Sie sind auf Seite 1von 9

Journal of Thrombosis and Haemostasis, 9 (Suppl. 1): 26–34 DOI: 10.1111/j.1538-7836.2011.04360.

INVITED REVIEW

Serpin structure, function and dysfunction


J. A. HUNTINGTON
Department of Haematology, Cambridge Institute for Medical Research, University of Cambridge, Cambridge, UK

To cite this article: Huntington JA. Serpin structure, function and dysfunction. J Thromb Haemost 2011; 9 (Suppl. 1): 26–34.

acronym SERPIN, since most members of this family were


Summary. Serpins have been studied as a distinct protein serine protease inhibitors. There are now well over 1500 serpin
superfamily since the early 80s. In spite of the poor sequence sequences identified in the genomes of organisms representing
homology between family members, serpins share a highly all kingdoms of life, and 36 confirmed human serpins [4]. Most
conserved core structure that is critical for their functioning as family members are indeed serine protease inhibitors, but
serine protease inhibitors. Therefore, discoveries made about several have additional cross-class inhibition functions and
one serpin can be related to the others. In this short review, I inhibit cysteine protease family members such as the caspases
introduce the serpin structure and general mechanism of and cathepsins, and others, such as OVA, are incapable of
protease inhibition, and illustrate, using recent crystallographic protease inhibition and serve other functions. In plants and
and biochemical data on antithrombin (AT), how serpin other basic organisms small ÔcanonicalÕ inhibitors (such as
activity can be modulated by cofactors. The ability of the Kunitz and Kazal type inhibitors) predominate [5]. In humans,
serpins to undergo conformational change is critical for their serpins have been selected to control processes that require
function, but it also renders them uniquely susceptible to tight regulation, such as blood coagulation, inflammation and
mutations that perturb their folding, leading to deficiency and fibrinolysis. Much has been learnt since 1980 about the
disease. A recent crystal structure of an AT dimer revealed that function of individual serpins in human health and disease,
serpins can participate in large-scale domain-swaps to form but a single question continues to define the field: Why has
stable polymers, and that such a mechanism may explain the evolution settled on a serpin to fulfil a particular function? The
accumulation of misfolded serpins within secretory cells. answer lies in the unique serpin structure, the conserved
Serpins play important roles in haemostasis and fibrinolysis, inhibitory mechanism, and the regulatory mechanisms that
and although each will have some elements specifically tailored govern serpin activity.
for its individual function, the mechanisms described here
provide a general conceptual framework.
Serpin structure and metastability
Keywords: antithrombin, crystallography, haemostasis, hepa- Since serpins were known to be inhibitors of serine proteases,
rin, polymerisation. it was clear that their structure would present a loop to
interact with the protease active site cleft. In other families of
serine protease inhibitors, these Ôreactive centre loopsÕ (RCL)
are short and are maintained in a rigid conformation ideal
Meet the serpin superfamily for protease docking. In this case, inhibition depends on the
inability of the ends of the scissile bond (P1–P1¢, as per the
In a paper published in 1980, Hunt and Dayhoff [1] noted the
nomenclature of Schechter and Berger [6]) to separate after
sequence similarity between two apparently unrelated proteins,
proteolytic attack. The inhibitor thus binds tightly, but
chicken ovalbumin (OVA) and human antithrombin III (AT),
reversibly to a protease, and neither protein undergoes
and suggested a provisional name of the Ôovalbumin-anti-
conformational change. In contrast, serpin RCLs are long
thrombin superfamilyÕ. Previous work had found sequence
(typically 20–24 residues) and flexible, resembling a substrate
homology between AT and another human protease inhibitor,
loop. The first crystal structure of a native serpin was that of
a1-antitrypsin (a1AT, also known as a1-proteinase inhibitor)
OVA [7], and its RCL was found to form a stable helix.
[2]. In a seminal review of this new family and its members in
Subsequent structures of inhibitory serpins showed a range of
1985, Carrell [3] coined the now-widely used descriptive
RCL conformations and a high degree of conformational
Correspondence: James A. Huntington, Department of Haematology,
flexibility (often reflected by high B-factors or lack of electron
Cambridge Institute for Medical Research, University of Cambridge, density). The native serpin is composed of approximately 400
Wellcome Trust/MRC Building, Hills Road, Cambridge CB2 0XY, amino acids that fold into an N-terminal helical domain and
UK. a C-terminal b-barrel domain (Fig. 1A). The two main
Tel.: +44 1223 763230; fax: +44 1223 336827. features related to function are the RCL and the five-
E-mail: jah52@cam.ac.uk

 2011 International Society on Thrombosis and Haemostasis


Serpin structure, function and dysfunction 27

stranded central b-sheet (b-sheet A). In the classic orienta- studies into the serpin mechanism of protease inhibition that the
tion, the RCL is on top and sheet A is facing (Fig. 1B). ability to rapidly and stably incorporate the RCL into b-sheet A
The unusual, if not unique, aspect of the serpins is that their distinguished the inhibitory (such as a1AT) from the non-
native fold is not their most stable. Serpins are able to inhibitory serpins (such as OVA) [10,11]. However, the nature of
incorporate the RCL into b-sheet A as the central fourth strand the inhibitory complex and the purpose of the conformational
(s4A), and in so doing, become hyperstable [8,9]. This confor- change to a hyperstable state were not known until a crystal
mational/topological change can occur upon proteolytic cleav- structure of the final serpin-protease complex was solved.
age within the RCL (Fig. 1C), or spontaneously to form the
intact ÔlatentÕ conformation (Fig. 1D). It was clear from early
The serpin mechanism
The long, flexible RCL of serpins and their ability to stably
A B P1 incorporate the RCL into b-sheet A suggested a mechanism
P1′ other than the well characterised canonical lock-and-key
mechanism used by other families of protease inhibitors, but
just what that new mechanism might be was unclear. One
proposal was that partial incorporation of an intact RCL
would lock it into a canonical (key-like) conformation that
would inhibit proteases via a non-covalent, reversible mecha-
nism [12]. However, the fact that serpin-protease complexes
could be observed on SDS-PAGE suggested a covalent
complex, somehow stabilised by the full incorporation of the
cleaved RCL into sheet A [13]. Fluorescence energy transfer
studies appeared to support of both mechanisms [14,15]. In
2000, we solved and published the first crystallographic
structure of a final serpin-protease complex [16] (Fig. 2A).
Consistent with the fluorescence studies performed by the
Gettins group [15], we found a fully incorporated RCL with
C D

A B

3
6 5 4
2

Lys328
P1
1
Asp194

Ser195

Fig. 1. Structures of native and stable serpins. (A) A ribbon diagram of


the native form of prototypical serpin a1AT (1QLP, [81]) is shown col-
oured from its N-to-C terminus (blue to red), illustrating the domain
structure of serpins, with the N-terminal helical region (predominantly
blue and green) and the C-terminal b-sheet region (mostly yellow,
orange and red). The RCL is the orange loop on the ÔtopÕ of the mole-
cule and the scissile bond is illustrated by showing rods for the P1 and Fig. 2. The structure of the final serpin-protease complex. (A) A ribbon
P1¢ side chains (Met and Ser). (B) The same structure as in (A) is col- diagram of the a1AT-trypsin complex is shown (1EZX, [16]), with a1AT
oured to highlight the main structural features of a serpin, with the RCL in coloured as previously, and trypsin in green. To illustrate the region of
yellow and b-sheet A in red. The P1 and P1¢ residues are indicated, and the trypsin that was not observed in the crystal structure, native trypsin is
N-terminal portion (hinge region) of the RCL is indicated by the oval. (C) shown as a semi-transparent cyan cartoon. The P3-P1 region of a1AT and
When cleaved at the P1–P1¢ bond, inhibitory serpins such as a1AT un- the catalytic loop of trypsin are shown as rods. (B) A close-up of the boxed
dergo a rapid conformational change where the RCL inserts as the fourth region from panel A illustrates the covalent bond between the serpin P1
strand in b-sheet A (strands are numbered, 8API, [17]). RCL residues that Met and Ser195 from trypsin, including original electron density for the
participate in the antiparallel b-sheet interactions include P15-P3 for most region (blue wire). The arrows indicate the nitrogen atoms that normally
serpins. This conformation is hyper-stable. (D) A similar increase in sta- form the oxyanion hole (note their distance from the carboxyl O of the P1
bility can be achieved without proteolytic cleavage to form the ÔlatentÕ residue, indicated by the star), and the new salt-bridge between Asp194 of
conformation (here of the serpin plasminogen-activator inhibitor 1, 1LJ5). trypsin and Lys328 of a1AT is also indicated.

 2011 International Society on Thrombosis and Haemostasis


28 J. A. Huntington

the bond between the P1 and P1¢ residue severed, exactly as in but that the extent of unfolding may depend on the serpin-
the first structure of an RCL cleaved serpin [17] (Fig. 1C). The protease pair. Factors that could affect the extent of unfolding
protease was found at the bottom of the serpin, covalently include the length of the RCL, the size and composition of
linked via an ester bond between the catalytic Ser195 of the loops on the ÔbottomÕ of the serpin and of loops flanking the
protease and main chain carbonyl carbon of the P1 residue active site of the protease, and the presence of ligands that
(Fig. 2B). This linkage represents the acyl-enzyme intermediate preferentially bind to the protease conformation.
of the serine protease catalytic cycle, after the expulsion of the
P¢ residues from the active site, but before the ingress of water
Heparin modulation of serpin activity
for deacylation. The catalytic loop of the protease was
distended by the pulling force exerted on it by the limited The activity of several serpins is modulated by glycosamino-
length of the RCL and the clash between the body of the glycans such as heparin [25]. In general, heparin serves to
protease and the body of the hyperstable serpin. This resulted ÔbridgeÕ the serpin to the protease, thus enhancing the rate of
in the destruction of the oxyanion hole, required for stabilisa- complex formation. This is true for protein C inhibitor with
tion of the tetrahedral transition state, thus preventing thrombin and activated protein C [26], for PAI-1 and thrombin
deacylation (Fig. 2B). In addition to the conformational [27], for protease nexin 1 and thrombin [28], for ZPI and fXa
change in the serpin (full-loop insertion) and changes in the and fXIa [29], and importantly for AT and thrombin [30]. This
catalytic site of the protease, about 40% of the protease also bridging effect also contributes to the heparin-acceleration of
appeared to have become disordered, based on lack of electron thrombin inhibition by HCII [31], and fIXa and fXa inhibition
density (Fig. 2A). The disordered region included residues 16– by AT [32], however, the majority of the rate enhancement in
41, 62-84, 110–120, 139–156, 186–190 and 223–224. Of these six these cases is provided by conformational change in the serpin
linear stretches, four (underlined) correspond to the serine (allostery). In order for allostery to contribute to heparin
protease zymogen activation domain [18]. This domain tran- acceleration of serpin function, the native state of the serpin
sitions from a disordered to ordered state upon conversion must be in a low activity conformation, and the new heparin-
from a zymogen to a protease due to the insertion of the N- bound conformation must somehow enable the formation of
terminus (Ile16 forms a salt-bridge with Asp194) into the productive recognition (Michaelis) complexes with target
activation pocket (formed by the three C-terminal stretches proteases. Over the years we and others have studied the
underlined above). The pulling force exerted on Ser195 by the activation of AT by heparin, and have successfully determined
serpin extends the loop 3.5Å from its normal catalytic position, the molecular basis of the relative inactivity of the native form,
thus destroying the oxyanion hole (main chain amides of how heparin binds, how its binding results in conformational
Gly193 and Ser195, Fig. 2B), and also breaking the salt-bridge change, and how the main targets of AT, thrombin, fXa and
between Ile16 and Asp194, effectively converting the protease fIXa, are recognised in the presence of heparin. AT is thus a
back into a zymogen-like state. Intriguingly, in our structure, thoroughly described system that provides a paradigm for the
Asp194 makes an alternate salt-bridge with Lys328 of the heparin binding serpins, as well as other serpins where binding
serpin at the bottom strand 5A (Fig. 2B). to cofactors and proteases is regulated by conformational
The extent of structural disorder engendered in the protease change (e.g. HCII [33], and even non-inhibitory serpins such as
by its complexation with serpins is a matter of some debate. thyroxine and corticosteroid binding globulins [34]).
Early studies revealed that discrete regions of the protease
became accessible to proteolytic attack when in complex with
AT conformations
serpins, and these regions were later mapped to the disordered
region in the serpin-protease complex structure. Similarly, AT circulates in a native state that is a poor inhibitor of its
trypsin was found to lose affinity for Ca2+ when in complex target proteases, namely thrombin, fXa and fIXa. The work of
with a1AT, suggesting disordering of the binding site (the 70s Steve Olson and Ingemar Björk (and others) elegantly demon-
loop) as seen in the crystal structure, and high Ca2+ strated that AT undergoes a conformational change upon
concentrations were shown to promote the dissociation of the heparin binding that results in a 500-fold acceleration of fXa
complex, presumably by populating the protease conformation and fIXa inhibition, but that thrombin is insensitive to the AT
[19]. Thrombin was shown to lose the ability to bind to exosite I conformational change (many papers, but most completely
ligands (the 30s and 70s-loops) [20,21], and we recently showed described in Ref. [32]). The conformational basis for the low
that both exosites I and II become dysfunctional due to the activity of the circulating form of AT was unknown until the
disordering induced by serpin complexation [22]. These results structure of native AT was solved in 1994 [35,36]. It was found
suggest a global disordering of the protease, and are consistent to have the normal serpin fold, but with two important
with NMR data implying a molten-globule-like conformation differences – the RCL was partly incorporated into b-sheet A
for proteases when complexed by serpins [23]. In striking (P15-P14 residues Gly379 and Ser380) and the P1 residue was
contrast, no disorder was observed in a recent crystal structure oriented toward the main body of AT and was thus unavailable
of porcine pancreatic elastase in complex with a1AT [24]. for interaction with a protease (Fig. 3A, centre panel).
However, on the weight of the currently available data, we can Subsequent biochemical and structural work has confirmed
conclude that some protease unfolding is induced by serpins, that the native conformation does have the RCL partially

 2011 International Society on Thrombosis and Haemostasis


Serpin structure, function and dysfunction 29

Fig. 3. The AT native equilibrium and pentasaccharide binding mechanism. (A) AT exists in a low activity native state as an ensemble of conformations.
The left panel is the structure of native AT in the monomeric state (i.e. not crystallised in complex with latent AT, 1T1F, [41]), and shows the hinge
region inserted (P15-P14) into sheet A, resulting in the approximation of the RCL to the body of AT. The P1 residue (sticks) interacts via a salt-
bridge with Glu237 on the surface of AT. The middle panel is the structure of native AT in complex with latent AT (1E04, [82]). It shares hinge
region insertion, but the RCL has moved into a more exposed position, and the P1 residue makes only a weak main-chain hydrogen bond with the ÔtopÕ
of AT. The right panel represents the P1-exposed version of the centre panel, and is poised for reaction with a protease (e.g. thrombin and fXa). (B) Binding
of the heparin pentasaccharide (green rods) to AT (coloured as before) occurs in a three-step mechanism. The first (left) is presumably just the
approximation of the negatively charged heparin with the positively charged region along helix D (cyan), and involves no conformational rearrangement
of native AT (1E04). The second step (centre) involves conformational changes in the N-terminal (lower) portion of AT, and contributes 40% of the
induced-fit binding energy (3EVJ, [37]). The final step (right) involves the expulsion of the hinge and closure of sheet A (2GD4, [55]), followed by the
elongation of helix D. This step contributes 60% of the energy of binding and essentially locks AT into a high-activity state [37]. All representations of AT
in this and other figures exclude the N-terminal 44 residues. (C) AT can be subdivided into four fragments that move as rigid bodies during the heparin
activation steps, and plastic regions that move independently (grey). The largest fragment (blue) is considered rigid, and the green and yellow frag-
ments undergo their conformational change in the first heparin binding step (in panel B, left to centre). The red region, composed of the tops of strands
2 and 3A, slides over to close b-sheet A to the 5-stranded form during the final step of heparin binding (panel B, centre to right). The inherently
flexible regions, including the N-terminus and part of the RCL, were excluded from the analysis.

incorporated, and that expulsion of the hinge is sufficient for low activity state was unclear, due in part to crystal contacts
full allosteric activation [37–40]. However, the orientation of involving the RCL. A structure of AT in another crystal form
the P1 residue and its role in maintaining circulating AT in a [41] again showed the P1 oriented toward the main body of AT,

 2011 International Society on Thrombosis and Haemostasis


30 J. A. Huntington

however, in this case it was making a different, more intimate and Xa. It has long been understood that thrombin inhibition
contact (Fig. 3A, left panel). The RCL of native AT was thus is insensitive to the conformation of AT [47], and thus the
capable of at least two conformations that sequestered the recognition complex should not require the expulsion of the
RCL away from an attacking protease, however, the appre- RCL from b-sheet A. The crystal structure of the ternary
ciable rate of thrombin inhibition in the absence of heparin complex (solved with a synthetic hexadecasaccharide) showed
(7000 M)1 s)1) suggested that the P1 residue must be accessible why, with thrombin positioned forward and towards the
to some degree (as illustrated in Fig. 3A, right panel). Native heparin binding site of AT (Fig. 4A) [48]. The hinge region
AT can thus be thought of as an ensemble of conformations, could not be fully resolved in electron density due to flexibility,
such as those shown in Fig. 3A, in rapid equilibrium. however, clear evidence for the pre-insertion of P15 was seen,
The effect of heparin binding on the conformation of AT can and the position of thrombin could easily be maintained with
be summarised as the expulsion of the RCL from b-sheet A, and the RCL inserted to the level seen in native AT. When the hinge
its subsequent closure so that the activated state of AT resembles region was prevented from being expelled from b-sheet A by an
the native conformation of most other serpins (such as a1AT, engineered disulphide bridge, the rate of thrombin inhibition in
Fig. 1B, right panel). This conformational change improves the presence of heparin was actually slightly increased [37].
affinity for the specific heparin pentasaccharide by 1000-fold Thus, the position of thrombin found in this structure is equally
through an induced-fit mechanism [42]. Recent structural and valid for heparin-activated and native AT. One important
biochemical data have revealed that conformational changes in
the lower helical part of the AT precede expulsion of the RCL
and contribute about 40% of the binding energy provided by the A B
conformational change [37,43]. Thus, the binding of AT to the
specific heparin pentasaccharide occurs through three steps
(Fig. 3B): the first is the non-specific association of the
negatively charged heparin with the basic heparin binding site
147-loop
on helix D (cyan in figure); the second step is the conformational
rearrangement of the helical domain of AT to accommodate the 60-loop
pentasaccharide and maximise ionic and hydrogen bonding
interactions; in the third step, the hinge is released from b-sheet
A, and the top portions of strands 2 and 3 slide over to make a
continuous five stranded b-sheet. Careful structural analysis of
these structures reveal a large constant domain, primarily
composed of the C-terminal b-barrel domain (blue in Fig. 3C),
three regions that move as rigid bodies (green, yellow and red), C
and loops that move independently (grey) [37]. In a reversal of
the heparin binding mechanism, full incorporation of the RCL
that accompanies formation of the final complex between AT
and a protease results in a 1000-fold decrease in heparin affinity
[43,44]. This analysis also provides a detailed paradigm for the
general mechanism of modulation of serpin activity through
conformational change.

Heparin activation of AT
Determining the mechanism of heparin activation of AT is Fig. 4. Structure of the recognition (Michaelis) complex between AT and
thrombin in the presence of an activating heparin chain. (A) A ribbon
crucial for understanding the mode of action of the numerous diagram of the ternary complex between AT (coloured as previously),
heparin fractionations and synthetic heparins currently on the S195A thrombin (heavy chain in magenta, light chain in light pink), and
market as anticoagulants. In all cases, specificity and high a synthetic hexadecasaccharide (green rods) containing the pentasaccha-
affinity binding to AT is conferred by the presence of the ride sequence and a highly sulphated thrombin binding site [48,83]. The
pentasaccharide sequence, present in about a third of unfrac- hinge region is in a similar conformation as that seen in native AT,
with P15 Gly making b-strand H-bonds (the rest of the hinge is not
tionated heparin chains. All such chains will allosterically resolved in electron density), due to the forward orientation of throm-
activate AT towards fXa and fIXa inhibition. However, chains bin on AT. (B) A surface representation of thrombin and AT without
must be 18 monosaccharide units in length (5.4 kDa) to the RCL (coloured and oriented as in panel A) illustrates the intimate
accelerate thrombin inhibition by a bridging mechanism [45], nature of the contact between the body of AT and the 60- and 147-
and about twice as long to bridge fXa and fIXa to AT [46]. insertion loops (indicated). (C) A close-up of the RCL is shown to illus-
trate the elongation of the P¢ insert (green) to allow the orientation of
Most low molecular weight heparin preparations contain a thrombin relative to AT and to accommodate the 60- and 147-insertion
distribution of sizes close to that capable of bridging thrombin, loops. The RCL of pentasaccharide-bound AT (1E03, [84]) is shown in
but only unfractionated heparin is likely to bridge factors IXa orange, with the P¢ insert in green.

 2011 International Society on Thrombosis and Haemostasis


Serpin structure, function and dysfunction 31

feature of thrombin is the depth of the active site cleft due to the A B
60 and 147-insertion loops, both of which make significant
contacts with the body of AT (Fig. 4B). AT has a unique three-
residue insert on the P¢ side of the RCL (shown in green in
Fig. 4C) that normally forms a tightly H-bonded turn, and that
is seen to unwind in the Michaelis complex to accommodate
the bulk of the 60 and 147-insertion loops and to allow the
forward orientation of thrombin (Fig. 4C). Consistent with
these findings, deletion of the three-residue insert reduces the
rate of thrombin inhibition [49]. Thus, recognition of thrombin
by AT in the absence and presence of heparin is independent of
the extension of the N-terminal portion of the RCL (the hinge
region), but does require the extension of the C-terminal C D
portion (the P¢ region).
In contrast, recognition of fXa and fIXa by AT is exquisitely
sensitive to the conformation of the hinge region. Disulphide
engineering studies have shown that fXa can form two separate
recognition complexes depending on the state of the hinge
region [37,50]. Presumably when AT is in the native state the
recognition complex would resemble that of thrombin. How-
ever, fIXa can only react with thrombin when the hinge region is Fig. 5. Crystal structures of the pentasaccharide-activated complexes be-
expelled and free to extend away from b-sheet A. It was also tween AT and factors Xa and IXa. (A) A ribbon diagram of the ternary
complex between AT (coloured as previously), fXa (epidermal growth-
clear from studies by Alireza Rezaie and Steve OlsonÕs groups factor [EGF] domain 2 [light blue] plus S195A variant of the catalytic
that the 147-loop (and Arg150 in particular) made a crucial domain [slate blue]) and the pentasaccharide (green sticks) (2GD4) reveals
exosite contact with the body of AT [51–54]. The structures of exosite contacts that necessitate the expulsion and extension of the hinge
the pentasaccharide-activated recognition complexes between region of the RCL. The critical contact involves residue Arg150 of the
AT and fXa [55] (Fig. 5A) and fIXa [56] (Fig. 5B) revealed a protease (sticks). (B) The same exosite contact (Arg150 in stick represen-
tation) was observed in the crystal structure of the AT, fIXa (magenta),
rotation in the protease (relative to the position observed for pentasaccharide complex (3KCG), although the fIXa was rotated by
thrombin) that resulted in an orientation requiring hinge region about 40. (C) The requirement for hinge region expulsion and extension is
extension (Fig. 5C). The rotation allowed the 147-loop of the best visualised by comparing the position of thrombin on AT to that of
protease to engage in extensive contacts with AT, and of fIXa on AT. The heavy chain of fIXa is shown coloured from N-to-C
particular importance, the burying of Arg150 in an acidic (and terminus from blue-to-red, and is placed in the position found in its
complex with AT (opaque) and in the position of thrombin (semitrans-
also hydrophobic) pocket formed by Arg235, Glu237, Met251, parent). The stereo representation helps to appreciate the radical move-
Tyr253, and His319. Intriguingly, this pocket is occupied in the ment of the protease domain, involving a 16Å translation and a 104
native monomeric state of AT by the P1 residue in a similar rotation. (D) To illustrate the exosite interaction involving Arg150 (cyan
fashion (Fig. 5D), supporting the idea that heparin binding for fXa and magenta for fIXa), the surface of AT (1T1F, without the
ÔexposesÕ the exosite utilised by factors IXa and Xa [57]. The RCL) is shown coloured according to electrostatics. The P1 Arg (yellow)
of AT competes for the same pocket when in the native form.
unique AT P¢ insert is partially unwound to allow fXa to rotate
into the observed position, but the tight H-bonded turn is
preserved in the complex with fIXa and makes several important
interactions to improve the stability of the complex [56]. Thus, signalling [59]. However, the inherent dependence of the serpin
the RCL of AT has unique features that are exploited to allow mechanism on conformational mobility and a metastable native
for heparin regulation of protease binding and inhibition. fold render serpins highly susceptible to mutations that perturb
function [60,61]. There are many things that can go wrong with
serpins that lead to loss or gain of function (for reviews see Refs
Serpin dysfunction
[60–63] and others). Classic examples are mutations in the RCL
The advantages of the serpin mechanism over the standard lock- that either change specificity (such as the Pittsburgh variant of
and-key mechanism are manifold, and include: the ability to a1AT that turns it into an inhibitor of thrombin [64]) or knock
regulate activity by altering the conformation or accessibility of out inhibitory function by slowing RCL incorporation into
the RCL (AT activation by heparin is a good example); the sheet A (such as the Cambridge II mutation in AT [65]). Another
covalent complex is irreversible; the change in serpin and common cause of serpin deficiency is the failure to secrete active
protease structure releases complexes from cofactors and protein due to mutations that affect the ability of serpins to fold
substrates (e.g. thrombin loses the ability to interact with exosite into the metastable native state. These mutations can occur
I and exosite II ligands [22], and AT loses the ability to bind with almost anywhere in the serpin, and are normally associated with
high affinity to heparin [58]); and change in serpin and/or the formation and accumulation of stable polymers within the
protease conformation can result in receptor binding and endoplasmic reticulum of secretory cells (for reviews see Refs

 2011 International Society on Thrombosis and Haemostasis


32 J. A. Huntington

[66,67] and others). Polymerisation always leads to lower levels A B


of secreted serpin, and in cases where the serpin concentration is
critical, heterozygous deficiency leads to disease (such as
thrombosis for AT). In other cases, only the homozygous
carriers of mutations develop a loss-of-function disease (such as
emphysema for a1AT). In rare cases, the accumulation of serpin
polymers is toxic to cells and leads to tissue damage through an
unknown gain-of-function mechanism [68]. Several homozy-
gous mutations in a1AT have been associated with liver disease,
including cirrhosis, and several heterozygous mutations in
neuroserpin have been described that lead to dementia and C
death. Although the incidence of disease caused by serpin
polymerisation (through both loss- and gain-of-function mech-
anisms) is quite rare, there has been a great deal of interest in the
phenomenon, due in part to its classification as a Ôconforma-
tional diseaseÕ, along with AlzheimerÕs, HuntingtonÕs and
ParkinsonÕs diseases, and the prion encephalopathies [69].

The mechanism of serpin polymerisation


Consistent with the polymerisation of a folding intermediate in
cells, serpins are capable of polymerising in vitro when
incubated under mildly denaturing conditions (heat, low pH
or chaotropic agents, or in combination) [70]. It has been
demonstrated for a1AT and for neuroserpin that unfolding Fig. 6. Structure and models related to the Ôdomain-swapÕ mode of serpin
and folding proceed via an intermediate that is prone to polymerisation. (A) A crystal structure of a dimer of AT (2ZNH, [79])
revealed that the basis of intermolecular linkage was the swapping of
polymerisation [71–75]. For wild-type protein this intermediate
strands 4 and 5 of b-sheet A from one monomer (yellow) into another
is short-lived and will thus form only a small amount of (green). (B) The closed dimer could be opened by simply uncoiling part of
polymers upon renaturation, but mutations that slow the the region following strand 6A, to form a head-to-tail dimer that could be
conversion from intermediate to native state result predomi- elongated indefinitely. (C) The dimer structure showed that strand 5A
nantly in polymeric serpin. Polymers obtained from cells or by extraction from its position in the native state is required to form the
polymeric linkage, therefore suggesting a polymerigenic folding interme-
partial denaturation have similar properties and are thought to
diate such as the model shown here (coloured as before).
share a common intermolecular linkage. Indeed, it has long
been believed that the molecular basis of polymerisation is the
same for all serpins, whether induced by mutations or by mild within cells, and if the regions that participate in domain
denaturation, and that the mechanism involved the incorpo- swapping depend on the serpin and the mutations involved.
ration of the RCL of one monomer into b-sheet A of another
(the Ôloop-sheetÕ mechanism). However, there is no direct
Conclusions
evidence for the formation of loop-sheet polymers by partial
denaturation or in vivo, and several experiments contradicting Serpins are found in all branches of life and appear time-and-
the loop-sheet hypothesis have been reported [76–78]. An again controlling proteolytic pathways related to human health
alternative mechanism of polymerisation was suggested by a and disease. It was perhaps serendipity that the most important
recent crystal structure of a stable AT dimer formed by anticoagulant serpin AT was a founding member of the
incubating native AT at pH 5.7, 37 C [79]. AT was found to superfamily, but nevertheless, AT provides a useful and well-
have undergone a large domain swap that included both the characterised model for understanding the possible range of
RCL (strand 4A) and strand 5A, along with other regions serpin structure and how it relates to serpin function and
(Fig. 6A). The structure could readily be extended to form an dysfunction.
open dimer with a flexible linker region (Fig. 6B), and
suggested a polymerigenic folding intermediate (or unfolding
Acknowledgements
intermediate) that had strand 5A as well as the RCL exposed
(Fig. 6C). Several experiments have been conducted to deter- JAH is a Senior Medical Research Council (MRC) Non-
mine if this mechanism is general for the serpins, and clinical Fellow.
preliminary data suggest that it may account for AT, a1-AT
and neuroserpin polymerisation under certain conditions
Disclosure of Conflict of Interests
[72,79,80]. More work is required to determine if this or other
domain swaps are responsible for polymerisation of serpins The author states that he has no conflict of interest.

 2011 International Society on Thrombosis and Haemostasis


Serpin structure, function and dysfunction 33

References 22 Li W, Johnson DJ, Adams TE, Pozzi N, De Filippis V, Huntington


JA. Thrombin inhibition by serpins disrupts exosite II. J Biol Chem
1 Hunt LT, Dayhoff MO. A surprising new protein superfamily con- 2010; 285: 38621–9.
taining ovalbumin, antithrombin-III, and alpha 1-proteinase inhibitor. 23 Peterson FC, Gettins PG. Insight into the mechanism of serpin-pro-
Biochem Biophys Res Commun 1980; 95: 864–71. teinase inhibition from 2D [1H-15N] NMR studies of the 69 kDa
2 Borsodi AD, Bradshaw RA, Rughani IK, Bruce RM. Structural and alpha 1-proteinase inhibitor Pittsburgh-trypsin covalent complex.
functional relationships of human antithrombin III and alpha-anti- Biochemistry 2001; 40: 6284–92.
trypsin. Adv Exp Med Biol 1975; 52: 249–53. 24 Dementiev A, Dobo J, Gettins PG. Active site distortion is sufficient
3 Carrell R, Travis J. [alpha]1-Antitrypsin and the serpins: variation and for proteinase inhibition by serpins: structure of the covalent complex
countervariation. Trends Biochem Sci 1985; 10: 20–4. of alpha1-proteinase inhibitor with porcine pancreatic elastase. J Biol
4 Law RH, Zhang Q, McGowan S, Buckle AM, Silverman GA, Wong Chem 2006; 281: 3452–7.
W, Rosado CJ, Langendorf CG, Pike RN, Bird PI, Whisstock JC. An 25 Rau JC, Beaulieu LM, Huntington JA, Church FC. Serpins in
overview of the serpin superfamily. Genome Biol 2006; 7: 216. thrombosis, hemostasis and fibrinolysis. J Thromb Haemost 2007; 5
5 Rawlings ND, Tolle DP, Barrett AJ. Evolutionary families of pepti- (Suppl. 1): 102–15.
dase inhibitors. Biochem J 2004; 378: 705–16. 26 Li W, Huntington JA. The heparin binding site of protein C inhibitor is
6 Schechter I, Berger A. On the size of the active site in proteases. I. protease-dependent. J Biol Chem 2008; 283: 36039–45.
Papain. Biochem Biophys Res Commun 1967; 27: 157–62. 27 Ehrlich HJ, Keijer J, Preissner KT, Gebbink RK, Pannekoek H.
7 Stein PE, Leslie AG, Finch JT, Turnell WG, McLaughlin PJ, Carrell Functional interaction of plasminogen activator inhibitor type 1 (PAI-
RW. Crystal structure of ovalbumin as a model for the reactive centre 1) and heparin. Biochemistry 1991; 30: 1021–8.
of serpins. Nature 1990; 347: 99–102. 28 Cunningham DD, Wagner SL, Farrell DH. Regulation of protease
8 Im H, Ahn HY, Yu MH. Bypassing the kinetic trap of serpin protein nexin-1 activity by heparin and heparan sulfate. Adv Exp Med Biol
folding by loop extension. Protein Sci 2000; 9: 1497–502. 1992; 313: 297–306.
9 Kaslik G, Kardos J, Szabo E, Szilagyi L, Zavodszky P, Westler WM, 29 Huang X, Rezaie AR, Broze GJ Jr, Olson ST. Heparin is a major
Markley JL, Graf L. Effects of serpin binding on the target proteinase: activator of the anticoagulant serpin, protein Z-dependent protease
global stabilization, localized increased structural flexibility, and con- inhibitor (ZPI). J Biol Chem 2011; 286: 8740–51.
served hydrogen bonding at the active site. Biochemistry 1997; 36: 30 Olson ST, Bjork I. Predominant contribution of surface approxima-
5455–64. tion to the mechanism of heparin acceleration of the antithrombin-
10 Stein PE, Tewkesbury DA, Carrell RW. Ovalbumin and angiotensi- thrombin reaction. Elucidation from salt concentration effects. J Biol
nogen lack serpin S-R conformational change. Biochem J 1989; 262: Chem 1991; 266: 6353–64.
103–7. 31 Verhamme IM, Bock PE, Jackson CM. The preferred pathway of
11 Hood DB, Huntington JA, Gettins PG. Alpha 1-proteinase inhibitor glycosaminoglycan-accelerated inactivation of thrombin by heparin
variant T345R. Influence of P14 residue on substrate and inhibitory cofactor II. J Biol Chem 2004; 279: 9785–95.
pathways. Biochemistry 1994; 33: 8538–47. 32 Olson ST, Swanson R, Raub-Segall E, Bedsted T, Sadri M, Petitou M,
12 Engh RA, Huber R, Bode W, Schulze AJ. Divining the serpin inhi- Herault JP, Herbert JM, Bjork I. Accelerating ability of synthetic ol-
bition mechanism: a suicide substrate ÔspringeÕ? Trends Biotechnol igosaccharides on antithrombin inhibition of proteinases of the clotting
1995; 13: 503–10. and fibrinolytic systems. Comparison with heparin and low-molecular-
13 Lawrence DA, Ginsburg D, Day DE, Berkenpas MB, Verhamme IM, weight heparin. Thromb Haemost 2004; 92: 929–39.
Kvassman JO, Shore JD. Serpin-protease complexes are 33 Baglin TP, Carrell RW, Church FC, Esmon CT, Huntington JA.
trapped as stable acyl-enzyme intermediates. J Biol Chem 1995; 270: Crystal structures of native and thrombin-complexed heparin cofactor
25309–12. II reveal a multistep allosteric mechanism. Proc Natl Acad Sci U S A
14 Wilczynska M, Fa M, Karolin J, Ohlsson PI, Johansson LB, Ny T. 2002; 99: 11079–84.
Structural insights into serpin-protease complexes reveal the inhibitory 34 Qi X, Loiseau F, Chan WL, Yan Y, Wei Z, Milroy LG, Myers RM,
mechanism of serpins. Nat Struct Biol 1997; 4: 354–7. Ley SV, Read RJ, Carrell RW, Zhou A. Allosteric modulation of
15 Stratikos E, Gettins PG. Formation of the covalent serpin-proteinase hormone release from thyroxine and corticosteroid binding-globulins.
complex involves translocation of the proteinase by more than 70 A J Biol Chem 2011; 286: 16163–73.
and full insertion of the reactive center loop into beta-sheet A. Proc 35 Schreuder HA, de Boer B, Dijkema R, Mulders J, Theunissen HJ,
Natl Acad Sci U S A 1999; 96: 4808–13. Grootenhuis PD, Hol WG. The intact and cleaved human anti-
16 Huntington JA, Read RJ, Carrell RW. Structure of a serpin- thrombin III complex as a model for serpin-proteinase interactions.
protease complex shows inhibition by deformation. Nature 2000; 407: Nat Struct Biol 1994; 1: 48–54.
923–6. 36 Carrell RW, Stein PE, Fermi G, Wardell MR. Biological implications
17 Loebermann H, Tokuoka R, Deisenhofer J, Huber R. Human alpha 1- of a 3 A structure of dimeric antithrombin. Structure 1994; 2: 257–70.
proteinase inhibitor. Crystal structure analysis of two crystal modifi- 37 Langdown J, Belzar KJ, Savory WJ, Baglin TP, Huntington JA. The
cations, molecular model and preliminary analysis of the implications critical role of hinge-region expulsion in the induced-fit heparin binding
for function. J Mol Biol 1984; 177: 531–57. mechanism of antithrombin. J Mol Biol 2009; 386: 1278–89.
18 Huber R, Bode W. Structural basis of the activation and action of 38 Futamura A, Gettins PG. Serine 380 (P14) fi glutamate mutation
trypsin. Acc Chem Res 1978; 11: 114–22. activates antithrombin as an inhibitor of factor Xa. J Biol Chem 2000;
19 Calugaru SV, Swanson R, Olson ST. The pH dependence of serpin- 275: 4092–8.
proteinase complex dissociation reveals a mechanism of complex sta- 39 Huntington JA, Gettins PG. Conformational conversion of anti-
bilization involving inactive and active conformational states of the thrombin to a fully activated substrate of factor Xa without need for
proteinase which are perturbable by calcium. J Biol Chem 2001; 276: heparin. Biochemistry 1998; 37: 3272–7.
32446–55. 40 Huntington JA, Olson ST, Fan B, Gettins PG. Mechanism of heparin
20 Fredenburgh JC, Stafford AR, Weitz JI. Conformational changes in activation of antithrombin. Evidence for reactive center loop prein-
thrombin when complexed by serpins. J Biol Chem 2001; 276: 44828– sertion with expulsion upon heparin binding. Biochemistry 1996; 35:
34. 8495–503.
21 Bock PE, Olson ST, Bjork I. Inactivation of thrombin by antithrombin 41 Johnson DJ, Langdown J, Li W, Luis SA, Baglin TP, Huntington JA.
is accompanied by inactivation of regulatory exosite I. J Biol Chem Crystal structure of monomeric native antithrombin reveals a novel
1997; 272: 19837–45. reactive center loop conformation. J Biol Chem 2006; 281: 35478–86.

 2011 International Society on Thrombosis and Haemostasis


34 J. A. Huntington

42 Olson ST, Srinivasan KR, Bjork I, Shore JD. Binding of high affinity 62 Devlin GL, Bottomley SP. A protein family under ÔstressÕ – serpin
heparin to antithrombin III. Stopped flow kinetic studies of the stability, folding and misfolding. Front Biosci 2005; 10: 288–99.
binding interaction. J Biol Chem 1981; 256: 11073–9. 63 Janciauskiene S. Conformational properties of serine proteinase
43 Schedin-Weiss S, Richard B, Olson ST. Kinetic evidence that allosteric inhibitors (serpins) confer multiple pathophysiological roles. Biochim
activation of antithrombin by heparin is mediated by two sequential Biophys Acta 2001; 1535: 221–35.
conformational changes. Arch Biochem Biophys 2010; 504: 169–76. 64 Owen MC, Brennan SO, Lewis JH, Carrell RW. Mutation of anti-
44 Skinner R, Chang WS, Jin L, Pei X, Huntington JA, Abrahams JP, trypsin to antithrombin. alpha 1-antitrypsin Pittsburgh (358 Met leads
Carrell RW, Lomas DA. Implications for function and therapy of a 2.9 to Arg), a fatal bleeding disorder. N Engl J Med 1983; 309: 694–8.
A structure of binary-complexed antithrombin. J Mol Biol 1998; 283: 65 Mushunje A, Zhou A, Carrell RW, Huntington JA. Heparin-induced
9–14. substrate behavior of antithrombin Cambridge II. Blood 2003; 102:
45 Bray B, Lane DA, Freyssinet JM, Pejler G, Lindahl U. Anti-thrombin 4028–34.
activities of heparin. Effect of saccharide chain length on thrombin 66 Belorgey D, Hagglof P, Karlsson-Li S, Lomas DA. Protein misfolding
inhibition by heparin cofactor II and by antithrombin. Biochem J 1989; and the serpinopathies. Prion 2007; 1: 15–20.
262: 225–32. 67 Knaupp AS, Bottomley SP. Serpin polymerization and its role in
46 Rezaie AR. Calcium enhances heparin catalysis of the antithrombin- disease – the molecular basis of alpha1-antitrypsin deficiency. IUBMB
factor Xa reaction by a template mechanism. Evidence that calcium Life 2009; 61: 1–5.
alleviates Gla domain antagonism of heparin binding to factor Xa. J 68 Davies MJ, Lomas DA. The molecular aetiology of the serpinopathies.
Biol Chem 1998; 273: 16824–7. Int J Biochem Cell Biol 2008; 40: 1273–86.
47 Olson ST, Bjork I. Role of protein conformational changes, surface 69 Carrell RW, Lomas DA. Conformational disease. Lancet 1997; 350:
approximation and protein cofactors in heparin-accelerated anti- 134–8.
thrombin-proteinase reactions. Adv Exp Med Biol 1992; 313: 155–65. 70 Belorgey D, Irving JA, Ekeowa UI, Freeke J, Roussel BD, Miranda E,
48 Li W, Johnson DJ, Esmon CT, Huntington JA. Structure of the Perez J, Robinson CV, Marciniak SJ, Crowther DC, Michel CH,
antithrombin-thrombin-heparin ternary complex reveals the anti- Lomas DA. Characterisation of serpin polymers in vitro and in vivo.
thrombotic mechanism of heparin. Nat Struct Mol Biol 2004; 11: 857– Methods 2010; 53: 255–66.
62. 71 Yu MH, Lee KN, Kim J. The Z type variation of human alpha 1-
49 Rezaie AR. Partial activation of antithrombin without heparin antitrypsin causes a protein folding defect. Nat Struct Biol 1995; 2:
through deletion of a unique sequence on the reactive site loop of the 363–7.
serpin. J Biol Chem 2002; 277: 1235–9. 72 Takehara S, Zhang J, Yang X, Takahashi N, Mikami B, Onda M.
50 Langdown J, Johnson DJ, Baglin TP, Huntington JA. Allosteric Refolding and polymerization pathways of neuroserpin. J Mol Biol
activation of antithrombin critically depends upon hinge region 2010; 403: 751–62.
extension. J Biol Chem 2004; 279: 47288–97. 73 Dafforn TR, Mahadeva R, Elliott PR, Sivasothy P, Lomas DA. A
51 Yang L, Manithody C, Olson ST, Rezaie AR. Contribution of basic kinetic mechanism for the polymerization of alpha1-antitrypsin. J Biol
residues of the autolysis loop to the substrate and inhibitor specificity Chem 1999; 274: 9548–55.
of factor IXa. J Biol Chem 2003; 278: 25032–8. 74 JamesEL,WhisstockJC,GoreMG,BottomleySP.Probing the unfolding
52 Manithody C, Yang L, Rezaie AR. Role of basic residues of the pathway of alpha1-antitrypsin. J Biol Chem 1999; 274: 9482–8.
autolysis loop in the catalytic function of factor Xa. Biochemistry 2002; 75 Devlin GL, Chow MK, Howlett GJ, Bottomley SP. Acid denaturation
41: 6780–8. of alpha1-antitrypsin: characterization of a novel mechanism of serpin
53 Izaguirre G, Olson ST. Residues Tyr253 and Glu255 in strand 3 of polymerization. J Mol Biol 2002; 324: 859–70.
beta-sheet C of antithrombin are key determinants of an exosite made 76 Huntington JA, Sendall TJ, Yamasaki M. New insight into serpin
accessible by heparin activation to promote rapid inhibition of factors polymerization and aggregation. Prion 2009; 3: 12–4.
Xa and IXa. J Biol Chem 2006; 281: 13424–32. 77 Huntington JA, Whisstock J. Molecular contortionism- on the phys-
54 Izaguirre G, Zhang W, Swanson R, Bedsted T, Olson ST. Localization ical limits of serpin loop-sheet polymers. Biol Chem 2010; 391: 973–82.
of an antithrombin exosite that promotes rapid inhibition of factors Xa 78 Yamasaki M, Sendall TJ, Harris LE, Lewis GM, Huntington JA.
and IXa dependent on heparin activation of the serpin. J Biol Chem Loop-sheet mechanism of serpin polymerization tested by reactive
2003; 278: 51433–40. center loop mutations. J Biol Chem 2010; 285: 30752–8.
55 Johnson DJ, Li W, Adams TE, Huntington JA. Antithrombin-S195A 79 Yamasaki M, Li W, Johnson DJ, Huntington JA. Crystal structure of
factor Xa-heparin structure reveals the allosteric mechanism of anti- a stable dimer reveals the molecular basis of serpin polymerization.
thrombin activation. EMBO J 2006; 25: 2029–37. Nature 2008; 455: 1255–8.
56 Johnson DJ, Langdown J, Huntington JA. Molecular basis of factor 80 Ordonez A, Martinez-Martinez I, Corrales FJ, Miqueo C, Minano A,
IXa recognition by heparin-activated antithrombin revealed by a 1.7-A Vicente V, Corral J. Effect of citrullination on the function and con-
structure of the ternary complex. Proc Natl Acad Sci U S A 2010; 107: formation of antithrombin. FEBS J 2009; 276: 6763–72.
645–50. 81 Elliott PR, Pei XY, Dafforn TR, Lomas DA. Topography of a 2.0 A
57 Olson ST, Chuang YJ. Heparin activates antithrombin anticoagulant structure of alpha1-antitrypsin reveals targets for rational drug design
function by generating new interaction sites (exosites) for blood clot- to prevent conformational disease. Protein Sci 2000; 9: 1274–81.
ting proteinases. Trends Cardiovasc Med 2002; 12: 331–8. 82 McCoy AJ, Pei XY, Skinner R, Abrahams JP, Carrell RW. Structure
58 Carlstrom AS, Lieden K, Bjork I. Decreased binding of heparin to of beta-antithrombin and the effect of glycosylation on antithrombinÕs
antithrombin following the interaction between antithrombin and heparin affinity and activity. J Mol Biol 2003; 326: 823–33.
thrombin. Thromb Res 1977; 11: 785–97. 83 Herbert JM, Herault JP, Bernat A, Savi P, Schaeffer P, Driguez PA,
59 Huntington JA. Shape-shifting serpins – advantages of a mobile Duchaussoy P, Petitou M. SR123781A, a synthetic heparin mimetic.
mechanism. Trends Biochem Sci 2006; 31: 427–35. Thromb Haemost 2001; 85: 852–60.
60 Kaiserman D, Whisstock JC, Bird PI. Mechanisms of serpin dys- 84 Jin L, Abrahams JP, Skinner R, Petitou M, Pike RN, Carrell RW. The
function in disease. Expert Rev Mol Med 2006; 8: 1–19. anticoagulant activation of antithrombin by heparin. Proc Natl Acad
61 Gooptu B, Lomas DA. Conformational pathology of the serpins: Sci U S A 1997; 94: 14683–8.
themes, variations, and therapeutic strategies. Annu Rev Biochem 2009;
78: 147–76.

 2011 International Society on Thrombosis and Haemostasis

Das könnte Ihnen auch gefallen