2,327
Views
9
CrossRef citations to date
0
Altmetric
Review Article

Systematic reviews and meta-analyses for more profitable strategies in peripheral artery disease

, , , , , , & show all
Pages 475-489 | Received 25 Mar 2014, Accepted 03 Jun 2014, Published online: 21 Jul 2014

Abstract

In the peripheral arteries, a thrombus superimposed on atherosclerosis contributes to the progression of peripheral artery disease (PAD), producing intermittent claudication (IC), ischemic necrosis, and, potentially, loss of the limb. PAD with IC is often undiagnosed and, in turn, undertreated. The low percentage of diagnosis (∼30%) in this setting of PAD is of particular concern because of the potential worsening of PAD (amputation) and the high risk of adverse vascular outcomes (vascular death, coronary artery disease, stroke). A Medline literature search of the highest-quality systematic reviews and meta-analyses of randomized controlled trials documents that, due to risk of bias, imprecision, and indirectness, the overall quality of the evidence concerning diagnostic tools and antithrombotic interventions in PAD is generally low. Areas of research emerge from the information collected. Appropriate treatments for PAD patients will only derive from ad-hoc studies. Innovative imaging techniques are needed to identify PAD subjects at the highest vascular risk. Whether IC unresponsive to physical exercise and smoking cessation identifies those with a heritable predisposition to more severe vascular events deserves to be addressed. Devising ways to improve prevention of vascular events in patients with PAD implies a co-ordinated approach in vascular medicine.

Key messages

  • The overall quality of the evidence concerning diagnostic tools and antithrombotic interventions in PAD is low in most cases.

  • New antithrombotic treatment is a major target to improve prevention of vascular events in patients with PAD.

  • Innovative imaging techniques should be explored to identify PAD subjects at the highest vascular risk.

Introduction

Atherothrombosis (thrombus formation superimposed upon atherosclerosis) is an unpredictable, sudden disruption (rupture or erosion/fissure) of an atherosclerotic plaque, which leads to platelet activation and thrombus formation (Citation1). Rupture/fissure of a plaque acts as a stimulus for atherothrombosis (Citation2). From a clinical point of view, atherothrombosis is a progressive, generalized disorder with manifestations—either acute or chronic and often multiple in any single patient—affecting the coronary, cerebral, and peripheral circulation (Citation3). In the case of a non-occlusive thrombus, ischemic symptoms are temporary and thrombosis can contribute to plaque growth through formation and resolution of subclinical platelet thrombi (thrombus formation is a dynamic process in which platelets aggregate but also spontaneously disaggregate, eventually leading to embolization of platelet aggregates from an evolving thrombus) (Citation4). In the case of an occlusive thrombus, there will be an acute ischemic syndrome in the coronary, cerebral, or peripheral vascular territory depending on the localization of the atherosclerotic plaque, potentially leading to permanent tissue damage. Because of the generalized and progressive nature of atherothrombosis, symptoms of the disease in one vascular bed are highly predictive of the risk of further ischemic events elsewhere. Cardiovascular and cerebrovascular events and peripheral artery disease (PAD) are thus part of a continuum of disease with the common underlying pathophysiology of atherothrombosis (Citation5).

In the peripheral arteries, a thrombus superimposed on atherosclerosis contributes to the progression of PAD, producing intermittent claudication (leg pain on walking that is relieved by rest), ischemic necrosis, and, potentially, loss of the limb. PAD affects about 8–10 million Americans, and every year it causes 500,000 hospitalizations and 100,000 angiograms in the US (Citation6). Significant coronary artery disease (in at least one coronary artery) has been documented in 60%–80% of patients with PAD, and hemodynamically significant carotid artery stenosis (by duplex ultrasound) has been found in 12%–25% (Citation7). Accordingly, the annual overall major vascular event rate (acute myocardial infarction, ischemic stroke and vascular death) is ∼5%–7% in PAD patients. The risk of AMI is increased by 20%–60%, whereas the risk of coronary death is increased 2–6-fold. PAD is associated with a 40% increase in the risk of stroke, and PAD severity is positively associated with the incidence of transient ischemic attacks (TIA) and stroke (Citation6).

The clinical spectrum of PAD is widely variable: patients may either be asymptomatic or, because of an impaired equilibrium between oxygen demand and supply, have pain as a result of a minimal exercise (e.g. walking) (Citation8). ACC/AHA recommendations for the diagnosis of PAD are reported in . Most asymptomatic patients with PAD will be identified through ankle-brachial index (ABI) screening (Citation9). The ABI is the ratio of the highest systolic blood pressure in the lower limb to that of the arm. The TASC-II (Citation10) guidelines also indicate the ABI as an easy, reliable means for the evaluation of PAD severity (). Resting ABI should be measured in both legs in all new patients with PAD of any severity to confirm the diagnosis and establish a baseline.

Table I. ACC/AHA 2013 recommendations for the diagnosis of PAD by non-invasive tools.

Table II. TASC II guidelines for ABI interpretation.

PAD (ABI < 0.9) with intermittent claudication (IC) is often undiagnosed and, in turn, undertreated. In 5298 Italian patients at moderate vascular risk, with no overt vascular diseases nor diabetes mellitus (DM) (Citation11), 0.02% had an ABI ≤ 0.4; 22.85% had an ABI ranging from 0.4 to 0.9; 23.9% had an ABI ranging from 0.91 to 0.99; 52.35% had an ABI ranging from 1.0 to 1.29; and 0.88% had an ABI ≥ 1.30. In 2027 Italian patients with non-valvular atrial fibrillation (AF), 21% had an ABI ≤ 0.9; 69% had an ABI ranging from 0.91 to 1.40; and ∼10% had an ABI ≥ 1.40 (Citation12). The low percentage of diagnosis (∼30%) among those with an ABI < 0.9 is of particular concern because of the high risk of adverse outcomes related to the worsening of PAD. Moreover, a low ABI (< 0.90) is a predictor of AMI, stroke, and vascular mortality independently of established vascular risk factors (Citation13–15). Among those with ABI < 0.9, ∼25% will experience worsening claudication necessitating surgical repair or amputation. During 5 years of follow-up (Citation16), 10%–20% of patients with IC would be expected to experience non-fatal AMI or stroke. In addition, death from coronary artery disease, other vascular diseases, and non-cardiovascular causes would be expected in 30%. In a longer (10-y) follow-up (Citation17), about 55% of PAD (ABI < 0.9) patients died of cardiovascular disease, 10% of cerebrovascular disease, and 25% of non-vascular reasons. Less than 10% died of other vascular events (mostly, aortic aneurysms in the abdomen). In another follow-up (Citation18), 10-y mortality was 61.8% in males with symptomatic PAD (ABI < 0.9); in comparison, in males without PAD (ABI < 0.9) mortality was 16.9%. Mortality rates in females were 33.3% and 11.6%, respectively. Less than 25% of patients with PAD (ABI < 0.9) survived for 10 y, vascular mortality being the dominant cause of death in that setting. After correction for established risk factors, PAD was an independent predictor of death. In the latter report, the subjects were classified as normal (no evidence of PAD), asymptomatic (no claudication), symptomatic (with claudication), and as subjects with severe PAD (claudication+ abnormal diagnostic tests). The risk of death was related to the severity of PAD and was as strong as that for cancer. The latter has emerged from 744 patients with PAD (Citation19): in a 5-y follow-up, those with severe PAD (ABI < 0.4) had 56% probability of surviving. Based on data collected in parallel (1986–1993), this figure was comparable with the 5-y survival curves (52%) in Caucasian patients with non-Hodgkin lymphomas.

These data set the stage for an extraordinary high morbidity and mortality in patients with PAD. As a matter of fact, over the last decades, while no currently available specific treatment for PAD is associated with a significant reduction in morbidity and mortality rates, screening programs for primary prevention of coronary artery disease and use of statins, ACE and platelet inhibitors, β-blockers, etc. have dramatically decreased the risk of major cardiovascular events (and of progression in arterial occlusion) in the PAD setting (Citation20).

Critical limb ischemia (CLI; ABI < 0.4), is observed in 12% of the PAD population (Citation21). In most cases, CLI is an advanced thrombotic complication of PAD, due to inadequate resting blood flow to the lower limbs, and marked by rest pain, ulceration, and eventually gangrene and loss of the limb (Citation22). The limb typically has developed a collateral blood supply, and the final occlusion of the vessel often is not immediately limb threatening with slow progression of disease (Citation23). CLI is seldom the result of an acute event (e.g. embolism, thrombosis, or trauma). Approximately 80% of emboli originate in the heart (e.g. left atrial appendage, left ventricular apex, cardiac valves) (Citation24). In the remaining cases, they originate from the aorta or peripheral vessels or from the veins (with migration through patent foramen ovale and atrial septal defects). Patients with CLI are candidates for prompt revascularization. CLI increases mortality: in the first year after the diagnosis 25% of patients (45% with amputation) will die, and 30% of them will have amputations, whereas only 45% will survive with both legs. After 5 years more than 60% of patients have died (Citation25).

Literature search method

In an attempt to identify potential reasons why PAD with IC is underdiagnosed and undertreated, the following questions have been raised: 1) Is the overall quality of the evidence concerning diagnostic tools for PAD as good as that for other atherothrombotic conditions? And 2) Is the overall quality of the evidence concerning antithrombotic interventions for PAD as good as that for other atherothrombotic conditions?

To estimate absolute benefits, harms, and limitations associated with a given treatment/diagnostic tool, a Medline literature search of the highest-quality published systematic reviews and meta-analyses of randomized controlled trials in the area was performed using the key words ‘diagnostic tools AND PAD’ and ‘treatments AND PAD’, and references of relevance were selected manually. Other references were either provided by the authors or obtained from the reference lists within the relevant selected articles. Databases were updated to August 2013.

Non-invasive and invasive tools for the diagnosis of PAD

Non-invasive imaging is mandatory to detect the anatomy, morphology (e.g. subclinical non-obstructive), and characteristics (e.g. instability) of the plaque in PAD.

By combining B-mode ultrasound and color Doppler ultrasound, duplex ultrasound (DUS) identifies the anatomical location and the degree of a stenosis. Peak systolic velocity (PSV) ratios (as determined within and beyond the obstruction and as compared with the adjacent upstream segment) are useful to estimate the rate of stenosis: a PSV ratio > 2:1 argues for > 50% stenosis, a PSV ratio > 4:1 for a > 75% stenosis, and a PSV ratio > 7:1 for a > 90% stenosis (Citation26). DUS helps identify patients with the need for endoluminal revascularization (Citation27,Citation28) and is currently employed to follow-up venous grafts (Citation29–36). The high sensitivity, specificity, diagnostic accuracy, and cost-effectiveness of DUS has been documented (Citation37–39).

Magnetic resonance angiography (MRA) very carefully defines the borders of the arterial wall (Citation40–42) and identifies the unstable fibrous cup of an atherosclerotic plaque (Citation43). Moreover, by allowing the identification of size, composition (i.e. the lipid-enriched necrotic core, hemorrhages, calcifications, etc.), and ulcerative components, MRA contributes to the morphological characterization of a plaque (Citation44–46). The morphological characterization of the plaque is improved by gadolinium contrast media that help differentiate the necrotic core from the surrounding fibrous tissue (Citation47,Citation48) and document neo-angiogenesis and the inflammatory burden (Citation49,Citation50). Meta-analysis and systematic reviews argue for the diagnostic accuracy of MRA in the PAD setting (Citation51–53). However, these studies also claim that MRA tends to overestimate the degree of stenosis. This is mostly due to turbulence and metal clips that, by mimicking vessel occlusions, can cause artifacts. Likewise, some metal stents may obscure vascular flow and, in turn, cause artifacts (Citation54).

Short acquisition time, very thin slices, high spatial resolution, and improved multi-detector computed tomography scanners enable scanning of the entire vascular tree in a limited period of time by computed tomography angiography (CTA), with a low amount of contrast medium employed and radiation burden (Citation55). These recent technical developments have made CTA one of the most important imaging techniques in PAD. A recent meta-analysis (Citation56) showed its high diagnostic accuracy in the PAD setting. The pooled sensitivity to detect a > 50% stenosis or occlusion was 95% (92%–97%) and the pooled specificity 96% (93%–97%). CTA correctly identified occlusions in 94% of segments, the presence of > 50% stenosis in 87% of segments, and absence of significant stenosis in 96% of segments. Nevertheless, similarly to MRA, CTA tends to overestimate the degree of stenosis (see above for details) (Citation54).

In spite of the current availability of less invasive imaging techniques, catheter angiography (CA) with or without digital subtraction angiography (DSA) is the gold-standard first-line imaging investigation for patients with PAD and the reference method for guiding percutaneous peripheral interventional procedures (Citation54). However, due to the need for contrast media (that may cause renal toxicity, arterial wall dissection, emboli, fistulae, pseudoaneurysms, and access site complications) and to its inherent limitations (no careful hemodynamic study of the stenotic segment nor of its length (overestimated) is possible), the clinical use of DSA is currently limited (Citation57).

Indications, limitations, and contraindications of each imaging technique are reported in and .

Table III. Recommendations for different imaging techniques to be employed in the diagnosis of PAD and of CLI.

Table IV. Comparison of different imaging techniques for patients with PAD: additional data.a

Antithrombotic treatment in asymptomatic and symptomatic PAD

Pharmacology of major antithrombotic agents

Antiplatelet agents in PAD

Following atherosclerotic plaque disruption/endothelial cell detachment, circulating platelets, exposed to a highly thrombogenic environment, become activated (Citation58). A series of soluble agonists (ADP, thromboxane A2 (T× A2), serotonin (5-HT), and thrombin) recruit and activate additional platelets. Upon activation, glycoprotein (Gp) IIb/IIIa (αIIbβ3 integrin) mediates platelet aggregation and spreading by means of fibrinogen bridges, which, once converted to fibrin, ultimately contribute to thrombus stabilization. This leads to the formation of platelet-rich thrombi that, occluding the arterial lumen and impairing blood-flow and oxygen supply, cause acute ischemia. The efficacy of aspirin lies in its ability irreversibly to inhibit platelet COX-1 (by acetylating a serine located near the active site of the enzyme) and, in turn, T× A2 formation (Citation59). The transduction of the ADP signal involves its interaction with two platelet receptors belonging to the P2 purinergic family, the Gαq-coupled receptor P2Y1 and the Gαi-coupled receptor P2Y12 (Citation60). The concomitant activation of both the Gαq and Gαi pathways by ADP is needed for platelet aggregation to occur. Signaling from the P2Y1 receptor causes platelet shape change and rapid transient aggregation, whereas the signaling from the P2Y12 receptor facilitates sustained irreversible aggregation and stimulates surface expression of the pro-inflammatory P-selectin. In addition, the P2Y12 receptor plays a critical role in the amplification of platelet aggregation induced by agents other than ADP, including 5-HT, T× A2, and thrombin. Together, these contribute to thrombus growth and stability. Two main classes of antiplatelet agents are licensed and widely used chronically in PAD: acetyl salicylic acid (aspirin) and P2Y12 inhibitors (ticlopidine, clopidogrel, prasugrel, cangrelor, ticagrelor) (Citation61).

Aspirin

Owing to its efficacy and favorable cost-effectiveness, aspirin is the mainstay treatment for all atherothrombotic conditions. A recent meta-analysis (Citation60) has assessed the role of aspirin in primary (95,000 subjects at low cardiovascular risk) and secondary (17,000 patients at medium/high risk) vascular prevention. While in high-risk conditions the advantage of aspirin outweigh the inherent bleeding hazard, in primary prevention aspirin is associated with an absolute benefit of 0.06%/year, too exiguous when compared to the 0.03% increase in major bleedings.

Ticlopidine

Ticlopidine was the first agent of the thienopyridine class shown both to prevent the interaction of ADP with its platelet purinergic receptor and to cause inhibition of fibrinogen binding to the αIIbβ3 integrin (Citation62). In subjects with a history of cerebrovascular events, ticlopidine was superior to placebo and to aspirin in the reduction of stroke, AMI, or vascular death (Citation63). In addition, the combination of ticlopidine with aspirin was successful in acute coronary syndrome (ACS) patients undergoing percutaneous coronary intervention (PCI) with stent implantation (Citation64). Diarrhea, aplastic anemia, thrombotic thrombocytopenic purpura, and neutropenia are the main limitations for a widespread use of ticlopidine.

Clopidogrel

The thienopyridine prodrug clopidogrel irreversibly binds the P2Y12 platelet receptor after a two-step activation by cytochrome P450 (CYP) liver isoenzymes. A variety of polymorphisms in the CYP2C19 gene (most often the CYP2C19*2), associated with a 20%–25% production of inactive metabolite, diminish the response to clopidogrel. Among subjects under treatment with clopidogrel for a previous vascular event, carriers of these polymorphisms have a 50% higher risk of cardiovascular death, AMI, or stroke (Citation58). Among clopidogrel-treated patients, carriers of at least one allele associated with the loss of or a reduced function in the CYP2C19 gene had a higher than normal occurrence of fatal and non-fatal coronary thrombotic events, as well as of stent thrombosis (Citation58). In the same study population, polymorphic alleles of a gene modulating clopidogrel absorption (ABCB1) have been associated with a higher rate of cardiovascular events at 1-year follow-up as compared to wild-type subjects.

Ticagrelor

Ticagrelor, an orally active cyclopentyl-triazolo-pyrimidine, binds to domains of the P2Y12 receptor other than those recognized by ADP (the 1, 2, and 7 transmembrane domains, the extracellular loop 2, and the N-terminal domain), determining a potent and rapid non-persistent receptor conformational change. After the occupancy of P2Y12, ADP-catalyzed conversion of cAMP from ATP, dephosphorylation of phosphorylated VASP, and activation of phosphoinositide 3-kinase are blocked. The net result is a reduced exposure of fibrinogen-binding sites on the αIIbβ3 integrin receptor and, in turn, the inhibition of platelet aggregation. Inhibition of ADP-mediated constriction of vascular smooth muscle and enhancement of adenosine-induced coronary blood-flow are also reported. After oral administration, ticagrelor is rapidly absorbed and does not require hepatic biotransformation to be pharmacologically active. However, ticagrelor is also metabolized to an equipotent, active metabolite (AR-C124910XX) by CYP3A4 enzymes. As both ticagrelor and AR-C124910XX are excreted by the intestinal route, no dose adjustment is needed in kidney failure. On the other hand, the concomitant use of CYP3A4 inhibitors/inducers as well as a significant liver dysfunction may be of concern for its use. After pharmacodynamic evaluations (Citation65,Citation66), a 90-mg twice-daily dose of ticagrelor has been chosen to optimize its efficacy, safety, and tolerability. A loading dose of 180–270 mg may minimize intersubject variability as to initial inhibition in platelet aggregation and may be appropriate in ticagrelor-naive patients with ACS or in preparation for PCI. In 174 subjects with a recent coronary artery disease receiving 75–100 mg/day aspirin (92 also under ticagrelor 180-mg load and 90 mg twice-daily maintenance dose, and 82 also under clopidogrel 600-mg load and 75 mg/d maintenance dose) the genotyping of the cytochrome P450 (CYP) 2C19 (*1,*2,*3,*4,*5,*6,*7,*8,*17) was performed. In addition, platelet function was measured (by aggregometry, VerifyNow P2Y12 assay, and VASP assay at pre-dose, 8 hours post-loading, and during maintenance). There was no significant effect of the genotype on platelet function during aspirin therapy alone. On the other hand, irrespective of the 2C19 genotype, of the metabolizer status, and of the assays employed, subjects on ticagrelor showed a lower platelet reactivity than did those on clopidogrel (P < 0.01). This is consistent with a genotype-independent better pharmacodynamic effect of ticagrelor as compared to clopidogrel (Citation67).

Cilostazol

Cilostazol, selectively targeting phosphodiesterase type 3 (PDE3) and, then, determining intracellular cAMP accumulation, inhibits platelet aggregation (Citation68). In diabetic patients on standard dual antiplatelet therapy, adjunctive treatment with cilostazol enhances inhibition of platelet P2Y12 signaling (Citation69). A Cochrane review (Citation70), in which two randomized studies on stroke prevention were summarized, documented that, compared with aspirin, cilostazol was associated with a significantly lower risk of vascular events (6.77% versus 9.39%; RR 0.72; 95% CI 0.57–0.91, composite outcome) and a lower risk of hemorrhagic stroke (0.53% versus 2.01%; RR 0.26; 95% CI 0.13–0.55). In terms of outcome of safety, cilostazol was associated with significantly fewer adverse events (8.22% versus 4.95%; RR 1.66; 95% CI 1.51–1.83) than aspirin. In the SILOAM phase IV study (ClinicalTrials.gov Identifier: NCT01261832), a triple antiplatelet therapy (cilostazol plus aspirin and clopidogrel) is compared (at 1 month and at 6 months) with the standard dual antiplatelet treatment (ASA and clopidogrel) in 951 ACS subjects (expected number) undergoing PCI and drug-eluting stent implantation. The primary efficacy end-point is the occurrence of major cardiovascular and cerebrovascular events (total death, non-fatal myocardial infarction, repeat revascularization, stroke). The end of the study is expected by July 2014.

Primary prevention of cardiovascular events in asymptomatic PAD ()

On the basis of an individual participant data meta-analysis for primary and secondary prevention of coronary artery disease (Citation60) and of data from a meta-analysis on cancer (Citation71), aspirin is expected to reduce total mortality and to increase major bleedings (Citation72). Since this is especially true in subjects taking more than one drug daily (most PAD patients), and since similar benefits are seen in patients with IC and those who had undergone peripheral vascular grafting or angioplasty, the use of aspirin over a prolonged time period should be encouraged on an individual basis. Regardless of this decision, intensive treatment for cardiovascular risk factor modifications (Citation73) is mandatory for primary prevention of cardiovascular events in patients with PAD. Tobacco cessation should be encouraged, eventually by behavior modification or pharmacologic strategies (Citation74). Cessation of tobacco use significantly reduces lower-extremity symptoms and progression of PAD, thus helping improve the maximal walking distance achieved by structured exercise programs (Citation75). As in other patients with high cardiovascular risk, LDL cholesterol levels should be lowered to < 70 mg/dL in PAD patients (Citation76). A sub-analysis of 6748 patients with PAD in the Heart Protection Study showed significant reductions in total mortality, vascular mortality, coronary heart disease events, strokes, and non-coronary revascularization in those treated with simvastatin (Citation77). There was no threshold cholesterol value below which statin therapy was not associated with benefit (Citation78–80). Statins improve the pain-free walking distance (Citation81) and reduce the progression of PAD, the overall cardiovascular risk, and the occurrence of complications needing invasive procedures (Citation82). The goal for blood pressure control in PAD is < 140/90 mmHg; in those with PAD and DM or with chronic kidney disease, it should be < 130/80 mmHg (Citation83). In a subgroup of 4046 patients with PAD, the Heart Outcomes Prevention Evaluation study showed that those randomly assigned to the angiotensin-converting enzyme inhibitor ramipril had a 22% reduction in risk, compared with the placebo group, that was independent of lowering of blood pressure (Citation84). Because of the additional cardio-protective effects, the use of β-blockers is important in patients with coexisting coronary artery disease. Although no trials have been designed or powered to examine glycemic control, a tight plasma glucose control reduces PAD progression in diabetic patients with PAD (Citation85–90) and improves the clinical outcome of percutaneous revascularization (Citation91). To this end, the reduction of the A1c hemoglobin levels < 7% should be associated with the control of all risk factors, proper foot care (e.g. cleansing of skin lesions), and appropriate footwear (Citation92).

Table V. Antithrombotic drugs for PAD: different strategies for different objectives.

Secondary prevention of cardiovascular events in symptomatic PAD

Main findings from the meta-analysis of aspirin therapy for primary and secondary prevention of coronary artery disease (Citation60) also show that, in a population at high risk for a serious vascular event (i.e. 8.2%/ year), the risk/benefit ratio between bleedings and prevention of total mortality, non-fatal MI, and non-fatal stroke is in favor of the prolonged use of aspirin. Clopidogrel, when administered alone in about 20,000 patients (all with a history of AMI, stroke, or PAD) in the randomized CAPRIE trial (Clopidogrel versus Aspirin in Patients at Risk for Ischemic Events) (Citation93), was only marginally superior to aspirin (relative risk reduction (RRR) 8.7; P = 0.043) in preventing non-fatal AMI and non-fatal extracranial bleeding with little or no effect on total mortality. This was the overall outcome in patients with a recent stroke and in patients with a recent AMI. However, when the chronic ‘PAD subset’ was analyzed alone (over 8000 patients), a 24% risk reduction in major adverse ischemic events was found. This was a pre-specified (pre-randomization) stratification, and the validity of such a conclusion in PAD patients led to the positive study outcome and FDA approval.

A significantly high efficacy of clopidogrel has also been shown when this drug was employed for the ‘dual antiplatelet therapy’. The CURE (Citation94) and the CREDO (Citation95) studies established the superiority of clopidogrel in combination with aspirin versus aspirin alone in ACS and in ACS with PCI, respectively. In high vascular risk patients with atherothrombosis (manifested as a recent stroke, recent MI, or symptomatic PAD), a meta-analysis of 10 studies examining the effects of thienopyridines (clopidogrel and ticlopidine) versus aspirin achieved results similar to those of the CAPRIE trial (Citation96). A reduction in non-fatal stroke and an increase in non-fatal extracranial bleeding with no effect on total mortality or non-fatal AMI was found in the long-term efficacy of clopidogrel+ aspirin versus aspirin alone in the randomized Clopidogrel for High Atherothrombotic Risk and Ischemic Stabilization, Management, and Avoidance (CHARISMA) trial (Citation97). A Cochrane systematic review that evaluated short- and long-term dual antiplatelet therapy in patients with established coronary artery disease reached similar conclusions (Citation98). Finally, in a meta-analysis of three studies there was no effect on mortality, non-fatal AMI, or non-fatal stroke, and a significant raise in major bleeding events in patients receiving warfarin (PT-INR 2-3)+ aspirin versus aspirin alone (Citation73).

Cilostazol, pentoxifylline, and prostanoids are used to improve the quality of life in patients with symptomatic PAD and with claudication. Since claudication itself is responsive to smoking cessation and exercise therapy, drugs for improving the quality of life should be considered only in patients who have limitations after smoking cessation and exercise therapy. Two systematic reviews have shown that patients with symptomatic PAD receiving cilostazol are more likely to experience important benefits as to physical health sub-scale (but not to health-related quality of life) than those receiving placebo (Citation99,Citation100) or pentoxifylline (Citation101). No difference in rates of AMI, stroke, or death was found in another review on 1374 participants randomized to cilostazol and 973 randomized to placebo (Citation102). Nor has a significant effect of cilostazol on major or minor bleeding rates been detected in another systematic review (Citation103).

In a meta-analysis (seven randomized studies) (Citation104), 643 patients were analyzed as to the use of prostaglandin E1 (PGE1) in the treatment of advanced PAD. At the end of treatment, PGE1 showed a significantly better response (ulcer healing and/or pain reduction) as compared to placebo (47.8% for PGE1 versus 25.2% for placebo, P = 0.0294). After a 6-month follow-up, a slight although still significant difference in favor of PGE1 was seen for the combined end-point ‘major amputation or death’ (22.6% for PGE1 versus 36.2% for placebo, P = 0.0150). That the benefit of PGE1 is apparent in the short term but decreases over time had been previously reported in a multicenter trial on 1560 patients with chronic critical leg ischemia (Citation105). The response rate (ulcer healing and/or pain relief) of the pooled treatment groups was 60.2% for PGE1, 25.2% for placebo, and 53.6% for iloprost, in that report. The adverse events rate of the pooled treatment groups showed a good tolerability for PGE1 with a rate of 39.6% in comparison to 73.9% for iloprost and 15.4% for placebo. In a recent Cochrane systematic review on 13 studies (Citation106), > 75% of subjects with critical limb ischemia without chance of rescue or reconstructive intervention that received prostanoids experienced at least one drug-related adverse event (i.e. headache, nausea, vomiting, diarrhea, and facial flushing) in addition to improved rest for pain and ulcer healing.

By preventing clot propagation and further embolism, short-term anticoagulation treatment is routinely used to reduce the extent of ischemia in acute limb ischemia. There are no formal studies demonstrating improved outcomes with such strategy. To restore flow to the occluded artery, either surgery or catheter-directed intra-arterial thrombolysis is employed. There are no data to support reperfusion therapy over anticoagulation alone in this setting. In a meta-analysis (Citation107) that has compared benefits and harms of intra-arterial thrombolysis versus surgery, only the risk of stroke and of major bleeding at 30 days was higher for thrombolysis as compared with surgery, no effect being found on amputation, death, or limb salvage.

However, we believe that caution is needed when commenting this conclusion. Among those analyzed, the only randomized comparative study which was designed and performed by investigators was the STILE trial (Citation108). In patients with true acute limb ischemia (1–14 days) the study reported a significant benefit of catheter-directed thrombolysis in terms of reducing amputation and improving amputation-free survival up to 1 year. These results were not reproduced in commercially written and performed studies.

In the past, thrombolysis for acute limb ischemia was administered i.v.; this strategy has been now replaced by catheter-directed thrombolysis. In a Cochrane systematic review on the management of acute lower-limb ischemia (Citation109), compared to thrombolysis by recombinant tissue-type plasminogen activator (rt-PA) or urokinase, i.v. streptokinase was associated with a high rate of amputation at 30 days (7/20 versus 1/20; RR 7.0; 95% CI 0.95–51.80) and no effect on limb salvage, major hemorrhage, or death.

Surgical revascularization in patients with CLI ()

When antiplatelet drugs and the control of DM, of dyslipidemia, of cigarette smoking, and of hypertension are not sufficient in the treatment of CLI, surgery (Citation110) or percutaneous transluminal angioplasty (PTA) with or without stenting (Citation111) are needed. The objective of revascularization is to establish adequate inflow to the distal vessels. Balloon devices are used to this end, either in the contralateral retrograde common femoral artery (CFA) access or in the ipsilateral anterograde CFA access. This intervention is aimed at establishing straight line flow in at least one tibial vessel, to supply the area of the foot with rest ischemia (typically, the forefoot) or with tissue loss. This treatment shows a very high immediate limb salvage rates (> 90%), with < 2% intervention-related mortality and < 5% risk of complications. In the first 12–24 months, the limb salvage rates are above 80%, and, when the therapeutic goal is achieved, the patency of treated arteries leads to resolution of ischemic lesions. The efficacy of angioplasty has been confirmed by several studies (Citation112,Citation113). A recent meta-analysis has shown that the efficacy and safety of endovascular techniques are comparable to surgical interventions (Citation114). By releasing anti-proliferative drugs such as paclitaxel, drug-eluting balloons efficiently maintain arterial patency after PTA and reduce the risk of restenosis (Citation115–117). A possible side effect of this procedure is wall dissection and restenosis. The use of stents is thought to reduce such risk (Citation118–120). However, presently it is unclear whether PTA with stent placement is superior to PTA alone with respect to patient-important outcomes. Compared with PTA without routine stenting, PTA plus routine stenting for superficial femoropopliteal arterial disease was associated with a reduction in restenosis (RR 0.85; 95% CI 0.69–1.06) but with no effect on the need for target vessel revascularization (RR 0.98; 95% CI 0.78–1.23) (Citation121). Among strategies regarding surgical revascularization and endovascular revascularization, great attention is currently paid to benefits of drug-eluting stents. In the prospective, multinational randomized controlled Zilver PTX trial, the 2-year safety and efficacy of a paclitaxel-coated drug-eluting stent (DES) was compared with PTA in patients with superficial femoral artery lesions. In patients who received the paclitaxel-coated DES, 2-year outcomes showed statistically significant differences in terms of event-free survival, primary patency, and clinical benefits (Citation122).

Table VI. TASC II classification of aortoiliac lesions.

The TASC-II classification of ischemic lesions is used to decide between PTA or surgical interventions—TASC A and B lesions being best treated with PTA, TASC C and D lesions being usually treated with surgical strategies. The prosthetic material used is dacron or ePTFE, autologous materials (e.g. femoral veins or cryoconserved allogenic veins or arteries) being used in cases with infections of the original graft (Citation123,Citation124). After 5 years, the primary patency rate of all surgical procedures is > 80%, significant differences being found when comparing claudicants with patients with CLI (Citation125). Such rates of success progressively decrease in patients with local co-morbidities. In a retrospective review (Citation126), there was a significant difference in primary patency after 36 months (better for surgery). However, it was severely diminished in patients with diabetes and distal PAD, and, compared with patients undergoing PTA, surgical patients needed more often additional interventions of reconstruction. Debulking procedures are taken into account in selected patients (i.e. those at a high risk of PTA-related complications) (Citation127). They include the excimer laser (to perform photoablation of occlusive material), the Rotablator (for the treatment of calcified plaques), and new techniques such as the Silverhawk system (designed for eccentric and not severely calcified infrainguinal lesions), the Rochawk system (for calcified plaques), and the newer Jetstream system (with an aspiration device to perform simultaneous thrombectomy and atherectomy).

Antithrombotic treatments in patients undergoing revascularization procedures ()

A Cochrane review (Citation128) on patients undergoing lower-extremity PTA without stent placement reported a reduction in reocclusion in patients taking aspirin+ dipyridamole compared with placebo. Another study randomized patients to i.v. unfractionated heparin versus subcutaneous nadroparin administered for 1 week post-procedure (Citation129). Nadroparin was associated with a reduction in the rate of vessel restenosis/occlusion but not in amputation. However, the overall quality of the evidence in these studies is low. Similarly poor is the evidence from studies on antithrombotic treatments in patients undergoing PTA with stent placement.

In a Cochrane systematic review in patients receiving post-infrainguinal bypass graft surgery compared to placebo, aspirin+ dipyridamole resulted in fewer graft occlusions (Citation130). In one study of that systematic review, compared with placebo, aspirin+ dipyridamole was associated with a reduction in amputations rates. Also in this case, the overall quality of the evidence is low. Thus, aspirin or clopidogrel are suggested as the treatment of choice in symptomatic PAD.

The Clopidogrel and Acetylsalicyclate Acid in Bypass Surgery for Peripheral Arterial Disease (CASPAR) study randomized patients undergoing unilateral below-knee bypass graft surgery for PAD to clopidogrel+ aspirin versus placebo+ aspirin (Citation131). There was a significant decrease in amputations in patients treated with clopidogrel+ aspirin, but only in the subgroup of patients undergoing prosthetic graft bypass. Finally, together with a reduction in non-fatal AMI and a significant increase in extracranial major bleedings, there was no advantage over aspirin of a high-intensity oral anticoagulation (target PT-INR 3-4.5) on all-cause mortality and non-fatal stroke in patients who had undergone infrainguinal bypass grafting (Citation132).

Open issues and areas of research

Due to imprecision, indirectness, and risk of bias, the overall quality of the evidence summarized above is low in most cases, and open issues and areas of research emerge.

Because of the paucity of studies regarding type and duration of antithrombotic therapy, in patients undergoing PTA with stent placement, the rationale for the current practice (aspirin and loading dose of clopidogrel pre-procedure; dual antiplatelet therapy for 1–3 months thereafter, particularly if a stent is placed in a small peripheral vessel) is largely based on indirect evidence from data in patients undergoing coronary stenting trials (Citation133). However, the risk of stent thrombosis is conceivably lower in stenting larger-caliber peripheral arteries than in smaller coronary arteries. Differences in stent types and differing outcomes should be also considered in PAD.

A meta-analysis (Citation134) has shown that, compared with those showing an optimal response to the drug, patients with persistent platelet reactivity despite clopidogrel treatment (∼30% of total) have a significantly higher risk of death and/or ischemic recurrence (Citation58). The same risk has been reported for aspirin (‘aspirin resistance’) (Citation135). The possibility has been documented that genetic variations in gut/liver enzymes that control biotransformation (and in turn the pharmacological activity) play a role in the low response to clopidogrel (Citation136). After oral administration, ticagrelor is rapidly absorbed and does not require hepatic biotransformation to be pharmacologically active. Compared to clopidogrel, a 16% RRR in major adverse cardiovascular events following acute coronary syndrome, a 21% RRR in cardiovascular mortality, and a numerical 22% RRR in all-cause mortality has been documented in 18,000 NSTEMI or STEMI patients in the multicenter, randomized, double-blind, double-dummy phase III PLATO trial (ticagrelor 90 mg b.i.d. versus clopidogrel 75 mg/d) (Citation137). In ∼11,500 male and female patients (> 50 years of age) with established PAD, the EUCLID (Examining Use of tiCagreLor In paD) study protocol (ClinicalTrials.gov Identifier: NCT01732822) is now comparing 90 mg b.i.d. oral ticagrelor with 75 mg/d oral clopidogrel and the corresponding placebo on the risk of cardiovascular death, AMI, ischemic stroke, and TIMI major (primary safety objective) and major or minor bleeding events. October 2015 is the estimated primary completion date for the EUCLID study protocol.

By preventing clot propagation and further embolism, short-term anticoagulation treatment with therapeutic doses of heparin is commonly used to reduce the extent of ischemic injury in the management of acute limb ischemia. No formal studies demonstrating improved outcomes with anticoagulation (nor side effects of heparin versus other anticoagulant agents) are available. Although effective, vitamin K antagonists have numerous limitations, which complicate their use. These limitations have prompted the introduction of new oral anticoagulants that overcome many of the problems associated with vitamin K antagonists. The new oral direct oral anticoagulants (DOACs) fall into two main classes: direct thrombin inhibitors (factor IIa inhibitors, dabigatran etexilate) and direct FXa inhibitors (rivaroxaban, apixaban, edoxaban) (Citation138). Although agents in the two classes have distinct mechanisms of action, targeting distinct enzymes in the coagulation pathway, all of the new drugs attenuate fibrin formation and have features in common that distinguish them from vitamin K antagonists, such as warfarin. These common features include a rapid onset of action, few drug–drug interactions, and a predictable anticoagulant response that enables fixed dosing for several indications and across a diverse range of patients with no need for routine coagulation monitoring. Major issues concerning the pharmacology including the efficacy/safety of DOACs in clinical trials on prevention and treatment of thromboembolism (in medical and surgical patients), as well as in AF have been recently reviewed (Citation139). Moreover, in a 14-mo follow-up in the everyday practice, 3.5 gastrointestinal bleedings and 2.4 intracranial hemorrhages/100,000 days at risk occurred in new users of warfarin, compared to 1.6 gastrointestinal bleedings and 0.8 intracranial hemorrhages/100,000 days at risk among new users of the DOAC dabigatran (Citation140). A randomized trial of DOACs versus low-molecular-weight heparin in ALI should be considered.

Claudication unresponsive to physical exercise and smoking cessation is present in a relevant group of patients with IC. Data argue for this phenotype as identifying those with a heritable predisposition to more severe vascular events (i.e. with unfavorable combination(s) of genes modulating lipid metabolism, arterial pressure, vascular function, inflammation, hemostasis, and/or leukocytes and endothelial cell function) (Citation141–144). This issue deserves to be thoroughly addressed.

Reports call attention to the occurrence of PAD in patients with AF (Citation145) or with chronic renal failure (CRF) (Citation146), a higher than normal risk of ischemic events and mortality being documented in both these settings (Citation147–149). The prevalence of PAD is higher in patients with DM than in the general population (Citation150–152); a distal location with an involvement of the infrapopliteal vessels is more common in DM (Citation153,Citation154), and the severity and outcome of PAD is worse in DM and is strongly related to duration and severity of the metabolic disorder (Citation155–157). In a meta-analysis of ∼48,000 healthy men and women, an ABI < 0.90 at baseline was associated with an approximate doubling of the 10-year mortality, cardiovascular mortality, and major coronary event rate after adjusting for the Framingham risk score (Citation158). When evaluating patients for primary prevention with aspirin, it has been suggested (Citation81) to use a risk stratification tool (e.g. the Framingham risk score) which provides estimates of low (< 10%), moderate (10%–20%), and high risk (> 20%) of cardiovascular events over 10 years, doubling the patient's risk score (from a low to moderate or a moderate to high category) when his/her ABI is < 0.90. Whether this is the case is unknown so far, and deserves to be addressed.

In the ICAI (Ischemia Critica degli Arti Inferiori) trial (Citation159), of 1560 patients enrolled, 298 died within 1 year, 187 were amputees at 6 months, and 746 continued to suffer from CLI. Prior major vascular events doubled the risk of dying within 1 year. Previous revascularization was associated with a lower mortality and a higher probability of amputation. Among cardiovascular risk factors, diabetes increased mortality and lowered the probability of recovery from CLI. Patients with tissue loss had a higher amputation rate and less probability of recovery. Ankle pressure was predictive of mortality and amputation only when it was not measurable. Compared to those in whom surgery was deemed unnecessary, patients requiring revascularization had better chances of recovering from CLI, but not of longer-term survival or limb salvage. Antiplatelet drugs caused resolution of CLI and decreased the amputation rate by about one-third, while the advantage of the test treatment with alprostadil-α-cyclodextrine was confined to CLI resolution only. Together, these data provide the rationale for defining stratification criteria for a severity score to estimate reliably the achievable benefit in routine practice and/or identify PAD subjects at the highest risk of amputation/vascular death.

By allowing the identification of size and ulcerative components and by differentiating the necrotic core from the surrounding fibrous tissue, neo-angiogenesis, and the inflammatory burden composition, MRA contributes to the morphological characterization of a plaque (see above). Whether and the extent to which early and more aggressive strategies in PAD patients at the highest risk of ischemic events should be based, at least in part, on non-invasive and invasive tools for the diagnosis of PAD is poorly understood.

Innovative treatments with growth factors or blood progenitor transfer (Citation160) are under evaluation to reduce pain and the rate of amputation in patients who have severe PAD and do not respond to PTA or surgery (e.g. those with DM) (Citation161). In addition to the obvious pathophysiological and therapeutic implications, information from these model systems should help clarify whether conditions (e.g. CLI) other than an acute coronary syndrome argue for a ‘pan-vascular destabilization’ status (162).

Acknowledgements

Portions of this review have been presented at the Pavia Spring Meeting on Thrombosis, 21–22 June 2012.

Declaration of interest: G.D.M. has served on advisory boards for and has received fees as a speaker at meetings organized by Boehringer-Ingelheim, Bristol-Myers Squibb, Pfizer, and Daiichi Sankyo. E.T. has served on advisory boards for and has received fees as a speaker at meetings organized by Bristol-Myers Squibb. All the other authors have nothing to declare.

References

  • Falk E, Shah PK, Fuster V. Coronary plaque disruption. Circulation. 1995;92:657–71.
  • Arbustini E, Dal Bello B, Morbini P, Burke AP, Bocciarelli M, Specchia G, et al. Plaque erosion is a major substrate for coronary thrombosis in acute myocardial infarction. Heart. 1999;82:269–72.
  • Drouet L. Atherothrombosis as a systemic disease. Cerebrovasc Dis. 2002;13:S1–6.
  • Badimon L, Vilahur G. Platelets, arterial thrombosis and cerebral ischemia. Cerebrovasc Dis. 2007;24 :S30–9.
  • Ogren M, Hedblad B, Isacsson SO, Janzon L, Jungquist G, Lindell SE. Non-invasively detected carotid stenosis and ischaemic heart disease in men with leg arteriosclerosis. Lancet. 1993;342:1138–41.
  • Davies MG. Critical limb ischemia: epidemiology. Methodist Debakey Cardiovasc J. 2012;8:10–14.
  • Lambert MA, Belch JJ. Medical management of critical limb ischaemia: where do we stand today? J Intern Med. 2013;274:295–307.
  • Meru AV, Mittra S, Thyagarajan B, Chugh A. Intermittent claudication: an overview. Atherosclerosis. 2006;187:221–37.
  • Leng GC, Fowkes FG, Lee AJ, Dunbar J, Housley E, Ruckley CV. Use of ankle brachial pressure index to predict cardiovascular events and death: a cohort study. BMJ. 1996;313:1440–4.
  • Norgren L, Hiatt WR, Dormandy JA, Nehler MR, Harris KA, Fowkes FG, et al. Inter-society consensus for the management of peripheral arterial disease (TASC-II). J Vasc Surg. 2007;45:S5–67.
  • Sanna G, Alesso D, Mediati M, Cimminiello C, Borghi C, Fazzari AL, et al. Prevalence of peripheral arterial disease in subjects with moderate cardiovascular risk: Italian results from the PANDORA study Data from PANDORA (Prevalence of peripheral Arterial disease in subjects with moderate CVD risk, with No overt vascular Diseases nor Diabetes mellitus). BMC Cardiovasc Disord. 2011;11:59.
  • Violi F, Davi G, Hiatt W, Lip GY, Corazza GR, Perticone F, et al. Prevalence of peripheral artery disease by abnormal ankle-brachial index in atrial fibrillation: implications for risk and therapy. J Am Coll Cardiol. 2013;62:2255–6.
  • Newman AB, Shemanski L, Manolio TA, Cushman M, Mittelmark M, Polak J, et al.; The Cardiovascular Health Study Group. Ankle-arm index as a predictor of cardiovascular disease and mortality in the Cardiovascular Health Study. Arterioscler Thromb Vasc Biol. 1999; 19:538–45.
  • Resnick HE, Lindsay RS, McDermott MM, Devereux RB, Jones KL, Fabsitz RR, et al. Relationship of high and low ankle brachial index to all-cause and cardiovascular disease mortality: the Strong Heart Study. Circulation. 2004;109:733–9.
  • Van der Meer IM, Bots ML, Hofman A, del Sol AI, van der Kuip DA, Witteman JC. Predictive value of noninvasive measures of atherosclerosis for incident myocardial infarction: the Rotterdam Study. Circulation. 2004;109:1089–94.
  • Dormandy JA. Natural history of intermittent claudication. Hospital Update. 1991;April:314–18.
  • Ouriel K. Peripheral arterial disease. Lancet. 2001;358:1257–64.
  • Criqui MH. Peripheral arterial disease: epidemiological aspects. Vasc Med. 2001;6:S3–7.
  • McKenna M, Wolfson S, Kuller L. The ratio of ankle and arm arterial pressure as an indipendent predictor of mortality. Atherosclerosis. 1991;87:119–28.
  • Beckera F, Robert-Ebadia H, Riccob JB, Setaccic C, Cao P, de Donato G, et al. Chapter I: definitions, epidemiology, clinical presentation and prognosis. Eur J Vasc Endovasc Surg. 2011;42:S4–12.
  • Street TK. What the primary care provider needs to know for limb salvage. Methodist Debakey Cardiovasc J. 2012;8:57–9.
  • O’Connell JB, Quiñones-Baldrich WJ. Proper evaluation and management of acute embolic versus thrombotic limb ischemia. Semin Vasc Surg. 2009;22:10–16.
  • Walker TG. Acute limb ischemia. Tech Vasc Interv Radiol. 2009;12: 117–29.
  • Lawall H, Zemmrich C, Bramlage P, Amann B. Health related quality of life in patients with critical limb ischemia. Vasa. 2012;41:78–88.
  • Huppert P, Tacke J, Lawall H; Deutschen Gesellschaft für Angiologie/Gefässmedizin. [S3 guidelines for diagnostics and treatment of peripheral arterial occlusive disease]. Radiologe. 2010;50:7–15. [In German]
  • Edwards JM, Coldwell DM, Goldman ML, Strandness DE Jr. The role of duplex scanning in the selection of patients for transluminal angioplasty. J Vasc Surg. 1991;13:69–74.
  • Van der Heijden FH, Legemate DA, van Leeuwen MS, Mali WP, Eikelboom BC. Value of duplex scanning in the selection of patients for percutaneous transluminal angioplasty. Eur J Vasc Surg. 1993; 7:71–6.
  • Mattos MA, van Bemmelen PS, Hodgson KJ, Ramsey DE, Barkmeier LD, Sumner DS. Does correction of stenoses identified with color duplex scanning improve infrainguinal graft patency? J Vasc Surg. 1993;17:54–64.
  • Bandyk DF, Schmitt DD, Seabrook GR, Adams MB, Towne JB. Monitoring functional patency of in situ saphenous vein bypasses: the impact of a surveillance protocol and elective revision. J Vasc Surg. 1989;9:286–96.
  • Mills JL, Harris EJ, Taylor LM Jr, Beckett WC, Porter JM. The importance of routine surveillance of distal bypass grafts with duplex scanning: a study of 379 reversed vein grafts. J Vasc Surg. 1990;12: 379–86.
  • Laborde AL, Synn AY, Worsey MJ, Bower TR, Hoballah JJ, Sharp WJ, et al. A prospective comparison of ankle/brachial indices and color duplex imaging in surveillance of the in situ saphenous vein bypass. J Cardiovasc Surg. 1992;33:420–5.
  • Taylor PR, Tyrrell MR, Crofton M, Bassan B, Grigg M, Wolfe JH, et al. Colour flow imaging in the detection of femoro-distal graft and native artery stenosis: improved criteria. Eur J Vasc Surg. 1992;6:232–6.
  • Golledge J, Beattie DK, Greenhalgh RM, Davies AH. Have the results of infrainguinal bypass improved with the widespread utilisation of postoperative surveillance? Eur J Vasc Endovasc Surg. 1996;11: 388–92.
  • Gerhard-Herman M, Gardin JM, Jaff M, Mohler E, Roman M, Naqvi TZ. Guidelines for noninvasive vascular laboratory testing: a report from the American Society of Echocardiography and the Society for Vascular Medicine and Biology. Vasc Med. 2006;11: 183–200.
  • Bandyk DF, Cato RF, Towne JB. A low flow velocity predicts failure of femoropopliteal and femorotibial bypass grafts. Surgery. 1985;98: 799–809.
  • Collins R, Burch J, Cranny G, Aguiar-Ibanez R, Craig D, Wright K, et al. Duplex ultrasonography, magnetic resonance angiography, and computed tomography angiography for diagnosis and assessment of symptomatic, lower limb peripheral arterial disease: systematic review. BMJ. 2007;334:1257.
  • Collins R, Cranny G, Burch J, Aguiar-Ibanez R, Craig D, Wright K, et al. A systematic review of duplex ultrasound, magnetic resonance angiography and computed tomography angiography for the diagnosis and assessment of symptomatic, lower limb peripheral arterial disease. Health Technol Assess. 2007;11:1–184.
  • Visser K, Hunink MG. Peripheral arterial disease: gadolinium enhanced MR angiography versus color-guided duplex US –a meta-analysis. Radiology. 2000;216:67–77.
  • Blum A, Nahir M. Future non-invasive imaging to detect vascular plaque instability and subclinical non-obstructive atherosclerosis. J Geriatr Cardiol. 2013;10:178–85.
  • Yuan C, Beach KW, Smith LH, Hatsukami TS. Measurement of atherosclerotic carotid plaque size in vivo using high resolution magnetic resonance imaging. Circulation. 1998;98:2666–71.
  • Zhang S, Hatsukami TS, Polissar NL, Han C, Yuan C. Comparison of vessel wall area measurement differences among three black blood contrast weighting images. Magn Reson Imaging. 2001;19:795–892.
  • Hatsukami TS, Ross R, Polissar NL, Yuan C. Identification of fibrous cap thickness and cap rupture in human atherosclerotic carotid plaque in-vivo with high resolution magnetic resonance imaging. Circulation. 2000;102:959–64.
  • Mitsumori LM, Hatsukami TS, Ferguson MS, Kervin WS, Cai J, Yuan C. In vivo accuracy of multisequence MR imaging for identifying unstable fibrous caps in advanced human carotid plaques. J Magn Reson Imaging. 2003;17:410–20.
  • Yuan C, Zhang SX, Polissar NL, Echelard D, Ortiz G, Dawis JW, et al. Identification of fibrous cap rupture with magnetic resonance imaging is highly associated with recent TIA or stroke. Circulation. 2002;105:2051–6.
  • Yuan C, Mitsumori LM, Ferguson MS, Polissar NL, Echelard D, Ortiz G, et al. In vivo accuracy of multispectral magnetic resonance imaging for identifying lipid-rich necrotic cores and intraplaque hemorrhage in advanced human carotid plaques. Circulation. 2001; 104:2051–6.
  • Yuan C, Kerwin WS, Ferguson MS, Polissar N, Zhang S, Cai J, et al. Contrast enhanced high resolution MRI for atherosclerotic carotid artery tissue characterization. J Magn Reson Imaging. 2002;15: 62–7.
  • Wasserman BA, Smith WI, Trout HH, Cannon RO, Balaban RS, Arai AE. Carotid artery atherosclerosis: in vivo morphologic characterization with gadolinium-enhanced double-oblique MR imaging initial results. Radiology. 2002;223:566–73.
  • Fayad ZA, Fuster V, Fallon JT, Jayasundera T, Worthley SG, Helft G, et al. Noninvasive in vivo human coronary artery lumen and wall imaging using black-blood magnetic resonance imaging. Circulation. 2000;102:506–10.
  • Botnar RM, Stuber M, Kim WY, Danias PG, Manning WJ. Magnetic resonance coronary lumen and vessel wall imaging. Rays. 2001;26: 291–303.
  • Lowery AJ, Hynes N, Manning BJ, Mahendran M, Tawfik S, Sultan S. A prospective feasibility study of duplex ultrasound arterial mapping, digital-subtraction angiography, and magnetic resonance angiography in management of critical lower limb ischemia by endovascular revascularization. Ann Vasc Surg. 2007;21:443–51.
  • Koelemay MJ, Lijmer JG, Stoker J, Legemate DA, Bossuyt PM. Magnetic resonance angiography for the evaluation of lower extremity arterial disease: a meta-analysis. JAMA. 2001;285:1338–45.
  • Nelemans PJ, Leiner T, de Vet HC, van Engelshoven JM. Peripheral arterial disease: meta-analysis of the diagnostic performance of MR angiography. Radiology. 2000;217:105–14.
  • Cao P, Ecksteinb HH, De Rangoc P, Setaccid C, Riccoe JB, de Donato G, et al. Chapter II: diagnostic methods. Eur J Vascular Endovasc Surg. 2011;42:S13–32.
  • Duran C, Bismuth J. Advanced imaging in limb salvage. Methodist Debakey Cardiovasc J. 2012;8:28–32.
  • Met R, Bipat S, Legemate DA, Reekers JA, Koelemay MJW. Diagnostic performance of computed tomography angiography in peripheral arterial disease: a systematic review and metaanalysis. JAMA. 2009;301:415–24.
  • Hessel SJ, Adams DF, Abrams HL. Complications of angiography. Radiology. 1981;138:273–81.
  • Di Minno MN, Guida A, Camera M, Colli S, Minno GD, Tremoli E. Overcoming limitations of current antiplatelet drugs: a concerted effort for more profitable strategies of intervention. Ann Med. 2011;43:531–44.
  • Davi G, Patrono C. Platelet activation and atherothrombosis. N Engl J Med. 2007;357:2482–94.
  • Baigent C, Blackwell L, Collins R, Emberson J, Godwin J, Peto R, et al.; Antithrombotic Trialists’ (ATT) Collaboration. Aspirin in the primary and secondary prevention of vascular disease: collaborative meta-analysis of individual participant data from randomised trials. Lancet. 2009;373:1849–60.
  • Jennings LK. Role of platelets in atherothrombosis. Am J Cardiol. 2009;103(suppl):4A–10A.
  • Di Minno G, Cerbone AM, Mattioli PL, Turco S, Iovine C, Mancini M. Functionally thrombasthenic state in normal platelets following the administration of ticlopidine. J Clin Invest. 1985;75:328–38.
  • Hass WK, Easton JD, Adams HP Jr, Pryse-Phillips W, Molony BA, Anderson S, et al. A randomized trial comparing ticlopidine hydrochloride with aspirin for the prevention of stroke in high-risk patients: Ticlopidine Aspirin Stroke Study Group. N Engl J Med. 1989; 321:501–7.
  • Urban P, Macaya C, Rupprecht HJ, Kiemeneij F, Emanuelsson H, Fontanelli A, et al. Randomized evaluation of anticoagulation versus antiplatelet therapy after coronary stent implantation in high-risk patients: the Multicenter Aspirin and Ticlopidine Trial after Intracoronary Stenting (MATTIS). Circulation. 1998;98:2126–32.
  • Storey RF, Husted S, Harrington RA, Heptinstall S, Wilcox RG, Peters G, et al. Inhibition of platelet aggregation by AZD6140, a reversible oral P2Y12 receptor antagonist, compared with clopidogrel in patients with acute coronary syndromes. J Am Coll Cardiol. 2007; 50:1852–6.
  • Husted S, Emanuelsson H, Heptinstall S, Sandset PM, Wickens M, Peters G. Pharmacodynamics, pharmacokinetics, and safety of the oral reversible P2Y12 antagonist AZD6140 with aspirin in patients with atherosclerosis: a double-blind comparison to clopidogrel with aspirin. Eur Heart J. 2006;27:1038–7.
  • Tantry US, Bliden KP, Wei C, Storey RF, Armstrong M, Butler K, et al. First analysis of the relation between CYP2C19 genotype and pharmacodynamics in patients treated with ticagrelor versus clopidogrel: the ONSET/ OFFSET and RESPOND genotype studies. Circ Cardiovasc Genet. 2010;3:556–66.
  • Goto S. Cilostazol: potential mechanism of action for antithrombotic effects accompanied by a low rate of bleeding. Atheroscler Suppl. 2005;6:3–11.
  • Angiolillo DJ, Capranzano P, Goto S, Aslam M, Desai B, Charlton RK, et al. A randomized study assessing the impact of cilostazol on platelet function profiles in patients with diabetes mellitus and coronary artery disease on dual antiplatelet therapy: results of the OPTIMUS-2 study. Eur Heart J. 2008;29:2202–11.
  • Kamal AK, Naqvi I, Husain MR, Khealani BA. Cilostazol versus aspirin for secondary prevention of vascular events after stroke of arterial origin. Cochrane Database Syst Rev. 2011;(1):CD008076.
  • Raju NC, Sobieraj-Teague M, Hirsh J, O’Donnell M, Eikelboom J. Effect of aspirin mortality in primary prevention of cardiovascular disease. Am J Med. 2011;124:621–9.
  • Alonso-Coello P, Bellmunt S, McGorrian C, Anand SS, Guzman R, Criqui MH, et al. Antithrombotic therapy in peripheral artery disease: Antithrombotic Therapy and Prevetion of Thrombosis, 9th ed: American College of Chest Physicians Evidence-based clinical practice Guidelines. Chest. 2012;141(2 Suppl):e669S–90S.
  • Gandhi S, Weinberg I, Margey R, Jaff MR. Comprehensive medical management of peripheral arterial disease. Prog Cardiovasc Dis. 2011;54:2–13.
  • Hirsch AT, Haskal ZJ, Hertzer NR, Bakal CW, Creager MA, Halperin JL, et al. Practice guidelines for the management of patients with peripheral arterial disease (lower extremity, renal, mesenteric, and abdominal aortic): a collaborative report from the American Association for Vascular Surgery/Society for Vascular Surgery, Society for Cardiovascular Angiography and Interventions, Society for Vascular Medicine and Biology, Society of Interventional Radiology, and the ACC/AHA Task Force on Practice Guidelines (Writing Committee to Develop Guidelines for the Management of Patients With Peripheral Arterial Disease): endorsed by the American Association of Cardiovascular and Pulmonary Rehabilitation; National Heart, Lung, and Blood Institute; Society for Vascular Nursing; TransAtlantic Inter-Society Consensus; and Vascular Disease Foundation. Circulation. 2006;113:e463–e654.
  • Girolami B, Bernardi E, Prins MH, Ten Cate JW, Hettiarachchi R, Prandoni P, et al. Treatment of intermittent claudication with physical training, smoking cessation, pentoxifylline, or nafronyl: a meta- analysis. Arch Intern Med. 1999;159:337–45.
  • Anderson JL, Halperin JL, Albert NM, Bozkurt B, Brindis RG, Curtis LH, et al. Management of patients with peripheral artery disease (compilation of 2005 and 2011 ACCF/AHA guideline recommendations): a report of the American College of Cardiology Foundation/American Heart Association Task Force on Practice Guidelines. J Am Coll Cardiol. 2013;61:1555–70.
  • Heart Protection Study Collaborative Group. Randomized trial of the effects of cholesterol-lowering with simvastatin on peripheral vascular and other major vascular outcomes in 20,536 people with peripheral arterial disease and other high-risk conditions. J Vasc Surg. 2007; 45:645–54.
  • McDermott MM, Guralnik JM, Greenland P, Pearce WH, Criqui MH, Liu K, et al. Statin use and leg functioning in patients with and without lower-extremity peripheral arterial disease. Circulation. 2003;107:757–61.
  • Mohler ER 3rd, Hiatt WR, Creager MA. Cholesterol reduction with atorvastatin improves walking distance in patients with peripheral arterial disease. Circulation. 2003;108:1481–6.
  • Pedersen TR, Kjekshus J, Pyorala K, Olsson AG, Cook TJ, Musliner TA, et al. Effect of simvastatin on ischemic signs and symptoms in the Scandinavian Simvastatin Survival Study (4S). Am J Cardiol. 1998;81: 333–5.
  • Beckman JA, Creager MA. The non-lipid effects of statins on endothelial function. Trends Cardiovasc Med. 2006;16:156–62.
  • Grundy SM, Cleeman JI, Merz CN, Brewer HB Jr, Clark LT, Hunninghake DB, et al. Implications of recent clinical trials for the National Cholesterol Education Program Adult Treatment Panel III guidelines. Circulation. 2004;110:227–39.
  • Mancia G, De Backer G, Dominiczak A, Cifkova R, Fagard R, Germano G, et al. 2007 Guidelines for the management of arterial hypertension: The Task Force for the Management of Arterial Hypertension of the European Society of Hypertension (ESH) and of the European Society of Cardiology (ESC). Eur Heart J. 2007;28: 1462–536.
  • Yusuf S, Sleight P, Pogue J, Bosch J, Davies R, Dagenais G. Effects of an angiotensin converting enzyme inhibitor, ramipril, on cardiovascular events in high-risk patients. The Heart Outcomes Prevention Evaluation Study Investigators. N Engl J Med. 2000;342:145–53.
  • Adler AI, Stratton IM, Neil HA, Yudkin JS, Matthews DR, Cull CA, et al. Association of systolic blood pressure with macrovascular and microvascular complications of type 2 diabetes (UKPDS 36): prospective observational study. BMJ. 2000;321:412–19.
  • Intensive blood glucose control with sulphonylureas or insulin compared with conventional treatment and risk of complications in patients with type 2 diabetes (UKPDS 33). UK Prospective Diabetes Study (UKPDS) Group. Lancet. 1998;352:837–53.
  • ; ADVANCE Collaborative GroupPatel A, MacMahon S, Chalmers J, Neal B, Billot L, Woodward M, et al. Intensive blood glucose control and vascular outcomes in patients with type 2 diabetes. N Engl J Med. 2008;358:2560–72.
  • Duckworth W, Abraira C, Moritz T, Reda D, Emanuele N, Reaven PD, et al. Glucose control and vascular complications in veterans with type 2 diabetes. N Engl J Med. 2009;360:129–39.
  • Nathan DM, Cleary PA, Backlund JY, Genuth SM, Lachin JM, Orchard TJ, et al. Intensive diabetes treatment and cardiovascular disease in patients with type 1 diabetes. N Engl J Med. 2005;353: 2643–53.
  • Carter RE, Lackland DT, Cleary PA, Yim E, Lopes-Virella MF, Gilbert GE, et al. Intensive treatment of diabetes is associated with a reduced rate of peripheral arterial calcification in the diabetes control and complications trial. Diabetes Care. 2007;30:2646–8.
  • De Rubertis BG, Pierce M, Ryer EJ, Trocciola S, Kent KC, Faries PL. Reduced primary patency rate in diabetic patients after percutaneous intervention results from more frequent presentation with limb-threatening ischemia. J Vasc Surg. 2008;47:101–8.
  • American Diabetes Association. Standards of medical care in diabetes—2010. Diabetes Care.2010;33:S11–61.
  • CAPRIE Steering Committee. A randomized, blinded trial of clopidogrel versus aspirin in patients at risk of ischaemic events (CAPRIE). Lancet. 1996;348:1329–39.
  • Peters RJ, Mehta SR, Fox KA, Zhao F, Lewis BS, Kopecky SL, et al. Effects of aspirin dose when used alone or in combination with clopidogrel in patients with acute coronary syndromes: observations from the Clopidogrel in Unstable angina to prevent Recurrent Events (CURE) study. Circulation. 2003;108:1682–7.
  • Aronow HD, Steinhubl SR, Brennan DM, Berger PB, Topol EJ; CREDO Investigators. Bleeding risk associated with 1 year of dual antiplatelet therapy after percutaneous coronary intervention: insights from the Clopidogrel for the Reduction of Events During Observation (CREDO) trial. Am Heart J. 2009;157:369–74.
  • Sudlow CL, Mason G, Maurice JB, Wedderburn CJ, Hankey GJ. Thienopyridine derivatives versus aspirin for preventing stroke and other serious vascular events in high vascular risk patients. Cochrane Database Syst Rev. 2009;(4):CD001246.
  • Bhatt DL, Fox KA, Hacke W, Berger PB, Black HR, Boden WE, et al. Clopidogrel and aspirin versus aspirin alone for the prevention of atherothrombotic events. N Engl J Med. 2006;354:1706–17.
  • Keller TT, Squizzato A, Middeldorp S. Clopidogrel plus aspirin versus aspirin alone for preventing cardiovascular disease. Cochrane Database Syst Rev. 2007;(3):CD005158.
  • Robless P, Mikhailidis DP, Stansby GP. Cilostazol for peripheral arterial disease. Cochrane Database Syst Rev. 2008;(1):CD003748.
  • Squires H, Simpson E, Meng Y, Harnan S, Stevens J, Wong R. Cilostazol, naftidrofuryl oxalate, pentoxifylline, and inositol nicotinate for intermittent claudication in people with peripheral arterial disease. London, UK: National Institute for Health and Clinical Excellence; 2011. pp. 1–252.
  • Dawson DL, Cutler BS, Hiatt WR, Hobson RW 2nd, Martin JD, Bortey EB, et al. A comparison of cilostazol and pentoxifylline for treating intermittent claudication. Am J Med. 2000;109:523–30.
  • Pratt CM. Analysis of the cilostazol safety database. Am J Cardiol. 2001;87:28D–33D.
  • Tamhane U, Meier P, Chetcuti S, Chen KY, Rha SW, Grossman MP, et al. Efficacy of cilostazol in reducing restenosis in patients undergoing contemporary stent based PCI: a meta-analysis of randomised controlled trials. EuroIntervention. 2009;5:384–93.
  • Creutzig A, Lehmacher W, Elze M. Meta-analysis of randomised controlled prostaglandin E1 studies in peripheral arterial occlusive disease stages III and IV. Vasa. 2004;33:137–44.
  • Prostanoids for chronic critical leg ischemia: a randomized, controlled, open-label trial with prostaglandin E1. The ICAI Study Group. Ann Intern Med. 1999;130:412–21.
  • Ruffolo AJ, Romano M, Ciapponi A. Prostanoids for critical limb ischaemia. Cochrane Database Syst Rev. 2010;(1):CD006544.
  • Berridge DC, Kessel D, Robertson I. Surgery versus thrombolysis for acute limb ischaemia: initial management. Cochrane Database Syst Rev. 2002;(3):CD002784.
  • Results of a prospective randomized trial evaluating surgery versus thrombolysis for ischemia of the lower extremity. The STILE trial. Ann Surg. 1994;220:251–66; discussion 266–8.
  • Robertson I, Kessel DO, Berridge DC. Fibrinolytic agents for peripheral arterial occlusion. Cochrane Database Syst Rev. 2010;(3):CD001099.
  • Balzer KM, Weis-Müller BT. Results of open vascular surgical therapy in chronic peripheral arterial disease. Vasa. 2011;40:359–67.
  • Dattiloa PB, Casserly IP. Critical limb ischemia: endovascular strategies for limb salvage. Prog Cardiovasc Dis. 2011;54:47–60.
  • Safian RD, Niazi K, Runyon JP, Dulas D, Weinstock B, Ramaiah V, et al. Orbital atherectomy for infrapopliteal disease: device concept and outcome data for the oasis trial. Catheter Cardiovasc Interv. 2009;73: 406–12.
  • Adam DJ, Beard JD, Cleveland T, Bell J, Bradbury AW, Forbes JF, et al. Bypass versus angioplasty in severe ischaemia of the leg (basil): multicentre, randomized controlled trial. Lancet. 2005;366:1925–34.
  • Romiti M, Albers M, Brochado-Neto FC, Durazzo AE, Pereira CA, De Luccia N. Meta-analysis of infrapopliteal angioplasty for chronic critical limb ischemia. J Vasc Surg. 2008;47:975–81.
  • De Vries JP, Karimi A, Fioole B, Van Leersum M, Werson DA, Van Den Heuvel DA. First- and second-generation drug-eluting balloons for femoro-popliteal arterial obstructions: update of technique and results. J Cardiovasc Surg. 2013;54:327–32.
  • Seedial SM, Ghosh S, Saunders RS, Suwanabol PA, Shi X, Liu B, et al. Local drug delivery to prevent restenosis. J Vasc Surg. 2013; 57:1403–14.
  • Minar E. Femoral stenting: what do the guidelines tell us. J Cardiovasc Surg. 2013;54:211–15.
  • Commeau P, Barragan P, Roquebert PO. Sirolimus for below the knee lesions: mid-term results of SiroBTK study. Catheter Cardiovasc Interv. 2006;68:793–8.
  • Feiring AJ, Krahn M, Nelson L, Wesolowski A, Eastwood D, Szabo A. Preventing leg amputations in critical limb ischemia with below- the-knee drug-eluting stents: the paradise (preventing amputations using drug eluting stents) trial. J Am Coll Cardiol. 2010;55:1580–9.
  • Martens JM, Knippenberg B, Vos JA, de Vries JP, Hansen BE, van Overhagen H; PADI Trial Group. Update on PADI trial: percutaneous transluminal angioplasty and drug-eluting stents for infrapopliteal lesions in critical limb ischemia. J Vasc Surg. 2009;50:687–9.
  • Kasapis C, Henke PK, Chetcuti SJ, Koenig GC, Rectenwald JE, Krishnamurthy VN, et al. Routine stent implantation vs. percutaneous transluminal angioplasty in femoropopliteal artery disease: a meta-analysis of randomized controlled trials. Eur Heart J. 2009;30:44–55.
  • Dake MD, Ansel GM, Jaff MR, Ohki T, Saxon RR, Smouse HB, et al.; Zilver PTX trial Investigators. Sustained safety and effectiveness of paclitaxel-eluting stents for femoropopliteal lesions: 2-year follow-up from the Zilver PTX randomized and single-arm clinical studies. J Am Coll Cardiol. 2013;61:2417–27.
  • Balzer KM, Sandmann W. Das biologische Verhalten von Homograft s zum arteriellen Gefäßersatz. Gefässchirurgie. 2010;15:101–7.
  • Clagett GP, Valentine RJ, Hagino RT. Autogenous aortoiliac/femoral reconstruction from superficial femoral-popliteal veins: feasibility and durability. J Vasc Surg. 1997;25:255–70.
  • Chiu KWH, Davies RSM, Nightingale PG, Bradbury AW, Adam DJ. Review of direct anatomical open surgical management of atherosclerotic aorto-iliac occlusive disease. Eur J Vasc Endovasc Surg. 2010; 39:460–71.
  • Kashyap VS, Pavkov ML, Bena JF, Sarac TP, O‘Hara PJ, Lyden SP, et al. The management of severe aortoiliac occlusive disease: endovascular therapy rivals open reconstruction. J Vasc Surg. 2008;48: 1451–7.
  • Schwarzwälder U, Zeller T. Debulking procedures: potential device specific indications. Tech Vasc Interv Radiol. 2010;13:43–53.
  • Dörffler-Melly J, Koopman MM, Prins MH, Büller HR. Antiplatelet and anticoagulant drugs for prevention of restenosis/reocclusion following peripheral endovascular treatment. Cochrane Database Syst Rev. 2005;(1):CD002071.
  • Schweizer J, Müller A, Forkmann L, Hellner G, Kirch W. Potential use of a low-molecular-weight heparin to prevent restenosis in patients with extensive wall damage following peripheral angioplasty. Angiology. 2001;52:659–69.
  • Brown J, Lethaby A, Maxwell H, Wawrzyniak AJ, Prins MH. Antiplatelet agents for preventing thrombosis after peripheral arterial bypass surgery. Cochrane Database Syst Rev. 2008;(4):CD000535.
  • Belch JJ, Dormandy J; CASPAR Writing Committee, Biasi GM, Cairols M, Diehm C, et al. Results of the randomized, placebo-controlled clopidogrel and acetylsalicylic acid in bypass surgery for peripheral arterial disease (CASPAR) trial. J Vasc Surg. 2010;52:825–33.
  • Efficacy of oral anticoagulants compared with aspirin after infrainguinal bypass surgery (The Dutch Bypass Oral Anticoagulants or Aspirin Study): a randomised trial. Lancet. 2000;355:346–51.
  • Vandvik PO, Lincoff AM, Gore JM, Gutterman DD, Sonnenberg FA, Alonso-Coello P, et al. Primary and secondary prevention of cardiovascular disease: antithrombotic therapy and prevention of thrombosis, 9th ed: American College of Chest Physicians evidence-based clinical practice guidelines. Chest. 2011;141:e637S–68S.
  • Sofi F, Marcucci R, Gori AM, Giusti B, Abbate R, Gensini GF. Clopidogrel non-responsiveness and risk of cardiovascular morbidity. An updated meta-analysis. Thromb Haemost. 2010;103:841–8.
  • Di Minno G. Aspirin resistance and platelet turnover: a 25-year old issue. Nutr Metab Cardiovasc Dis. 2011;21:542–5.
  • Simon T, Verstuyft C, Mary-Krause M, Quteineh L, Drouet E, Méneveau N, et al. Genetic determinants of response to clopidogrel and cardiovascular events. N Engl J Med. 2009;360:363–75.
  • Wallentin L, Becker RC, Budaj A, Cannon CP, Emanuelsson H, Held C, et al. Ticagrelor versus clopidogrel in patients with acute coronary syndromes. N Engl J Med. 2009;361:1045–57.
  • Di Minno G, Russolillo A, Gambacorta C, Di Minno A, Prisco D. Improving the use of direct oral anticoagulants in atrial fibrillation. Eur J Intern Med. 2013;24:288–94.
  • De Caterina R, Husted S, Wallentin L, Andreotti F, Arnesen H, Bachmann F, et al. New oral anticoagulants in atrial fibrillation and acute coronary syndromes: ESC Working Group on Thrombosis-Task Force on Anticoagulants in Heart Disease position paper. J Am Coll Cardiol. 2012;59:1413–25.
  • Southworth MR, Reichman ME, Unger EF. Dabigatran and postmarketing reports of bleeding. New Engl J Med. 2013;368:1272–4.
  • Knowles JW, Assimes TL, Li J, Quertermous T, Cooke JP. Genetic susceptibility to peripheral arterial disease: a dark corner in vascular biology. Arterioscler Thromb Vasc Biol. 2007;27:2068–78.
  • Katwal AB, Dokun AO. Peripheral arterial disease in diabetes: is there a role for genetics? Curr Diab Rep. 2011;11:218–25.
  • Zintzaras E, Zdoukopoulos N. A field synopsis and meta-analysis of genetic association studies in peripheral arterial disease: the CUMAGAS-PAD database. Am J Epidemiol. 2009;170:1–11.
  • Leeper NJ, Kullo IJ, Cooke JP. Genetics of peripheral artery disease. Circulation. 2012;125:3220–8.
  • Violi F, Lip GYH, Basili S. Peripheral artery disease and atrial fibrillation: a potentially dangerous combination. Intern Emerg Med. 2012;7:213–18.
  • Garimella PS, Hart PD, O’Hare A, DeLoach S, Herzog CA, Hirsch AT. Peripheral artery disease and CKD: a focus on peripheral artery disease as a critical component of CKD care. Am J Kidney Dis. 2012; 60:641–54.
  • O’Hare A, Johansen K. Lower-extremity peripheral arterial disease among patients with end-stage renal disease. J Am Soc Nephrol. 2001;12:2838–47.
  • Violi F, Criqui M, Longoni A, Castiglioni C. Relation between risk factors and cardiovascular complications in patients with peripheral vascular disease. Results from the A.D.E.P. study. Atherosclerosis. 1996;120:25–35.
  • Sarnak M, Levey AS, Schoolwerth AC, Coresh J, Culleton B, Hamm LL, et al. Kidney disease as a risk factor for development of cardiovascular disease: a statement from the American Heart Association Councils on Kidney in Cardiovascular Disease, High Blood Pressure Research, Clinical Cardiology, and Epidemiology and Prevention. Circulation. 2003;108:2154–69.
  • Marso SP, Hiatt WR. Peripheral arterial disease in patients with diabetes. J Am Coll Cardiol. 2006;47:921–9.
  • Selvin E, Erlinger TP. Prevalence of and risk factors for peripheral arterial disease in the United States: results from the National Health and Nutrition Examination Survey, 1999-2000. Circulation.2004;110:738–43.
  • Kannel WB, McGee DL. Update on some epidemiologic features of intermittent claudication: the Framingham Study. J Am Geriatr Soc. 1985;33:13–18.
  • Mackaay AJ, Beks PJ, Dur AH, Bischoff M, Scholma J, Heine RJ, et al. The distribution of peripheral vascular disease in a Dutch Caucasian population: comparison of type II diabetic and non-diabetic subjects. Eur J Vasc Endovasc Surg. 1995;9:170–5.
  • Dosluoglu HH, Lall P, Nader ND, Harris LM, Dryjski ML. Insulin use is associated with poor limb salvage and survival in diabetic patients with chronic limb ischemia. J Vasc Surg. 2010;51:1178–89.
  • Jude EB, Oyibo SO, Chalmers N, Boulton AJ. Peripheral arterial disease in diabetic and nondiabetic patients: a comparison of severity and outcome. Diabetes Care. 2001;24:1433–7.
  • Bild DE, Selby JV, Sinnock P, Browner WS, Braveman P, Showstack JA. Lower-extremity amputation in people with diabetes. Epidemiology and prevention. Diabetes Care. 1989;12:24–31.
  • Wattanakit K, Folsom AR, Selvin E, Weatherley BD, Pankow JS, Brancati FL, et al. Risk factors for peripheral arterial disease incidence in persons with diabetes: the Atherosclerosis Risk in Communities (ARIC) Study. Atherosclerosis. 2006;180:389–97.
  • Ankle Brachial Index Collaboration; Fowkes FG, Murray GD, Butcher I, Heald CL, Lee RJ, Chambless LE, et al. Ankle brachial index combined with Framingham Risk Score to predict cardiovascular events and mortality: a meta-analysis. JAMA. 2008;300:197–208.
  • Bertelè V, Roncaglioni MC, Pangrazzi J, Terzian E, Tognoni EG. Clinical outcome and its predictors in 1560 patients with critical leg ischaemia. Chronic Critical Leg Ischaemia Group. Eur J Vasc Endovasc Surg. 1999;18:401–10.
  • Berger JS, Hiatt WR. Medical therapy in peripheral artery disease. Circulation. 2012;126:491–500.
  • Davies MG. Critical limb ischemia: cell and molecular therapies for limb salvage. Methodist Debakey Cardiovasc J. 2012;8:20–7.
  • Rioufol G, Finet G, Ginon I, André-Fouët X, Rossi R, Vialle E, et al. Multiple atherosclerotic plaque rupture in acute coronary syndrome: a three-vessel intravascular ultrasound study. Circulation. 2002; 106:804–8.