You are on page 1of 19

ARTICLE IN PRESS

Biomaterials 28 (2007) 5009–5027


www.elsevier.com/locate/biomaterials

Leading Opinion

Prosthetic vascular grafts: Wrong models,


wrong questions and no healing$
Peter Zilla, Deon Bezuidenhout, Paul Human
Christian Barnard Department of Cardiothoracic Surgery/Cardiovascular Research Unit, University of Cape Town Medical School, Cape Town, South Africa
Received 5 June 2007; accepted 6 July 2007
Available online 3 August 2007

Abstract

In humans, prosthetic vascular grafts remain largely without an endothelium, even after decades of implantation. While this
shortcoming does not affect the clinical performance of large bore prostheses in aortic or iliac position, it contributes significantly to the
high failure rate of small- to medium-sized grafts (SMGs). For decades intensive but largely futile research efforts have been under way
to address this issue. In spite of the abundance of previous studies, a broad analysis of biological events dominating the incorporation of
vascular grafts was hitherto lacking. By focusing on the three main contemporary graft types, expanded polytetrafluoroethylene
(ePTFE), Dacron and Polyurethane (PU), accumulated clinical and experimental experience of almost half a century was available. The
main outcome of this broad analysis—supported by our own experience in a senescent non-human primate model—was twofold: Firstly,
inappropriate animal models, which addressed scientific questions that missed the point of clinical relevance, were largely used. This led
to a situation where the vast majority of investigators unintentionally studied transanastomotic rather than transmural or blood-borne
endothelialization. Given the fact that in patients transanastomotic endothelialization (TAE) covers only the immediate perianastomotic
region of sometimes very long prostheses, TAE is rather irrelevant in the clinical context. Secondly, transmural endothelialization seems
to have a time window of opportunity before a build-up of an adverse microenvironment. In selecting animal models that prematurely
terminate this build-up through the early presence of an endothelium, the most significant ‘impairment factor’ for physiological tissue
regeneration in vascular grafts remained ignored.
By providing insight into mechanisms and experimental designs which obscured the purpose and scope of several decades of vascular
graft studies, future research may better address clinical relevance.
r 2007 Elsevier Ltd. All rights reserved.

Keywords: Vascular grafts; Endothelialization; Polytetrafluoroethylene; Polyethylene terephthalate; Polyurethane; Foreign body response

1. Introduction and iliac surgery, smaller diameter grafts became the


nemesis of research and a symbol for the limitations of
For almost half a century synthetic vascular grafts have modern biotechnology. The main reasons for the poor
been an integral tool of vascular surgery. However, while performance of small- (p4 mm) to medium-sized (p7 mm)
large bore prostheses added an impressive scope to aortic grafts (SMGs) are anastomotic intimal hyperplasia and
ongoing surface thrombogenicity. In spite of these perpe-
$
Note: Leading Opinions: This paper provides evidence-based scientific
tual shortcomings of contemporary vascular prostheses, no
opinions on topical and important issues in biomaterials science. They alternative concept has yet emerged that promises to
have some features of an invited editorial but are based on scientific facts, replace the current generation of synthetic grafts soon.
and some features of a review paper, without attempting to be While a small spearhead of researchers pursues sophisti-
comprehensive. These papers have been reviewed for factual, scientific cated tissue engineering approaches [1–4] the majority of
content.
Corresponding author. Cape Heart Centre, University of Cape Town/ surgeons continue to implant the well-established products
Faculty of Health Services, Observatory 7925, Cape Town, South Africa. of the past decades. Similarly, commercial grafts continue
E-mail address: peter.Zilla@uct.ac.za (P. Zilla). to be manufactured on the basis of material choices,

0142-9612/$ - see front matter r 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.biomaterials.2007.07.017
ARTICLE IN PRESS
5010 P. Zilla et al. / Biomaterials 28 (2007) 5009–5027

mechanical strength, regulatory compliance and surgical


preferences rather than with a view towards biological
integration and functional tissue regeneration. Neverthe-
less, regenerative medicine has made significant inroads in
recent years and it seems only a matter of time until
synthetic vascular grafts will also benefit from this
development. Before embarking on concepts stimulating
the in situ generation of functional vascular tissue,
however, it seems paramount to understand why such a
physiological tissue formation would not spontaneously
occur in contemporary grafts. For reasons still incomple-
tely understood, neither transmural ingrowth through the
graft wall nor transanastomotic ingrowth from the
adjacent artery seem to be capable of endothelializing
more than a narrow zone confined to the immediate
anastomotic area in humans [5,6]. Given the high number
of annual implants, it is surprising how poorly investigated
these limitations have remained throughout the years.
While the initial enthusiasm for the availability of polyester
grafts prompted several major studies in the 1960s and
1970s [7–10], relatively little has been added to our
understanding since then. Typical for subsequent prosthe-
tic vascular graft research, intensive efforts went into ever
new permutations of prostheses without a sound baseline Fig. 1. Schematic representation of rapid trans-anastomotic endothelia-
lization (TAE) in animal models (bottom) compared to limited TEA in
knowledge of pathological events behind the healing humans (top). Typically, the transanastomotic ingrowth stoppage in
impairment of grafts which have been clinically implanted humans results in permanently non-endothelialized midgraft sections. By
for decades. Furthermore, by choosing inadequate animal prematurely reaching confluence in experimental grafts through transa-
models and too short graft lengths, an involuntary nastomotic ingrowth, neither blood-born endothelial cells nor ingrowing
emphasis of the majority of these studies was on capillaries from outside can unfold their potential to populate the blood
surface.
transanastomotic endothelial ingrowth—a biological phe-
nomenon of utter irrelevance for the clinical set-up.
the time of retrieval (Fig. 1) [11]. Until 10 years ago,
2. Transanastomotic endothelialization (TAE): barking up approximately 90% of prostheses in large animal models
the wrong tree were as short as 5.571.2 cm (Dacron and expended
polytetrafluoroethylene (ePTFE)) and 4.571.7 cm polyur-
Given the key role endothelium plays in preventing a ethanes (PU), with a median implantation periods of
blood vessel from occluding, it is understandable that the 91.8717.3 days for Dacron grafts, 60.079.36 days for
main focus of vascular graft studies was on endothelializa- polytetrafluoroethylene (PTFE) grafts [11] and 907127.1
tion. It is therefore even more surprising that the majority days for PU grafts. In more recent years, slightly longer
of investigators chose largely inadequate models for grafts were implanted (64% were still shorter than 7 cm)
assessing midgraft endothelialization although the clinical and observation periods became shorter (medians:
background to all these studies has been unambiguous 28.078.9 and 14.0718.7 for ePTFE and Dacron, respec-
throughout the years. It has been known for more than tively) [12–29]. In the small proportion of studies where
four decades that in humans, transanastomotic endothelial truly long prostheses were implanted, they were either
ingrowth does not exceed more than 1–2 cm [6] even after aorto–aortic [30–35], aorto–iliac [36–39] or aorto–femoral
years of implantation. At the same time peripheral bypass [40]. As much as these grafts allowed one to distinguish
grafts are usually between 40 and 60 cm long. Thus, TAE is transanastomotic from transmural endothelium, their
obviously an irrelevant mechanism for achieving confluent surrounding environment differed so profoundly from that
surface endothelialization on clinically implanted grafts. of clinically implanted grafts that deductions regarding
Yet, TAE was the very phenomenon that was inadvertently transmural endothelialization in humans need to be seen
studied in a majority of animal implants. By using animal with caution.
models that characteristically show rapid and extensive Against this background, it seems logical that a prime
TAE, ‘‘spontaneous’’ endothelialization was studied on paradigm for future vascular graft studies should be the
grafts that would never have endothelialized in humans. separation of TAE from other sources of endothelializa-
Moreover, by choosing particularly short graft lengths and tion. A prerequisite for this is a better insight into TAE.
long implantation periods, it does not come as a surprise In principle, there are four factors influencing
that the prostheses were frequently fully endothelialized at TAE: species, senescence, anatomical dimensions and graft
ARTICLE IN PRESS
P. Zilla et al. / Biomaterials 28 (2007) 5009–5027 5011

surfaces. Amongst these four main determinants, species is the peri-anastomotic region of prosthetic grafts in humans.
certainly most influential. The same ingrowth distance, for The simple reason for this lies in the fact that intimal
instance, for which humans need 56 weeks [6] is reached proliferation can only occur on the back of existing tissue
after approximately 4 weeks in young yellow baboons [41] and in man, most of the blood surface of a synthetic
and 3.5 weeks in dogs [33,34,42–44]. The second most vascular graft remains uncovered by tissue. This also
influential determinant of TAE is whether an animal is explains the clinical failure mode of prosthetic grafts: it is
juvenile or senescent. Even in the non-human primate model either midgraft thrombosis for a lack of endothelium or
senescence leads to a distinctly slower rate of endothelial intimal narrowing at the anastomosis, which may prompt
outgrowth [45] than infancy [39]. Senescence also needs to be occlusion. Aetiology and pathogenesis of this often-termed
considered in cross-species comparisons. Pigs [32], sheep ‘pannus’ tissue are multi-factorial, from surgical trauma to
[46–49] and calves [50], for instance, are usually large lack of endothelium, from compliance mismatch to flow
enough for vascular implants at young age but dog puppies conditions. Of these, flow conditions and shear stress seem
are too small. Therefore, dogs are usually senescent at to be overriding forces. However, as much as Alexander
implantation [51] whereas porcine, ovine and bovine models Clowes’ group [38,59] and others showed that variations in
tend to be juvenile. To a certain degree, senescence may even flow [60,61] and shear stress [61] result in an adaptive
outweigh species differences. After several months, the reduction or increase in the thickness of the subendothelial
senescent dog model covers a comparable ingrowth distance tissue, almost all the studies of the past four decades used
[33,34,42–44,52] to the juvenile yellow baboon [41,53]. To models where the shear stress conditions did not in the least
further complicate the issue of interpreting TAE, anatomical resemble those in clinical bypass grafts (Fig. 3). In animal
dimensions may make it challenging to compare results. It is models where distal arteriosclerotic narrowings are not
impossible, for example, to relate the transanastomotic present, the main determinant of shear stress in vascular
ingrowth distance of 2.2 mm after 6 weeks in the rat grafts is the cross-sectional quotient (Qc) between prosthe-
[41,54–56] and 1–2 mm (2 weeks) in the rabbit model [57] sis and run-off vessel. In a typical clinical small-graft
(ePTFE/30 mm internodal distance (IND)) to the 7.8– situation, Qc is estimated to lie somewhere in the range of
13.6 mm seen on the same grafts in large animal models 0.25–0.35 [62]. From Qc40.5 onwards (cross-sectional area
[11]. Last, but not least, the type of prosthetic graft of the graft less than twice that of the run-off artery), shear
implanted does seem to influence TAE. On low-porosity forces seem to become inhibitory for the development of
ePTFE, for example, TAE is twice as pronounced intimal hyperplasia [62]. Typically, almost all the vascular
[33,34,42–44] as on Dacron [6,58]. Nevertheless, if one puts graft studies in the past chose graft diameters that were
this ‘augmentation’ of TAE by certain graft surfaces into largely size matched or even smaller than the artery and
clinical perspective, the futility of TAE as a means of thus did not undercut the critical cross-sectional ratio from
achieving confluent graft endothelialization is highlighted which point intimal hyperplasia develops. For example, the
once again: even a doubling of TAE is a far cry from the estimated mean Qcs for interposition grafts ranged from
30-fold increase required under clinical conditions (Fig. 2). 0.63 to 42.0 in dogs [14–17,23,42,63–66], 0.56 to 42.0 in

3. Anastomotic intimal hyperplasia: clinical villain,


experimental void

Intimal hyperplasia occurs in vein grafts throughout


their length. In contrast, intimal hyperplasia is confined to

Fig. 2. Schematic comparison of graft lengths typically used for clinical


bypass grafts (60 cm, top) and experimental grafts in animals (in the Fig. 3. Representation of anastomotic pannus formation in clinical (left)
majority of studies o6 cm; bottom). In most animal models, transanas- and experimental grafts (right). Since anastomotic intimal hyperplasia
tomotic endothelial outgrowth (red) leads to complete surface endothe- (AIH; red) is augmented under low flow conditions, the typical clinical
lialization within weeks making it impossible to study transmural or scenario of a graft (blue) larger than the artery (yellow) leads to flow
blood-born endothelialization. In clinically implanted bypass grafts, in conditions that are conducive for AIH. In contrast, the overwhelming
contrast, ingrowth stoppage permanently leaves over 90% of the graft majority of experimental grafts have a smaller luminal diameter than the
non-endothelialized (blue). Therefore, animal studies generally investigate artery and thus experience increased flow. Therefore, experimental
a mode of surface endothelialization that is of utter irrelevance to the conditions are largely ignorant of a phenomenon that is behind a
human situation. significant proportion of clinical graft occlusions.
ARTICLE IN PRESS
5012 P. Zilla et al. / Biomaterials 28 (2007) 5009–5027

rats [14–17,23,42,63–66], 1.65 in pigs [32] and 0.77 to 1.00 intimal hyperplasia [38] only underlines the importance of
in baboons [45,67,68]. Assuming normal anatomy for each getting the flow-dynamics right. Considering the overall
species and weight group, the estimated average Qc for all design of the studies, however, it should not come as a
studies considered in this review (n ¼ 70) was as high as surprise that flow-restricting anastomotic intimal hyper-
1.5270.11(SEM) (Fig. 4). The resulting shear forces were plasia (AIH) hardly developed in any vascular graft study.
therefore approximately 1.8 times that of the artery (as This does not mean that intimal tissue did not reach across
opposed to 1/8th for a Qc of 0.25), far above the threshold the anastomosis or that other mechanisms triggering and
level below which intimal thickening occurs [69] (Fig. 5). augmenting AIH were not at work. One such mechanism is
Some of the studies even used aorto–aortic or aorto–iliac an almost template function of the initial fibrinous inner
end-to-end or end-to-side models, where an increased capsule. Two weeks after implantation for instance, the
inflow most likely produced even more pronounced thickness of the surface thrombus on ePTFE grafts
supra-normal flow conditions [30–34,36–40,70–75]. In measured about 18 to 14 of that observed on Dacron grafts
some studies, additional down-stream arterio–venous [33,34,76]. Similarly, the thickness of the ‘pannus’ tissue
fistulae were created to further increase the flow through replacing this thrombus in the anastomotic region not only
the graft [38]. The fact that even in these grafts relative measured 18 to 14 on ePTFE grafts compared with Dacron
differences in flow velocity did have an impact on diffuse prostheses [34,70] but also actually had the same thickness
as the pre-existing fibrin layer [34,60,77]. Another deter-
mining factor for the thickness of anastomotic pannus
tissue seems to be the existence or the lack of an
endothelium. Similar to post-PTCA- and stent-injuries,
the absence of a mature and confluent endothelium at the
time of subintimal tissue development seems to augment an
overshooting hyperplastic proliferation of poorly differ-
entiated connective tissue. As early as 1975, Mansfield et al.
[50] described the prevention of undifferentiated pannus
ingrowth in calves through pre-endothelialization of the
grafts. In the mid-1980s, clinical in vitro endothelialization
showed a similar phenomenon [45].

4. Graft-structure-related limitations to tissue regeneration

Initial attempts to replace arteries with solid tubes of


Fig. 4. Relative vessel (yellow) and graft (blue) diameters used in clinical
synthetic material [78] soon made it clear that porosity is a
applications and in experimental animal models. The cross-sectional
quotient (Qc) between run-off artery and graft is typically in the range of prerequisite for graft patency [79,80]. Therefore, ranking
Qc0.25–0.35 in clinical small-diameter grafts (left) as opposed by structural porosity above material properties emerged as
Qc1.4–1.6 in the vast majority of experimentally implanted grafts (right). the early creed of synthetic vascular graft research. To date,
this dogma is still unchallenged. Efforts to determine
porosity requirements for graft healing, however, are
complicated by material specific characteristics, structural
uniqueness and the patho-physiological microenvironment.
The complex three-dimensional structure of contemporary
grafts results in a wide range of voids between the synthetic
components, hardly allowing a precise definition of
ingrowth spaces. For manufacturers, a way out of this
dilemma was the introduction of water permeability as a
means of defining porosity. For example, materials such as
Dacron were not only defined on the basis of being
knitted or woven, but also further characterized as being
of high porosity (1500–4000 ml/cm2/min) or low porosity
(200–1000 ml/cm2/min) [81]. The same principle applied to
ePTFE on the basis of distances between the nodes of the
expanded material. However, although these rough para-
meters served as useful means to characterize the materials,
Fig. 5. Schematic representation of longitudinal cross-sections of end-to- they were only of limited value in determining structural
end (top) and end-to-side (bottom) anastomoses showing shear force
differences resulting from the typical diameter mismatch between graft
limitations of healing. Researchers were long puzzled as to
(blue) and artery (yellow) under clinical circumstances (left) and in why interstitial graft spaces of particular prostheses were
experimental conditions (right). sufficient to allow macrophages and other inflammatory
ARTICLE IN PRESS
P. Zilla et al. / Biomaterials 28 (2007) 5009–5027 5013

cells to infiltrate, whereas connective tissue cells of similar


dimensions were often conspicuously absent [34,82–84].
Today, we understand that certain of these connective
tissue cells, like smooth muscle cells, may follow ingrowing
endothelial cells [85,86]. Therefore, one important deter-
mining dimension for complete graft healing is that which
permits capillaries to sprout into the interstices of the
prosthesis and still leave some space for accompanying
cells. Since the average diameter of a capillary is 10 mm
[87–89], it would seem as if the minimum area required for
ingrowth of capillaries is at least 20–80 mm2. The diameter
of a functional arteriole, i.e. an endothelium and at least
one layer of smooth muscle cells, is approximately
23.1713.1 mm [11]. However, with future regenerative
vascular prostheses in mind, space orientation needs to
be additionally considered beyond mere dimensional
concerns. As the different orientation of smooth muscle Fig. 6. Scanning electron micrograph of a cross-section through a high-
cells in arteries and veins shows, biomechanical forces porosity ePTFE vascular graft with 150 mm internodal distance (IND). It
determine smooth muscle orientation. The only geometric is obvious that internodal and interfibrillar distances determine ingrowth
spaces. The non-elastic fibrils allow only a limited degree of stretching,
structure, which allows free orientation in response to a making disproportional increases in IND necessary to achieve moderate
particular biomechanical environment, is the open dode- increases in ingrowth dimensions. Furthermore, available ingrowth
cahedron of a foam [90]. channels diminish in size with increasing penetration depth due to the
random obstruction by adjacent fibrils.
4.1. ePTFE grafts

ePTFE grafts consist of circumferentially aligned, thin


and irregular-shaped solid membranes (the so-called
‘‘nodes’’) and a dense meshwork of fine fibrils stretching
between the nodes. Although the porosity of ePTFE grafts
is traditionally defined on the basis of the distance between
the nodes (IND), the actual dimension of voids within the
prosthetic wall is defined by the spaces between the fibrils.
Thus, the available ingrowth spaces in ePTFE grafts are
manifold smaller than the IND suggests. Although the
process of expanding Teflon material to the typical node-
and-fibre structure of ePTFE grafts allows for a wide range
of INDs, the resulting changes in ingrowth spaces are
moderate. With well-defined anchor points and a rigid
material like Teflon, the dispensability of the fibrils is
limited, thus making disproportional increases in IND
necessary to achieve moderate increases of ingrowth
dimensions of a few micrometers (Fig. 6). Apart from
differences in fibril length between equally defined grafts of Fig. 7. Demonstration of the dependence of ingrowing capillaries on
available ingrowth spaces. Dual-layer ePTFE graft with high-porosity
different diameters (e.g. 24.3 versus 20.7 mm for 6 and outer two thirds and low-porosity inner third. (Femo–femoral interposi-
10 mm grafts), there are distinct differences in the IND of tion graft; 4 mm ID; 4 weeks; senescent Chacma Baboon; CD 31).
‘30 mm’ ePTFE grafts between different manufacturers (e.g.
17.875.6 and 23.876.2 mm for two market leaders,
respectively) [11]. Furthermore, theoretically available, IND, in contrast, transmural capillarization would be
continual transmural ingrowth spaces in ‘low-porosity’ permitted by 100% of the transmural spaces and even some
(30 mm IND) grafts diminish in size over the first ten layers arterioles would be able to reach the blood surface [11].
of fibrils out of an overall number of approximately 2000 Even before these calculations were available, a similar cut-
layers/wall thickness [11]. This is a result of the random off point of porosity emerged empirically over the years
obstruction of interfibrillar spaces of one layer by fibrils of below which tissue ingrowth is impermissible (Fig. 7). In
the next. The available channel size would therefore not spite of manufacturer-related differences between contem-
accommodate capillaries from as early as layer 200 porary ePTFE grafts, this cut-off point was identified to lie
onwards, the equivalent of the outer 1/10th of wall within a relatively narrow range somewhere between 30
thickness. In ‘high- porosity’ ePTFE grafts of 60 mm and 45 mm of IND. Therefore, it became customary to refer
ARTICLE IN PRESS
5014 P. Zilla et al. / Biomaterials 28 (2007) 5009–5027

to grafts with an IND of less than or equal to 30 mm as low degree of variability in this microporosity. Other techni-
porosity ePTFE grafts and to those of 45 mm or more as ques to achieve porosity include foam flotation [100] and
high- porosity ePTFE grafts. This distinction is of course dip-coating [96] (laminar structures with ill-defined and
invalid in grafts where the manufacturer adds an impene- poorly interconnected voids), gas expansion [101] (closed
trable external wrap. porosity), and laser perforation [102] (radial channels
through a solid tube rather than three-dimensional pores).
4.2. Dacron grafts Once the focus shifted from material choice, surface
structure and ‘water porosity’ to transmural tissue regen-
In contrast to the two distinctly different structures of eration, continuity in void dimensions became a desirable
ePTFE grafts (nodes and fibrils), Dacron prostheses consist goal. Two methods that did result in open-porosity foam
of one basic element, namely polyester fibres. However, structures capable of allowing tissue ingrowth were
since the fibres are mostly bundled into multifilament yarns replamineform [103] (where porous sea-urchin spikes were
that are either woven or knitted into a fabric, there are in used as a template to produce pores ranging from 18 to
effect again two different structures determining the shape 180 mm), and reticulation (opening of closed-pored foams
and the dimensions of voids. Generally speaking, weaving to yield open porosity) [104]. The first method is limited by
results in narrow spaces for tissue ingrowth whereas issues pertaining to supply, machining and the porosity of
knitting leads to looser patterns containing larger voids. naturally occurring structures, while the second involved
Woven material is often preferred by surgeons, as it is the rolling of reticulated sheets into tubular structures;
known for its high bursting strengths, low permeability to both methods were not further pursued. The most obvious
liquids and minimal tendency to deform under stress. The practical way of achieving ingrowth-permissive porosity
almost complete impermeability of woven Dacron is was by superimposing macropores onto the microporous
opposed by a range of limited ingrowth spaces in knitted structure obtained by phase inversion. The addition and
Dacron. Still, in the lower range of water permeability subsequent extraction of porogens to the phase inversion
capillaries are also space-restricted to the outer 23 of the wall technique was suggested [105–107] leading to pores of
and spaces are generally too narrow for arterioles [11]. 30–75 [105], 40 [106,108] and 200 mm [109] as well as pore
size gradients (10–100 mm) [110]. However, the use of salts
4.3. PU grafts and other irregular porogens, together with the pre-mixing
of the porogen and the polymer solution, generally resulted
The enthusiasm with which PUs were initially adopted as in irregular-shaped pores, limiting connectivity between
vascular graft materials was based on two main advantages pores while trapping residual porogens in the structure
over polyester and Teflon: elasticity and ease of handling. [111]. In order to overcome these limitations, a variation of
Porosity was an integral part of PU grafts from the the phase inversion/porogen extraction method has been
beginning, but was mainly understood as a principle of developed, in which a PU solution is infiltrated into a
patency rather than tissue ingrowth. As a result of this, tightly prepacked column of spherical porogens. Subse-
defined inter-connectivity of the pores was largely absent in quent precipitation and extraction yields a well-defined,
both fibrillar and foamy PU grafts. Although the pore size open porosity comprising uniformly sized dodecahedrons
was somehow changeable in either of the two structures it with even and equally well-defined interconnections
was not defined around ingrowth dimensions. In fibrillar [90,112]. Such an ‘open-cell’ foam represents the ideal
designs porosity was varied through fibre thickness, wind- ingrowth scaffold if alignment along stress and strain needs
ing angle, regularity of the deposition and extent of [4] or nutrient and oxygen gradients is desired because it is
coagulation of fibres before solidification. With fibre the only structure which allows cell and tissue forma-
thickness varying between 1 and 30 mm [72,91], for tions to freely orientate themselves in any given direction
instance, pore sizes could be achieved between 10 and (Fig. 8).
60 mm [72,91–93]. However, regardless of whether fibrillar
grafts were produced through weaving [94,95], knitting 5. Midgraft tissue response: no healing, no regeneration
[94], electrostatic spinning [91] or winding processes [72,73]
the presence of communicating transmural spaces permis- Even after prolonged periods of implantation, a persis-
sible for blood vessel ingrowth was coincidental rather than tent foreign body response dominates the interstices of
intentional. permanent (non-resorbable) synthetic arterial prostheses.
Similar to fibrillar grafts, foamed prostheses mostly Simultaneously, thrombotic appositions build up on the
lacked spaces that reproducibly allowed blood vessels to luminal surface. Thus, healing—defined as the end point of
reach through the entire wall thickness. Phase inversion a pacified repair process—does not occur. Moreover, since
alone, for example, led to micropores that were generally not even traces of vascular tissue are being formed in the
poorly interconnected and smaller than 15 mm [92,96–98] interstices or on the surface of these grafts, regeneration
therefore not allowing more than a shallow, superficial remains permanently absent, too. Yet, in spite of many
penetration by capillaries. Differences between various millions of clinical implants during the course of the past
phase/temperature inversion techniques [99] allowed some decades the degree of misconception about the presumed
ARTICLE IN PRESS
P. Zilla et al. / Biomaterials 28 (2007) 5009–5027 5015

Fig. 8. Comparitive schematic representation of available ingrowth spaces in ePTFE, Dacron and well-defined PU scaffolds. While both ePTFE and
Dacron grafts have structures that predetermine ingrowth directions and pose some degree of spatial limitation, an open dodecahedron structure allows
cells to freely orientate themselves in any given direction (Insert: single cell of polyurethane graft with well-defined, spherical 157 mm spaces and equally
well-defined interconnecting windows of 76 mm). In a graft concept that includes compliance, the cell orientation follows biomechanical strain
requirements. This principle is not only realized in the different alignment-angle of smooth muscle cells (SMCs) in arteries and veins, but was also
demonstrated in tissue engineering constructs where SMC were embedded in a gel and where strain resulted in the artery-like orientation of the cells (upper
right corner) [4].

degree of ‘healing’ of vascular grafts amongst surgeons is prostheses from outside [54,82,114–116]. Even after 6
often surprising. months, however, this connective tissue ingrowth has
hardly increased at all, remaining mostly limited to the
5.1. Low-porosity ePTFE grafts (o30 mm IND) outer part of the graft wall [41,63,64,83,84,117] with only
scanty capillaries [117,118]. In wrapped grafts, the ePTFE
There is hardly any difference in the midgraft response of membrane reinforcing the prosthesis acts as an effective
low-porosity ePTFE grafts between most recipient models. barrier against cellular infiltration. These grafts show an
At the blood surface, a thin, 10–20 mm thick fibrin layer almost acellular fibrin matrix extending from the centre of
develops during the initial 2 weeks of implantation [33,34]. the prosthetic wall all the way to the outside surface while
Over the subsequent weeks the thickness of this layer stays eliciting a much stronger foreign body giant cell reaction
constant [113] but its composition becomes more diverse on the outside than unwrapped grafts.
ranging from loose [45,113] or compacted acellular fibrin
[114] to platelet carpets with interspersed granulocytes 5.2. High-porosity ePTFE grafts (445 mm IND)
[33,45]. Even after prolonged observation periods, midgraft
regions lack all forms of cellular coverage in low-porosity In contrast to low-porosity ePTFE grafts, there is a
ePTFE [5,6,36,57]. distinct difference between the healing responses of various
Similarly, there is no transmural tissue ingrowth. During animal models in high-porosity ePTFE grafts particularly
the initial 2 weeks of implantation, most of the interstices with regards to the timing of events.
of the prosthetic wall are devoid of any recognizable Initially, a thin fibrin or amorphous protein layer
cellular material [33]. Only a few macrophages begin to covering a significant proportion of the graft coexists next
migrate into the micropores—almost exclusively from to exposed areas, where the nodal structure of the ePTFE
outside [54]. After 3 weeks we find a typical triple layered remains visible. Over time, the fibrin attracts white blood
picture with macrophages invading from the blood—and cells and platelets [119] on its surface. In the depth, it
to a much larger extent—the adventitial surface into an continues to be widely acellular except for a mild rim of
otherwise almost acellular central fibrin matrix. Eventually, macrophages lining the ePTFE surface. Underneath these
after 6 weeks of implantation, connective tissue cells begin macrophages, the fibrin filling the interstices of the ePTFE
to sporadically invade the interstices of non-wrapped is initially also acellular but over time gets progressively
ARTICLE IN PRESS
5016 P. Zilla et al. / Biomaterials 28 (2007) 5009–5027

popularized with macrophages and polymorphnuclear surface, this is accomplished in the 2nd week of implanta-
leukocytes [120,121] particularly in the outer third of the tion in the juvenile yellow baboon. In the latter, patches of
graft wall . In more senescent animal models like the adult endothelial cells [68] and capillary orifices—approximately
dog or chacma baboon this stage persists into the 4th to 6th 100–500 mm apart [59]—appear on the blood surface and
week of implantation [119] as opposed to less than 2 weeks confluence is often reached before the 3rd or 4th week of
in juvenile models such as the yellow baboon [121]. At that implantation [59,68,70,121–123]. Soon thereafter a layer of
time, sprouting capillaries begin to reach into the outer one actin positive cells develops underneath the endothelium
third to one half of the graft wall [119,120]. Although [59,68,70,121–123]. These cells—which exhibit the ultra-
fibroblasts from the peri-graft region soon follow the structural characteristics of arterial smooth muscle cells
capillaries [55], the majority of the interstices remain [68]—only begin to appear and proliferate after the
dominated by macrophages [121], while foreign body giant endothelium has formed, corresponding with the sequence
cells (FBGCs) are conspicuously absent. Distinctly differ- of events in angiogenesis, where endothelial-derived PDGF
ent from the widely dilated, sinus-like endothelial cell chemotactically attracts pericytes to follow and proliferate
formations found in Dacron grafts (see below), however, [124]. The actual discovery of PDGF–A mRNA in the
the microvessels soon resemble mature arterioles. They are endothelium of high-porosity ePTFE grafts [121] increases
small in diameter (23.06713.11 mm) [11] and often the likelihood of this scenario. It would also explain why
accompanied by at least one layer of smooth muscle cells the density of actin-positive cells is highest close to the
(Fig. 9). In contrast to senescent animal models where it endothelium [121]. PDGF is not only one of the strongest
may take months for these microvessels to reach the blood growth factors for smooth muscle cells but also is known to
be secreted by endothelial cells into the abluminal basal
compartment [125]. It’s strong presence in the endothelium
covering high-porosity ePTFE grafts may also explain the
5–100 times higher proliferative activity of SMCs and
10–100 times higher mitotic activity of endothelial cells
[126] in grafts than in normal arteries. Although it remains
open whether this is an adaptive response or a primary
phenomenon, the manifold thicker neo-intima found on
high-porosity ePTFE grafts exposed to low flow conditions
as compared to high flow conditions [127] points at least
partly to an adaptive process.

5.3. Woven Dacron grafts

Immediately after implantation, a thin layer of fibrin,


erythrocytes, white blood cells and platelets is deposited on
the blood surface of the prosthesis [10,58]. During the first
few hours to days this thrombus layer slowly starts
thickening until it reaches a stable equilibrium of com-
pacted fibrin [128]. In dogs this equilibrium stage is reached
within 6 months [128] whereas in humans it is only
observed after 18 months. In the depth of the graft, fibrin
also instantly fills the narrow interstices of the graft wall
and the space between the graft and the surrounding tissue
[129]. At the interface between this fibrous capsule and the
Dacron yarns, a variable degree of FBGCs begins to build
up [129]. Subsequently, a few capillaries and fibroblasts
start growing into the tight interstitial spaces of the woven
Dacron [58,129]. This process is quite variable in its extent
and time course. In dogs it can be seen as early as 2–3
weeks after implantation [129], whereas in humans it may
be observed after 5 months [58] or not at all [10,130,131].
Exceptionally, the tissue reaches the inner fibrin capsule
and begins to organize the basis of the fibrin capsule in a
Fig. 9. High-porosity (90 mm IND) ePTFE grafts implanted into the
few scanty areas of the graft [130]. In the rare event that
femoral arteries of senescent Chacma baboons (4 weeks). A significant
number of factor-VIII-positive vessels (A) is supported by an accompany- such tissue forms, it was found to occur after 4–8 weeks in
ing layer of actin-positive SMC (B), indicating that continual ingrowth dogs [58,128] and after 18 months in humans [130]. In even
spaces of more than 23 mm are available. fewer cases, these tissue areas replace the inner fibrin
ARTICLE IN PRESS
P. Zilla et al. / Biomaterials 28 (2007) 5009–5027 5017

capsule in its entire thickness and form a pseudo-neointima prostheses, [65,126] comparable to that in highly porous
[58,130,131]. Both the pseudo-neointima and the com- ePTFE grafts, is seen. However, in contrast to high-
pacted fibrin usually reach an equilibrium at a thickness of porosity ePTFE grafts it still remains unclear whether this
approximately 400–500 mm [130]. In humans these rare and endothelium is derived from transmural tissue ingrowth,
localized neointimal spots extend over less than 1 mm and facilitated transanastomotic ingrowth or fall-out healing
are collagen rich, with increasing cellularity towards the [144,145]. The latter is a phenomenon predominantly
graft surface [10]. If these islands of connective tissue are observed in Dacron grafts where—apparently independent
rare, areas that are covered by an endothelium are even from transmural tissue ingrowth and with some delay—
rarer. Usually no endothelial cells are found at all within endothelial islands emerge on the surface of the compacted
the first 1–2 years [58,130,131]. After prolonged implanta- fibrin. These islands are well-separated from the transmural
tion periods of up to 11 years—which are naturally tissue ingrowth by a distinct layer of compacted fibrin
restricted to humans—small islands of endothelial cells [6,7,76,131]. They either rest directly on the fibrin [6,131,
may occasionally appear [130,131]. These endothelial 140] or on a few layers of isolated actin-positive cells that
islands always rest on the compacted fibrin rather than have no other tissue connection [131,142]. Such endothelial
the scanty islands of connective tissue. islands can be observed in dogs after 2–6 months [140] and
In summary, tissue ingrowth is very limited in woven in humans perhaps in every 4th graft after 1–11 years
Dacron grafts, seldom penetrating the entire wall thickness. [6,131,142]. Nevertheless, although sporadic endotheliali-
Most importantly, even if it extends to the inner capsule, it zation occasionally occurs in limited areas of clinically
hardly manages to break through the compacted fibrin to implanted grafts, the majority of the surface remains
reach the blood surface [132]—even after decades of covered by fibrin only.
implantation. If one focuses on the transmural tissue ingrowth itself, it
also begins with the infiltration of the surrounding outer
5.4. Knitted Dacron grafts fibrin matrix by granulation tissue [7,34]. Initially it is
primarily macrophages that begin to invade the interstitial
In knitted Dacron grafts, the initial surface meshwork of fibrin [115], popularizing the entire depth of the graft at the
fibrin, white blood cells, erythrocytes and platelets [7] end of the 1st month of implantation [76]. Early giant cell
increases to a thickness of 100–120 mm in the course of the formation surrounding the yarns commences after 7 days.
1st week of implantation [76]. During the subsequent 5 During the following months a distinct foreign body
weeks, the inner fibrin capsule has not reached an reaction occurs predominantly in the outer half of the
equilibrium yet holding a wide range of thicknesses: graft wall [66,133]. Between this outer part of the prosthetic
100–500 mm in humans [76], 30-210 mm in dogs [34,66,77, wall that is packed with FBGC, and the surrounding
133–135] and 290-300 mm in the yellow baboon [67,70]. connective tissue capsule with its longitudinally aligned
After a further insignificant increase thereafter [66,76,133], collagen bundles [76,131,146], hugely dilated capillary
the thickness of the internal capsule eventually stabilizes sinuses are usually present during the early weeks of
before the 6th month of implantation [66,133]. Although a implantation. These sinuses typically consist of a single
superficial screening of the literature suggests that it is only endothelial cell layer that contains no smooth muscle cells,
a matter of time until this entire layer of compacted fibrin but does contain macrophages and FBGC that tightly
gets organized by ingrowing tissue, deeper analysis makes snuggle against the basal lamina from outside. The few
it likely that this may well be a misinterpretation based on capillaries that sometimes extend through the entire graft
the confusion of midgraft healing with anastomotic wall and reach the inner capsule also remain free of smooth
ingrowth. It is true that the vast majority of studies describe muscle cells.
a mature endothelium and a well-developed intimal smooth In summary, the porosity of knitted and to some extent
muscle layer, resting directly on the graft surface [136–138]. even woven Dacron prostheses eventually allows a certain
However, the graft length and implantation period in these degree of tissue ingrowth from outside. At the same time,
studies make it almost certain that this mature ‘‘neointi- however, it seems again that the compacted inner fibrin
mal’’ tissue represents nothing but extended anastomotic capsule on the blood surface of Dacron grafts represents an
ingrowth [34,70,82,122,137–139]. This suspicion is con- ingrowth barrier for transmural connective tissue regard-
firmed by descriptions of true midgraft regions in particu- less of whether the grafts are knitted or woven. Although
larly long grafts in which the inner fibrin capsule appears to this barrier may eventually be overcome by ingrowth of
be both impenetrable for the ingrowing interstitial graft undifferentiated fibrous tissue, capillaries remain uncon-
tissue [34,70,82,122,137–139] and non-endothelialized nected to the often poorly endothelialized blood surface
[76,140,141]. This lack of surface healing occurs [132]. Moreover, it is also interesting that the fibrin
[6,7,76,131,132,140,142] although transmural tissue reaches coverage on Dacron grafts seems to have a higher
the compacted fibrin of the blood surface within 3–4 weeks propensity to capture blood borne cells than that on
in the calf [76], the yellow baboon [70] and the dog [34,143] ePTFE grafts [145]. In this context it is noteworthy that the
and 3–6 months in humans [7,76]. Very occasionally, resulting isolated endothelial islands often rest on equally
though, even complete endothelialization of very long isolated smooth muscle layers [6,131]. However, in spite of
ARTICLE IN PRESS
5018 P. Zilla et al. / Biomaterials 28 (2007) 5009–5027

the scientific excitement arising from this observation, the determine that interdependence of inflammation and
rare and late occurrence of the phenomenon make it rather vascularization was less pronounced than assumed, [90]
irrelevant in the healing process of contemporary Dacron as well as a reproducible chronology of the tissue
grafts. responses. Most importantly, the inflammatory response
and the FBGC index could be reduced by 56% and 21%,
5.5. Fibrillar PU grafts respectively, when the pore size was increased from 66 to
157 mm, while vascularization did not diminish. The fast
With very rare exceptions fibrillar PU grafts are largely and massive early vessel penetration of this well-defined
impenetrable for transmural tissue ingrowth [91,147,148]. graft structure [112] from outside (Fig. 10) highlighted the
Only particularly large inter-fibrillar spaces allow complete basic dilemma of transmural endothelialization: while
fibro-connective tissue penetration throughout the wall blood vessels rush into the graft wall, their crossing of a
thickness [49,73,93]. At the same time, events on the blood distance of as short as 0.6 mm towards the blood surface
surface resemble more those on woven Dacron than on was distinctly protracted in spite of widely open traverse-
ePTFE. Particularly, anastomotic pannus tissue tends to be spaces. The build-up of an adverse microenvironment was
thicker on PU [149] than on expanded Teflon with capillary best studied in these foam grafts when femorally implanted
orifices opening onto the anastomotic endothelium in senescent Chacma baboons. Immediately upon implan-
[148,150]. From day one the midgraft region is continu- tation, the entire graft wall was filled with a loose fibrin
ously covered by a fibrin layer [151] hardly diminishing in matrix allowing sprouting capillaries from outside to reach
thickness during the course of the 1st year [71,72,91, as deep as half into the graft wall after 7 days. After 2
147,148,150–152]. The fibrin layer itself is not as mighty as weeks, transmural vessel ingrowth had hardly progressed
on knitted Dacron, some times giving a ‘ghost impression’
of the underlying prosthetic fibres. The outside capsule and
scar formation is usually thin and dense [147,148,150] with
occasional FBGC wrapping around PU fibres [73]. At 2
weeks the FBGC become more prevalent increasing until
week 6 [153]. The inflammatory reaction does not further
increase between week 6 and 12 [153].

5.6. Foamy PU grafts

In microporous PU foams with less than 15 mm pore size,


there is relatively little ingrowth, even over an extended
period of time [108,109,154–156]. Below this cut-off point,
only plasma-like insudation with some erythrocytes and a
few clusters of white blood cells are found in the depth of
the wall [47,155]. Macrophages, giant cells and a few
fibroblasts may penetrate a short distance into the implant
[156] if the pore size is borderline. The foreign body giant
cell response along the outer graft interface tends to get
augmented until week 6–9 [153]. From an interconnecting
pore size of 25–40 mm onwards macrophages, giant cells,
fibroblasts and capillaries rapidly populate the wall during
the initial 3–4 weeks [153,156,157]. The degree of ingrowth
is initially more pronounced in larger porosity, but
equalizes for all porosities after 3 months [156]. Compar-
able to fibrillar PU grafts, the blood surface also shows a
significantly more pronounced fibrin layer than on ePTFE
[47,158]. Similarly, anastomotic pannus tissue is strongly
developed with capillary openings to the blood surface.
This anastomotic intimal hyperplasia was distinctly miti-
gated when compliant PU grafts were compared with non-
compliant ones [159]. Due to a fast build-up of collagen
within the pores of cell-populated PU grafts [74,157], Fig. 10. Corrosion casts of PU grafts showing vessel ingrowth into well-
defined pores. The vigorous vessel-penetration of the graft from outside
however, compliance decreases by 50% over the initial 6
demonstrates how attracted capillaries initially are by the scaffold and
weeks [159]. Going beyond ingrowth permissible porosity how increasingly adverse the micro-environment of a vascular graft must
by introducing well-defined, equally sized spherical pores become to make the crossing of the remaining 0.6 mm distance to the
and large, equally well defined interconnectivity, we could blood surface such a protracted and challenging affair.
ARTICLE IN PRESS
P. Zilla et al. / Biomaterials 28 (2007) 5009–5027 5019

further towards the blood surface, which showed early content of fibrinogen. Since TGF-b is also liberated from
signs of a dense fibrin layer. Another week later, the a-granules and its incorporation into fibrin matrices is
capillaries had reached the inner third of the wall, but the proportional, a platelet-augmented, fibrinogen-rich fibrin
fibrin layer had become even denser and ‘squeezed’ into matrix is rich in TGF-b. High concentrations of TGF-b are
the pores. While it took 4–6 weeks for a few capillaries to known to be inhibitory for angiogenesis [161]. Since it is
reach the surface, open up through a microorifice and known that the fibrin structure influences cell growth, the
begin to proliferate on top of a very dense fibrinous surface main question is whether high concentrations of fibrinogen
layer, the entire distance of 0.6 mm was rapidly crossed if and other a-granule and dense granule release products
grafts occluded in the first few days. such as calcium affect the structure and/or composition of
fibrin in a way that it becomes hostile to tissue ingrowth.
6. Biological events: adversity rather than facilitation Although none of the investigators has yet analysed the
composition of fibrin on vascular grafts, there is evidence
The complexity of an unabated chronic foreign body that the slow growing fibrin layer on blood surfaces does
reaction at the blood–tissue interface is certainly beyond indeed differ from rapidly generated fibrin found in wound
the scope of a short review. Yet, the two seemingly trivial injuries or acute vessel occlusion (Fig. 11). It is recognized
main components of the tissue incorporation response— that an increase in fibrinogen concentration results in both
fibrin deposition and macrophage infiltration—are suffi-
ciently contentious to offer themselves as potential keys to
understanding main mechanisms behind the mitigated
tissue incorporation of prosthetic vascular grafts. Almost
from the time of implantation, fibrin becomes an integral
part of prosthetic vascular grafts both in the interstices and
on the blood surface. Fibrin is known to be an ideal
attachment and ingrowth matrix for endothelial cells. Yet,
circulating endothelial and progenitor cells hardly populate
the blood surface of grafts, making ‘fall out healing’ a late
and very scarce event. Similarly, transmurally ingrowing
capillaries hardly ever manage to fully cross a stretch of less
than 1 mm of a fibrin-containing graft wall, even if
ingrowth spaces are permissible and grafts were implanted
for years. Similarly, macrophages—which persist in their
presence from the earliest time of implantation—have been
the subject of much speculation with the spectrum ranging
from being quintessential for vascularization to being the
main culprits behind ingrowth inhibition.

6.1. The potential role of fibrin

The phenomenon of a seemingly impenetrable inner


fibrin capsule has been reported for more than 20 years in
porous Dacron grafts [6,7,51,131,132,140,142]. Likewise,
the ‘rubbery’ inner fibrin layer in aortic aneurysms [160] is
equally cell-poor to acellular, as are intra-atrial clots. It
seems reasonable to consider the possibility that ingrowth
inhibition may be linked to the particular nature of fibrin
forming on a blood surface of synthetic implants. In
contrast to fibrin plugs forming in wound spaces or Fig. 11. Schematic presentation of the hypothetical build-up of an adverse
occluding arteries, the continual apposition of fibrin on ingrowth environment over time: The fibrin clot initially filling the
interstices of the graft and the surgical wound on the adventitial side of the
the blood surface of vascular grafts occurs in an environ-
graft is formed on the basis of the limited fibrinogen supply present at the
ment of undiminishing supply. Ongoing platelet activation time of implantation and the surgical trauma. Typically, lower fibrinogen
provides high concentrations of a-granule products like concentrations lead to thicker and fewer fibrin strands that strongly
platelet factor 3, fibrinogen and von Willebrand factor as stimulate angiogenesis and cell infiltration. In contrast, the continual
well as dense granule contents such as calcium. At the same replenishment of clotting factors and calcium from the blood surface
create conditions where thinner and much denser fibres, which are
time, graft surfaces are continually rinsed by blood with its
inhibitory for vessel ingrowth, are being formed, Furthermore, continual
undiminishing concentrations of fibrinogen and other platelet replenishment and degranulation makes this fibrin rich in TGF-b,
clotting factors. Thus, the fibrin generated on the inner which at high release-concentrations is additionally inhibitory to capillary
surface of vascular prostheses has a particularly high formation.
ARTICLE IN PRESS
5020 P. Zilla et al. / Biomaterials 28 (2007) 5009–5027

a decrease in fibre bundle thickness and an increase in fibre the release of the FGFs. This suggests a paracrine
bundle density [162]. While a fibrin matrix strongly mechanism in which hypoxia stimulates macrophages to
stimulates angiogenesis at lower fibre density [163], it release mitogenic factors for hypoxically preconditioned
inhibits capillary formation at higher fibre density. endothelial cells thereby promoting angiogenesis. Addi-
A similar inverse correlation between cell growth and fibre tionally, TNF-a, which is also upregulated in macrophages
density was found on an ultrastructural level [162]. The by hypoxia [172], is known to play a role in macrophage-
higher the density of fibrin fibres and the lower the average mediated angiogenesis presumably through the activation
thickness of these fibre bundles was, the lower was its and recruitment of further monocytes [173]. The role of
penetrability for cells [162]. The better healing of graft TNF-a in angiogenesis, however, is controversial and
scaffolds in subcutaneous than vascular position seems to seems to depend on the biological context [173]. These
support this explanation. effects are further potentiated by upregulation of MIP-1
during oxygen deprivation [172], a potent macrophage
6.2. Potential role of macrophages chemoattractant. Therefore, macrophages residing in the
outer half of synthetic grafts together with VEGF-releasing
In attempting to answer the often recurring question of infiltrating neutrophils [174] are likely to continue to
whether macrophages are friends or foes in prosthetic graft contribute to a pro-angiogenic milieu at this stage. The
healing, it seems as if they are both: friends at the beginning low concentration of TGF-b released with migration-
and foes at the end. Their active secretion of cytokines acts associated degradation of the adventitial ‘wound’ fibrin
as a key chemotactic and mitogenic factor for capillary and should not be inhibitory for endothelial cell proliferation
connective tissue ingrowth in the early phases of wound [161]. However, once the dense inner fibrin capsule is
healing. Their continual presence, however, increasingly reached, released TGF-b concentrations would most likely
derails the delicate chronology of cytokines required for be inhibitory.
the successful accomplishment of healing. During the As the organization of the macrophages within the graft
initial infiltration of the fresh fibrin matrix, macrophage wall changes and transmural endothelial cell ingrowth
migration is facilitated by a low level of basic uPA activity. begins, the situation where the outer wall of the graft is
This is sufficient to liberate and activate platelet-derived dominated by single macrophages also slowly begins to
TGF-b from the fibrin [164]. Since TGF-b upregulates the change by the increasing appearance of FBGCs [41,54,77].
expression of integrins on the surface of the blood These may persist for the life-time of the implant. Giant
monocytes thereby promoting adhesion and migration cell formation depends on the material surface as well as
[165], it acts as a strong chemotactic agent to recruit structure. It has been shown that relatively hydrophilic
macrophages to the inflammatory site [165]. Moreover, it material surfaces induce more fusion than relatively
also upregulates the uPA activity of macrophages [166] to hydrophobic ones [175–178]. Theoretical analyses show
further facilitate migration. Macrophages are also capable that the time-dependent foreign body giant cell formation
of binding and internalizing fibrin through the Mac-1 is also influenced by the initial density of adherent
receptor, which augments fibrinolysis by a plasmin- macrophages [179]. Therefore, any surface characteristic
independent mechanism [167]. The degradation products that initially influences the number of adherent cells will
of fibrinolysis in return also act as chemoattractants eventually also influence the extent of giant cell formation.
recruiting still more macrophages into the implant site The most conspicuous deviation from this sequence is the
[168]. In addition, routing of monocytes to inflammatory absence of FBGCs inside the wall structure of ePTFE
sites also involves the superfamily of chemoattractant grafts compared to Dacron and PU, in spite of a distinct
cytokines (chemokines ) and their receptors [169]. Specifi- presence of macrophages. The absence of giant cells on the
cally the C–C chemokine MCP-1 has been demonstrated to inner nodal surfaces which are subdivided by numerous
be expressed by macrophages residing at the implant site as fibrils indicates that giant cell formation may be suppressed
early as 48 h post-implantation [170]. This sequence of by geometrical space limitations. However, given (i) the
events illustrates how this initial phase of the so-called in vitro evidence that fewer monocytes adhere to ePTFE,
healing is dominated by self-augmentation of macrophage (ii) that ePTFE is less inflammatory than Dacron and PU
chemotaxis. Until macrophages actually contact the [180–183], and (iii) that a more hydrophobic surface
material surface, the recruitment mode into the wound decreases the propensity of single macrophages to fuse
fibrin of the prosthetic graft does not deviate from that [175,177–179], the inflammatory potential of the surface
taking place in normal wound healing. As the macrophage may also play a role in giant cell formation insofar as a
population migrates into the depth of the graft structure it more chemically inert, more hydrophobic ePTFE surface
must eventually overstep the maximum distance a cell can reduces fusion. Not only do mononuclear cells gradually
migrate away from a capillary before getting mildly fuse to form more giant cells in a material-dependent
hypoxic. Macrophages exposed to hypoxia synthesize and manner but further complexity is added by the changing
release increased amounts of PDGF as well as aFGF and nature of the mononuclear macrophage population over
bFGF [171]. Medium conditioned by hypoxic macrophages time. Macrophages found in intimate contact with
in culture is mitogenic for hypoxic endothelial cells through synthetic materials in situ were ED1 and MHCII positive,
ARTICLE IN PRESS
P. Zilla et al. / Biomaterials 28 (2007) 5009–5027 5021

whereas ED2 positive macrophages were observed lying that inflammation associated with normal wound healing is
behind the ED1 positive cells [184]. This indicates a pattern resolved because of the transient nature of the macrophage
of distribution of not only mono versus multinuclear response, it would appear that the persistence of the
macrophages, but also immature (ED1 positive) versus inflammatory stimulus in the form of a non-degradable
mature (ED2 positive) and activated (MHCII positive) biomaterial confounds even the best intentions of these
versus non-activated macrophages within the implant site. cells to mediate healing.
Theoretical analysis shows that up to 5 weeks of What is becoming increasingly obvious is that a
implantation, FBGCs are in fact formed from the fusion macrophage is not always a macrophage, and that the
of a small percentage of the adherent macrophages which contribution of each of these subsets of macrophage
are already present on the third day of implantation populations to the lack of healing of the vascular graft
[175,179]. This raises the issue of the fusogenic potential of remains to be investigated. Overall, the macrophage
different subsets of cells. Expression of cell surface remains the Jekyll and Hyde of graft healing but as our
receptors may be a prerequisite. For example, inhibitors insight into the mechanisms of healing grows so does our
of mannose receptor activity have been shown to inhibit understanding of which critical questions remain to be
IL-4-induced fusion in vitro suggesting that this receptor answered in the elucidation of the role of this enigmatic
may play a role in giant cell formation [185]. cell.
If it is difficult to decide whether macrophages are
beneficial for graft healing or not, it is even more difficult 7. Half a century of vascular graft implantation: the essence
to decide whether the presence of FBGCs represents a first
step towards pacification or a particularly detrimental The key to new and emerging concepts for replacement
variant of chronic inflammation. One indication for the arteries lies in understanding the obstacles of contemporary
FBGC’s role as a mediator of pacification of inflammation vascular prostheses. By analyzing studies of the past
arises from the fact that their fusion can be induced by IL-4 half century, certain principles emerged which confirm or
[186,187]. IL-4 is secreted by Th2 lymphocytes and is defy prevailing stereotypes—some of them more than
known to downregulate several macrophage inflammatory others:
functions [188]. In this context it has been proposed that
giant cells may represent an attempt to ‘‘mop-up’’ the  The vast majority of contemporary synthetic vascular
deleterious effect of a persistent chronic foreign body grafts are so impervious that transmural tissue ingrowth
reaction. In contrast, a detrimental role for the giant cell is is impossible.
suggested by the fact that material surface cracking occurs  None of the clinically used ePTFE, Dacron or PU
directly beneath adherent FBGCs, implicating the secre- prostheses spontaneously develop a neo-intima, except
tory products of these cells in the degradation of for sporadically observed small islands of endothelium.
biosynthetics [189]. It is also peculiar that the accumulation  The overwhelming majority of animal experiments
of giant cells in Dacron regularly coincides with the studied TAE, a rather irrelevant mechanism for the
development of huge convolutes of capillary sinuses in clinical situation.
the vicinity. These widely dilated and irregularly shaped  The cross-sectional quotients of the vast majority of
endothelial sacks are principally devoid of any second cell experimentally implanted grafts exceeded the clinical
layer. In situ hybridization of human explants, however, situation by more than a factor 3, thereby seriously
has demonstrated that interstitial macrophages and distorting both the shear stress-linked potential for
FBGCs are still secreting IL-1b even after numbers of developing significant anastomotic intimal hyperplasia
years [190]. Although IL-1b is known to stimulate both as well as endothelial cell function and metabolism.
smooth muscle cell proliferation and fibronectin deposition  ‘‘Fall-out healing’’ leading to the formation of isolated
in atherosclerotic plaque formation [191], it has been endothelial islands with no connection to either transa-
suggested that this interleukin can also inhibit vascular nastomotic or transmural tissue formations occurs
smooth muscle proliferation in an autocrine fashion more often on Dacron than ePTFE grafts. All together,
through nitric oxide upregulation [192]. The giant cells, however, this is a scanty and late-occurring phenomen-
therefore, may be responsible for not only affecting on that does not play any significant role in contem-
capillarization but also inhibiting smooth muscle cell porary vascular prostheses.
ingrowth through persistent IL-1b secretion.  If interstitial spaces are ingrowth-permissible, dimensions
In summary, the true character of the FBGC remains AND structures seem to determine the tissue response.
elusive. On the one hand, their persistence at the implant While minimal dimensions are required to allow tissue
site, their destructive role in biodegradation as well as their ingrowth in all, shape and sub-structure of the voids
possible role in the mitigation of vessel formation seems to appear to significantly affect the type of healing reaction.
label them as villains of incomplete graft healing. Yet, the  In senescent animal models and humans vessel,
evidence supporting a role for a typically anti-inflamma- ingrowth shows progressive transmural inhibition. In
tory cytokine-IL-4 in stimulating giant cell formation is contrast, juvenile models often allow complete trans-
fairly convincing both in vitro and in vivo [186,187]. Given mural endothelialization, suggesting that a competitive
ARTICLE IN PRESS
5022 P. Zilla et al. / Biomaterials 28 (2007) 5009–5027

situation between the alacrity of capillary ingrowth and


the build-up of an adverse biological environment exists.
 The inner fibrinous capsule—particularly pronounced on
Dacron prostheses—seems to be the main barrier against
tissue ingrowth. Although both connective tissue and
capillaries reach this capsule from underneath, they
hardly ever connect to the blood surface, even after
years of implantation.
 The mere fact that transmural endothelialization did
occur in some juvenile animal models refutes the Fig. 12. Schematic representation of an isolation graft showing artery
possibility of explaining ingrowth inhibition with a (yellow), isolation segments (blue), and test segment (red). In order to
avoid the confusion of transanastomotic with transmural endothelializa-
reverse oxygen gradient from outside to inside.
tion, experimental grafts are isolated from the adjacent artery through
low-porosity isolation segments (ePTFE o30 mm IND). The length of the
isolation segments can be adjusted to the transanastomotic outgrowth
8. Conclusions for vascular graft research situation of the respective animal model as well as the planned
implantation period.
Many of the experimental concepts for the vascular
grafts of tomorrow promise to overcome the main A prerequisite for this is a profound knowledge of the
obstacles plaguing those prostheses currently available for TAE rates of a particular model.
clinical implantation. At the same time, resolving one  The build-up of adverse ingrowth conditions does not
problem may introduce another. By eliminating the occur in the presence of an endothelium. Therefore,
underlying cause of the chronic foreign body response by rapidly endothelializing animal models pre-empt this
using resorbable materials, for example, the resulting scar build-up and thus fail to mimic a clinically relevant
tissue may deprive us of that very compliance which we situation. Senescent animal models are therefore prefer-
tried to introduce through new elastomers. Similarly, by able to juvenile ones.
implanting dense, tissue-engineered tubes of smooth
muscle cells, we may deprive ourselves of the possibility In summary, without a serious focus on developing new
of allowing transmural endothelialization—should the animal models which address the real issues which have
engineered endothelium undergo the initial detachment plagued half a century of clinical graft implantation, l’art
often seen in veingrafts. However, as much as all these new pour l’art laboratory research will continue to chase virtual
problems will need to be addressed in a dynamic way as rather than the real problems that led to the failure of
they come up, the overall questions asked and models used millions of bypass grafts in patients.
need to be of relevance. Moreover, understanding where
we were misled in the past will be as helpful as under- References
standing the obstacles which made half a century of SMG
research so frustrating and unrewarding. [1] Weinberg C, Bell E. A blood vessel model constructed from collagen
and cultured vascular cells. Science 1986;231:397–400.
 The main issue seems to be the need to clearly separate [2] Niklason L. Techview: medical technology. Replacement arteries
trans-anastomotic from transmural or blood-borne made to order. Science 1999;286(5444):1493–4.
[3] Lheureux N, Germain L, Labbe R, Auger F. Construction of a
endothelialization. Isolation models, where an experi- human blood vessel from cultured vascular cells; a morphologic
mental graft was separated from the artery by an study. J Vasc Surg 1993;17:499–509.
interposition segment, have been suggested before [16]. [4] Hirai J, Matsuda T. Self-organized, tubular hybrid vascular tissue
The externally sealed segments, however, would add composed of vascular cells and collagen for low-pressure-loaded
venous system. Cell Transplant 1995;4(6):597–608.
to the already high occlusion rate of small diameter
[5] Crombez M, Chevallier P, Gaudreault RC, Petitclerc E, Mantovani
graft. In our hands the use of standard low-porosity D, Laroche G. Improving arterial prosthesis neo-endothelialization:
ePTFE instead sufficiently prevented transmural in- application of a proactive VEGF construct onto PTFE surfaces.
growth without worsening overall thrombogenicity [112] Biomaterials 2005;26(35):7402–9.
(Fig. 12). [6] Berger K, Sauvage L, Rao A, Wood S. Healing of arterial prostheses
 In as much as a smaller graft diameter means a higher in man: its incompleteness. Ann Surg 1972;175(1):118–27.
[7] Wesolowski S, Fries C, Gennigar G, Fox L, Sawyer P, Sauvage L.
occlusion rate, vascular graft studies need to choose Factors contributing to long-term failures in human vascular
animal models that emulate flow conditions of clinical prosthetic grafts. Cardiovas Surg 1964;38:544–67.
bypass grafts. [8] Takebayashi J, Kamatani M, Namba S, Katagami Y, Hayashi K.
 The surrounding environment needs to be representative Early reparative process of implanted arterial prosthesis: a
for the clinical situation of SMGs. In order to avoid morphological and hematological study. J Surg Res 1974;17(2):
102–10.
abdominal implantations because of the need for too [9] Ghidoni J, Liotta D, Hall C, Adams J, Lechter A, Barrionueva M,
long grafts, the length of the ‘isolation segments’ will et al. Healing of pseudointimas in velour-lined, impermeable arterial
need to be optimized for each model and each question. prostheses. Am J Pathol 1968:375–83.
ARTICLE IN PRESS
P. Zilla et al. / Biomaterials 28 (2007) 5009–5027 5023

[10] Szilagyi D, Smith R, Elliott J, Allen H. Long-term behavior of a ePTFE grafts effectively prevents pseudointima and intimal
Dacron arterial substitute: clinical, roentgenologic and histologic hyperplasia development. Eur J Vasc Endovasc Surg 2006;32(4):
correlations. Ann Surg 1965;162(3):453–77. 418–24.
[11] Davids L, Dower T, Zilla P. The lack of healing in convenitonal [29] Sun LB, Utoh J, Moriyama S, Tagami H, Okamoto K, Kitamura N.
vascular grafts. In: Zilla P, Greisler H, editors. Tissue engineering of Pretreatment of a Dacron graft with tissue factor pathway inhibitor
prosthetic vascular grafts. Austin: R G Landes; 1999. p. 3–44. decreases thrombogenicity and neointimal thickness: a preliminary
[12] Lin P, Chen C, Bush R, Yao Q, Lumsden A, Hanson S. Small- animal study. ASAIO J 2001;47(4):325–8.
caliber heparin-coated ePTFE grafts reduce platelet deposition and [30] Sterpetti A, Hunter W, Schultz R, Farina C. Healing of high-
neointimal hyperplasia in a baboon model. J Vasc Surg 2004;39(6): porosity polytetrafluoroethylene arterial grafts is influenced by the
1322–8. nature of the surrounding tissue. Surgery 1992;111(6):677–82.
[13] Murray-Wijelath J, Lyman DJ, Wijelath ES. Vascular graft healing. [31] Kusaba A, Fischer CD, Matulewski T, Matsumoto T. Experimental
III. FTIR analysis of ePTFE graft samples from implanted bigrafts. study of the influence of porosity on development of neointima in
J Biomed Mater Res B Appl Biomater 2004;70(2):223–32. Gore-Tex grafts: a method to increase long-term patency rate. Am
[14] Pfeiffer T, Kever M, Grabitz K, Reiher L, Muller KM, Hildebrand Surg 1981;47(8):347–54.
A, et al. Healing characteristics of small-calibre vascular prostheses [32] Hollier L, Fowl R, Pennell R, Heck C, Winter K, Fass D, et al. Are
coated with plasmin-treated fibrin—an experimental study. Vasa seeded endothelial cells the origin of neointima on prosthetic
2000;29(2):117–24. vascular grafts? J Vasc Surg 1986;3(1):65–73.
[15] Nishibe T, Okuda Y, Kumada T, Tanabe T, Yasuda K. Enhanced [33] Graham L, Burkel W, Ford J, Vinter D, Kahn R, Stanley J.
graft healing of high-porosity expanded polytetrafluoroethylene Expanded polytetrafluoroethylene vascular prostheses seeded with
grafts by covalent bonding of fibronectin. Surg Today 2000;30(5): enzymatically derived and cultured canine endothelial cells. Surgery
426–31. 1982;91(5):550–9.
[16] Bhattacharya V, McSweeney PA, Shi Q, Bruno B, Ishida A, Nash [34] Herring M, Baughman S, Glover J, Kesler K, Jesseph J, Campbell J,
R, et al. Enhanced endothelialization and microvessel formation in et al. Endothelial seeding of Dacron and polytetrafluoro-
polyester grafts seeded with CD34(+) bone marrow cells. Blood ethylene grafts: the cellular events of healing. Surgery 1984;96(4):
2000;95(2):581–5. 745–55.
[17] Lyman DJ, Murray-Wijelath J, Ambrad-Chalela E, Wijelath ES. [35] Whitehouse WJ, Wakefield T, Vinter D, Ford J, Swanson D, Thrall
Vascular graft healing. II. FTIR analysis of polyester graft samples J, et al. Indium-111-oxine labeled platelet imaging of endothelial
from implanted bi-grafts. J Biomed Mater Res 2001;58(3):221–37. seeded dacron thoracoabdominal vascular prostheses in a canine
[18] Kuzuya A, Matsushita M, Oda K, Kobayashi M, Nishikimi N, model. Trans Am Soc Artif Intern Organs 1983;29:183–7.
Sakurai T, et al. Healing of implanted expanded polytetrafluor- [36] Golden M, Hanson S, Kirkman T, Schneider P, Clowes A. Healing
oethylene vascular access grafts with different INDs: a histologic of polytetrafluoroethylene arterial grafts is influenced by graft
study in dogs. Eur J Vasc Endovasc Surg 2004;28(4):404–9. porosity. J Vasc Surg 1990;11(6):838–44 [discussion 845].
[19] Shimada T, Nishibe T, Miura H, Hazama K, Kato H, Kudo F, et al. [37] Hussain S, Glover J, Augelli N, Bendick P, Maupin D, McKain M.
Improved healing of small-caliber, long-fibril expanded polytetra- Host response to autologous endothelial seeding. J Vasc Surg
fluoroethylene vascular grafts by covalent bonding of fibronectin. 1989;9(5):656–63 [discussion 663-654].
Surg Today 2004;34(12):1025–30. [38] Geary R, Kohler T, Vergel S, Kirkman T, Clowes A. Time course of
[20] Ueberrueck T, Meyer L, Zippel R, Nestler G, Wahlers T, Gastinger flow-induced smooth muscle cell proliferation and intimal thicken-
I. Healing characteristics of a new silver-coated, gelatine impreg- ing in endothelialized baboon vascular grafts. Circ Res 1994;74(1):
nated vascular prosthesis in the porcine model. Zentralbl Chir 14–23.
2005;130(1):71–6. [39] Clowes A, Kirkman T, Reidy M. Mechanisms of arterial graft
[21] Kenagy RD, Fischer JW, Lara S, Sandy JD, Clowes AW, Wight healing. Rapid transmural capillary ingrowth provides a source of
TN. Accumulation and loss of extracellular matrix during shear intimal endothelium and smooth muscle in porous PTFE prostheses.
stress-mediated intimal growth and regression in baboon vascular Am J Pathol 1986;123(2):220–30.
grafts. J Histochem Cytochem 2005;53(1):131–40. [40] Zamora J, Navarro L, Ives C, Weilbaecher D, Gao Z, Noon G.
[22] Phaneuf MD, Dempsey DJ, Bide MJ, Quist WC, LoGerfo FW. Seeding of arteriovenous prostheses with homologous endothelium.
Coating of Dacron vascular grafts with an ionic polyurethane: a A preliminary report. J Vasc Surg 1986;3(6):860–6.
novel sealant with protein binding properties. England, 2001. [41] Clowes A, Gown A, Hanson S, Reidy M. Mechanisms of arterial
[23] Miura H, Nishibe T, Yasuda K, Shimada T, Hazama K, Katoh H, graft failure. 1. Role of cellular proliferation in early healing of
et al. The influence of node-fibril morphology on healing of high- PTFE prostheses. Am J Pathol 1985;118(1):43–54.
porosity expanded polytetrafluoroethylene grafts. Eur Surg Res [42] Hanel K, McCabe C, Abbott W, Fallon J, Megerman J. Current
2002;34(3):224–31. PTFE grafts: a biomechanical, scanning electron, and light
[24] Begovac PC, Thomson RC, Fisher JL, Hughson A, Gallhagen A. microscopic evaluation. Ann Surg 1982;195(4):456–63.
Improvements in GORE-TEX vascular graft performance by [43] Koveker G, Burkel W, Graham L, Wakefield T, Stanley J.
Carmeda BioActive surface heparin immobilization. England, 2003. Endothelial cell seeding of expanded polytetrafluoroethylene vena
[25] Hazama K, Nishibe T, Shimada T, Miura H, Watanabe S, Kondoh cava conduits: effects on luminal production of prostacyclin, platelet
S, et al. Effects of omental wrap on performance of small-caliber adherence, and fibrinogen accumulation. J Vasc Surg 1988;7(4):
high-porosity expanded polytetrafluoroethylene grafts. J Surg Res 600–5.
1999;81(2):174–80. [44] Ombrellaro M, Stevens S, Kerstetter K, Freeman M, Goldman M.
[26] Lin PH, Bush RL, Yao Q, Lumsden AB, Chen C. Evaluation of Healing characteristics of intra-arterial stented grafts: effect of intra-
platelet deposition and neointimal hyperplasia of heparin-coated luminal position on prosthetic graft healing. Surgery 1996;120(1):
small-caliber ePTFE grafts in a canine femoral artery bypass model. 60–70.
J Surg Res 2004;118(1):45–52. [45] Zilla P, Preiss P, Groscurth P, Rosemeier F, Deutsch M, Odell J,
[27] Cagiannos C, Abul-Khoudoud OR, DeRijk W, Shell DHt, Jennings et al. In vitro-lined endothelium: initial integrity and ultrastructural
LK, Tolley EA, et al. Rapamycin-coated expanded polytetrafluor- events. Surgery 1994;116(3):524–34.
oethylene bypass grafts exhibit decreased anastomotic neointimal [46] Christenson J, Thulesius O, Owunwanne A, Nazzal M. Forskolin
hyperplasia in a porcine model. J Vasc Surg 2005;42(5):980–8. impregnation of small calibre PTFE grafts lowers early platelet graft
[28] Heise M, Schmidmaier G, Husmann I, Heidenhain C, Schmidt J, sequestration and improves patency in a sheep model. Eur J Vasc
Neuhaus P, et al. PEG-hirudin/iloprost coating of small diameter Surg 1991;5(3):271–5.
ARTICLE IN PRESS
5024 P. Zilla et al. / Biomaterials 28 (2007) 5009–5027

[47] Aldenhoff Y, van DVF, ter WJ, Habets J, Poole - WL, Koole L. [67] Shepard A, Eldrup - JJ, Keough E, Foxall T, Ramberg K, Connolly
Performance of a polyurethane vascular prosthesis carrying a R, et al. Endothelial cell seeding of small-caliber synthetic grafts in
dipyridamole (Persantin) coating on its lumenal surface. J Biomed the baboon. Surgery 1986;99(3):318–26.
Mater Res 2001;54(2):224–33. [68] Clowes A, Kirkman T, Clowes M. Mechanisms of arterial graft
[48] Fleser P, Nuthakki V, Malinzak L, Callahan R, Seymour M, failure. II. Chronic endothelial and smooth muscle cell proliferation
Reynolds M, et al. Nitric oxide-releasing biopolymers inhibit in healing polytetrafluoroethylene prostheses. J Vasc Surg 1986;3(6):
thrombus formation in a sheep model of arterio-venous bridge 877–84.
grafts. J Vasc Surg 2004;40(4):803–11. [69] Sho E, Nanjo H, Sho M, Kobayashi M, Komatsu M, Kawamura K,
[49] Poole - WL, Schindhelm K, Graham A, Slowiaczek P, Noble K. et al. Arterial enlargement, tortuosity, and intimal thickening in
Performance of small diameter synthetic vascular prostheses with response to sequential exposure to high and low wall shear stress.
confluent autologous endothelial cell linings. J Biomed Mater Res J Vasc Surg 2004;39(3):601–12.
1996;30(2):221–9. [70] Zacharias R, Kirkman T, Clowes A. Mechanisms of healing in
[50] Mansfield P, Wechezak A, Sauvage L. Preventing thrombus on synthetic grafts. J Vasc Surg 1987;6(5):429–36.
artificial vascular surfaces: true endothelial cell linings. Trans Am [71] Therrien M, Guidoin R, Adnot A, Paynter R. Hydrophobic and
Soc Artif Intern Organs 1975;21:264–72. fibrillar microporous polyetherurethane urea prosthesis: an ESCA
[51] DeBakey M, Jordan GJ, Beall A, O’Neal R, Abbott J, Halpert B. study on the internal and external surfaces of explanted grafts.
Basic biologic reactions to vascular grafts and prostheses. Surg Clin Biomaterials 1989;10(8):517–20.
North Am 1965;45:477. [72] Leidner J, Wong E. A novel process for the manufacturing of
[52] Herring M, Gardner A, Peigh P, Madison D, Baughman S, Brown porous grafts: Process description and product evaluation. J Biomed
J, et al. Patency in canine inferior vena cava grafting: effects of graft Mater Res 1983;17:229–47.
material, size, and endothelial seeding. J Vasc Surg 1984;1(6): [73] Wilson GJ, MacGregor DC, Klement P, Lee JM, del Nido PJ,
877–87. Wong EW, et al. Anisotropic polyurethane nonwoven conduits: a
[53] Keough E, Callow A, Connolly R, Weinberg K, Aalberg J, O DTJ. new approach to the design of a vascular prosthesis. Trans Am Soc
Healing pattern of small caliber Dacron grafts in the baboon: an Artif Intern Organs 1983;29:260–8.
animal model for the study of vascular prostheses. J Biomed Mater [74] Seifalian AM, Salacinski HJ, Tiwari A, Edwards A, Bowald S,
Res 1984;18(3):281–92. Hamilton G. In vivo biostability of a poly(carbonate–urea)urethane
[54] van der Lei B, Wildevuur C. Microvascular polytetrafluoroethylene graft. Biomaterials 2003;24(14):2549–57.
prostheses: the cellular events of healing and prostacyclin produc- [75] Berkowitz HD, Perloff LJ, Roberts B. Pseudointimal development
tion. Plast Reconstr Surg 1988;81(5):735–41. on microporous polyurethane lattices. Trans Am Soc Artif Intern
[55] Stronck J, van dLB, Wildevuur C. J Thorac Cardiovasc Surg Organs 1972;18(0):25–9.
1992;103(1):146–52. [76] Sauvage L, Berger K, Wood S, Nakagawa Y, Mansfield P. An
[56] van der Lei B, Wildevuur C. Improved healing of microvascular external velour surface for porous arterial prostheses. Surgery 1971;
PTFE prostheses by induction of a clot layer: an experimental study 70(6):940–53.
in rats. Plast Reconstr Surg 1989;84(6):960–8. [77] Margolin D, Kaufman B, DeLuca D, Fox P, Graham L. Increased
[57] Friedman E, Hamilton A. Polytetrafluoroethylene grafts in the platelet-derived growth factor production and intimal thickening
peripheral venous circulation of rabbits. Am J Surg 1983;146(3):355–9. during healing of Dacron grafts in a canine model. J Vasc Surg 1993;
[58] Stewart G, Essa N, Chang K, Reichle F. A scanning and 17(5):858–66 [discussion 866-857].
transmission electron microscope study of the luminal coating on [78] Haimovici H. History of vascular surgery. In: Haimovici H, editor.
Dacron prostheses in the canine thoracic aorta. J Lab Clin Med Vascular surgery principles and techniques. 2nd ed. Norwark,
1975;85(2):208–26. Connecticut: Appleton-Century-Crofts; 1984. p. 3–18.
[59] Kohler T, Kirkman T, Kraiss L, Zierler B, Clowes A. Increased [79] Voorhees A, Jaretzki A, Blakemore A. The use of tubes constructed
blood flow inhibits neointimal hyperplasia in endothelialized from vinyon ‘‘N’’ cloth in bridging arterial defects. A preliminary
vascular grafts. Circ Res 1991;69(6):1557–65. report. Ann Surg 1952:332–6.
[60] Binns R, Ku D, Stewart M, Ansley J, Coyle K. Optimal graft [80] Wesolowski S. Foundations of modern vascular grafts. In: Sawyer
diameter: effect of wall shear stress on vascular healing. J Vasc Surg P, Kaplitt M, editors. Vascular grafts. New York: Appleton Century
1989;10(3):326–37. Croft; 1978.
[61] Kraiss L, Geary R, Mattsson E, Vergel S, Au Y, Clowes A. Acute [81] Tabbara M, White R. Biologic and prosthetic materials for vascular
reductions in blood flow and shear stress induce platelet-derived conduits. In: Veith F, Hobson R, Williams R, Wilson S, editors.
growth factor-a expression in baboon prosthetic grafts. Circ Res Vascular surgery: principles and practice. USA: McGraw-Hill, Inc;
1996;79(1):45–53. 1994. p. 523–35.
[62] Zilla P, Human P, Wolf M, Lichtenberg W, Raffie N, Bezuidenhout [82] Mathisen S, Wu H, Sauvage L, Usui Y, Walker M. An experimental
D, et al. Clinically relevant downsizing of veingrafts through study of eight current arterial prostheses. J Vasc Surg 1986;4(1):
external nitinol stenting in non-human primates. 2007. Submitted 33–41.
for publication [83] Campbell C, Goldfarb D, Roe R. A small arterial substitute:
[63] Florian A, Cohn L, Dammin G, Collins JJ. Small vessel replacement expanded microporous polytetrafluoroethylene: patency versus
with gore-tex (expanded polytetrafluoroethylene). Arch Surg 1976; porosity. Ann Surg 1975;182(2):138–43.
111(3):267–70. [84] Bellon J, Bujan J, Contreras L, Hernando A, Jurado F. Simi-
[64] Matsumoto H, Hasegawa T, Fuse K, Yamamoto M, Saigusa M. A larity in behavior of polytetrafluoroethylene (ePTFE) prostheses im-
new vascular prosthesis for a small caliber artery. Surgery 1973; planted into different interfaces. J Biomed Mater Res 1996;31(1):
74(4):519–23. 1–9.
[65] Shi Q, Wu M, Hayashida N, Wechezak A, Clowes A, Sauvage L. [85] Hirschi K, Rohovsky S, D AP. Cell–cell interactions in vessel
Proof of fallout endothelialization of impervious Dacron grafts in assembly: a model for the fundamentals of vascular remodelling.
the aorta and inferior vena cava of the dog. J Vasc Surg 1994; Transpl Immunol 1997;5(3):177–8.
20(4):546–56 [discussion 556-547]. [86] Beck LJ, D AP. Vascular development: cellular and molecular
[66] Baitella - EG, Groscurth P, Zilla P, Lachat M, Muller - GW, regulation. FASEB J 1997;11(5):365–73.
Schneider J, et al. Long-term results of tissue development and cell [87] Cheresh D, Berliner S, Vicente V, Ruggeri Z. Recognition of distinct
differentiation on Dacron prostheses seeded with microvascular cells adhesive sites on fibrinogen by related integrins on platelets and
in dogs. J Vasc Surg 1993;18(6):1019–28. endothelial cells. Cell 1989;58(5):945–53.
ARTICLE IN PRESS
P. Zilla et al. / Biomaterials 28 (2007) 5009–5027 5025

[88] Dvorak H, Nagy J, Berse B, Brown L, Yeo K, Yeo T, et al. Vascular [109] Williams SK, Carter T, Park PK, Rose DG, Schneider T, Jarrell BE.
permeability factor, fibrin, and the pathogenesis of tumor stroma Formation of a multilayer cellular lining on a polyurethane vascular
formation. Ann N Y Acad Sci 1992;667:101–11. graft following endothelial cell sodding. J Biomed Mater Res
[89] Gamble J, Matthias L, Meyer G, Kaur P, Russ G, Faull R, et al. 1992;26(1):103–17.
Regulation of in vitro capillary tube formation by anti-integrin [110] van der Lei B, Wildevuur C, Dijk F, Blaauw E, Molenaar I,
antibodies. J Cell Biol 1993;121(4):931–43. Nieuwenhuis P. Sequential studies of arterial wall regeneration in
[90] Bezuidenhout D, Davies N, Zilla P. Effect of well defined microporous, compliant, biodegradable small-caliber vascular grafts
dodecahedral porosity on inflammation and angiogenesis. ASAIO in rats. J Thorac Cardiovasc Surg 1987;93(5):695–707.
J 2002;48(5):465–71. [111] Nam Y, Park T. Porous biodegradable polymeric scaffolds prepared
[91] Annis D, Bornat A, Edwards RO, Higham A, Loveday B, Wilson J. by thermally induced phase separation. J Biomed Mater Res 1999;
An elastomeric vascular prosthesis. Trans Am Soc Artif Intern 47(1):8–17.
Organs 1978;24:209–14. [112] Bezuidenhout D, Vici M, Davies N, Schilling L, Fittkau M, Zilla P.
[92] Eberhart A, Zhang Z, Guidoin R, Laroche G, Guay L, De LFD, Small diameter vascular grafts: a Meaningful model at last. Nineth
et al. A new generation of polyurethane vascular prostheses: rara biennial meeting of the ISACB: Savannah, GA; 2004.
avis or ignis fatuus? J Biomed Mater Res 1999;48(4):546–58. [113] Boyd K, Schmidt S, Pippert T, Sharp W. Endothelial cell seeding of
[93] Wilson GJ, MacGregor DC, Klement P, Dereume JP, Weber BA, ULTI carbon-coated small-diameter PTFE vascular grafts. ASAIO
Binnington AG, et al. The composite Corethane/Dacron vascular Trans 1987;33(3):631–5.
prosthesis, canine in vivo evaluation of, 4, mm diameter grafts with [114] Sterpetti A, Lepidi S, Cucino A, Patrizi A, Palumbo R, Taranta A,
1 year follow-up. ASAIO Trans 1991;37(3, M475–476). et al. Growth factor production after polytetrafluoroethylene and
[94] Wintermantel E, Mayer J, Blum J, Eckert K, Luscher P, Mathey M. vein arterial grafting: an experimental study. J Vasc Surg 1996;
Tissue engineering scaffolds using superstructures. Biomaterials 23(3):452–60.
1996;17(2):83–91. [115] Bartels H, van dLB, Robinson P. Contrib Title: Prosthetic
[95] Gupta B, Kasyanov V. Biomechanics of human common carotid microvenous grafting in the rat femoral vein. Lab Anim 1993;27(1):
artery and design of novel hybrid textile compliant vascular grafts. J 47–54.
Biomed Mater Res 1997;34(3):341–9. [116] Van der Lei B, Stronck J, Wildevuur C. Enhanced healing of 30
[96] Hayashi K, Takamizawa K, Saito T, Kira K, Hiramatsu K, Kondo microns Gore-Tex PTFE microarterial prostheses by alcohol-
K. Elastic properties and strength of a novel small-diameter, pretreatment. Br J Plast Surg 1991;44(6):428–33.
compliant polyurethane vascular graft. J Biomed Mater Res [117] Kenney D, Tu R, Peterson R. Evaluation of compliant and
1989;23(A2 Suppl):229–44. noncompliant PTFE vascular prostheses. ASAIO Trans 1988;34(3):
[97] Chen J, Laiw R, Jiang S, Lee Y. Microporous segmented 661–3.
polyetherurethane vascular graft: I. Dependency of graft morphol- [118] Herring M, Gardner A, Glover J. Endothelial seeding on vascular
ogy and mechanical properties on compositions and fabrication prostheses. Arch Surg 1979;114:679.
conditions. J Biomed Mater Res 1999;48(3):235–45. [119] Douville E, Kempczinski R, Birinyi L, Ramalanjaona G. Impact of
[98] Martz H, Beaudoin G, Paynter R, King M, Marceau D, Guidoin R. endothelial cell seeding on long-term patency and subendothelial
Physicochemical characterization of a hydrophilic microporous proliferation in a small-caliber highly porous polytetrafluoroethy-
polyurethane vascular graft. J Biomed Mater Res 1987;21(3): lene graft. J Vasc Surg 1987;5(4):544–50.
399–412. [120] Bull D, Hunter G, Holubec H, Aguirre M, Rappaport W, Putnam
[99] Zdrahala R. Small caliber vascular grafts. Part II: polyurethanes C. Cellular origin and rate of endothelial cell coverage of PTFE
revisited. J Biomater Appl 1996;11(1):37–61. grafts. J Surg Res 1995;58(1):58–68.
[100] Kowligi RR, von Maltzahn WW, Eberhart RC. Synthetic vascular [121] Golden M, Au Y, Kirkman T, Wilcox J, Raines E, Ross R, et al.
graft fabrication by a precipitation-flotation method. ASAIO Trans Platelet-derived growth factor activity and mRNA expression in
1988;34(3):800–4. healing vascular grafts in baboons. Association in vivo of platelet-
[101] Kowligi RR, von Maltzahn WW, Eberhart RC. Fabrication and derived growth factor mRNA and protein with cellular prolifera-
characterization of small-diameter vascular prostheses. J Biomed tion. J Clin Invest 1991;87(2):406–14.
Mater Res 1988;22(3 Suppl):245–56. [122] Clowes A, Zacharias R, Kirkman T. Early endothelial coverage of
[102] Doi K, Nakayama Y, Matsuda T. Novel compliant and tissue- synthetic arterial grafts: porosity revisited. Am J Surg 1987;153(5):
permeable microporous polyurethane vascular prosthesis fabricated 501–4.
using an excimer laser ablation technique. J Biomed Mater Res [123] Lado M, Knighton D, Cavallini M, Fiegel V, Murray C, Phillips G.
1996;31(1):27–33. Induction of neointima formation by platelet derived angiogenesis
[103] Hiratzka LF, Goeken JA, White RA, Wright CB. In vivo fraction in a small diameter, wide pore, PTFE graft. Int J Artif
comparison of replamineform, silastic, and bioelectric polyurethane Organs 1992;15(12):727–36.
arterial grafts. Arch Surg 1979;114(6):698–702. [124] Hirschi KK, Rohovsky SA, D’Amore PA. PDGF, TGF-beta, and
[104] Berkowitz H, Perloff L, Roberts B. Pseudointimal development on heterotypic cell–cell interactions mediate endothelial cell-induced
microporous polyurethane lattices. Surgery 1972;72(6):888–96. recruitment of 10T1/2 cells and their differentiation to a smooth
[105] Hinrichs WL, Kuit J, Feil H, Wildevuur CR, Feijen J. In vivo muscle fate. J Cell Biol 1998;141(3):805–14.
fragmentation of microporous polyurethane- and copolyesterether [125] Koyama N, Hart C, Clowes A. Different functions of the platelet-
elastomer-based vascular prostheses. Biomaterials 1992;13(9):585–93. derived growth factor-alpha and -beta receptors for the migration
[106] Murabayashi S, Kambic H, Harasaki H, Morimoto T, Yozu R, and proliferation of cultured baboon smooth muscle cells. Circ Res
Nose Y. Fabrication and long-term implantation of semi-compliant 1994;75(4):682–91.
small vascular prosthesis. Trans Am Soc Artif Intern Organs [126] Noishiki Y. Pattern of arrangement of smooth muscle cells in
1985;31:50–4. neointimae of synthetic vascular prostheses. J Thorac Cardiovasc
[107] Uchida N, Kambic H, Emoto H, Chen JF, Hsu S, Murabayshi S, Surg 1978;75(6):894–901.
et al. Compliance effects on small diameter polyurethane graft [127] Bellon J, Bujan J, Hernando A, Honduvilla N, Jurado F. Arterial
patency. J Biomed Mater Res 1993;27(10):1269–79. autografts and PTFE vascular microprostheses: similarities in the
[108] Kambic H, Murabayashi S, Yozu R, Morimoto T, Furuse M, healing process. Eur J Vasc Surg 1994;8(6):694–702.
Harasaki H, et al. Small vessel replacement with elastomeric protein [128] Harrison H. Synthetic materials as vascular prostheses. Ia. A
composite materials: preliminary studies. Trans Am Soc Artif Intern comparative study in small vessels of Nylon, Dacron, Orlon, Ivalon
Organs 1984;30:406–10. sponge and Teflon. Amer J Surg 1958;95:3–15.
ARTICLE IN PRESS
5026 P. Zilla et al. / Biomaterials 28 (2007) 5009–5027

[129] Weselow A. Biological behaviour of tissue and prosthetic grafts. In: [149] Walpoth B, Rogulenko R, Tikhvinskaia E, Gogolewski S, Schaffner
Haimovici H, editor. Vascular surgery: principles and techniques. T, Hess O, et al. Improvement of patency rate in heparin-coated
New York: Appleton-Century-Crofts; 1984. p. 93–118. small synthetic vascular grafts. Circulation 1998;98(19 Sup-
[130] Wu M, Shi Q, Wechezak A, Clowes A, Gordon I, Sauvage L. pl):II319–23 [discussion II324].
Definitive proof of endothelialization of a Dacron arterial prosthesis [150] Bull P, Denck H, Guidoin R, Gruber H. Preliminary clinical
in a human being. J Vasc Surg 1995;21(5):862–7. experience with polyurethane vascular prostheses in femoro-
[131] Shi Q, Wu M, Onuki Y, Ghali R, Hunter G, Johansen K, et al. popliteal reconstruction. Eur J Vasc Surg 1992;6(2):217–24.
Endothelium on the flow surface of human aortic Dacron vascular [151] Hess F, Jerusalem C, Braun B. The endothelialization process of a
grafts. J Vasc Surg 1997;25(4):736–42. fibrous polyurethane microvascular prosthesis after implantation in
[132] Herring M, Dilley R, Jersild RJ, Boxer L, Gardner A, Glover J. the abdominal aorta of the rat. A scanning electron microscopic
Seeding arterial prostheses with vascular endothelium. The nature of study. J Cardiovasc Surg (Torino) 1983;24(5):516–24.
the lining. Ann Surg 1979;190(1):84–90. [152] Zhang Z, Marois Y, Guidoin R, Bull P, Marois M, How T, et al.
[133] Burkel W, Ford J, Vinter D, Kahn R, Graham L, Stanley J. Fate of Vascugraft polyurethane arterial prosthesis as femoro-popliteal and
knitted dacron velour vascular grafts seeded with enzymatically femoro-peroneal bypasses in humans: pathological, structural and
derived autologous canine endothelium. Trans Am Soc Artif Intern chemical analyses of four excised grafts. Biomaterials 1997;18(2):
Organs 1982;28:178–84. 113–24.
[134] Schmidt S, Hunter T, Hirko M, Belden T, Evancho M, Sharp W, [153] Huang B, Marois Y, Roy R, Julien M, Guidoin R. Cellular reaction
et al. Small-diameter vascular prostheses: two designs of PTFE and to the Vascugraft polyesterurethane vascular prosthesis: in vivo
endothelial cell-seeded and nonseeded Dacron. J Vasc Surg 1985; studies in rats. Biomaterials 1992;13(4):209–16.
2(2):292–7. [154] Ives CL, Zamora JL, Eskin SG, Weilbaecher DG, Gao ZR,
[135] Schmidt S, Monajjem N, Evancho M, Pippert T, Sharp W. Noon GP, et al. In vivo investigation of a new elastomeric vascular
Microvascular endothelial cell seeding of small-diameter Dacron graft (mitrathane). Trans Am Soc Artif Intern Organs 1984;30:
vascular grafts. J Invest Surg 1988;1(1):35–44. 587–90.
[136] Criado E, Marston W, Reddick R, Woosley J. Contrib Title: [155] Ota K, Sasaki Y, Nakagawa Y, Teraoka S. A completely new
Endothelial coverage of endovascular Dacron grafts in dogs [letter; poly(ether–urethane) graft ideal for hemodialysis blood access.
comment]. J Vasc Surg 1996;23(4):736–7. ASAIO Trans 1987;33(3):129–35.
[137] Sottiurai V, Sue S, Rau D, Tran A. Comparative analysis of [156] Pollock E, Andrews EJ, Lentz D, Sheikh K. Tissue ingrowth and
pseudointima biogenesis in Gelseal coated Dacron knitted graft porosity of biomer. Trans Am Soc Artif Intern Organs 1981;27:
versus crimped and noncrimped graft. J Cardiovasc Surg (Torino) 405–9.
1989;30(6):902–9. [157] White RA. The effect of porosity and biomaterial on the healing and
[138] Nomura Y. The ultra-structure of the pseudointima lining synthetic long-term mechanical properties of vascular prostheses. ASAIO
arterial grafts in the canine aorta with special reference to the origin Trans 1988;34(2):95–100.
of the endothelial cell. J Cardiovasc Surg (Torino) 1970;11(4): [158] Brothers TE, Stanley JC, Burkel WE, Graham LM. Small-caliber
282–91. polyurethane and polytetrafluoroethylene grafts: a comparative
[139] Mesh C, Majors A, Mistele D, Graham L, Ehrhart L. Graft smooth study in a canine aortoiliac model. J Biomed Mater Res 1990;24(6):
muscle cells specifically synthesize increased collagen. J Vasc Surg 761–71.
1995;22(2):142–9. [159] Uchida N, Emoto H, Kambic H, Harasaki H, Chen JF, Hsu SH,
[140] Hertzer N. Regeneration of endothelium in knitted and velour et al. Compliance effect on patency of small diameter vascular
dacron vascular grafts in dogs. J Cardiovasc Surg (Torino) grafts. ASAIO Trans 1989;35(3):556–8.
1981;22(3):223–30. [160] Crane C. Contrib Title: Arteriosclerotic aneurysm of the abdominal
[141] Rosenfeld J, Savarese R, McCombs P, DeLaurentis D. Endothelial aorta; some pathological and clinical correlations. N Engl J Med
infiltration and lining of knitted dacron arterial grafts. Surgical 1955;253(22):954–8.
forum 1981. [161] Pepper M, Vassalli J, Orci L, Montesano R. Biphasic effect of
[142] DeBakey M, Jordan GJ, Abbott J, Halbert B, O’Neal R. The fate of transforming growth factor-beta 1 on in vitro angiogenesis. Exp Cell
Dacron vascular grafts. Arch Surg 1964;89:757–82. Res 1993;204(2):356–63.
[143] Burkel W, Vinter D, Ford J, Kahn R, Graham L, Stanley J. [162] Herbert C, Nagaswami C, Bittner G, Hubbell J, Weisel J. Effects of
Sequential studies of healing in endothelial seeded vascular fibrin micromorphology on neurite growth from dorsal root ganglia
prostheses: histologic and ultrastructure characteristics of graft cultured in three-dimensional fibrin gels. J Biomed Mater Res 1998;
incorporation. J Surg Res 1981;30(4):305–24. 40(4):551–9.
[144] Stumb M, Jordan G, DeBakey M. Endothelium growth from [163] Nehls V, Herrmann R. The configuration of fibrin clots determines
circulating blood on isolated intravascular Dacron hub. Amer J capillary morphogenesis and endothelial cell migration. Microvasc
Path 1963;43:361–8. Res 1996;51(3):347–64.
[145] Hammond W. Surface population with blood-borne cells. In: Zilla [164] Nunes I, Shapiro R, Rifkin D. Characterization of latent TGF-beta
P, Greisler H, editors. Tissue engineering of prosthetic vascular activation by murine peritoneal macrophages. J Immunol 1995;
grafts. Austin: Landes Bioscience; 1998. 155(3):1450–9.
[146] Wu M, Shi Q, Kouchi Y, Onuki Y, Ghali R, Yoshida H, et al. [165] Wahl S, Allen J, Weeks B, Wong H, Klotman P. Transforming
Implant site influence on arterial prosthesis healing: a comparative growth factor beta enhances integrin expression and type IV
study with a triple implantation model in the same dog. J Vasc Surg collagenase secretion in human monocytes. Proc Natl Acad Sci U
1997;25(3):528–36. S A 1993;90(10):4577–81.
[147] Hess F, Jerusalem C, Steeghs S, Reijnders O, Braun B, Grande P. [166] Falcone D, McCaffrey T, Haimovitz - FA, Garcia M. Transforming
Development and long-term fate of a cellular lining in fibrous growth factor-beta 1 stimulates macrophage urokinase expression
polyurethane vascular prostheses implanted in the dog carotid and femo- and release of matrix-bound basic fibroblast growth factor. J Cell
ral artery. A scanning and light microscopical study up to 53 months Physiol 1993;155(3):595–605.
after implantation. J Cardiovasc Surg (Torino) 1992;33(3):358–65. [167] Loscalzo J. The macrophage and fibrinolysis. Semin Thromb
[148] Kogel H, Vollmar JF, Cyba-Altunbay S, Mohr W, Frosch D, Hemost 1996;22(6):503–6.
Amselgruber W. New observations on the healing process in [168] Gross T, Leavell K, Peterson M. CD11b/CD18 mediates the
prosthetic substitution of large veins by microporous grafts—animal neutrophil chemotactic activity of fibrin degradation product D
experiments. Thorac Cardiovasc Surg 1989;37(2):119–24. domain. Thromb Haemost 1997;77(5):894–900.
ARTICLE IN PRESS
P. Zilla et al. / Biomaterials 28 (2007) 5009–5027 5027

[169] Premack B, Schall T. Chemokine receptors: gateways to inflamma- expression by biomedical polymers in the cage implant system.
tion and infection. Nat Med 1996;2(11):1174–8. J Biomed Mater Res 1994;28(5):635–46.
[170] Rhodes N, Hunt J, Williams D. Macrophage subpopulation [181] Swartbol P, Truedsson L, Parsson H, Norgren L. Surface adhesion
differentiation by stimulation with biomaterials. J Biomed Mater molecule expression on human blood cells induced by vascular graft
Res 1997;37(4):481–8. materials in vitro. J Biomed Mater Res 1996;32(4):669–76.
[171] Kuwabara K, Ogawa S, Matsumoto M, Koga S, Clauss M, Pinsky [182] Bonfield T, Colton E, Anderson J. Protein adsorption of biomedical
D, et al. Hypoxia-mediated induction of acidic/basic fibroblast polymers influences activated monocytes to produce fibroblast
growth factor and platelet-derived growth factor in mononuclear stimulating factors. J Biomed Mater Res 1992;26(4):457–65.
phagocytes stimulates growth of hypoxic endothelial cells. Proc Natl [183] Bonfield T, Colton E, Marchant R, Anderson J. Cytokine and
Acad Sci U S A 1995;92(10):4606–10. growth factor production by monocytes/macrophages on protein
[172] Van Otteren G, Standiford T, Kunkel S, Danforth J, Strieter R. preadsorbed polymers. J Biomed Mater Res 1992;26(7):837–50.
Alterations of ambient oxygen tension modulate the expression of [184] Hunt J, Meijs G, Williams D. Hydrophilicity of polymers and soft
tumor necrosis factor and macrophage inflammatory protein-1 tissue responses: a quantitative analysis. J Biomed Mater Res 1997;
alpha from murine alveolar macrophages. Am J Respir Cell Mol 36(4):542–9.
Biol 1995;13(4):399–409. [185] McNally A, DeFife K, Anderson J. Interleukin-4-induced macro-
[173] Leibovich S, Polverini P, Shepard H, Wiseman D, Shively V, Nuseir phage fusion is prevented by inhibitors of mannose receptor activity.
N. Macrophage-induced angiogenesis is mediated by tumour Am J Pathol 1996;149(3):975–85.
necrosis factor-alpha. Nature 1987;329(6140):630–2. [186] Kaufman B, DeLuca D, Folsom D, Mansell S, Gorman M, Fox P,
[174] Taichman N, Young S, Cruchley A, Taylor P, Paleolog E. Human et al. Elevated platelet-derived growth factor production by aortic
neutrophils secrete vascular endothelial growth factor. J Leukoc grafts implanted on a long-term basis in a canine model. J Vasc Surg
Biol 1997;62(3):397–400. 1992;15(5):806–15 [discussion 815-806].
[175] Kao W, Zhao Q, Hiltner A, Anderson J. Theoretical analysis of in [187] McNally A, Anderson J. Interleukin-4 induces FBGCs from human
vivo macrophage adhesion and foreign body giant cell formation on monocytes/macrophages. Differential lymphokine regulation of
polydimethylsiloxane, low density polyethylene, and polyetherur- macrophage fusion leads to morphological variants of multi-
ethanes. J Biomed Mater Res 1994;28(1):73–9 [published erratum nucleated giant cells. Am J Pathol 1995;147(5):1487–99.
appears in J Biomed Mater Res 1994 Jun;28(6):761]. [188] Essner R, Rhoades K, McBride W, Morton D, Economou J. IL-4
[176] Mathur A, Collier T, Kao W, Wiggins M, Schubert M, Hiltner A, down-regulates IL-1 and TNF gene expression in human mono-
et al. In vivo biocompatibility and biostability of modified cytes. J Immunol 1989;142(11):3857–61.
polyurethanes. J Biomed Mater Res 1997;36(2):246–57. [189] Zhao Q, Topham N, Anderson J, Hiltner A, Lodoen G, Payet C.
[177] Defife K, Jenney C, Colton E, Anderson J. Confocal and light Foreign-body giant cells and polyurethane biostability: in vivo
microscopic evaluation of silane surface-dependent macrophage correlation of cell adhesion and surface cracking. J Biomed Mater
development and IL-4 induced foreign body giant cell formation. Res 1991;25(2):177–83.
Fifth world biomaterials congress, 1996. [190] Anderson J. Inflammatory reaction: the nemesis of implants. In:
[178] McNally A, Anderson J. The lymphokine IL-4 induces very large Zilla P, Greisler H, editors. Tissue engineering of prosthetic vascular
FBGC and syncytia from human macs in a material surface property grafts. Austin: Landes Bioscience; 1998.
dependent manner in vitro. Fifth world biomaterials congress, 1996. [191] Forsyth E, Aly H, Neville R, Sidawy A. Proliferation and
[179] Kao W, Hiltner A, Anderson J, Lodoen G. Theoretical analysis of extracellular matrix production by human infragenicular smooth
in vivo macrophage adhesion and foreign body giant cell formation muscle cells in response to interleukin-1 beta. J Vasc Surg
on strained poly(etherurethane urea) elastomers. J Biomed Mater 1997;26(6):1002–7 [discussion 1007–1008].
Res 1994;28(7):819–29. [192] Makita S, Nakamura M, Yoshida H, Hiramori K. Autocrine growth
[180] Petillo O, Peluso G, Ambrosio L, Nicolais L, Kao W, Anderson J. inhibition of IL-1 beta-treated cultured human aortic smooth muscle
In vivo induction of macrophage Ia antigen (MHC class II) cells: possible role of nitric oxide. Heart Vessels 1996;11(5):223–8.

You might also like