with regression of atherosclerosis, the observed benefit of (particularly statin) therapy begins within months after its initiation, making regression an unlikely cause at this early stage. (See 'Regression of atherosclerosis' above.)

Other mechanisms thought to be involved include plaque stabilization, reduction of inflammation, reversal of endothelial dysfunction, and decreased thrombogenicity. (See 'Reduced inflammation' above and 'Reversal of endothelial dysfunction' above and 'Decreased thrombogenicity' above and 'Plaque stabilization' above.)

Reduction in monocyte adhesion to the endothelium, reduced oxidative modification of LDL, and increases in mobilization and differentiation of endothelial progenitor cells have also been considered as potential benefits from lipid lowering.

  1. Thompson GR, Hollyer J, Waters DD. Percentage change rather than plasma level of LDL-cholesterol determines therapeutic response in coronary heart disease. Curr Opin Lipidol 1995; 6:386.
  2. Sacks FM, Gibson CM, Rosner B, et al. The influence of pretreatment low density lipoprotein cholesterol concentrations on the effect of hypocholesterolemic therapy on coronary atherosclerosis in angiographic trials. Harvard Atherosclerosis Reversibility Project Research Group. Am J Cardiol 1995; 76:78C.
  3. Jukema JW, Bruschke AV, van Boven AJ, et al. Effects of lipid lowering by pravastatin on progression and regression of coronary artery disease in symptomatic men with normal to moderately elevated serum cholesterol levels. The Regression Growth Evaluation Statin Study (REGRESS). Circulation 1995; 91:2528.
  4. Baseline serum cholesterol and treatment effect in the Scandinavian Simvastatin Survival Study (4S). Lancet 1995; 345:1274.
  5. Influence of pravastatin and plasma lipids on clinical events in the West of Scotland Coronary Prevention Study (WOSCOPS). Circulation 1998; 97:1440.
  6. Davignon J. Beneficial cardiovascular pleiotropic effects of statins. Circulation 2004; 109:III39.
  7. Effects of pravastatin in patients with serum total cholesterol levels from 5.2 to 7.8 mmol/liter (200 to 300 mg/dl) plus two additional atherosclerotic risk factors. The Pravastatin Multinational Study Group for Cardiac Risk Patients. Am J Cardiol 1993; 72:1031.
  8. Furberg CD, Byington RP, Crouse JR, Espeland MA. Pravastatin, lipids, and major coronary events. Am J Cardiol 1994; 73:1133.
  9. Go AS, Iribarren C, Chandra M, et al. Statin and beta-blocker therapy and the initial presentation of coronary heart disease. Ann Intern Med 2006; 144:229.
  10. Vaughan CJ, Gotto AM Jr, Basson CT. The evolving role of statins in the management of atherosclerosis. J Am Coll Cardiol 2000; 35:1.
  11. Maron DJ, Fazio S, Linton MF. Current perspectives on statins. Circulation 2000; 101:207.
  12. Corti R, Fayad ZA, Fuster V, et al. Effects of lipid-lowering by simvastatin on human atherosclerotic lesions: a longitudinal study by high-resolution, noninvasive magnetic resonance imaging. Circulation 2001; 104:249.
  13. Schartl M, Bocksch W, Koschyk DH, et al. Use of intravascular ultrasound to compare effects of different strategies of lipid-lowering therapy on plaque volume and composition in patients with coronary artery disease. Circulation 2001; 104:387.
  14. Smilde TJ, van Wissen S, Wollersheim H, et al. Effect of aggressive versus conventional lipid lowering on atherosclerosis progression in familial hypercholesterolaemia (ASAP): a prospective, randomised, double-blind trial. Lancet 2001; 357:577.
  15. Callister TQ, Raggi P, Cooil B, et al. Effect of HMG-CoA reductase inhibitors on coronary artery disease as assessed by electron-beam computed tomography. N Engl J Med 1998; 339:1972.
  16. Effect of simvastatin on coronary atheroma: the Multicentre Anti-Atheroma Study (MAAS). Lancet 1994; 344:633.
  17. Brown BG, Zhao XQ, Chait A, et al. Simvastatin and niacin, antioxidant vitamins, or the combination for the prevention of coronary disease. N Engl J Med 2001; 345:1583.
  18. Corti R, Fuster V, Fayad ZA, et al. Lipid lowering by simvastatin induces regression of human atherosclerotic lesions: two years' follow-up by high-resolution noninvasive magnetic resonance imaging. Circulation 2002; 106:2884.
  19. Williams JK, Sukhova GK, Herrington DM, Libby P. Pravastatin has cholesterol-lowering independent effects on the artery wall of atherosclerotic monkeys. J Am Coll Cardiol 1998; 31:684.
  20. Crisby M, Nordin-Fredriksson G, Shah PK, et al. Pravastatin treatment increases collagen content and decreases lipid content, inflammation, metalloproteinases, and cell death in human carotid plaques: implications for plaque stabilization. Circulation 2001; 103:926.
  21. Ambrose JA, Martinez EE. A new paradigm for plaque stabilization. Circulation 2002; 105:2000.
  22. Libby P. Molecular bases of the acute coronary syndromes. Circulation 1995; 91:2844.
  23. Falk E, Shah PK, Fuster V. Coronary plaque disruption. Circulation 1995; 92:657.
  24. Aikawa M, Rabkin E, Sugiyama S, et al. An HMG-CoA reductase inhibitor, cerivastatin, suppresses growth of macrophages expressing matrix metalloproteinases and tissue factor in vivo and in vitro. Circulation 2001; 103:276.
  25. Bellosta S, Via D, Canavesi M, et al. HMG-CoA reductase inhibitors reduce MMP-9 secretion by macrophages. Arterioscler Thromb Vasc Biol 1998; 18:1671.
  26. Cipollone F, Fazia M, Iezzi A, et al. Suppression of the functionally coupled cyclooxygenase-2/prostaglandin E synthase as a basis of simvastatin-dependent plaque stabilization in humans. Circulation 2003; 107:1479.
  27. Rosenson RS. Pluripotential mechanisms of cardioprotection with HMG-CoA reductase inhibitor therapy. Am J Cardiovasc Drugs 2001; 1:411.
  28. Knapp AC, Huang J, Starling G, Kiener PA. Inhibitors of HMG-CoA reductase sensitize human smooth muscle cells to Fas-ligand and cytokine-induced cell death. Atherosclerosis 2000; 152:217.
  29. Shiomi M, Ito T, Tsukada T, et al. Reduction of serum cholesterol levels alters lesional composition of atherosclerotic plaques. Effect of pravastatin sodium on atherosclerosis in mature WHHL rabbits. Arterioscler Thromb Vasc Biol 1995; 15:1938.
  30. Aikawa M, Rabkin E, Okada Y, et al. Lipid lowering by diet reduces matrix metalloproteinase activity and increases collagen content of rabbit atheroma: a potential mechanism of lesion stabilization. Circulation 1998; 97:2433.
  31. Aikawa M, Voglic SJ, Sugiyama S, et al. Dietary lipid lowering reduces tissue factor expression in rabbit atheroma. Circulation 1999; 100:1215.
  32. Burke AP, Farb A, Malcom GT, et al. Coronary risk factors and plaque morphology in men with coronary disease who died suddenly. N Engl J Med 1997; 336:1276.
  33. Ridker PM, Rifai N, Pfeffer MA, et al. Long-term effects of pravastatin on plasma concentration of C-reactive protein. The Cholesterol and Recurrent Events (CARE) Investigators. Circulation 1999; 100:230.
  34. Jialal I, Stein D, Balis D, et al. Effect of hydroxymethyl glutaryl coenzyme a reductase inhibitor therapy on high sensitive C-reactive protein levels. Circulation 2001; 103:1933.
  35. Albert MA, Danielson E, Rifai N, et al. Effect of statin therapy on C-reactive protein levels: the pravastatin inflammation/CRP evaluation (PRINCE): a randomized trial and cohort study. JAMA 2001; 286:64.
  36. Plenge JK, Hernandez TL, Weil KM, et al. Simvastatin lowers C-reactive protein within 14 days: an effect independent of low-density lipoprotein cholesterol reduction. Circulation 2002; 106:1447.
  37. McCarey DW, McInnes IB, Madhok R, et al. Trial of Atorvastatin in Rheumatoid Arthritis (TARA): double-blind, randomised placebo-controlled trial. Lancet 2004; 363:2015.
  38. Rosenson RS, Brown AS. Statin use in acute coronary syndromes: cellular mechanisms and clinical evidence. Curr Opin Lipidol 2002; 13:625.
  39. Nissen SE. High-dose statins in acute coronary syndromes: not just lipid levels. JAMA 2004; 292:1365.
  40. Schwartz GG, Olsson AG, Ezekowitz MD, et al. Effects of atorvastatin on early recurrent ischemic events in acute coronary syndromes: the MIRACL study: a randomized controlled trial. JAMA 2001; 285:1711.
  41. Cannon CP, Braunwald E, McCabe CH, et al. Intensive versus moderate lipid lowering with statins after acute coronary syndromes. N Engl J Med 2004; 350:1495.
  42. de Lemos JA, Blazing MA, Wiviott SD, et al. Early intensive vs a delayed conservative simvastatin strategy in patients with acute coronary syndromes: phase Z of the A to Z trial. JAMA 2004; 292:1307.
  43. Ridker PM, Rifai N, Pfeffer MA, et al. 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.
  44. Horne BD, Muhlestein JB, Carlquist JF, et al. Statin therapy, lipid levels, C-reactive protein and the survival of patients with angiographically severe coronary artery disease. J Am Coll Cardiol 2000; 36:1774.
  45. Chan AW, Bhatt DL, Chew DP, et al. Relation of inflammation and benefit of statins after percutaneous coronary interventions. Circulation 2003; 107:1750.
  46. Ridker PM, Rifai N, Clearfield M, et al. Measurement of C-reactive protein for the targeting of statin therapy in the primary prevention of acute coronary events. N Engl J Med 2001; 344:1959.
  47. Horne BD, Muhlestein JB, Carlquist JF, et al. Statin therapy interacts with cytomegalovirus seropositivity and high C-reactive protein in reducing mortality among patients with angiographically significant coronary disease. Circulation 2003; 107:258.
  48. Weitz-Schmidt G, Welzenbach K, Brinkmann V, et al. Statins selectively inhibit leukocyte function antigen-1 by binding to a novel regulatory integrin site. Nat Med 2001; 7:687.
  49. Frenette PS. Locking a leukocyte integrin with statins. N Engl J Med 2001; 345:1419.
  50. Kwak B, Mulhaupt F, Myit S, Mach F. Statins as a newly recognized type of immunomodulator. Nat Med 2000; 6:1399.
  51. Ferro D, Parrotto S, Basili S, et al. Simvastatin inhibits the monocyte expression of proinflammatory cytokines in patients with hypercholesterolemia. J Am Coll Cardiol 2000; 36:427.
  52. Egashira K, Hirooka Y, Kai H, et al. Reduction in serum cholesterol with pravastatin improves endothelium-dependent coronary vasomotion in patients with hypercholesterolemia. Circulation 1994; 89:2519.
  53. Anderson TJ, Meredith IT, Yeung AC, et al. The effect of cholesterol-lowering and antioxidant therapy on endothelium-dependent coronary vasomotion. N Engl J Med 1995; 332:488.
  54. Treasure CB, Klein JL, Weintraub WS, et al. Beneficial effects of cholesterol-lowering therapy on the coronary endothelium in patients with coronary artery disease. N Engl J Med 1995; 332:481.
  55. Tsunekawa T, Hayashi T, Kano H, et al. Cerivastatin, a hydroxymethylglutaryl coenzyme a reductase inhibitor, improves endothelial function in elderly diabetic patients within 3 days. Circulation 2001; 104:376.
  56. Vita JA, Yeung AC, Winniford M, et al. Effect of cholesterol-lowering therapy on coronary endothelial vasomotor function in patients with coronary artery disease. Circulation 2000; 102:846.
  57. Yokoyama I, Momomura S, Ohtake T, et al. Improvement of impaired myocardial vasodilatation due to diffuse coronary atherosclerosis in hypercholesterolemics after lipid-lowering therapy. Circulation 1999; 100:117.
  58. Lefer AM, Campbell B, Shin YK, et al. Simvastatin preserves the ischemic-reperfused myocardium in normocholesterolemic rat hearts. Circulation 1999; 100:178.
  59. Dupuis J, Tardif JC, Cernacek P, Théroux P. Cholesterol reduction rapidly improves endothelial function after acute coronary syndromes. The RECIFE (reduction of cholesterol in ischemia and function of the endothelium) trial. Circulation 1999; 99:3227.
  60. Landmesser U, Bahlmann F, Mueller M, et al. Simvastatin versus ezetimibe: pleiotropic and lipid-lowering effects on endothelial function in humans. Circulation 2005; 111:2356.
  61. Liu PY, Liu YW, Lin LJ, et al. Evidence for statin pleiotropy in humans: differential effects of statins and ezetimibe on rho-associated coiled-coil containing protein kinase activity, endothelial function, and inflammation. Circulation 2009; 119:131.
  62. John S, Schlaich M, Langenfeld M, et al. Increased bioavailability of nitric oxide after lipid-lowering therapy in hypercholesterolemic patients: a randomized, placebo-controlled, double-blind study. Circulation 1998; 98:211.
  63. Laufs U, Liao JK. Post-transcriptional regulation of endothelial nitric oxide synthase mRNA stability by Rho GTPase. J Biol Chem 1998; 273:24266.
  64. Laufs U, La Fata V, Plutzky J, Liao JK. Upregulation of endothelial nitric oxide synthase by HMG CoA reductase inhibitors. Circulation 1998; 97:1129.
  65. Kaesemeyer WH, Caldwell RB, Huang J, Caldwell RW. Pravastatin sodium activates endothelial nitric oxide synthase independent of its cholesterol-lowering actions. J Am Coll Cardiol 1999; 33:234.
  66. Hernández-Perera O, Pérez-Sala D, Navarro-Antolín J, et al. Effects of the 3-hydroxy-3-methylglutaryl-CoA reductase inhibitors, atorvastatin and simvastatin, on the expression of endothelin-1 and endothelial nitric oxide synthase in vascular endothelial cells. J Clin Invest 1998; 101:2711.
  67. Hernández-Perera O, Pérez-Sala D, Soria E, Lamas S. Involvement of Rho GTPases in the transcriptional inhibition of preproendothelin-1 gene expression by simvastatin in vascular endothelial cells. Circ Res 2000; 87:616.
  68. Laufs U, Liao JK. Targeting Rho in cardiovascular disease. Circ Res 2000; 87:526.
  69. van Nieuw Amerongen GP, Vermeer MA, Nègre-Aminou P, et al. Simvastatin improves disturbed endothelial barrier function. Circulation 2000; 102:2803.
  70. Kalinowski L, Dobrucki LW, Brovkovych V, Malinski T. Increased nitric oxide bioavailability in endothelial cells contributes to the pleiotropic effect of cerivastatin. Circulation 2002; 105:933.
  71. Evans M, Anderson RA, Graham J, et al. Ciprofibrate therapy improves endothelial function and reduces postprandial lipemia and oxidative stress in type 2 diabetes mellitus. Circulation 2000; 101:1773.
  72. Rosenson RS, Tangney CC. Antiatherothrombotic properties of statins: implications for cardiovascular event reduction. JAMA 1998; 279:1643.
  73. Eto M, Kozai T, Cosentino F, et al. Statin prevents tissue factor expression in human endothelial cells: role of Rho/Rho-kinase and Akt pathways. Circulation 2002; 105:1756.
  74. Undas A, Brummel KE, Musial J, et al. Simvastatin depresses blood clotting by inhibiting activation of prothrombin, factor V, and factor XIII and by enhancing factor Va inactivation. Circulation 2001; 103:2248.
  75. Szczeklik A, Musiał J, Undas A, et al. Inhibition of thrombin generation by simvastatin and lack of additive effects of aspirin in patients with marked hypercholesterolemia. J Am Coll Cardiol 1999; 33:1286.
  76. Dangas G, Badimon JJ, Smith DA, et al. Pravastatin therapy in hyperlipidemia: effects on thrombus formation and the systemic hemostatic profile. J Am Coll Cardiol 1999; 33:1294.
  77. Lacoste L, Lam JY, Hung J, et al. Hyperlipidemia and coronary disease. Correction of the increased thrombogenic potential with cholesterol reduction. Circulation 1995; 92:3172.
  78. Mayer J, Eller T, Brauer P, et al. Effects of long-term treatment with lovastatin on the clotting system and blood platelets. Ann Hematol 1992; 64:196.
  79. Pignatelli P, Carnevale R, Pastori D, et al. Immediate antioxidant and antiplatelet effect of atorvastatin via inhibition of Nox2. Circulation 2012; 126:92.
  80. Henn V, Slupsky JR, Gräfe M, et al. CD40 ligand on activated platelets triggers an inflammatory reaction of endothelial cells. Nature 1998; 391:591.
  81. Lindmark E, Tenno T, Siegbahn A. Role of platelet P-selectin and CD40 ligand in the induction of monocytic tissue factor expression. Arterioscler Thromb Vasc Biol 2000; 20:2322.
  82. Aukrust P, Müller F, Ueland T, et al. Enhanced levels of soluble and membrane-bound CD40 ligand in patients with unstable angina. Possible reflection of T lymphocyte and platelet involvement in the pathogenesis of acute coronary syndromes. Circulation 1999; 100:614.
  83. Schönbeck U, Varo N, Libby P, et al. Soluble CD40L and cardiovascular risk in women. Circulation 2001; 104:2266.
  84. Cipollone F, Mezzetti A, Porreca E, et al. Association between enhanced soluble CD40L and prothrombotic state in hypercholesterolemia: effects of statin therapy. Circulation 2002; 106:399.
  85. Sanguigni V, Pignatelli P, Lenti L, et al. Short-term treatment with atorvastatin reduces platelet CD40 ligand and thrombin generation in hypercholesterolemic patients. Circulation 2005; 111:412.
  86. Mitchell LB, Powell JL, Gillis AM, et al. Are lipid-lowering drugs also antiarrhythmic drugs? An analysis of the Antiarrhythmics versus Implantable Defibrillators (AVID) trial. J Am Coll Cardiol 2003; 42:81.
  87. De Sutter J, Tavernier R, De Buyzere M, et al. Lipid lowering drugs and recurrences of life-threatening ventricular arrhythmias in high-risk patients. J Am Coll Cardiol 2000; 36:766.
  88. Weber C, Erl W, Weber KS, Weber PC. HMG-CoA reductase inhibitors decrease CD11b expression and CD11b-dependent adhesion of monocytes to endothelium and reduce increased adhesiveness of monocytes isolated from patients with hypercholesterolemia. J Am Coll Cardiol 1997; 30:1212.
  89. Hussein O, Schlezinger S, Rosenblat M, et al. Reduced susceptibility of low density lipoprotein (LDL) to lipid peroxidation after fluvastatin therapy is associated with the hypocholesterolemic effect of the drug and its binding to the LDL. Atherosclerosis 1997; 128:11.
  90. Girona J, La Ville AE, Solà R, et al. Simvastatin decreases aldehyde production derived from lipoprotein oxidation. Am J Cardiol 1999; 83:846.
  91. Dimmeler S, Aicher A, Vasa M, et al. HMG-CoA reductase inhibitors (statins) increase endothelial progenitor cells via the PI 3-kinase/Akt pathway. J Clin Invest 2001; 108:391.
  92. Llevadot J, Murasawa S, Kureishi Y, et al. HMG-CoA reductase inhibitor mobilizes bone marrow--derived endothelial progenitor cells. J Clin Invest 2001; 108:399.
  93. Rosenson RS, Tangney CC. Beneficial effects of statins. Lancet 1996; 348:1583.
  94. Marais AD, Firth JC, Bateman ME, et al. Atorvastatin: an effective lipid-modifying agent in familial hypercholesterolemia. Arterioscler Thromb Vasc Biol 1997; 17:1527.
  95. Durrington PN, Mackness MI, Bhatnagar D, et al. Effects of two different fibric acid derivatives on lipoproteins, cholesteryl ester transfer, fibrinogen, plasminogen activator inhibitor and paraoxonase activity in type IIb hyperlipoproteinaemia. Atherosclerosis 1998; 138:217.
  96. Iakoubova OA, Tong CH, Rowland CM, et al. Association of the Trp719Arg polymorphism in kinesin-like protein 6 with myocardial infarction and coronary heart disease in 2 prospective trials: the CARE and WOSCOPS trials. J Am Coll Cardiol 2008; 51:435.
  97. Iakoubova OA, Robertson M, Tong CH, et al. KIF6 Trp719Arg polymorphism and the effect of statin therapy in elderly patients: results from the PROSPER study. Eur J Cardiovasc Prev Rehabil 2010; 17:455.
  98. Assimes TL, Hólm H, Kathiresan S, et al. Lack of association between the Trp719Arg polymorphism in kinesin-like protein-6 and coronary artery disease in 19 case-control studies. J Am Coll Cardiol 2010; 56:1552.
  99. Stewart AF, Dandona S, Chen L, et al. Kinesin family member 6 variant Trp719Arg does not associate with angiographically defined coronary artery disease in the Ottawa Heart Genomics Study. J Am Coll Cardiol 2009; 53:1471.
  100. Ridker PM, MacFadyen JG, Glynn RJ, Chasman DI. Kinesin-like protein 6 (KIF6) polymorphism and the efficacy of rosuvastatin in primary prevention. Circ Cardiovasc Genet 2011; 4:312.
Topic 4546 Version 24.0

References

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