You are on page 1of 22

ARTICLE IN PRESS

Biomaterials 28 (2007) 16891710 www.elsevier.com/locate/biomaterials

Review

Coronary stents: A materials perspective


Gopinath Mania, Marc D. Feldmanb,c, Devang Patelb, C. Mauli Agrawala,
Department of Biomedical Engineering, College of Engineering, The University of Texas at San Antonio, One UTSA Circle, San Antonio, TX 78249 0619, USA b Division of Cardiology, Department of Medicine, The University of Texas Health Science Center at San Antonio, 7703 Floyd Curl Drive, San Antonio, TX 78229 3900, USA c The Department of Veterans Affairs South Texas Health Care System, San Antonio, TX 78229 4404, USA Received 5 September 2006; accepted 29 November 2006 Available online 22 December 2006
a

Abstract The objective of this review is to describe the suitability of different biomaterials as coronary stents. This review focuses on the following topics: (1) different materials used for stents, (2) surface characteristics that inuence stentbiology interactions, (3) the use of polymers in stents, and (4) drug-eluting stents, especially those that are commercially available. r 2006 Elsevier Ltd. All rights reserved.
Keywords: Stent; Surface treatment; Surface modication; Drug delivery

Contents 1. 2. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1690 Metallic stents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1690 2.1. 316L SS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1691 2.2. PtIr alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1691 2.3. Ta . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1691 2.4. Ti . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1692 2.5. NiTi. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1692 2.6. CoCr alloy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1693 2.7. Biodegradable metallic stents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1693 2.7.1. Pure Fe . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1693 2.7.2. Mg alloys. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1693 Surface characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1694 3.1. Surface energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1694 3.2. Surface texture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1694 3.3. Surface potential. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1694 3.4. Stability of surface oxide layer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1695 Rationale for coatings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1695 4.1. Types of coatings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1695 4.1.1. Inorganic coatings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1695 4.1.2. Endothelial cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1696 4.1.3. Porous materials. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1696 Polymers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1697

3.

4.

5.

Corresponding author. Tel.: +1 210 458 5526; fax: +1 210 458 5556.

E-mail address: mauli.agrawal@utsa.edu (C.M. Agrawal). 0142-9612/$ - see front matter r 2006 Elsevier Ltd. All rights reserved. doi:10.1016/j.biomaterials.2006.11.042

ARTICLE IN PRESS
1690 G. Mani et al. / Biomaterials 28 (2007) 16891710

6.

7.

Biostable polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1697 Biodegradable polymers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1697 Copolymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1698 Biological polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1698 5.4.1. PC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1698 5.4.2. HA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1698 5.4.3. Fibrin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1699 Rationale for DES. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1699 6.1. Techniques for drug-loading and release kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1699 6.2. DES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1699 6.2.1. Heparin-coated stents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1699 6.2.2. Sirolimus-eluting stents (SES) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1700 6.2.3. Paclitaxel-eluting stents (PES) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1701 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1702 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1703

5.1. 5.2. 5.3. 5.4.

1. Introduction Percutaneous transluminal coronary angioplasty (PTCA) is an invasive procedure performed to reduce blockages in coronary arteries [1,2]. However, restenosis follows PTCA in 3040% of coronary lesions within 6 months [3,4]. Although providing intra-arterial support with bare metal stents (BMS) dramatically improves the angiographic and clinical outcome of patients to a restenosis rate of 2030% [3,4], in-stent restenosis still remains a major limitation for this approach with exaggerated intimal hyperplasia [5]. The biology of restenosis in stents includes plaque redistribution, thrombosis, and neointimal hyperplasia [6]. The basic mechanisms [79] underlying thrombus formation and neointimal muscle cell proliferation, followed by extracellular expansion are understood to some extent, but the basic biology of restenosis still remains an active area of research. As a result of the inadequacies of BMS, different kinds of materials, designs, and techniques have been explored to further optimize stent design. Coronary stents developed to date can be grouped in four categories: bare metallic stents, coated metallic stents, biodegradable stents and drugeluting stents (DES). The advent of DES, which release drugs such as sirolimus and paclitaxel for localized delivery, is a major advancement in the evolution of stents. However, there is a risk of late stent thrombosis (LST) associated with DES [10,11]. This review evaluates the pros and cons of choosing different materials for the manufacture of coronary stents. The physical properties of each material that are relevant for this application are discussed. The inuence of a materials surface characteristics on the biology of restenosis will be discussed as well. A variety of coating materials are commonly used in an attempt to improve the performance of stents; including inorganic materials, polymers, endothelial cells, and porous ceramics. The role of these different types of coatings is described in detail. The materials and the coating techniques used in commer-

cially available DES are described. A list of ideal characteristics for coronary stents and the materials and processes that best meet these requirements are tabulated in the concluding section. The physical design of a stent, another important parameter, is not covered here as the discussion is conned to biomaterials.

2. Metallic stents Balloon expandable stents should have the ability to undergo plastic deformation and then maintain the required size once deployed [12]. Self-expanding stents, on the other hand, should have sufcient elasticity to be compressed for delivery and then expanding in the target area [12]. The characteristics of an ideal stent have been described in numerous reviews [1315]. In general, it should have (1) low proleability to be crimped on the balloon catheter supported by a guide wire; (2) good expandability ratioonce the stent is inserted at the target area and the balloon is inated, the stent should undergo sufcient expansion and conform to the vessel wall; (3) sufcient radial hoop strength and negligible recoilonce implanted, the stent should be able to overcome the forces imposed by the atherosclerotic arterial wall and should not collapse; (4) sufcient exibilityit should be exible enough to travel through even the smaller diameter atherosclerotic arteries; (5) adequate radiopacity/magnetic resonance imaging (MRI) compatibilityto assist clinicians in assessing the in-vivo location of the stent; (6) thromboresistivitythe material should be blood compatible and not encourage platelet adhesion and deposition; and (7) drug delivery capacitythis has become one of the indispensable requirements for stents of the modern era to prevent restenosis. Generally, the metals commonly used for manufacturing stents are 316L stainless steel (316L SS), platinumiridium (PtIr) alloy, tantalum (Ta), nitinol (NiTi), cobalt chromium (CoCr) alloy, titanium (Ti), pure iron (Fe),

ARTICLE IN PRESS
G. Mani et al. / Biomaterials 28 (2007) 16891710 Table 1 Mechanical properties of the metals that are used for making stents Metal Elastic modulus (GPa) Yield strength (MPa) Tensile strength (MPa) 586 207 760 895 Density (g/cm3) 7.9 16.6 4.5 6.7 References 1691

316L stainless steel (ASTM F138 and F139; annealed) Tantalum (annealed) CpTitanium (F67; 30% cold worked) Nitinol

190 185 110 83 (Austenite phase) 2841 (Martensite phase)

331 138 485 195690 (Austenite phase) 70140 (Martensite phase) 448648 120150 162

[23,59] [23] [23,59] [23,37]

Cobaltchromium (ASTM F90) Pure iron Mg alloy (WE43)

210 211.4 44

9511220 180210 250

9.2 7.87 1.84

[23,37,59] [269] [71,270]

and magnesium (Mg) alloys. These are briey discussed below: 2.1. 316L SS Whether it is a bare stent or with a coating material, 316L SS is the most commonly used metal for stents. It has wellsuited mechanical properties (Table 1) and excellent corrosion resistance (carbon content o0.030 wt%), making it the preferred material for this application [12]. However, the clinical limitations of using 316L SS are its ferromagnetic nature (6065 wt% pure Fe) and low density. These properties make SS a non-MRI compatible and poorly visible uoroscopic material. Also, biocompatibility is an issue with bare SS stents. The weight percentage of nickel, chromium, and molybdenum in 316L SS are 12, 17, and 2.5, respectively [16]. Allergic reactions to the release of nickel can occur among SS implants [17]. In particular, the release of nickel, chromate, and molybdenum ions from SS stents may trigger local immune response and inammatory reactions, which in turn may induce intimal hyperplasia and in-stent restenosis [18]. SS stents made of grades with lower nickel content can reduce the concern over allergic reactions to nickel. A number of SS grades with low nickel concentration (4.59%) are available [16]. However, it has been shown that higher nickel content (1014%) can be advantageous in decreasing the ferromagnetic properties of SS by stabilizing Fe in a non-magnetic state [19]. Hence, the SS grades (321 and 321H) with optimal nickel (10.5%) and carbon (0.08%) concentrations are promising. The addition of Ti (0.4%) in these grades makes them more attractive. A variety of materials have been used as a coating for 316L SS stents, mainly to circumvent its visibility limitations and to improve its biocompatibility by preventing the release of ions from the metal surface. However, the supremacy of 316L SS platforms for making stents is evident from Table 2. Out of the eight coronary stents approved by the US Food and Drug Administration (FDA), seven are made from 316L SS.

2.2. PtIr alloys An alloy of 90% platinum and 10% iridium was used for making bare stents and successfully implanted in animal models [20,21]. These stents showed excellent radiopacity [20] and it is even possible to take the three-dimensional image of the lumen of these stents using MRI [22]. The artifacts produced by the PtIr alloy in MRI are much lower when compared with 316L SS stents [19,22]. The presence of iridium in the alloy could pave the way for potential applications in radioactive stents [21]. In general, these alloys show excellent corrosion resistance but poor mechanical properties [23,24]. Although a reduction in both thrombosis and neointimal proliferation with less inammatory reactions was observed for these stents, their recoiling percentage was higher (16%) than the 316L SS stents (5%) [20,21]. Though a human clinical trial [25] encouraged the use of these stents as safe and effective, the literature on the biocompatibility and haemocompatibility of PtIr (90/10) alloys remains limited. 2.3. Ta Ta has excellent corrosion resistance because of its highly stable surface oxide layer, which prevents electron exchange between the metal and the adsorbed biological species [26,27]. It has been coated on a 316L SS surfaces to improve corrosion properties, thereby enhancing the biocompatibility of 316L SS [28]. It has excellent uoroscopic visibility because of its high density. It is an MRI compatible material as it produces no signicant artifacts because of its non-ferromagnetic properties [29,30]. Ta is also known for its good biocompatibility [31,32]. Enhanced hemocompatibility was achieved by adding Ta to Ti oxide and the lms showed improved endothelialization rate as the percentage concentration of Ta increased [33,34]. Though the biocompatibility and visibility properties of Ta are superior to 316L SS, the commercial availability of Ta stents is lower than 316L SS stents. This is mainly

ARTICLE IN PRESS
1692 G. Mani et al. / Biomaterials 28 (2007) 16891710

FDA approval date

September 2000

October 2000 April 2003

July 2003 October 2003 March 2004

April 2005

April 2005

because of its poor mechanical properties. Since the yield strength of Ta is closer to its tensile strength (Table 1), these stents have a higher possibility of breaking during deployment. Hence, the pressure applied for the deployment of these stents is usually low and this might result in recoiling. The recoiling percentage was signicantly higher for Ta stents when compared with 316L SS stents and resulted in enhanced neointimal formation [35]. Although no Ta-based stents have been approved by the FDA for general use to date, Cordis (Johnson & Johnson, USA) has used a bare Ta stent in clinical trials and released this stent commercially in Japan, Canada and Europe [36]. 2.4. Ti Ti and its alloys have been extensively used in orthopedic and dental biomedical applications because of their excellent biocompatibility [24,37]. The highly stable surface oxide layer provides excellent corrosion resistance [24,37]. However, Ti is not commonly used for making stents. Although Ti and CoCr both have high yield strength in approximately the same range, Ti has a signicantly lower tensile strength (Table 1). Thus, there is a higher probability of tensile failure of the Ti stents when expanded to stresses beyond their yield strengths, which is the norm in balloon expandable stent deployment. This is one of the reasons why CoCr is used for making stents and not Ti. Alloying Ti with materials that reduce its yield strength while retaining tensile properties might prove to be optimum. Because of its low ductility, Ti stents are more prone to fracture. Because of these inadequate mechanical properties, commercially pure Ti failed to make an impact as the sole stent material. However, the applications of Ti are not limited to coronary stent applications. Ti-nitrideoxide coating on 316L SS was found to be biologically inert with reduced platelet and brinogen deposition and thereby reducing neointimal hyperplasia [38]. The Titan stent (Hexacath, France), which has implemented this coating technique has shown promising results in human clinical trials [39,40]. Also, Ti-based Ta and niobium alloys, which have potential applications for stents, showed excellent haemocompatibility [41]. One of the Ti alloys which is extensively used for making stents is NiTi. 2.5. NiTi NiTi constitutes 49.557.5 at% nickel and the remaining is Ti [42]. It is used for fabricating self-expanding stents mainly because of its shape memory effect. Self-expanding stents have a smaller diameter at room temperature and expand to their preset diameter at body temperature [43]. NiTi is plastically deformed at room temperature (martensitic phase) and crimped on to the delivery system [37,42,43]. After implantation it regains its original shape (already memorized austenite phase according to the diameter of the target vessel) and conforms to the vessel wall because of the increase in temperature inside the body

References

[271]

[272] [230]

[65] [273] [274]

[275] TM MonorailTM Liberte 316L stainless steel

Nil First coat: Parylene C; second coat: mixture of polyethylene-covinyl acetate, poly n-butyl methacrylate, and Sirolimus; third coat: mixture of polyethylene-covinyl acetate, poly n-butyl methacrylate Nil Nil Mixture of poly(styrene-bisobutylene-b-styrene) triblock copolymer and paclitaxel Nil L-605 cobalt chromium alloy 316L stainless steel 316L stainless steel 316L stainless steel 316L stainless steel MULTI-LINK VISIONTM NIRexTM TAXUSTM Express2TM BeStentTM 2 CYPHERTM Guidant Corporation, CA Medinol Ltd., Israel Boston Scientic Corporation

Cross-linked phosphorylcholine

Bare stent material

Coating

Biocompatibles Cardiovascular Inc. CA Medtronic, Inc., Minnesota Cordis Corporation, FL

316L stainless steel

Table 2 FDA approved coronary stents

BiodivYsioTM AS

Rithron-XR

Stent name

Boston Scientic Corporation, MN Biotronik GmbH, Germany

Manufacturer

316L stainless steel

Amorphous silicon-carbide

[276]

ARTICLE IN PRESS
G. Mani et al. / Biomaterials 28 (2007) 16891710 1693

[37,43]. The maximum strain recovery is 8.5% after plastic deformation [37]. NiTi also has suitable mechanical properties [37] (Table 1). However, the corrosion resistance of NiTi is actively debated. Though the literature generally portrays NiTi as a corrosion resistant material [37,42,44], the release of nickel ions and their toxic effects to tissues have been reported in many cases [45,46]. In order to overcome this problem, the surface is passivated to increase the Ti oxide concentration at the surface and thereby reduce the nickel concentration [47,48]. This can be achieved by plasma-immersion ion implantation [48], nitric acid treatments [49], heat treatments [50], and electropolishing [50]. Also, some of the materials like polyurethane [51], Ti nitride [52], and polycrystalline oxides [53] have been coated on NiTi stents mainly to improve the corrosion resistance. NiTi stents are not adequately visible by uoroscopy and this is an issue [54]. Although MRI visualizes the stent [55], most stent deployment is performed under uoroscopy. ACT-OneTM (Progressive Angioplasty Systems, USA) [56], Paragon (Progressive Angioplasty Systems, USA) [57], and Radius (Scimed, USA) [58] are some of NiTi stents used in clinical trials. 2.6. CoCr alloy CoCr alloys, which conform to ASTM standards F562 and F90, have been used in dental and orthopedic applications for decades [59] and recently have been used for making stents. These alloys have excellent radial strength because of their high elastic modulus (Table 1). The thickness of the struts is a critical issue in designing a stent [6062], hence, the ability to make ultra-thin struts with increased strength using these alloys is one of their main attractions [63]. In addition to this, they are radiopaque [63] and MRI-compatible [64]. The cobalt alloy platform DRIVER stents (Medtronic Inc, USA) are commercially available in Europe. Recently, the FDA approved the L-605 CoCr alloy Guidant Multi-Link Vision stent [65]. 2.7. Biodegradable metallic stents Pure Fe [66] and Mg alloys [67] are the two metals that have been used for making biodegradable coronary stents recently. 2.7.1. Pure Fe Pure Fe (more than 99.5%) is the major component in degradable Fe stents [66,68]. Fe has superior radial strength because of its higher elastic modulus (Table 1). This can be helpful in making stents with thinner struts. Since the yield strength and tensile strength of pure Fe are close to each other (Table 1), theoretically, these stents may fracture during deployment. However, these stents were successfully deployed in rabbit and porcine arteries with balloon pressures of 3.5 and 10 atmospheres, respectively [66,68]. Fluoroscopy was used to view these stents (strut thickness varied from 100 to 120 mm) [66,68]. The biodegradation involves the oxidation of Fe into ferrous

and ferric ions and these ions dissolve into biological media [69]. Ferrous ions reduce the proliferation of smooth muscle cells in in-vitro conditions, and thus may inhibit neointimal hyperplasia [69]. Thrombogenicity and neointimal proliferation were reduced and no local toxicity was observed [66,68]. Endothelialization of Fe stents was also observed in animal models [66,68]. These studies were limited by the small study groups [66] and slow degradation kinetics of Fe [66,68]. 2.7.2. Mg alloys Mg and its alloys have been previously used for biodegradable orthopedic implants [70]. However, these materials are novel in their application to coronary stents [67]. The mechanical and corrosion properties of pure Mg [7173] do not suit the requirements for stent material. Hence, Mg alloys with improved mechanical and corrosion properties [7173] were chosen for the purpose. AE21 [67] and WE43 [74] are the two Mg-based alloys reported in the literature for making stents. AE21 contains aluminum (2%) and rare earth metals (1%) [67]. WE43 contains 4% yttrium, 0.6% zirconium, and 3.4% rare earth metals [71]. The remaining component is Mg in both of these alloys. The typical mechanical properties of commercially available WE43 are tabulated in Table 1. It has poor radial strength because of its low elastic modulus. In order to provide proper vessel wall support, the struts have to be thicker and this increases the area of metalartery interaction. Mg alloy stents may fracture because of their low ductility. Also, these stents are radiolucent and cannot be imaged by X-rays. However, intravascular ultrasound and MRI techniques have used to visualize these stents [75]. In the physiologic environment, Mg corrodes into soluble Mg hydroxide, Mg chloride, and hydrogen gas [76]. However, it is vital to investigate the corrosion products of Mg alloys that are actually used for making stents. The Lekton Magic stent (Biotronik, Switzerland) is made from WE43 and implanted in porcine models [74]. Reduced smooth muscle cell growth and enhanced endothelialization were observed [74]. A Biotroniks Mg absorbable metal stent (AMS) was implanted in a baby and was well tolerated [77], but not in another baby [78]. These mixed results show the need for further research. The biodegradable metallic stent looks promising for the growing artery in children. However, the types of degradation products, size of these products, and their biocompatibility still need to be studied. Theoretically, the mechanical properties of Mg are poor for a coronary stent. Also, the degradation behavior of these stents is not controllable. Local toxicity of the degradation products of these stents is unlikely because Fe and Mg are present naturally in the human body [66,74]. However, the impacts of elevated local concentration of these elements are unknown. A detailed investigation is needed in this area based on large clinical trials. The materials for metal stents are often chosen with an emphasis on their engineering properties. Hence, there is

ARTICLE IN PRESS
1694 G. Mani et al. / Biomaterials 28 (2007) 16891710

a need to emphasize the materialbiology interactions early on in the technology development process. 3. Surface characteristics Surface characteristics of a stent material, which inuence thrombosis and neointimal hyperplasia, include surface energy, surface texture, surface potential, and the stability of the surface oxide layer [79,80]. In many circumstances a combination of one or more of these listed factors predicts the outcome. The surface properties of a material may depend on the surface treatment of the material. For example, microblasting produced a rough Ta surface with particle contaminants [81]. Reactive ion etching on the other hand, produced an even rougher surface on the same material but removed all the contaminants [81], thus demonstrating that different surface treatments can produce different surface textures and surface chemistries on the same material surface. This can provide different surface energies to the material surface. An in-vitro study showed that the adherence, growth and proliferation of endothelial cells on Ta lms were much better than on 316L SS and Ti lms [32]. This result can be attributed to the technique that the Ta lms were actually deposited (pulsed metal vacuum arc source deposition) and processed (annealed to 700 1C) rather than the nature of the material (Ta oxide) itself. In another study, the sputter coating of a TiTa target produced a surface that showed better endothelialization because of the changes in the microstructure of the natural Ti-oxide lm produced [34]. To improve the corrosion resistance, Ta is coated on a 316L SS substrate by physical vapor deposition sputtering [28]. However, when the material is plastically deformed, cracks appear in the coating [28]. Though it was claimed that the crack surfaces were repassivated by treating with the physiological saline, this kind of acute change (cracking) in surface morphology might pose a serious threat when the material is exposed to in-vivo conditions. The nature of the coating may be biocompatible, but, if the coating loses its integrity during the stent placement and expansion, it can cause adverse effects. 3.1. Surface energy The surface energy of a material as well as its surface chemistry affects its wettability. The thrombogenicity of a material surface increases with increasing surface energy [82,83]. The thrombogenic potentials of PET and PTFE were compared and the thrombogenicity was signicantly higher for PET (high surface energy) than PTFE (low surface energy) [84]. This result has initiated the use of coating metal surfaces that usually have higher surface energy with polymeric materials having low surface energy. A polyurethane coating on Ta and SS reduced the surface energy which resulted in the signicant reduction of thrombosis [85,86]. One limitation of these thromboresistant coatings (hydrophobic polymers) is that they not only

prevent the adherence of platelets but also endothelial cells [87]. This problem can be prevented by coating the polymer surfaces with bronectin, which enhances the endothelialization process [87,88]. A combination of human plasma proteins and heparin coating on copolymers, which are derived from hydrophilic 1-vinylpyrrolidinone and hydrophobic n-butyl-methacrylate, provided better endothelialization and thromboresistance to the material, respectively [89]. Seeger et al. [90] coated hydrophilic polymers such as N-vinylpyrrolidone and potassium sulfopropyl acrylate on SS to reduce the accumulation of platelets. These coatings indeed smoothened the irregularities of underlying SS. In spite of their similar hydrophilicity, there was a signicant difference in the platelet accumulation between the polymers. This can be attributed to different surface charges of the polymers [90]. Hence, when polymers are coated on a metal stent, the surface energy is not the only parameter that is altered but also the other parameters like surface texture and surface charge. 3.2. Surface texture Thrombogenicity is usually higher for rougher surfaces [9193], thus polishing is essential for stent materials. Acid pickling followed by electrochemical polishing has been used to remove the slag (formed on the stent during its laser cut) and to polish the stents, respectively [94]. It has been shown that polishing of coronary stents resulted in decreased thrombogenicity as well as neointimal hyperplasia in different animal models [95,96]. Also, the coating techniques used for surface deposition can directly inuence the surface texture. Hehrlein et al. [97] investigated the effect of two surface deposition methods, galvanization and ion implantation, on the biocompatibility of endovascular stents. Both thrombogenicity and neointimal hyperplasia were higher for the stents that are coated by galvanization because of the pores and cracks created during expansion [97]. Close control over the surface texture of stents is relatively difcult due to morphological changes during its expansion. Hence, when a stent is polished or coated, sufcient care should be taken to evaluate the effects of expansion. Recently, Sprague et al. [98] observed that the grooved surfaces double the migration rate of endothelial cells over polished and smooth controls. Larger grooves (on the order of 22 mm) resulted in greater migration rates and faster endothelialization times. Thus, clearly the surface roughness of the stent is an important parameter in their clinical success. This needs to be taken into consideration as new coatings are being developed, especially for drug elution, as coatings may have different surface textures compared to the bare metal. 3.3. Surface potential The net electrical charge on a material surface is also critical to the success of a stent [26,91]. Zitter et al. [26] investigated the current densities of different metals in

ARTICLE IN PRESS
G. Mani et al. / Biomaterials 28 (2007) 16891710 1695

in-vitro conditions and ranked the current densities in the following order: Au4316L SS4CoCr4Ti6Al4V4 Ti4Nb4Ta. The reasons for the good and poor biocompatibilities of Ta and Au, respectively were discussed with respect to the decreasing order of current densities [26]. Most metals are electropositive while blood elements tend to be electronegative, which accentuates the thrombogenicity problem [91]. Ta has a net negative electrical charge and thus has a theoretical advantage over other metals [26,91]. However, experimental results have been contradictory; Scott et al. [99] compared the thrombogenicity of 316L SS and Ta stents of identical design in baboon and porcine models. The platelet and brin deposition on these stents were determined and it was concluded that there was no signicant difference in thrombogenicity between 316L SS and Ta stents [99]. In another study, the least electropositive metals like copper induced more neointimal hyperplasia than the most electropositive metals like gold and platinum [97]. This illustrates that thrombogenicity and biocompatibility depend on multiple factors and that surface charge is just one such parameter. 3.4. Stability of surface oxide layer The stability of surface oxide layer directly inuences the biocompatibility of a material as the surface layer acts as a barrier to the release of ions from the bulk materials underneath the surface. For example, the percentage of nickel in NiTi and SS is 50% and 12%, respectively [16,100]. Atomic adsorption spectrophotometry analyses revealed a signicant release of nickel and chromium metal ions from non-coated SS stents over 96 h in human plasma [101]. The release of nickel ions from NiTi has been reported in few cases [45,46]. The released metal ions induced platelet and leukocyte activation [101,102], which resulted in thrombogenicity of the stents. Also, the endothelial cell damage caused by the release of very low concentration metal ions may be considered as a potentially toxic effect [102]. The stability of oxide layer is also key to several of the surface characteristics discussed above. It inuences surface energy by providing hydrophilicity to a material surface and surface potential by preventing the release of electrons. Since the stability of the natural surface oxide layer in 316L SS and NiTi is not very high, the possibility of metal ions being released is greater. Coating of stents is the most common approach to prevent this effect. 4. Rationale for coatings Since the basic mechanisms underlying the interaction between a metal and tissue/blood are still not completely understood, the biocompatibility and the hemocompatibility of metallic stents still remains an issue. Thrombosis and neointimal hyperplasia were commonly reported among bare metallic stents [103107]. Coating the metallic surface with other materials to alter its surface character-

istics without interfering with the bulk properties of the metal stent has been one rational approach to address this issue [108,109]. The coating of a stent can improve the surface properties signicantly; surface energy can be reduced, surface texture can be smoothened, surface potential can be neutralized and the stability of the surface oxide layer can be enhanced. These modications could directly inuence thrombosis and neointimal proliferation, which both can reduce restenosis. 4.1. Types of coatings Galvanization [97], sputtering followed by ion bombardment [97], pulsed biased arc ion plating [110], dipping [111,112], spraying [113,114], and plasma-based depositions [115,116] are some of the techniques that have been commonly used for coating stents. Initially, the coatings were used to increase the biocompatibility of stents, but later this technique became a platform for the controlled delivery of drug to inhibit intimal hyperplasia. The coating materials for stents can be broadly classied into four types: inorganic materials, polymers, porous metals, and endothelial cells. Inorganic materials and endothelial cells are exclusively used as coating materials, while the polymers are used as both as a coating material as well as the sole stent material. Porous metal coatings can be of the same metal as the stent or an alternative metal. 4.1.1. Inorganic coatings There are many inorganic coating materials which are potentially suitable for the treatment of medical implant surfaces. Gold, silicon-carbide, iridium oxide, and diamond-like carbon are some of the commonly used inorganic-coating materials on stents. 4.1.1.1. Gold. At one juncture gold coating was a preferred coating on SS stents to enhance uoroscopic visibility for reduced strut thicknesses of 5080 mm [117]. Since gold has six times the radiopacity of steel, a 5 mm coating on each side doubles the radiopacity of an 80 mmthick steel stent [117]. Edelman et al. [118] investigated the vascular response in porcine coronary arteries by comparing the standard gold coating with thermally processed gold coating. The reduction in neointimal hyperplasia and inammation for the thermally processed coating over standard coating was mainly attributed to smoothened gold surface and removal of embedded impurities in the gold coating. This study indicated that surface properties and material purity may play a signicant role in tissuematerial interactions. However, the human clinical trials on gold-coated stents were not satisfactory. Dahl et al. [119] reported that the neointimal proliferation was more in patients who received gold-coated stents. Danzi et al. [120] reported that the morphology of the restenosis was proliferative in 83% of the cases and the remaining 17% were occluded.

ARTICLE IN PRESS
1696 G. Mani et al. / Biomaterials 28 (2007) 16891710

4.1.1.2. Iridium oxide. Iridium oxide, generally accepted as a highly biocompatible inert ceramic material has been used for stent coatings [121,122]. It was found that hydrogen peroxide is produced at a metal (cobalt, zinc, nickel, copper, silver, chromium, and some of their alloys) surface when it is corroded [123]. Hydrogen peroxide, a strong oxidizing agent, can be harmful to the artery and can cause inammatory reactions. It is claimed that a metal coated with iridium oxide promotes the immediate conversion of hydrogen peroxide into water and oxygen, so it was expected that it might reduce inammatory reactions and promote endothelialization [124]. Initial studies in the porcine model showed that this coating can reduce the neointimal thickness from 118 mm on a bare SS stent to 55 mm on an iridium-coated stent [121]. The Lunar stent (Inow Dynamics, Germany) is a 316L SS with a thin inner layer of gold for increasing visibility and an outer layer of iridium oxide for improving biocompatibility [124]. A clinical trial evaluated the immediate and long-term outcome of these stents and reported that the overall angiographic restenosis rate was 13.8% [124]. It was reported that the iridium oxide promoted fast endothelialization because of its capability to prevent the production of free oxygen radicals, which can affect the adhesion and proliferation of endothelial cells [124]. The surface of SS produced little hydrogen peroxide unlike its alloying metals (nickel and chromium) [123]. Hence, a detailed investigation is needed to evaluate the amount of H2O2 actually produced on a 316L SS surface and the role of iridium oxide in the conversion of H2O2 into H2O and O2. 4.1.1.3. Silicon-carbide (SiC). Amorphous hydrogenated SiC, a semiconductor, has been well known for its antithrombogenic properties [125]. The reduction in the deposition of platelets, leukocytes, and monocytes over a stent made of SiC offers a promising coating material for reducing restenosis [126,127]. Although in-vitro studies [128] provided encouraging results, the outcomes from the various human trials produced contradictory results. For instance, the presence of endothelialization was reported in a 6-month clinical follow-up of the Tenax coronary stent [129]. In contrast, results showing greater neointimal hyperplasia were observed in a 6-month clinical follow-up for SiC-coated stents in another study [130]. A clinical trial compared the SiC-coated stents (Biotronik, Germany) with 316L NIR stents (Boston Scientic, USA) and concluded that both stents had a low rate of major adverse coronary events at 81712 weeks of follow-up, with no denite superiority [131]. These mixed results indicate the need for further research in this area. Bickel et al. [132] compared the platelet adhesion on different coating materials and found that SiC coating has a signicant effect in reducing the number of adhesion platelets than the uncoated stents but the effect is inferior to heparin and/or carbon coating. 4.1.1.4. Carbon. Coating of stents with diamond-like carbon, a chemically inert hydrocarbon, have shown

improved biocompatibility [133,134]. The results of a 6month follow-up of Carbostent (Sorin Biomedica, Italy) showed that the carbon coating can signicantly reduce stent thrombosis and restenosis in relatively high-risk patients [135]. The observed angiographic restenosis rate was 11% with no indication of subacute thrombosis. However, the results were not consistent with other studies. More recently, carbon coating has been considered inactive because they showed no improvements in angiographic restenosis [136]. Another clinical study compared uncoated MAC (AMG Raesfeld-Erle, Germany) stents with carbon-coated MAC stents [137] and showed that the carbon coating did not inuence the inammatory response. Other studies showed similar rates of binary restenosis, 31.8% for carbon-coated stents and 35.9% for bare metallic stents [138]. Thus carbon-coating stents have not yielded a signicant clinical improvement. A variety of these coating materials have been claimed to have improved biocompatibility, hemocompatibility, and antithrombogenicity properties for stents. However, these claims are usually not based on thorough comparative experiments and more denitive work remains to be done. 4.1.2. Endothelial cells Endothelial cell damage and exposure of subendothelial matrix at the site of arterial injury is a basis for both thrombus and neointima formation [7,9]. This clearly illustrates the importance of re-endothelialization, an approach in which endothelial cells are placed on stents before being implanted, and the cells are expected to proliferate, differentiate and release growth factors, which in turn inhibit thrombosis and neointimal hyperplasia [139]. Van der Giessen et al. [140] were the rst ones to seed endothelial cells on stents and study their in-vitro behavior. Several attempts have been made to seed endothelial cells on stents, but most of them have been limited by the quick loss of seeded cells, endothelial cell damage upon stent expansion, and the inability to maintain cell adherence to the artery wall during the blood ow [9,139]. 4.1.3. Porous materials One method that has been attempted to promote rapid endothelialization is to create micropores on the walls of vascular grafts [141,142]. Nakayama et al. [143] implemented this technique by covering stents with a segmented polyurethane lm with micropores of diameter 30 mm prepared by laser ablation technique. Increased pore density resulted in better endothelialization and a thinner neointimal layer [143]. However, this technique was limited by two factors [144]: (i) the non-at luminal surface design leads to thrombus formation, and (ii) when the edges of the polyurethane were overlapped by gluing, the pore density at these spots becomes zero which results in high neointimal hyperplasia. Later this technique was modied by dip coating the stent in polyurethane twice to have a at luminal surface and a microporous outer surface [145]. Also, heparin and tacrolimus were immobilized on the

ARTICLE IN PRESS
G. Mani et al. / Biomaterials 28 (2007) 16891710 1697

luminal and outer surface of the PU matrix, respectively by using photoreactive gelatin followed by UV irradiation [145]. However, after these techniques were reported for immobilizing the drugs, there have been no published reports available on drug release proles from these systems. Wieneke et al. [146] created a nanoporous aluminum oxide coating on a SS stent for loading tacrolimus. Though the surface modication by ceramic coating did not show signicant effect in reducing the neointima in a rabbit model, the release of tacrolimus reduced neointimal thickness in a dose dependent (52% and 56% reduction for 60 and 120 mg, respectively) manner [146]. Recently, it was reported that this kind of ceramic coating may liberate particle debris which in turn affects the antiproliferative effect of tacrolimus and results in a signicant increase in the neointima as compared to the control uncoated stents [147]. These studies further stress the importance of coating integrity. The coating is applied on a stent surface in its crimped state; i.e. the surface area of the stent is minimal. When the stent is expanded, the surface area increases and can result in ssures, cracks and pores on the coating. The damaged coating can also release particulate debris. Such changes in surface morphology and localized particle delivery can increase the chances of restenosis. Also it was reported that 80% of loaded drug was released from nanoporous coatings within 100 h and the remaining 20% was not released [147]. This type of release prole is indicative of a burst effect. Since it is preferable to release the drug over at least a 30-day period, open porous coating may either require smaller pores to trap the drugs longer or a second coating, which retards quick drug release. 5. Polymers Polymers used for coating stents can be broadly classied into biostable (non-biodegrable) polymers, biodegradable polymers, copolymers, and biological polymers. Several polymers with previous medical or dental applications have been used for coating stents or for making the entire stent. Although a wide range of polymers have been used to coat the stent, only a few, like polyethylene terepthalate (PET), poly-L-lactic acid (PLLA), and poly-L-glycolic acid (PLGA) have been tested as a lone stent material. 5.1. Biostable polymers The principle of biostable polymeric stents is very similar to metallic stents: the stent should have sufcient mechanical properties for providing stable support in maintaining the lumen gain [148]. Besides that, it should be biocompatible, and should not initiate thrombus formation and inammatory reactions. The elastic modulus of biomedical polymers usually lies in the range of 15 GPa [149], which raises concerns whether they posses sufcient mechanical properties for use as stents. However, Van der Giessen

et al. [150] have shown that the radial pressure exerted by PET braided mesh stents is the same as that of SS stents. PET has been investigated for making stents because of its good mechanical properties [149] and its reputation as a successful material for cardiovascular grafts [149,151]. Murphy et al. [152] deployed PET stents in the coronary arteries of porcine model. Though this study demonstrated the possibility of percutaneous deployment of polymeric stents in coronary arteries, the use of PET was associated with a chronic foreign body inammatory reaction and an intense proliferative neointimal response that resulted in the complete occlusion of the vessel. In another study in a porcine model, the PET-stented vessels were endothelialized and the neointimal thickening was 44113 mm 4 weeks post implantation [153]. Though it was stated that the extent of neointimal proliferation was limited when compared to the responses induced by bare metallic stents, the foreign body reaction was signicant and highlighted the inammatory reactions elicited by PET. Lack of radiopacity is also a concern for PET stents [154]. 5.2. Biodegradable polymers Due to the above-mentioned limitations of bare metallic stents and biostable polymer stents, biodegradable stents have been considered as an option. Biodegradable stents have the theoretical advantage of no longer being present as a foreign material in arteries once they have scaffolded the vessel for a relevant period of time [148,155]. The other signicant advantage is that drugs can be released in a controlled manner [148,155]. Stack et al. [156] developed the rst biodegradable stent made of PLLA and deployed it in canine model. The initial results showed the occurrence of limited thrombosis and minimal neointimal proliferation in the short term and also at 18 months. The mechanical behavior of these stents was also investigated [157]. However, the implantation of biodegradable polymers in a porcine model showed extensive inammation and neointimal proliferation [158]. To investigate the issues in detail, Van der Giessen et al. [159] implanted 3 different biodegradable polymers (polyorthoester, polycaprolactone, and polyethylene oxide/polybutylene terepthalate (PEO/PBTP)) and 3 biostable polymers (polyurethane, silicone, and PET) in porcine arteries. After 30 days of implantation, histological examination strongly conrmed the presence of neointimal thickening at the polymercoated side of each stent [159]. Extensive bro-muscular proliferation, multinucleated giant cell formation, mononuclear and eosinophilic smooth muscle cell proliferation were seen adjacent to PEO/PBTP and PET samples [159]. Lincoff et al. [160] showed that high molecular weight PLLA was well tolerated in a porcine model while the low molecular weight PLLA was not. This study showed that molecular weight of polymer also has an impact on neointimal hyperplasia. In spite of the controversies of using biodegradable stents, PLLA stents were implanted in a small clinical trial and the results were encouraging [161].

ARTICLE IN PRESS
1698 G. Mani et al. / Biomaterials 28 (2007) 16891710

Though it has been speculated that biodegradable polymers induce inammatory reactions because of an immune response to degradation products and non-reacted monomer compounds [158], the basic mechanism is still not completely understood. 5.3. Copolymers Several copolymers which include polyhydroxy butyrate/valerate [159], PEO/PBTP [159], methyl methacrylate/2-hydroxy ethyl methacrylate [162], ethylene-vinyl acetate [163], laurylmethacrylate/methacryloylphosphorylcholine [164], PLGA [159,165] and polyurethanes (PU) have been evaluated either as a coating material or as a base stent material. However, PLGA and PU are the most investigated copolymers for coronary stents. PLGA, the copolymer of polylactic acid and polyglycolic acid, has been widely used in bioresorbable sutures, drug delivery devices and orthopedic implants [149]. The degradation behavior of PLGA is crucial for its use in controlled drug delivery stent systems. For example, it has been observed that the rate of heparin release was slowest for PLLA, followed by PLGA (80/20) and nally PLGA (53/47) for coronary stent applications [166]. Thus the rate of drug release from PLGA stents would likely depend on the copolymer ratio. Recently, Venkatraman et al. [167] imparted self-expanding capability to biodegradable PLLA stents (at 37 1C) by adding PLGA (53/47). This effect was induced in the bilayered stents by fabricating them at 37 1C. Interestingly, PLGA stents have been investigated more in urological applications than coronary applications [168,169]. PU have been extensively used for medical device applications because of their excellent biocompatibility [83]. These polymers have been used as a coating material on stents to improve the antithrombogenic properties of Ta [85], corrosion resistance of NiTi [51], and biocompatibility of SS [170]. It was also reported that PU coatings can improve endothelialization [171]. PU has also been investigated in drug delivery systems [172174]. Lambert et al. [172] successfully studied the drug release kinetics and distribution of the model drug Forskolin delivered from the PU coated metallic stents. However, some studies have shown that PU coating can be accompanied by extensive inammatory reactions [175]. Thus while some studies advocated the use of PU for stents, others show contradictory evidence. Hence it is difcult to categorize PU as a good/poor stent material. Although PU has successfully been used in many cardiovascular devices (pacemaker lead wires, vascular grafts, articial heart pumps, and inner surface coatings of articial heart [149,176,177]), it does not necessarily mean that the coating may be benecial for stents. 5.4. Biological polymers Natural polymers are derived from natural resources and can be broadly classied into those of plant and animal

origin. Phosphorylcholine (PC), hyaluronic acid (HA), and brin are some of the biological polymers that were extensively explored for coating stents. 5.4.1. PC Phospholipid PC, an essential part of the red blood cell membrane, is structurally composed of both hydrophilic and hydrophobic components. This has been coated on metallic stents mainly to prevent the adhesion of coagulation-inducing cells [178]. An initial study of PC as a stentcoating material in porcine models showed its excellent bio- and hemo-compatibility [179]. Thereafter, extensive literature is available on the hemocompatibility and tissue compatibility of PC-coated stents [179,180]. The BiodivYsio stent, coated with PC, was evaluated in a human clinical study and the results showed that the restenosis decreased from 8977% to 5.676% [181]. Many other human clinical studies conrmed the antithrombogenic properties and decreased restenosis rates of PC-coated stents [182,183]. Also, the PC coating is stable up to 6 months of implantation [184186]. These characteristics together with its ability to deliver drugs make this material an attractive choice of coating for DES [187]. A human clinical trial studied the PC coating-based elution of an antiproliferative agent, ABT-578, using Endeavor stents in humans [188]. The restenosis rate of endeavor stents (13.3%) was almost three times less than the bare cobalt alloy driver stent (34.2%) [188]. A non-randomized trial investigated the PC coating-based elution of dexamethasone and reported the binary restenosis rate of 13.3% for 60 patients at 6 months [189]. Recently, a human clinical trial successfully evaluated the release of angiopeptin from a PC-coated stent and showed encouraging results [190]. 5.4.2. HA HA, a linear polysaccharide non-sulfated glycosaminoglycan present in various tissues of the body, has been found to improve the thrombo-resistance of stents. Verheye et al. [191] reported a signicant reduction of platelet deposition on HA-coated SS stents in a baboon model. In order to extend the antiproliferative and antithrombogenic properties of biodegradable HA, it can be made insoluble by self-cross-linking with N-(3-dimethylaminopropyl)-N0 -ethyl carbodiimide [192]. Reduced inammatory responses were found for periods up to a month when compared with uncoated SS stents in undiseased pig coronary arteries [192]. In another study, HA was covalently attached to SS [193]. Epoxy silane was covalently attached to SS and then the epoxy group was converted to aldehyde group to react it with HA. The approaches which involve chemical modication of the coating polymer need to be thoroughly characterized before implantation. Even traces of the chemical used for chemical modication can be non-biocompatible and eventually leads to erroneous conclusions about the coating material. Though the available literature on HA

ARTICLE IN PRESS
G. Mani et al. / Biomaterials 28 (2007) 16891710 1699

coatings is meager, coating the stents with these biopolymers seems promising. 5.4.3. Fibrin Fibrin is an insoluble protein produced during the coagulation of blood. This biopolymer is well known for its biocompatible, biodegradable, and viscoelastic properties [194,195]. Holmes et al. [196] demonstrated the potential role of exogenous brin as a better coating material in a porcine model. They compared a circumferential brin sleeve-coated coil wire stent with a PET stent and a PU-coated stent. Markedly less vessel occlusion and foreign body reaction was observed with exogenous brin than with PU and PET stents. Another major advantage of the brin-lm stent is that it provides complete endoluminal paving by covering 100% of the arterial surface, compared with the partial coverage achieved with bare metallic stents [197]. This strategy particularly facilitates site-specic therapies, such as delivering the drugs to the entire lesion surface [197]. It may also lead to rapid endothelization [198]. The application of all these biological polymers as stent coatings appears very promising. Since such polymers facilitate re-endothelialization and show negligible inammatory reactions, human clinical trials for stents with coatings of biological polymers is the logical next step. 6. Rationale for DES Endothelial and smooth muscle cell damage, unavoidable in PTCA and stent placement, is a cause of restenosis [9]. The optimization of the architecture and mechanical characteristics of stents has lead to a decrease in restenosis but using drug delivery platforms remains a promising way to further reduce restenosis. The main reason for the failure of systemic pharmacological therapy is the inability to deliver an adequate drug dose at the site of injury [199]. Earlier approaches for local drug delivery by using catheter-mounted balloons and needles were not successful due to rapid washout of the drugs by the blood stream [199,200]. Currently, the treatment available for preventing restenosis is the implantation of DES. Heparin was the rst therapeutic agent attached directly to a stent. The concept of delivering medications at the injury site has evolved from heparin-coated stents to stents with drugs that inhibit neointimal hyperplasia. For preventing neointimal hyperplasia, an appropriate drug concentration has to be delivered for at least 30 days during which the biology of restenosis is known to occur. 6.1. Techniques for drug-loading and release kinetics The techniques for loading drugs on a stent can be categorized into three major types: (i) attaching the drug directly onto the metal surface; (ii) loading the drug into the pores of porous metal stents; and (iii) incorporating the drug in a polymer that is then used as a stent-coating

material. The drug release depends on the way the drug is loaded on the stent. If the drug is physically adsorbed on the metal surface or in the porous surface, it can be released by simple diffusion. Here, the porous surface offers the possibility of incorporating more drugs than the metal surface because of the greater surface area. The amount of drug release can also be controlled by the size and density of the pores. If the drugs are trapped inside nonbiodegradable polymers (techniques used in CYPHERTM and TAXUSTM Express2TM), they are released by diffusion. In this case, the amount of drug released depends on the thickness of the outer coating, as it modulates the amount of drug that can be released per unit time. When the drugs are chemically attached to the surface, the drug release depends on the rate at which the chemical bonds are cleaved. The rate of chemical bond cleavage depends on the orientation of drug molecules, which determines the triggers access to the bond. Drug delivery through biodegradation is the most common phenomenon and it has been extensively reviewed in the literature for orthopedic [201,202], ocular [203,204], neuro [205,206], and cardiovascular applications [155,207]. The same concept applies in case of DES as the drug-incorporated matrix is coated on the metal surface and the rate of drug release depends on the rate at which the matrix is degraded. 6.2. DES The three drugs that have been investigated in depth for treating restenosis are heparin, sirolimus, and paclitaxel (Fig. 1). Heparin has been effective in reducing both thrombosis and neointimal proliferation while sirolimus and paclitaxel were mainly used for their anti-proliferative effects in blocking neointimal hyperplasia. 6.2.1. Heparin-coated stents Heparin, a heterogeneous group of unbranched, acidic glycosaminoglycans, has been widely used for modifying the surfaces of vascular implants because of its anticoagulant properties [208]. Heparin activity depends on the interaction between its active sites (carbohydrate sequences) and the circulating antithrombin III. The antithrombin which binds to the active sites catalyzes the inhibition of thrombin and the resultant inactive antithrombin/thrombin complex is released into the blood stream [209]. There are various ways in which heparin can bind to a stent surface and these include physical adsorption, ionic bonding, copolymerization, and polymer encapsulation. Physical adsorption has been attained by coating the stent with a solution of water-insoluble benzalkonium chloride complex [210]. For ionic bonding, the material surface was cationically charged through quarternization (treatment with tridodecylmethylammonium chloride ammonium salt or ethyl bromide) treatment and then the anionic heparin molecules are ionically bonded on to the cationic surfaces [211,212]. The stability of both physically adsorbed and

ARTICLE IN PRESS
1700 G. Mani et al. / Biomaterials 28 (2007) 16891710

Fig. 1. Chemical structure of (A) heparin, (B) sirolimus and (C) paclitaxel.

ionically bound heparin is low as they are easily removed from the surface when exposed to plasma [213]. The stability issue raised a question about the long-term anticoagulation therapy. To improve the stability, heparin was copolymerized with a variety of polymers like poly(methylmethacrylate) [214], poly(vinyl alcohol) [215], and PU [216]. Though the copolymerization techniques provided a more stable binding of heparin to the surfaces compared to physical adsorption and ionic binding, these techniques tend to alter the chemical sequences of heparin, which are essential for its therapeutic effect. Larm et al. [213] created aldehyde groups in heparin through its reaction with nitrous acid. Then, the aldehyde groups were covalently bonded to the amine terminated stent surface. This endpoint attachment technique has the unique advantage of having secured the active carbohydrate sequences, which are essential for binding antithrombin. Heparin delivery has been achieved using biodegradable PLGA microspheres [217]. In order to control the heparin delivery from the biodegradable polymers (PLLA, PLLGA, PLGA), polyethylene glycol, which is a plasticizer, was added to the polymerheparin lms [166]. The effect of plasticizers on the release proles of heparin was found to be dependent on the copolymer ratios of PLA and PGA. Though several techniques have been tried to immobilize and/or deliver heparin at the target site, each technique has its own advantages and disadvantages and none has proved to be optimum. Bonan et al. [218] were the rst to use heparin coated zigzag stents in canine coronary arteries. Several other animal studies [219,220] and clinical trials [221225] investigated the efcacy of heparin-coated stents and

strongly conrmed the absence of thrombosis. A reduction in neointimal formation was also reported in few animal studies [163,226]. 6.2.2. Sirolimus-eluting stents (SES) Sirolimus, an immunosuppressive agent, binds to an intracellular receptor protein and ultimately induces cellcycle arrest [227]. It inhibits vascular smooth cell migration, proliferation and growth [227,228]. A variety of biodegradable and non-biodegradable polymers have been used as coatings for SES [229]. The FDA approved sirolimus-coated BXTM Velocity balloon expandable stent (CYPHERTM) [230] is made from electropolished laser-cut 316L SS. The stent is coated with a three layers of polymer coating: parylene C, an inert, hydrophobic and biocompatible polymer is initially coated on the metallic stent. Then, a mixture of polyethylene-covinyl acetate (PEVA) and poly n-butyl methacrylate (PBMA) in a ratio of 67:33 is mixed with sirolimus and coated on the parylene C coating. Finally, a mixture of PEVA and PBMA is then applied as the third layer without sirolimus. The main purpose of this nal coating is to prevent the fast leaching of drugs from the polymer coating during the initial period post-implantation. Recently Chen et al. [113] spray-coated collagen and sirolimus layer-bylayer alternatively on to a SS stent. The collagen matrices were used for releasing the sirolimus in a controlled fashion without having any burst effect. As the collagen is already known for its better blood compatibility, this kind of biocompatible coatings without the use of polymers seems promising. Several clinical studies have addressed the usefulness of SES [231235].

ARTICLE IN PRESS
G. Mani et al. / Biomaterials 28 (2007) 16891710 1701

6.2.3. Paclitaxel-eluting stents (PES) Paclitaxel is a drug used in the treatment of cancer. This drug binds to the tubulin protein of microtubules, which are the components of cells that provide structural framework and enable cells to divide and grow. The abnormality, paclitaxel/microtubule complex, in vascular smooth muscle cells inhibits cellular replication and ultimately causes cellular death [236]. The coating of paclitaxel on stents can be broadly classied into two types: polymer-based and non-polymerbased coatings. Heldman et al. [111] coated paclitaxel directly on the stent surface by dipping the stent in the ethanolic solution of paclitaxel followed by evaporating the alcohol. This technique is particularly advantageous in that there are no concerns over inammatory reactions induced by the polymer-based drug delivery systems. Also, the tissue can be in direct touch with the drug coating. However, the disadvantage of dip coating is that a signicant amount of drug is lost during the stent placement and expansion similar to the problems that were encountered during catheter-based drug elution attempts [111]. This shows the signicance of having a carrier, which can hold the drugs and release them in a controlled fashion. In one study, three different doses (0.2, 15, and 187 mg/stent) of paclitaxel were coated directly on to the stent [111]. The higher dose showed signicant reduction in neointimal hyperplasia compared to the lower doses in a pig model after 28 days. The amount of drug that is loaded on the stent is a very crucial factor. In the case of dip coating, the amount of drug that can be coated depends on the surface area of the stent. The surface area is determined by the basic design of the stent and thickness of the struts. A low prole design with thinner struts is usually preferred for successful implantation. However this signicantly limits the amount of drug that can be coated on the stent. A multicenter study evaluated the ability of Supra-G (Cook Inc), a 316L SS stent coated with a polymer-free formulation of paclitaxel, to inhibit restenosis [237,238]. This study showed that a paclitaxel-coated stent could signicantly reduce restenosis at 6 months after intervention. The dose of 3.1 mg/mm2 was more effective than the dose of 1.3 mg/mm2. Intravascular-ultrasound evaluation demonstrated a dose-dependent reduction in the volume of intimal hyperplasia. In another clinical study, researchers evaluated the V-Flex Plus (Cook Inc), a 316L SS stent, coated with increasing doses of paclitaxel (0.2, 0.7, 1.4, and 2.7 mg/mm2 stent surface area) [238]. The drug was applied directly to the albuminal surface of the stent with polymer-free formulation. The authors concluded that the in-stent restenosis was signicantly reduced for a paclitaxel dose density of 2.7 mg/mm2 without short- or medium-term side effects. These studies clearly demonstrate the importance of the amount of drug loading in preventing restenosis. Polymer coatings, in general, can carry higher loads of drug compared to direct drug adsorption on the metal surface. A vascular compatible poly(styrene-b-isobutylene-b-styrene) triblock copolymer

(SIBS) is used as a paclitaxel carrier in the FDA approved TAXUSTM Express2TM paclitaxel-eluting coronary stent [239]. Unlike CYPHERTM, this stent does not have additional outer coating (polymer coating without drug) to prevent the burst effect. This may be one of the reasons why the drug elution proles for TAXUSTM stents are for 30-day periods while compared to the 60-day periods for CYPHERTM stents. Ranade et al. [240] found that the paclitaxel solubility in SIBS matrix is extremely low and it exists as nanoparticles in the polymer matrix. There was no observable change in the surface morphology of the polymer matrix after the incorporation of paclitaxel. AFM images conrmed the morphology changes only during the drug elution period and supported a burst effect of greater than 8.8% of the total amount of the loaded drug. This conrmed that the paclitaxel release is directly dependent on the paclitaxel loading. In order to increase the miscibility of paclitaxel in SIBS, Sipos et al. [241] chemically modied the styrenic portion of the SIBS polymer system and thereby modulating the drug release prole. It was reported that the modulation is due to the improved hydrophilicity and polarity of the polymer systems [241]. The efcacy of NIRxTM (Boston Scientic) PES were evaluated by several clinical trials [242244]. A recent meta-analysis of 6 randomized trials comparing PES with SES was performed [245]. Superiority of SES was observed with angiographic restenosis rates of 9.3% when compared to 13.1% for PES. The difference between the two DES is multifactorial, and may be related to the underlying stent design, polymeric coating, mechanism of drug action, drug-release kinetics, and drug distribution across the vessel wall [245]. The hydrophobic/hydrophilic nature of drug molecules is also a vital parameter in this application [246]. The stent surface is exposed to blood from the time it is mounted on a balloon catheter till it is expanded at the tissue in the target area. If the drug is hydrophilic, it can possibly be lost in the blood due to its high solubility [246]. The loss of hydrophobic paclitaxel is less than 5% for a 30 s exposure time because of its low solubility in blood [111]. Though the material aspects contribute the most to the success of DES in reducing the restenosis percentage, the physicochemical characteristics of drugs have their own contribution [246248]. Though the sirolimus and paclitaxel eluting stents have shown promising short-to-medium term clinical results, recently published reports on the occurrence of LST (occurs after 30 days) with adverse clinical events have raised concerns. A clinical trial that compared the efcacy of BMS and DES reported the occurrence of early stent thrombosis (occurs within 30 days) as 11.5% for both stent categories and no difference was observed in the occurrence of early stent thrombosis [249]. The occurrence of late thrombosis in patients treated with BMS and DES vary from 0.65% to 0.76% [250,251] and 0.35% to 0.7% [10,11], respectively. Though the reason(s) for the occurrence of LST in DES is still unknown, the factors that could contribute may be the following: (a) delayed

ARTICLE IN PRESS
1702 G. Mani et al. / Biomaterials 28 (2007) 16891710

endothelialization [10,252,253]the reasons for the delay in vessel wall healing after the implantation of DES are not yet clearly known. However, there is a concern that the nature of therapeutics used and their concentration/ distribution across the vessel wall may affect the healing [252,253]; (b) adverse effects of the polymer coatings [254]polymer coatings like parylene, PEVA and PBMA have been traditionally used for coating blood-contacting devices for their quality of adherence to the metal surfaces and/or their hydrophobicity. It is puzzling that such coating materials did not provide optimal results. This exposes the lack of knowledge about diseased tissuebiomaterial interactions; (c) discontinuation of antiplatelet therapyLST was observed among the patients treated with DES when they stop taking antiplatelet medications [252]; (d) neointimal growth for a longer periodthe growth of neointima reaches the peak at 6 months for BMS and regresses after that [255]. On the contrary, the neointima grows up to 4 years in DES [255]; (f) increased length of DES [256258]long DES were implanted to treat the entire diseased portion of the artery. This was reported to cause problems during deployment and positioning of stents in the arteries, which may eventually result in abnormal shear stress and cause thrombosis [259]. Also, the longer stents increase the area of polymer coatingtissue interactions. These studies clearly show the need for further research in this area and that the currently available DES are far from optimal. 7. Conclusion From a review of the literature it is evident that the material used for making stents has to have appropriate mechanical properties, suitable surface characteristics, excellent haemocompatibility, good biocompatibility, and drug delivery capacity. Every material has its own pros and cons. Table 3 provides a list of materials which posses the ideal for a specic material property (Table 3). It may not
Table 3 Materials with ideal characteristics for coronary stent applications Properties Elongation modulus Tensile strength Yield strength Surface energy Biocompatibility Surface potential Surface texture Stability of surface oxide layer Therapeutics Radiopacity MRI compatibility Preferred way of drug loading Preferred way of drug elution Preferred category of polymers Materials 316L stainless steel CoCr CoCr PTFE Ti Ta Electropolishing Ta/Ti Paclitaxel Gold Ta/Ti/Nitinol Polymer based Biodegradable Biopolymers

be possible for a single material to posses all the desired requirements. So, the success lies in choosing the optimum combination of materials and properties for the coronary stent applications. Though DES emerged recently, they appear to be the future of coronary stents. Ever since the FDA approved DES, the commercial availability of these stents has increased rapidly. However, it will take several years for this approach to become optimized once the long-term outcomes of the clinical trials are reported. The occurrence of late stent thrombosis in the patients treated with DES has raised concerns about these stents. Additionally, several cases have been reported recently on hypersensitivity reactions to DES [254,260262]. In a pathological study of stent-related hypersensitivity reactions, it was noted that the polymer-coated stents released polymer fragments which were surrounded by giant cells and eosinophils [254]. Stents were also found to induce inammatory reactions predominantly consisted of T lymphocytes and eosinophils with extensive inammation of the arterial wall [254,260,262]. The FDA has posted a cautionary view about the adverse and hypersensitive reactions following deployment of sirolimus-eluting CYPHER stents [261,263,264]. In conclusion, in its present form, percutaneous transluminal coronary angioplasty cannot be performed without damaging blood vessels and eliciting restenosis. Drug elution at the target site is a clear solution to this problem. However, the present methods for drug elution are still plagued with problems. Most commercially available DES use polymer matrices for coating and releasing the drugs. Increasing evidence suggests that some adverse reactions may be caused by these polymers. Hence, research should be carried out in designing and developing new polymer materials and should include essential features like hemocompatibility, hydrophobicity, anti-inammatory, conformability to the stent surface, aking resistance, sterilizability, and biodegradability. Other approaches such

Rationale Optimal value for a balloon expandable stent Higher value Much lesser when compared to its own tensile strength Lower value Extensive literature Presence of stable oxide layer Stability of surface oxide layer Best polishing technique to-date Excellent stability among the implant materials Hydrophobicity High density No Fe content Amount of drug can be increased to the need just by increasing the thickness of the coating No polymer material will be present once the process is nished Minimal inammatory and hypersensitive reactions

ARTICLE IN PRESS
G. Mani et al. / Biomaterials 28 (2007) 16891710 1703

as nanoporous coatings and using self-assembled monolayers [265268] for drug delivery also have potential applications in the next generation of stents.

References
[1] Shepherd RFJ, Vlietstra RE. The history of balloon angioplasty. In: Vlietstra RE, Holmes DR, editors. Percutaneous transluminal coronary angioplasty. Philadelphia: F.A. Davis Company; 1987. p. 117. [2] Myler RK, Stertzer SH. Coronary and peripheral angioplasty: historical perspective. In: Topol EJ, editor. Textbook of interventional cardiology. 2nd ed. Philadelphia: W.B. Saunders Company; 1994. p. 17185. [3] Serruys PW, Jaegere PD, Kiemeneij F, Macaya C, Rutsch W, Heyndrickx G, et al. A comparison of balloon-expandable-stent implantation with balloon angioplasty in patients with coronary artery disease. New Engl J Med 1994;331(8):48995. [4] Fischman DL, Leon MB, Baim DS, Schatz RA, Savage MP, Penn I, et al. A randomized comparison of coronary-stent placement and balloon angioplasty in the treatment of coronary artery disease. New Engl J Med 1994;331(8):496501. [5] Holmes J. State of the art in coronary intervention. Am J Cardiol 2003;91(3A):50A3A. [6] Wolf MG, Moliterno D, Lincoff A, Topol E. Restenosis: an open le. Clin Cardiol 1996;19(5):34756. [7] Newby AC, Zaltsman AB. Molecular mechanisms in intimal hyperplasia. J Pathol 2000;190:3009. [8] Mak KH, Belli G, Ellis SG, Moliterno DJ. Subacute stent thrombosis: evolving issues and current concepts. JACC 1996;27(2): 494503. [9] Kipshidze N, Dangas G, Tsapenko M, Moses J, Leon MB, Kutryk M, et al. Role of the endothelium in modulating neointimal formation: vasculoprotective approaches to attenuate restenosis after percutaneous coronary interventions. J Am Coll Cardiol 2004;44(4):7339. [10] Ong ATL, McFadden EP, Regar E, deJaegere PPT, vanDomburg RT, Serruys PW. Late angiographic stent thrombosis (LAST) events with drug-eluting stents. J Am Coll Cardiol 2005;45(12):208892. [11] Iakovou I, Schmidt T, Bonizzoni E, Ge L, Sangiorgi G, Stankovic G, et al. Incidence, predictors, and outcome of thrombosis after successful implantation of drug-eluting stents. J Am Med Assoc 2005;293(17):212630. [12] Taylor A. Metals. In: Sigwart U, editor. Endoluminal stenting. London: W.B. Saunders Company Ltd; 1996. p. 2833. [13] Schatz RA. A view of vascular stents. Circulation 1989;79(2): 44557. [14] Wong SC, Schatz RA. Developmental background and design of the Palmaz-Schatz coronary stents. In: Herrmann Jr HC, Hirshfeld JW, editors. Clinical use of the Palmaz-Schatz intracoronary stent. New York: Futura Publishing Company, Inc.; 1993. p. 319. [15] Rundback JH, Leonardo R, Rozenbilt GN. Peripheral vascular stents. In: Dolmatch BL, Blum U, editors. Stent-grafts current clinical practice. New York: Thieme; 2000. p. 122. [16] Cardarelli F. Ferrous metals and their alloys. In: Materials handbook. London: Springer London Limited; 2000. p. 201. [17] Haudrechy P, Foussereau J, Mantout B, Baroux B. Nickel release from 304 and 316 stainless steels in synthetic sweat. Comparison with nickel and nickel-plated metals. Consequences on allergic contact dermatitis. Corros Sci 1993;35(14):32936. [18] Ko ster R, Sommerauer M, Ka hler J, Baldus S, Meinertz T, Hamm CW, et al. Nickel and molybdenum contact allergies in patients with coronary in-stent restenosis. Lancet 2000;356(9245):18957. [19] Teitelbaum G, Bradley W, Klein B. MR imaging artifacts, ferromagnetism, and magnetic torque of intravascular lters, stents, and coils. Radiology 1988;166(3):65764.

[20] Hijazi Z, Homoud M, Aronovitz M, Smith J, Faller G. A new platinum balloon-expandable stent (Angiostent) mounted on a high pressure balloon: acute and late results in an atherogenic swine model. J Invas Cardiol 1995;7(5):12734. [21] Bhargava B, Scheerder ID, Ping Q, Yanming H, Chan R, Kim HS, et al. A novel platinumiridium, potentially gamma radioactive stent: evaluation in a porcine model. Cathet Cardiovasc Interv 2000;51(3):3648. [22] Trost D, Zhang H, Prince M, Winchester P, Wang Y, Watts R, et al. Three-dimensional MR angiography in imaging platinum alloy stents. J Magn Reson Imag 2004;20(6):97580. [23] Park JB, Kim YK. Metallic biomaterials. In: Park JB, Bronzino JD, editors. Biomaterials principles and applications. Boca Raton: CRC Press; 2003. p. 120. [24] Park JB. Metallic implant materials. In: Biomater Sci Eng. New York: Plenum Press; 1987. p. 193233. [25] Foti R, Tamburino C, Galassi A, Russo G, Nicosia A, Grassi R, et al. Safety, feasibility and efcacy of a new single-wire stent in the treatment of complex coronary lesions: the angiostent. Cardiologia 1998;43(7):72530. [26] Zitter H, Plenk H. The electrochemical behavior of metallic implant materials as an indicator of their biocompatibility. J Biomed Mater Res 1987;21(7):88196. [27] Johnson P, Bernstein J, Hunter G, Dawson W, Hench L. An in vitro and in vivo analysis of anodized tantalum capacitive electrodes: corrosion response, physiology, and histology. J Biomed Mater Res 1977;11(5):63756. [28] Macionczyk F, Gerold B, Thull R. Repassivating tantalum/ tantalum oxide surface modication on stainless steel implants. Surf Coat Technol 2001;142144:10847. [29] Teitelbaum G, Raney M, Carvlin M, Matsumoto A, Barth K. Evaluation of ferromagnetism and magnetic resonance imaging artifacts of the Strecker tantalum vascular stent. Cardiovasc Intervent Radiol 1989;12(3):1257. [30] Matsumoto A, Teitelbaum G, Barth K, Carvlin M, Savin M, Strecker E. Tantalum vascular stents: in vivo evaluation with MR imaging. Radiology 1989;170:7535. [31] Matsuno H, Yokoyama A, Watari F, Uo M, Kawasaki T. Biocompatibility and osteogenesis of refractory metal implants, titanium, hafnium, niobium, tantalum and rhenium. Biomaterials 2001;22(11):125362. [32] Leng YX, Chen JY, Yang P, Sun H, Wang J, Huang N. The biocompatibility of the tantalum and tantalum oxide lms synthesized by pulse metal vacuum arc source deposition. Nucl Instrum Methods Phys Res Sec B: Beam Interact Mater Atoms 2006;242(12):302. [33] Chen JY, Leng YX, Tian XB, Wang LP, Huang N, Chu PK, et al. Antithrombogenic investigation of surface energy and optical bandgap and hemocompatibility mechanism of Ti(Ta+5)O2 thin lms. Biomaterials 2002;23(12):254552. [34] Chen JY, Leng YX, Zhang X, Yang P, Sun H, Wang J, et al. Effect of tantalum content of titanium oxide lm fabricated by magnetron sputtering on the behavior of cultured human umbilical vein endothelial cells. Nucl Instrum Methods Phys Res Sec B: Beam Interact Mater Atoms 2006;242(12):269. [35] Barth K, Virmani R, Froelich J, Takeda T, Lossef S, Newsome J, et al. Paired comparison of vascular wall reactions to Palmaz stents, Strecker tantalum stents, and Wallstents in canine iliac and femoral arteries. Circulation 1996;93(12):21619. [36] Ozaki Y, Keane D, Nobuyoshi M, Hamasaki N, Popma JJ, Serruys PW. Coronary lumen at six-month follow-up of a new radiopaque cordis tantalum stent using quantitative angiography and intracoronary ultrasound. Am J Cardiol 1995;76(16):1103212. [37] Davis JR. Metallic materials. In: Handbook of medical devices. Materials Park: ASM International; 2003. p. 2150. [38] Windecker S, Mayer I, Pasquale GD, Maier W, Dirsch O, Groot PD, et al. Stent coating with titanium-nitride-oxide for reduction of neointimal hyperplasia. Circulation 2001;104(8):92833.

ARTICLE IN PRESS
1704 G. Mani et al. / Biomaterials 28 (2007) 16891710 materials in medicine. 2nd ed. San Diego: Elsevier Academic Press; 2004. p. 13753. Briguori C, Sarais C, Pagnotta P, Liistro F, Montorfano M, Chieffo A, et al. In-stent restenosis in small coronary arteries: impact of strut thickness. J Am Coll Cardiol 2002;40(3):4039. Kastrati A, Mehilli J, Dirschinger J, Dotzer F, Schuhlen H, Neumann F-J, et al. Intercoronary stenting and angiographic results: strut thickness effect on restenosis outcome. Circulation 2001;103(23):281621. Rittersma S, Winter Rd, Koch K, Bax M, Schotborgh C, Mulder K, et al. Impact of strut thickness on late luminal loss after coronary artery stent placement. Am J Cardiol 2004;93(4):47780. Kereiakes D, Cox D, Hermiller J, Midei M, Bachinsky W, Nukta E, et al. Usefulness of a cobalt chromium coronary stent alloy. Am J Cardiol 2003;92(4):4636. Klocke A, Kemper J, Schulze D, Adam G, Kahl-Nieke B. Magnetic eld interactions of orthodontic wires during magnetic resonance imaging (MRI) at 1.5 Tesla. J Orofac Orthop 2005;66(4):27987. US Food and Drug Administration, Center for Devices and Radiological Health, MULTI-LINK VISIONTM RX & OTW coronary stent systemP020047. Available from /http:// www.fda.gov/cdrh/mda/docs/p020047.htmlS updated 16 September 2003. Peuster M, Wohlsein P, Bru gmann M, Ehlerding M, Seidler K, Fink C, et al. A novel approach to temporary stenting: degradable cardiovascular stents produced from corrodible metal-results 618 months after implantation into new zealand white rabbits. Heart 2001;86(5):5639. Heublein B, Rohde R, Kaese V, Niemeyer M, Hartung W, Haverich A. Biocorrosion of magnesium alloys: a new principle in cardiovascular implant technology? Heart 2003;89(6):6516. Peuster M, Hesse C, Schloo T, Fink C, Beerbaum P, Schnakenburg Cv. Long-term biocompatibility of a corrodible peripheral iron stent in the porcine descending aorta. Biomaterials 2006;27(28):495562. Mueller PP, May T, Perz A, Hauser H, Peuster M. Control of smooth muscle cell proliferation by ferrous iron. Biomaterials 2006;27(10):2193200. Staiger M, Pietak A, Huadmai J, Dias G. Magnesium and its alloys as orthopedic biomaterials: a review. Biomaterials 2006;27(9): 172834. Cardarelli F. Less common non-ferrous metals. In: Materials handbook. London: Springer London Limited; 2000. p. 99107. Busk RS. Magnesium and its alloys. In: Kutz M, editor. Handbook of materials selection. New York: Wiley; 2002. p. 25965. Eliezer D, Alves H. Corrosion and oxidation of magnesium alloys. In: Kutz M, editor. Handbook of materials selection. New York: Wiley; 2002. p. 26791. Mario CD, Grifths H, Goktekin O, Peeters N, Verbist J, Bosiers M, et al. Drug-eluting bioabsorbable magnesium stent. J Interv Cardiol 2004;17(6):3915. Eggebrecht H, Rodermann J, Hunold P, Schmermund A, Bose D, Haude M, et al. Images in cardiovascular medicine. Novel magnetic resonance-compatible coronary stent: the absorbable magnesiumalloy stent. Circulation 2005;112(18):e3034. Covino BS, Cramer SD. Corrosion: fundamentals, testing, and protection in ASM handbook, 10th ed. ASM International; 2003. Zartner P, Cesnjevar R, Singer H, Weyand M. First successful implantation of a biodegradable metal stent into the left pulmonary artery of a preterm baby. Catheter Cardiovasc Interv 2005;66(4):5956. Schranz D, Zartner P, Michel-Behnke I, Akinturk H. Bioabsorbable metal stents for percutaneous treatment of critical recoarctation of the aorta in a newborn. Catheter Cardiovasc Interv 2006;67(5): 6713. Palmaz J. New advances in endovascular technology. Tex Heart Inst J 1997;24(3):1569. Huang Y, Verbeken E, Schacht E, Scheerder ID. Local drug delivery using drug-eluting stents. In: Serruys PW, Leon MB, Colombo A, [39] Mosseri M, Tamari I, Plich M, Hasin Y, Brizines M, Frimerman A, et al. Short- and long-term outcomes of the titanium-NO stent registry. Cardiovasc Revasc Med 2005 2005;6(1):26. [40] Windecker S, Simon R, Lins M, Klauss V, Eberli F, Rof M, et al. Randomized comparison of a titanium-nitride-oxide-coated stent with a stainless steel stent for coronary revascularization: the TiNOX trial. Circulation 2005;111(20):261722. [41] Biehl V, Wack T, Winter S, Seyfert U, Breme J. Evaluation of the haemocompatibility of titanium based biomaterials. Biomol Eng 2002;19(26):97101. [42] Sumita M, Teoh SH. Durability of metallic implant materials. In: Teoh SH, editor. Engineering materials for biomedical applications. Singapore: World Scientic Publishing Co; 2004. p. 2-12-31. [43] Stoeckel D, Pelton A, Duerig T. Self-expanding nitinol stents: material and design considerations. Eur Radiol 2004;14(2):292301. [44] Trepanier C, Venugopalan R, Pelton AR. Corrosion resistance and biocompatibility of passivated NiTi. In: Yahia LH, editor. Shape memory implants. New York: Springer; 2000. p. 3545. [45] Heintz C, Riepe G, Birken L, Kaiser E, Chakfe N, Morlock M, et al. Corroded nitinol wires in explanted aortic endografts: an important mechanism of failure? J Endovasc Ther 2001;8(3):24853. [46] Berger-Gorbet M, Broxup B, Rivard C, Yahia L. Biocompatibility testing of NiTi screws using immunohistochemistry on sections containing metallic implants. J Biomed Mater Res 1996;32(2):2438. [47] Shabalovskaya S. On the nature of the biocompatibility and on medical applications of NiTi shape memory and superelastic alloys. Biomed Mater Eng 1996;6(4):26789. [48] Maitz M, Shevchenko N. Plasma-immersion ion-implanted nitinol surface with depressed nickel concentration for implants in blood. J Biomed Mater Res A 2006;76(2):35665. [49] OBrien B, Carroll W, Kelly M. Passivation of nitinol wire for vascular implantsa demonstration of the benets. Biomaterials 2002;23(8):173948. [50] Trepanier C, Tabrizian M, Yahia L, Bilodeau L, Piron D. Effect of modication of oxide layer on NiTi stent corrosion resistance. J Biomed Mater Res 1998;43(4):43340. [51] Mazumder M, De S, Trigwell S, Ali N, Mazumder M, Mehta J. Corrosion resistance of polyurethane-coated nitinol cardiovascular stents. J Biomater Sci Polym Ed 2003;14(12):135162. [52] Starosvetsky D, Gotman I. Corrosion behavior of titanium nitride coated NiTi shape memory surgical alloy. Biomaterials 2001;22(13):18539. [53] Shih C, Lin S, Chung K, Chen Y, Su Y. Increased corrosion resistance of stent materials by converting current surface lm of polycrystalline oxide into amorphous oxide. J Biomed Mater Res 2000;52(2):32332. [54] Schurmann K, Vorwerk D, Kulisch A, Stroehmer-Kulisch E, Biesterfeld S, Stopinski T, et al. Experimental arterial stent placement. Comparison of a new nitinol stent and wallstent. Invest Radiol 1995;30(7):41220. [55] Adams G, Baltazar U, Karmonik C, Bordelon C, Lin P, Bush R, et al. Comparison of 15 different stents in supercial femoral arteries by high resolution MRI ex vivo and in vivo. J Magn Reson Imag 2005;22(1):12535. [56] Nakamura S, Degawa T, Nishida T, Anzai H, Mitsuo K, Sakatani H, et al. Preliminary experience of Act-OneTM coronary stent implantation. J Am Coll Cardiol 1996;27(2, Suppl. 1):53. [57] Holmes Jr DR, Lansky A, Kuntz R, Bell MR, Buchbinder M, Fortuna R, et al. The PARAGON stent study: a randomized trial of a new martensitic nitinol stent versus the Palmaz-Schatz stent for treatment of complex native coronary arterial lesions. Am J Cardiol 2000;86(10):10739. [58] Isshiki T, Eto K, Ochini M, Milani H, Kondo K, Takoshita S, et al. Nitinol radius stent induces less platelet aggregation than stainless steel Palmaz-Schatz/Glanturco-Rubin II stents. J Am Coll Cardiol 1998;31(Suppl. 1):312. [59] Brunski JB. Metals. In: Ratner BD, Hoffman AS, Schoen FJ, Lemons JE, editors. Biomaterials science an introduction to

[60]

[61]

[62]

[63]

[64]

[65]

[66]

[67]

[68]

[69]

[70]

[71] [72] [73]

[74]

[75]

[76] [77]

[78]

[79] [80]

ARTICLE IN PRESS
G. Mani et al. / Biomaterials 28 (2007) 16891710 Kutryk MJ, editors. Coronary lesions a pragmatic approach. London: Martin Dunitz Ltd; 2002. p. 31935. Pypen C, Plenk H, Ebel M, Svagera R, Wernisch J. Characterization of microblasted and reactive ion etched surfaces on the commercially pure metals niobium, tantalum and titanium. J Mater Sci Mater Med 1997;8(12):7814. Ruckenstein E, Gourisankar SV. A surface energetic criterion of blood compatibility of foreign surfaces. J Colloid Interf Sci 1984;101(2):43651. Bhat SV. Synthetic polymers. In: Biomaterials. 2nd ed. Harrow: Alpha Science International Ltd; 2005. p. 645. Hamlin G, Rajah S, Crow M, Kester R. Evaluation of the thrombogenic potential of three types of arterial graft studied in an articial circulation. Br J Surg 1978;65(4):2726. Fontaine A, Koelling K, Clay J, Spigos D, Passos SD, Christoforidis G, et al. Decreased platelet adherence of polymer-coated tantalum stents. J Vasc Interv Radiol 1994;5(4):56772. Scheerder ID, Wilczek K, Verbeken E, Vandorpe J, Lan P, Schacht E, et al. Biocompatibility of polymer-coated oversized metallic stents implanted in normal porcine coronary arteries. Atherosclerosis 1995;114(1):10514. Wachem PV, Vreriks C, Beugeling T, Feijen J, Bantjes A, Detmers J, et al. The inuence of protein adsorption on interactions of cultured human endothelial cells with polymers. J Biomed Mater Res 1987;21(6):70118. Budd J, Bell P, James R. Attachment of indium-111 labelled endothelial cells to pretreated polytetrauoroethylene vascular grafts. Br J Surg 1989;76(12):125961. Knetsch M, Aldenhoff Y, Schraven M, Koole L. Human endothelial cell attachment and proliferation on a novel vascular graft prototype. J Biomed Mater Res A 2004;71(4):61524. Seeger J, Ingegno M, Bigatan E, Klingman N, Amery D, Widenhouse C, et al. Hydrophilic surface modication of metallic endoluminal stents. J Vasc Surg 1995;22(3):32735. DePalma VA, Baier RE, Ford JW, Gott VL, Furuse A. Investigation of three-surface properties of several metals and their relation to blood compatibility. J Biomed Mater Res 1972;6(4):3775. Goldberg L, Bosco P, Shors E, Klein S, Nelson R, White R. Effect of surface porosity on early thrombogenicity using vascular grafts with two surfaces in sequence. Trans Am Soc Artif Intern Organs 1981;27:51721. Hecker J, Scandrett L. Roughness and thrombogenicity of the outer surfaces of intravascular catheters. J Biomed Mater Res 1985;19(4): 38195. Zhao H, Humbeeck JV, Sohier J, Scheerder ID. Electrochemical polishing of 316L stainless steel slotted tube coronary stents. J Mater Sci Mater Med 2002;13(10):9116. Scheerder ID, Sohier J, Wang K, Verbeken E, Zhou XR, Frooyen L, et al. Metallic surface treatment using electrochemical polishing decreases thrombogenicity and neointimal hyperplasia after coronary stent implantation in a porcine model. JACC 1998;31(Suppl. 1):227A. Sheth S, Litvack F, Fishbein MC, Forrester JS, Eigler NL. Reduced thrombogenicity of polished and unpolished nitinol vs. stainless steel slotted-tube stents in a pig coronary artery model. JACC; 1996. Abstracts-Poster:197A. Hehrlein C, Zimmermann M, Metz J, Ensinger W, Kubler W. Inuence of surface texture and charge on the biocompatibility of endovascular stents. Coron Artery Dis 1995;6(7):5816. Palmaz J, Benson A, Sprague E. Inuence of surface topography on endothelialization of intravascular metallic material. J Vasc Interv Radiol 1999;10(4):43944. Scott NA, Robinson KA, Nunes GL, Thomas CN, Viel K, King SB, et al. Comparison of the thrombogenicity of stainless steel and tantalum coronary stents. Am Heart J 1995;129(5):86672. Yahia LH, Ryhanen J. Bioperformance of shape memory alloys. In: Yahia LH, editor. Shape memory implants. New York: Springer; 2000. p. 323. 1705 [101] Gutensohn K, Beythien C, Bau J, Fenner T, Grewe P, Koester R, et al. In vitro analyses of diamond-like carbon coated stents: reduction of metal ion release, platelet activation, and thrombogenicity. Thromb Res 2000;99:57785. [102] Klien CL, Kohler H, Kirkpatrick CJ. Increased adhesion and activation of polymorphonuclear neutrophil granulocytes to endothelial cells under heavy metal exposure in vitro. Pathobiology 1994;62:908. [103] Vogt P, Sigwart U, Urban P, Kaufmann U, Goy J, Stauffer J, et al. Immediate and late complications secondary to the implantation of a coronary endoprosthesis. Schweiz Med Wochenschr 1989;119(43): 15214. [104] Giessen WVD, Serruys P, Woerkens LV, Beatt K, Visser W, Jongkind J, et al. Arterial stenting with self-expandable and balloonexpandable endoprostheses. Int J Card Imag 1990;5(23): 16371. [105] Strauss B, Serruys P, Bertrand M, Puel J, Meier B, Goy J, et al. Quantitative angiographic follow-up of the coronary wallstent in native vessels and bypass grafts (European experienceMarch 1986 to March 1990). Am J Cardiol 1992;69(5):47581. [106] Karas S, Gravanis M, Santoian E, Robinson K, Anderberg K, King S. Coronary intimal proliferation after balloon injury and stenting in swine: an animal model of restenosis. J Am Coll Cardiol 1992;20(2):46774. [107] Ribeiro P, Gallo R, Antonius J, Mimish L, Sriram R, Bianchi S, et al. A new expandable intracoronary tantalum (Strecker) stent: early experimental results and follow-up to twelve months. Am Heart J 1993;125(2 Pt 1):50110. [108] Gunn J, Cumberland D. Stent coatings and local drug delivery. Eur Heart J 1999;20:1693700. [109] Hofma S. Recent developments in coated stents. Curr Intervent Cardiol Rep 2001;3:2836. [110] Liu CL, Chu PK, Lin GQ, Qi M. Anti-corrosion characteristics of nitride-coated AISI 316L stainless steel coronary stents. Surf Coat Technol 2006;201(6):28026. [111] Heldman A, Cheng L, Jenkins G, Heller P, Kim D, Ware M, et al. Paclitaxel stent coating inhibits neointimal hyperplasia at 4 weeks in a porcine model of coronary restenosis. Circulation 2001;103(18): 228995. [112] Nakayama Y, Ji-Youn K, Nishi S, Ueno H, Matsuda T. Development of high-performance stent: gelatinous photogel-coated stent that permits drug delivery and gene transfer. J Biomed Mater Res 2001;57(4):55966. [113] Chen M, Liang H, Chiu Y, Chang Y, Wei H, Sung H. A novel drugeluting stent spray-coated with multi-layers of collagen and sirolimus. J Control Release 2005;108(1):17889. [114] Huang Y, Wang L, Verweire I, Qiang B, Liu X, Verbeken E, et al. Optimization of local methylprednisolone delivery to inhibit inammatory reaction and neointimal hyperplasia of coated coronary stents. J Invasive Cardiol 2002;14(9):50513. [115] Huang N, Leng YX, Yang P, Chen JY, Sun H, Wang J, et al. Surface modication of coronary artery stent by TiO/TiN complex lm coating prepared with plasma immersion ion implantation and deposition. Nucl Instrum Methods Phys Res Sec B: Beam Interact Mater Atoms 2006;242(12):1821. [116] Thierry B, Winnik FM, Merhi Y, Silver J, Tabrizian M. Radionuclideshyaluronan-conjugate thromboresistant coatings to prevent in-stent restenosis. Biomaterials 2004;25(17):3895905. [117] Herman RA, Rybnikar A, Resch A, Marki B, Alt E, Stemberger A. Thrombogenicity of stainless steel coronary stents with a completely gold coated surface. JACC 1998;Suppl.(Feb):413A. [118] Edelman E, Seifert P, Groothuis A, Morss A, Bornstein D, Rogers C. Gold-coated NIR stents in porcine coronary arteries. Circulation 2001;103(3):42934. [119] Dahl JRV, Haager PK, Grube E, Gross M, Beythien C, Kromer EP, et al. Effects of gold coating of coronary stents on neointimal proliferation following stent implantation. Am J Cardiol 2002;89(7): 8015.

[81]

[82]

[83] [84]

[85]

[86]

[87]

[88]

[89]

[90]

[91]

[92]

[93]

[94]

[95]

[96]

[97]

[98]

[99]

[100]

ARTICLE IN PRESS
1706 G. Mani et al. / Biomaterials 28 (2007) 16891710 versus uncoated stainless steel stents in coronary artery disease. Am J Cardiol 2004;93:4747. Consigny P. Endothelial cell seeding on prosthetic surfaces. J Long Term Eff Med Implants 2000;10(12):7995. Giessen WVD, Serruys P, Visser W. Endothelialization of intravascular stents. J Interv Cardiol 1988;1:10920. Doi K, Matsuda T. Enhanced vascularization in a microporous polyurethane graft impregnated with basic broblast growth factor and heparin. J Biomed Mater Res 1997;34(3):36170. Doi K, Matsuda T. Signicance of porosity and compliance of microporous, polyurethane-based microarterial vessel on neoarterial wall regeneration. J Biomed Mater Res 1997;37(4):57384. Nakayama Y, Nishi S, Ishibashi-Ueda H, Matsuda T. Surface microarchitectural design in biomedical applications: in vivo analysis of tissue ingrowth in excimer laser-directed micropored scaffold for cardiovascular tissue engineering. J Biomed Mater Res 2000;51(3):5208. Nakayama Y, Nishi S, Ishibashi-Ueda H, Okamoto Y, Nemoto Y. Development of microporous covered stents: geometrical design of the luminal surface. Int J Artif Organs 2005;28(6):6008. Nakayama Y, Nishi S, Ishibashi-Ueda H. Fabrication of drugeluting covered stents with micropores and differential coating of heparin and FK506. Cardiovasc Radiat Med 2003;4(2):7782. Wieneke H, Dirsch O, Sawitowski T, Gu Y, Brauer H, Dahmen U, et al. Synergistic effects of a novel nanoporous stent coating and tacrolimus on intima proliferation in rabbits. Catheter Cardiovasc Interv 2003;60(3):399407. Kollum M, Farb A, Schreiber R, Terfera K, Arab A, Geist A, et al. Particle debris from a nanoporous stent coating obscures potential antiproliferative effects of tacrolimus-eluting stents in a porcine model of restenosis. Catheter Cardiovasc Interv 2005;64(1): 8590. Peng T, Gibula P, Yao K, Goosen M. Role of polymers in improving in improving the results of stenting in coronary arteries. Biomaterials 1996;17(7):68594. Cooper SL, Visser SA, Hergenrother RW, Lamba NMK. Polymers. In: Ratner BD, Hoffman AS, Schoen FJ, Lemons JE, editors. Biomaterials science an introduction to materials in medicine. 2nd ed. San Diego: Elsevier Academic Press; 2004. p. 6779. Giessen WVD, Slager C, Beusekom HV, Schenau DVI, Huijts R, Schuurbiers J, et al. Development of a polymer endovascular prosthesis and its implantation in porcine arteries. J Interv Cardiol 1992;5(3):17585. Davis JR. Polymeric materials. In: Handbook of materials for medical devices. Materials Park: ASM International; 2003. p. 15170. Murphy JG, Schwartz RS, Edwards WD, Camrud AR, Vlietstra RE, Holmer DR. Percutaneous polymeric stents in porcine coronary artery initial experience with polyethylene terepthalate stents. Circulation 1992;86(5):1596604. Beusekom HMMV, Giessen WJVD, Schenau DVI, Slager CJ. Synthetic polymers as an alternative to metal in stents? In vivo and mechanical behavior of polyethylene-terepthalate. Circulation 1992;86(Suppl. 4):I731 (Abstract # 2912). Wilczek KSI, Wang K, Verbeken E, Piessens J. Comparison of selfexpanding polyethylene terephthalate and metallic stents implanted in porcine iliac arteries. Cardiovasc Intervent Radiol 1996;19(3): 17680. Tanguay J, Zidar J, Phillips H, Stack R. Current status of biodegradable stents. Cardiol Clin 1994;12(4):699713. Stack R, Califf R, Phillips H, Pryor D, Quigley P, Bauman R, et al. Interventional cardiac catheterization at Duke medical center. Am J Cardiol 1988;62:3F24F. Agrawal CM, Haas KF, Leopold DA, Clark HG. Evaluation of poly(L-lactic acid) as a material for intravascular polymeric stents. Biomaterials 1992;13(3):17682. Lincoff AM, Schwartz RS, Giessen WJVD, Beusekom HMMV, Serruys PW, Holmes DR, et al. Biodegradable polymers can evoke a [120] Danzi G, Capuano C, Sesana M, Blasi AD, Predolini S, Antoniucci D. Patterns of in-stent restenosis after placement of NIR goldcoated stents in unselected patients. Catheter Cardiovasc Interv 2002;55(2):15762. [121] Seliger C, Schwennicke K, Schaffar C. Inuence of a rough, ceramic-like stent surface made of iridium oxide on neointimal structure and thickening. Eur Heart J 2000;21(Suppl.):286. [122] Sgura F, Mario CD, Liistro F, Montorfano M, Colombo A, Grube E. The lunar stent characteristics and clinical results. Herz 2002;27(6):5147. [123] Zhao Z-H, Sakagami Y, Osaka T. Toxicity of hydrogen peroxide produced by electroplated coatings to pathogenic bacteria. Can J Microbiol 1998;44(5):4417. [124] Mario CD, Grube E, Nisanci Y, Reifart N, Colombo A, Rodermann J, et al. MOONLIGHT: a controlled registry of an iridium oxide-coated stent with angiographic follow-up. Int J Cardiol 2004;95(23):32931. [125] Bolz A, Schaldach M. Articial heart valves: improved blood compatibility by PECVD a-SiC:H coating. Artif Organs 1990;14(4): 2609. [126] Unverdorben M, Sippel B, Degenhardt R, Sattler K, Fries R, Abt B, et al. Comparison of a silicon carbide-coated stent versus a noncoated stent in human beings: the Tenax versus Nir stent studys long-term outcome. Am Heart J 2003;145(4):E17. [127] Schuler P, Assefa D, Ylanne J, Basler N, Olschewski M, Ahrens I, et al. Adhesion of monocytes to medical steel as used for vascular stents is mediated by the integrin receptor Mac-1 (CD11b/CD18; alphaM beta2) and can be inhibited by semiconductor coating. Cell Commun Adhes 2003;10(1):1726. [128] Monnink S, Boven V, Peels, Tigchelaar, Kam D, Crijns, et al. Silicon-carbide coated coronary stents have low platelet and leukocyte adhesion during platelet activation. J Investig Med 1999;47(6):30410. [129] Kalnins U, Erglis A, Dinne I, Kumsars I, Jegere S. Clinical outcomes of silicon carbide coated stents in patients with coronary artery disease. Med Sci Monit 2002;8(2):11620. [130] Tanajura LF, Abizaid AA, Feres F, Pinto I, Mattos L, Staico R, et al. Randomized intravascular ultrasound comparison between patients that underwent amorphous hydrogenated silicon-carbide coated stent deployment versus uncoated stent. JACC; 2003. AbstractsAngiography & Interventional Cardiology:58A. [131] Unverdorben M, Sattler K, Degenhardt R, Fries R, Abt B, Wagner E, et al. Comparison of silicon carbide coated stent versus a noncoated stent in humans: the Tenax-versus Nir-stent Study. J Interv Cardiol 2003;16(4):32533. [132] Bickel C, Rupprecht H, Darius H, Binz C, Hauroder B, Krummenauer F, et al. Substantial reduction of platelet adhesion by heparin-coated stents. J Interv Cardiol 2001;14(4):40713. [133] Gutensohn K, Beythien C, Bau J, Fenner T, Grewe P, Koester R, et al. In vitro analyses of diamond-like carbon coated stents. Reduction of metal ion release, platelet activation, and thrombogenicity. Thromb Res 2000;99(6):57785. [134] Linder S, Pinkowski W, Aepfelbacher M. Adhesion, cytoskeletal architecture and activation status of primary human macrophages on a diamond-like carbon coated surface. Biomaterials 2002;23(3): 76773. [135] Antoniucci D, Bartorelli A, Valenti R, Montorsi P, Santoro GM, Fabbiocchi F, et al. Clinical and angiographic outcome after coronary arterial stenting with the carbostent. Am J Cardiol 2000;85:8215. [136] Colombo A, Airoldi F. Passive coating: the dream does not come true. J Invasive Cardiol 2003;15(10):5667. [137] Korkmaz M, Tayfun E, Muderrisoglu H, Yildirir A, Ozin B, Ulucam M, et al. Carbon coating of stents has no effect on inammatory response to primary stent deployment. Angiology 2002;53(5):5638. [138] Airoldi F, Colombo A, Tavano D, Stankovic G, Klugmann S, Paolillo V, et al. Comparison of diamond-like carbon-coated stents

[139] [140] [141]

[142]

[143]

[144]

[145]

[146]

[147]

[148]

[149]

[150]

[151]

[152]

[153]

[154]

[155] [156]

[157]

[158]

ARTICLE IN PRESS
G. Mani et al. / Biomaterials 28 (2007) 16891710 unique inammatory response when implanted in the coronary artery. Circulation 1992;86(Suppl.4):I801 (Abstract # 3186). Giessen WV, Lincoff A, Schwartz R, Beusekom HV, Serruys P, Holmes D, et al. Marked inammatory sequelae to implantation of biodegradable and nonbiodegradable polymers in porcine coronary arteries. Circulation 1996;94(7):16907. Lincoff A, Furst J, Ellis S, Tuch R, Topol E. Sustained local delivery of dexamethasone by a novel intravascular eluting stent to prevent restenosis in the porcine coronary injury model. J Am Coll Cardiol 1997;29:80816. Tamai H, Gaki K, Kyo E, Kosuga K, Kawashima A, Matsui S, et al. Initial and 6-month results of biodegradable poly-L-lactic acid coronary stents in humans. Circulation 2000;102(4): 399404. Bar F, Veen FVD, Benzina A. New biocompatible polymer surface coating for stents results in a low neointimal response. J Biomed Mater Res 2000;52:1938. Edelman E, Adams D, Karnovsky M. Effect of controlled adventitial heparin delivery on smooth muscle cell proliferation following endothelial injury. Proc Natl Acad Sci 1990;87(10): 37737. Whelan D, Giessen WVD, Krabbendam S, Vliet EV, Verdouw P, Serruys P, et al. Biocompatibility of phosphorylcholine coated stents in normal porcine coronary arteries. Heart 2000;83: 33845. Banai S, Gertz S, Gavish L, Chorny M, Perez L, Lazarovichi G, et al. Tyrphostin AGL-2043 eluting stent reduces neointima formation in porcine coronary arteries. Cardiovasc Res 2004;64(1):16571. Tan LP, Venkatraman SS, Sung PF, Wang XT. Effect of plasticization on heparin release from biodegradable matrices. Int J Pharm 2004;283:8996. Venkatraman S, Tan L, Joso J, Boey Y, Wang X. Biodegradable stents with elastic memory. Biomaterials 2006;27(8):15738. Valimaa T, Laaksovirta S. Degradation behaviour of self-reinforced 80L/20G PLGA devices in vitro. Biomaterials 2004;25(78): 122532. Talja M, Lumiaho J, Heino A, Alhava E, Tormala P. Drainage and antireux characteristics of a partial self-degradable self-expanding braided SR-PLGA ureteral stents. An experimental study. Eur Urol Suppl 2004;3(2):190. Severini A, Mantero S, Tanzi M, Cigada A, Addis F, Cozzi G, et al. In vivo study of polyurethane-coated Gianturco-Rosch biliary Zstents. Cardiovasc Intervent Radiol 1999;22(6):5104. Muller-Hulsbeck S, Walluscheck K, Priebe M, Grimm J, Cremer J, Heller M. Experience on endothelial cell adhesion on vascular stents and stent-grafts: rst in vitro results. Invest Radiol 2002;37(6): 31420. Lambert T, Dev V, Rechavia E, Forrester J, Litvack F, Eigler N. Localized arterial wall drug delivery from a polymer-coated removable metallic stent. Kinetics, distribution, and bioactivity of forskolin. Circulation 1994;90(2):100311. Dev V, Eigler N, Sheth S, Lambert T, Forrester J, Litvack F. Kinetics of drug delivery to the arterial wall via polyurethane-coated removable nitinol stent: comparative study of two drugs. Cathet Cardiovasc Diagn 1995;34(3):2728. Kondyurin A, Maitz M, Romanova V, Begishev V, Kondyurina I, Guenzel R. Drug release from polyureaurethane coating modied by plasma immersion ion implantation. J Biomater Sci Polym Ed 2004;15(2):14559. Rechavia E, Litvack F, Fishbien M, Nakamura M, Eigler N. Biocompatibility of polyurethane-coated stents: tissue and vascular aspects. Cathet Cardiovasc Diagn 1998;45(2):2027. Davis JR. In: Handbook of medical devices. Materials Park: ASM International; 2003. Park JB. Soft tissue replacement implants. In: Biomaterials science and engineering. New York: Plenum Press; 1987. p. 338. Cumberland D, Gunn J, Malik N, Holt C. Biomimicry 1: PC. Semin Interv Cardiol 1998;3(34):14950. 1707 [179] Whelan D, Giessen WVD, Krabbendam S, Vliet EV, Verdouw P, Serruys P, et al. Biocompatibility of phosphorylcholine coated stents in normal porcine coronary arteries. Heart 2000;83(3):33845. [180] Atalar E, Haznedaroglu I, Aytemir K, Aksoyek S, Ovunc K, Oto A, et al. Effects of stent coating on platelets and endothelial cells after intracoronary stent implantation. Clin Cardiol 2001;24(2): 15964. [181] Grenadier E, Roguin A, Hertz I, Peled B, Boulos M, Nikolsky E, et al. Stenting very small coronary narrowings (o2 mm) using the biocompatible phosphoryline coated coronary stent. Catheter Cardiovasc Interv 2002;55(3):3038. [182] Galli M, Bartorelli A, Bedogni F, DeCesare N, Klugmann S, Maiello L, et al. Italian BiodivYsio open registry (BiodivYsio PCcoated stent): study of clinical outcomes of the implant of a PCcoated coronary stent. J Invas Cardiol 2000;12(9):4528. [183] Zheng H, Barragan P, Corcos T, Simeoni J, Favereau X, Roquebert P, et al. Clinical experience with a new biocompatible phosphorylcholine-coated coronary stent. J Invas Cardiology 1999;11(10): 60814. [184] Lewis A, Tolhurst L, Stratford P. Analysis of a phosphorylcholinebased polymer coating on a coronary stent pre- and postimplantation. Biomaterials 2002;23(7):1697706. [185] Lewis A, Furze J, Small S, Robertson J, Higgins B, Taylor S, et al. Long-term stability of a coronary stent coating post-implantation. J Biomed Mater Res 2002;63(6):699705. [186] Lewis A, Willis S, Small S, Hunt S, Obyrne V, Stratford P. Drug loading and elution from a phosphorylcholine polymer-coated coronary stent does not affect long-term stability of the coating in vivo. Biomed Mater Eng 2004;14(4):35570. [187] Lewis A, Stratford P. Phosphorylcholine-coated stents. J Long Term Eff Med Implants 2002;12(4):23150. [188] Buellesfeld L, Grube E. ABT-578-eluting stents. The promising successor of sirolimus- and paclitaxel-eluting stent concepts? Herz 2004;29(2):16770. [189] Liu X, Huang Y, Hanet C, Vandormael M, Legrand V, Dens J, et al. Study of antirestenosis with the BiodivYsio dexamethasone-eluting stent (STRIDE): a rst-in-human multicenter pilot trial. Catheter Cardiovasc Interv 2003;60(2):1728. [190] Kwok O, Chow W, Law T, Chiu A, Ng W, Lam W, et al. First human experience with angiopeptin-eluting stent: a quantitative coronary angiography and three-dimensional intravascular ultrasound study. Catheter Cardiovasc Interv 2005;66(4):5416. [191] Verheye S, Markou C, Salame M, Wan B, King S, Robinson K, et al. Reduced thrombus formation by hyaluronic acid coating of endovascular devices. Arterioscler Thromb Vasc Biol 2000;20(4): 116872. [192] Heublein B, Evagorou E, Rohde R, Ohse S, Meliss R, Barlach S, et al. Polymerized degradable hyaluronana platform for stent coating with inherent inhibiting effects on neointimal formation in a porcine coronary model. Int J Artif Organs 2002;25(12):116673. [193] Pitt W, Morris R, Mason M, Hall M, Luo Y, Prestwich G. Attachment of hyaluronan to metallic surfaces. J Biomed Mater Res A 2004;68(1):95106. [194] Radosevich M, Goubran H, Burnouf T. Fibrin sealant: scientic rationale, production methods, properties, and current clinical use. Vox Sang 1997;72(3):13343. [195] Weisel JW. The mechanical properties of brin for basic scientists and clinicians. Biophys Chem 2004;112(23):26776. [196] Holmes D, Camrud A, Jorgenson M, Edwards W, Schwartz R. Polymeric stenting in the porcine coronary artery model: differential outcome of exogenous brin sleeves versus polyurethane-coated stents. J Am Coll Cardiol 1994;24(2):52531. [197] McKenna CJ, Camrud AR, Sangiorgi G, Kwon HM, Edwards WD, Holmes DR, Jr. et al. Fibrin-lm stenting in a porcine coronary injury model: Efcacy and safety compared with uncoated stents. JACC 1998;31(6):14341438. [198] Byer A, Peters S, Settepani F, Pagliaro M, Galletti G. Fibrin sealant coated stents compared with non-coated stents in a porcine carotid

[159]

[160]

[161]

[162]

[163]

[164]

[165]

[166]

[167] [168]

[169]

[170]

[171]

[172]

[173]

[174]

[175]

[176] [177] [178]

ARTICLE IN PRESS
1708 G. Mani et al. / Biomaterials 28 (2007) 16891710 artery model. Preliminary study report. J Cardiovasc Surg (Torino) 2001;42(4):5439. Wilensky RL, March KL, Gradus-Pizlo I, Spaedy AJ, Hathaway DR. Methods and devices for local drug delivery in coronary and peripheral arteries. Trends Cardiovasc Med 1993;3(5):16370. Hertzog BA, Thanos C, Sandor M, Raman V, Edelman ER. Cardiovascular drug delivery systems. In: Mathiowitz E, editor. Encyclopedia of controlled drug delivery. New York: Wiley; 1999. p. 16172. Kanellakopoulou K, Giamarellos-Bourboulis E. Carrier systems for the local delivery of antibiotics in bone infections. Drugs 2000;59(6): 122332. Saito N, Murakami N, Takahashi J, Horiuchi H, Ota H, Kato H, et al. Synthetic biodegradable polymers as drug delivery systems for bone morphogenetic proteins. Adv Drug Deliv Rev 2005;57(7): 103748. Kimura H, Ogura Y. Biodegradable polymers for ocular drug delivery. Ophthalmologica 2001;215(3):14355. Yasukawa T, Ogura Y, Kimura H, Sakurai E, Tabata Y. Drug delivery from ocular implants. Expert Opin Drug Deliv 2006;3(2):26173. Domb A, Maniar M, Bogdansky S, Chasin M. Drug delivery to the brain using polymers. Crit Rev Ther Drug Carrier Syst 1991;8(1): 117. Wang P, Frazier J, Brem H. Local drug delivery to the brain. Adv Drug Deliv Rev 2002;54(7):9871013. Tsuji T, Tamai H, Igaki K, Kyo E, Kosuga K, Hata T, et al. Biodegradable stents as a platform to drug loading. Int J Cardiovasc Intervent 2003;5(1):136. Day JRS, Landis RC, Taylor KM. Heparin is much more than just an anticoagulant. J Cardiothor Vascu Anesth 2004;18(1):93100. Hardhammar P, Beusekom HV, Emanuelsson H, Hofma S, Albertsson P, Verdouw P, et al. Reduction in thrombotic events with heparin-coated Palmaz-Schatz stents in normal porcine coronary arteries. Circulation 1996;93(3):42330. Breckwoldt W, Belkin M, Gould K, Allen M, Connolly R, Termin P. Modication of the thrombogenicity of a self-expanding vascular stent. J Invest Surg 1991;4(3):26978. Grode G, Anderson S, Grotta H, Falb R. Nonthrombogenic materials via a simple coating process. Trans Am Soc Artif Intern Organs 1969;15:16. Tanzawa H, Mori Y, Harumiya N, Miyama H, Hori M, Ohshima N, et al. Preparation and evaluation of a new athrombogenic heparinized hydrophilic polymer for use in cardiovascular system. Trans Am Soc Artif Intern Organs 1973;19:18894. Larm O, Larssom R, Olsson P. A new non-thrombogenic surface prepared by selective covalent binding of heparin via a modied reducing terminal residue. Biomater Med Devices Artif Organs 1983;11(23):16173. Labarre D, Jozefowicz M, Boffa MC. Properties of heparin-poly (methyl methacrylate) copolymers. J Biomed Mater Res 1977;11(2): 28395. Goosen M, Sefton M. Properties of a heparin-poly(vinyl alcohol) hydrogel coating. J Biomed Mater Res 1983;17(2):35973. Mazid MA, Scott E, Li N-H. New biocompatible polyurethane-type copolymer with low molecular weight heparin. Clin Mater 1991;8(12):7180. Yang Z, Birkenhauer P, Julmy F, Chickering D, Ranieri J, Merkle H, et al. Sustained release of heparin from polymeric particles for inhibition of human vascular smooth muscle cell proliferation. J Control Release 1999;60(23):26977. Bonan R, Bhat K, Lefevre T, Lemarbre L, Paiement P, Wolff R, et al. Coronary artery stenting after angioplasty with selfexpanding parallel wire metallic stents. Am Heart J 1991;121(5): 152230. Stratienko A, Zhu D, Lambert C. Improved thromboresistance of heparin coated Palmaz-Schatz coronary stents in an animal model. Circulation 1993;88(Abstracts):I-596. [220] Sheth S, Dev V, Jacobs H, Forrester J, Litvack F, Eigler N. Prevention of subacute stent thrombosis by polymerpolyethylene oxide-heparin coating in the rabbit carotid artery. J Am Coll Cardiol 1995;25(Abstracts):348A. [221] Kiemeneij F, Serruys P, Macaya C, Rutsch W, Heyndrickx G, Albertsson P, et al. Continued benet of coronary stenting versus balloon angioplasty: ve-year clinical follow-up of Benestent-I trial. J Am Coll Cardiol 2001;37(6):1598603. [222] Serruys P, Hout BV, Bonnier H, Legrand V, Garcia E, Macaya C, et al. Randomised comparison of implantation of heparincoated stents with balloon angioplasty in selected patients with coronary artery disease (Benestent II). Lancet 1998;352(9129): 67381. [223] Stone G, Brodie B, Grifn J, Morice M, Costantini C, Goar FS, et al. Prospective, multicenter study of the safety and feasibility of primary stenting in acute myocardial infarction: in-hospital and 30-day results of the PAMI stent pilot trial. Primary angioplasty in myocardial infarction stent pilot trial investigators. J Am Coll Cardiol 1998;31(1):2330. [224] Dzavik V, Carere R, Mancini G, Cohen E, Catellier D, Anderson T, et al. Predictors of improvement in left ventricular function after percutaneous revascularization of occluded coronary arteries: a report from the total occlusion study of Canada (TOSCA). Am Heart J 2001;142(2):3018. [225] Vrolix M, Legrand V, Reiber J, Grollier G, Schalij M, Brunel P, et al. Heparin-coated wiktor stents in human coronary arteries (MENTOR trial). MENTOR trial investigators. Am J Cardiol 2000;86(4):3859. [226] Matsumoto Y, Shimokawa H, Morishige K, Eto Y, Takeshita A. Reduction in neointimal formation with a stent coated with multiple layers of releasable heparin in porcine coronary arteries. J Cardiovasc Pharmacol 2002;39(4):51322. [227] Marx S, Jayaraman T, Go L, Marks A. Rapamycin-FKBP inhibits cell cycle regulators of proliferation in vascular smooth muscle cells. Circul Res 1995;76(3):4127. [228] Poon M, Marx S, Gallo R, Badimon J, Taubman M, Marks A. Rapamycin inhibits vascular smooth muscle cell migration. J Clin Invest 1996;98(10):227783. [229] Kipshidze N, Leon M, Tsapenko M, Falotico R, Kopia G, Moses J. Update on sirolimus drug-eluting stents. Curr Pharm Des 2004; 10(4):33748. [230] US Food and Drug Administration; Center for Devices and Radiological Health, CYPHERTM Sirolimus-eluting coronary stentP020026. Available from /http://www.fda.gov/cdrh/mda/ docs/p020026.htmlS updated 10 June 2003. [231] Schampaert E, Cohen EA, Schluter M, Reeves F, Traboulsi M, Title LM, et al. The Canadian study of the sirolimus-eluting stent in the treatment of patients with long de novo lesions in small native coronary arteries (C-SIRIUS). J Am Coll Cardiol 2004;43(6): 11105. [232] Moussa I, Leon M, Baim D. Impact of sirolimus-eluting stents on outcome in diabetic patients: a SIRIUS substudy. Circulation 2004;109:22738. [233] Morice M-C, Serruys P, Costantini C. Three-year follow-up of the RAVEL study: a randomized study with the Sirolimuseluting Bx VelocityTM stent in the treatment of patients with De Novo native coronary lesions. JACC 2004;43(5; Suppl. 1): 87A8A. [234] Schofer J, Schluter M, Gershlick A. Sirolimus-eluting stents for treatment of patients with long atherosclerotic lesions in small coronary arteries: double-blind, randomized controlled trial (E-SIRIUS). Lancet 2003;362:10939. [235] Wijns W. The European multicenter, randomized, double-blind study of the sirolimus-eluting stent in the treatment of patients with de novo coronary artery lesions (E-SIRIUS): 1-year clinical outcomes. Eur Heart J 2003;24(5 Suppl 1):267. [236] Liistro F, Bolognese L. Drug-eluting stents. Heart Drug 2003;3: 20313.

[199]

[200]

[201]

[202]

[203] [204]

[205]

[206] [207]

[208] [209]

[210]

[211]

[212]

[213]

[214]

[215] [216]

[217]

[218]

[219]

ARTICLE IN PRESS
G. Mani et al. / Biomaterials 28 (2007) 16891710 [237] Katuza GL, Gershlick AH, Park S-J, Scheerder ID, Chevalier B, Camenzind E, et al. Comparison of neointimal formation in polymer-free paclitaxel stents versus stainless steel stents from the ASPECT and ELUTES randomized clinical trial. Am J Cardiol 2004;94:199201. [238] Gershlick A, Scheerder I, Chevalier B. Inhibition of restenosis with a paclitaxel-eluting, polymer-free coronary stent: the European evaluation of paclitaxel eluting stent (ELUTES) trial. Circulation 2004;109:48793. [239] TAXUSTM Express2TM Coronary Stent System (P030025); Summary of safety and effectiveness data. /http://wwwfdagov/cdrh/ pdf3/P030025bpdfS /http://wwwfdagov/ohrms/dockets/dailys/04/ sep04/090904/04m-0403-aav0001-03-SSED-vol1pdfS 2003. [240] Ranade S, Miller K, Richard R, Chan A, Allen M, Helmus M. Physical characterization of controlled release of paclitaxel from the TAXUS Express2 drug-eluting stent. J Biomed Mater Res A 2004;71(4):62534. [241] Sipos L, Som A, Faust R, Richard R, Schwarz M, Ranade S, et al. Controlled delivery of paclitaxel from stent coatings using poly(hydroxystyrene-b-isobutylene-b-hydroxystyrene) and its acetylated derivative. Biomacromolecules 2005;6(5):257082. [242] Finci L, Silber S, Grube E, Belardi J. Paclitaxel-eluting stent in complicated lesions of patient subsets: a subanalysis from the TAXUS II study. JACC 2004;43(5; Suppl. 1):46A. [243] Drachman DE. Clinical experience with drug-eluting Stents. Rev Cardiovasc Med 2002;3(Suppl. 5):S317. [244] Stone G, Ellis S, Cox D, Hermiller J, OShaughnessy C, Mann J, et al. One-year clinical results with slow-release, polymer-based, paclitaxel-eluting TAXUS stent: the TAXUS-IV trial. Circulation 2004;109(16):19427. [245] Kastrati A, Dibra A, Eberle S, Mehilli J, Lezo JSD, Goy J-J, et al. Sirolimus-eluting stents vs Paclitaxel-eluting stents in patients with coronary artery disease: meta-analysis of randomized trials. JAMA 2005;294:81925. [246] Creel C, Lovich M, Edelman E. Arterial paclitaxel distribution and deposition. Circ Res 2000;86(8):87984. [247] Lovich M, Edelman E. Mechanisms of transmural heparin transport in the rat abdominal aorta after local vascular delivery. Circ Res 1995;77(6):114350. [248] Hwang C, Wu D, Edelman E. Impact of transport and drug properties on the local pharmacology of drug-eluting stents. Int J Cardiovasc Intervent 2003;5(1):712. [249] Ong ATL, Hoye A, Aoki J, vanMieghem CA, Granillo GAR, Sonnenschein K, et al. Thirty-day incidence and six-month clinical outcome of thrombotic stent occlusion after bare-metal, sirolimus, or paclitaxel stent implantation. J Am Coll Cardiol 2005;45(6): 94753. [250] Heller LI, Shemwell KC, Hug K. Late stent thrombosis in the absence of prior intracoronary brachytherapy. Catheter Cardiovasc Interv 2001;53(1):238. [251] Wang F, Stouffer G, Waxman S, Uretsky B. Late coronary stent thrombosis: early vs. late stent thrombosis in the stent era. Catheter Cardiovasc Interv 2002;55(2):1427. [252] McFadden EP, Stabile E, Regar E, Cheneau E, Ong ATL, Kinnaird T, et al. Late thrombosis in drug-eluting coronary stents after discontinuation of antiplatelet therapy. Lancet 2004;364(9444): 151921. [253] Joner M, Finn A, Farb A, Mont E, Kolodgie F, Ladich E, et al. Pathology of drug-eluting stents in humans: delayed healing and late thrombotic risk. J Am Coll Cardiol 2006;48(1):2035. [254] Virmani R, Guagliumi G, Farb A, Musumeci G, Grieco N, Motta T, et al. Localized hypersensitivity and late coronary thrombosis secondary to a sirolimus-eluting stent: should we be cautious? Circulation 2004;109(6):7015. [255] Ong ATL, Serruys PW. Drug-eluting stents: current issues. Tex Heart Inst J 2005;32(3):3727. [256] Moreno R, Fernandez C, Hernandez R, Alfonso F, Angiolillo DJ, Sabate M, et al. Drug-eluting stent thrombosis: results from a 1709 pooled analysis including 10 randomized studies. J Am Coll Cardiol 2005;45(6):9549. Mauri L, OMalley AJ, Cutlip DE, Ho KK, Popma JJ, Chauhan MS, et al. Effects of stent length and lesion length on coronary restenosis. Am J Cardiol 2004;93(11):13406. Chieffo A, Bonizzoni E, Orlic D, Stankovic G, Rogacka R, Airoldi F, et al. Intraprocedural stent thrombosis during implantation of sirolimus-eluting stents. Circulation 2004;109(22):27326. Fujii K, Carlier SG, Mintz GS, Yang Y-M, Moussa I, Weisz G, et al. Stent underexpansion and residual reference segment stenosis are related to stent thrombosis after sirolimus-eluting stent implantation: an intravascular ultrasound study. J Am Coll Cardiol 2005;45(7):9958. Virmani R, Farb A, Guagliumi G, Kolodgie F. Drug-eluting stents: caution and concerns for long-term outcome. Coron Artery Dis 2004;15(6):3138. Virmani R, Kolodgie F, Farb A. Drug-eluting stents: are they really safe? Am Heart Hosp J 2004;2(2):858. Nebeker JR, Virmani R, Bennet CL, Hoffman JM, Samore MH, Alvarez J, et al. Hypersensitivity cases associated with drug-eluting coronary stents: a review of available cases from the Research on Adverse Drug Events and Reports (RADAR) project. J Am Coll Cardiol 2006;47(1):1823. McFadden E, Stabile E, Regar E, Cheneau E, Ong A, Kinnaird T, et al. Late thrombosis in drug-eluting coronary stents after discontinuation of antiplatelet therapy. Lancet 2004;364:14667. Iakovou I, Schmidt T, Bonizzoni E, Ge L, Sangiorgi G, Stankovic G, et al. Incidence, predictors, and outcome of thrombosis after successful implantation of drug-eluting stents. JAMA 2005;293(17): 21546. Mani G, Mahapatro A, Johnson DM, Patel DN, Feldman MD, Ayon AA, et al. Therapeutic self-assembled monolayers. Baltimore, MD: Biomedical Engineering Society; 2005 Abstract no: 652. Mahapatro A, Johnson DM, Patel DN, Feldman MD, Ayon AA, Agrawal CM. Surface modication of functional self-assembled monolayers on 316L stainless steel via lipase catalysis. Langmuir 2006;22(3):9015. Mani G, Johnson DM, Marton D, Mahapatro A, Feldman M, Patel D, et al. Drug elution using self-assembled monolayers. In: Transactions of the 31st annual meeting of the Society For Biomaterials. Society For Biomaterials, Pittsburgh, PA, USA; 2006. Abstract no: 307. Mani G, Johnson DM, Marton D, Mahapatro A, Feldman M, Patel D, et al. Stability of therapeutic self assembled monolayers during drug loading and in-vitro drug delivery. In: Transactions of the 31st annual meeting of the Society for Biomaterials, Society for Biomaterials, Pittsburgh, PA, USA; 2006. Abstract no: 327. Goodfellow Corporation, Devon, PA. Iron (Fe). Available from /http://www.goodfellow.com/csp/active/gfHome.cspS July 2006. Magnesium Elektron, Manchaster, England. ELEKTRON WE43. Available from /http://www.magnesium-elektron.com/ data/downloads/DS467WE43.pdfS July 2006. US Food and Drug Administration; Center for Devices and Radiological Health.BiodivYsioTMAS PC (phosphorylcholine) coated stent delivery systemP000011. Available from /http:// www.fda.gov/cdrh/mda/docs/p000011.htmlS updated 8 March 2001. US Food and Drug Administration; Center for Devices and Radiological Health.BeStentTM 2 with Discrete TechnologyTM Over-the-Wire and rapid exchange coronary stent delivery systemsP000022. Available from /http://www.fda.gov/cdrh/mda/ docs/p000022.htmlS updated 31 January 2001. US Food and Drug Administration, Center for Devices and Radiological Health. NIRexTM premounted coronary stent systemP020040. Available from /http://www.fda.gov/cdrh/mda/ docs/p020040.htmlS updated 27 January 2004. US Food and Drug Administration, Center for Devices and Radiological Health. TAXUSTM Express2TM paclitaxel-eluting

[257]

[258]

[259]

[260]

[261] [262]

[263]

[264]

[265]

[266]

[267]

[268]

[269] [270]

[271]

[272]

[273]

[274]

ARTICLE IN PRESS
1710 G. Mani et al. / Biomaterials 28 (2007) 16891710 /http://www.fda.gov/cdrh/mda/docs/p040016.htmlS updated 13 April 2005. [276] US Food and Drug Administration, Center for Devices and Radiological Health. Rithron-XR coronary stent system P030037. Available from /http://www.fda.gov/cdrh/mda/docs/ p030037.htmlS updated 23 May 2005. coronary stent systemP030025. Available from /http:// www.fda.gov/cdrh/mda/docs/p030025.htmlS updated 9 September 2004. [275] US Food and Drug Administration, Center for Devices and Radiological Health. Boston Scientic LiberteTM MonorailTM and Over-the-Wire coronary stent systemsP040016. Available from

You might also like