throbber
Protein aggregation: folding aggregates, inclusion bodies
`and amyloid
`Anthony L Fink
`
`Review R9
`
`Aggregation results in the formation of inclusion
`bodies, amyloid fibrils and folding aggregates.
`Substantial data support the hypothesis that partially
`folded intermediates are key precursors to aggregates,
`that aggregation involves specific intermolecular
`interactions and that most aggregates involve βb sheets.
`
`Address: Department of Chemistry and Biochemistry, University of
`California, Santa Cruz, CA 95064, USA.
`E-mail: enzyme@cats.ucsc.edu
`
`Folding & Design 01 February 1998, 3:R9–R23
`http://biomednet.com/elecref/13590278003R0009
`
`© Current Biology Ltd ISSN 1359-0278
`
`Introduction
`Protein aggregation can be merely a nuisance factor in
`many in vitro studies of proteins or it can cause major
`economic and technical problems in the biotechnology
`and pharmaceutical industries. Its effects can be lethal in
`patients who suffer from a variety of diseases involving
`protein aggregation, such as the amyloidoses, prion dis-
`eases and other protein deposition disorders [1,2]. This
`review focuses on the basic mechanism(s) of protein aggre-
`gation, the factors that determine whether it will occur,
`and the conformation of the protein molecules in the
`aggregate. Protein aggregation is intimately tied to protein
`folding and stability, and also, in the cell, to molecular
`chaperones. The prevalence of protein aggregation is
`probably much higher than generally realized — it is often
`ignored or worked around, and in protein folding experi-
`ments its presence may not even be realized [3]. The
`growing recognition of the critical importance of protein
`aggregation has resulted in a number of reviews [4–10].
`
`Unless specifically noted to the contrary, in this review
`the term aggregation will apply to aggregated protein
`involving the formation of insoluble precipitates that may
`be considered ‘pathological’ in nature. This is in contrast
`to the insolubility of the native state due to protein con-
`centrations exceeding the solubility limit (e.g. ‘salting
`out’), or the intermolecular association involved in the for-
`mation of native oligomers. It should be noted that in
`many such cases of pathological aggregation the initial
`material formed may be soluble aggregates, but these
`become insoluble when they exceed a certain size.
`
`It is convenient to classify protein aggregation according
`to the following categories: in vivo and in vitro, and
`ordered and disordered. Amyloid fibrils (both in vivo and
`
`in vitro) are examples of ordered aggregates, whereas
`inclusion bodies are examples of in vivo disordered aggre-
`gates. Corresponding disordered in vitro aggregates are
`those formed during the refolding of denaturant-unfolded
`protein at high protein concentrations, or under weakly
`native conditions at high protein concentration; these will
`be referred to as folding aggregates.
`
`Native, folded proteins may aggregate under certain con-
`ditions, most notably salting out and isoelectric precipita-
`tion (when the net charge on the protein is zero). Such
`precipitates of native protein are readily distinguished
`from pathological aggregates by their solubility in buffer
`under native-like conditions. In contrast, pathological
`aggregates dissociate and dissolve only in the presence of
`high concentrations of denaturant or detergent. In my lab-
`oratory, it has been shown that the native conformation is
`retained in salting out precipitates (Figure 1).
`
`Protein aggregation has usually been assumed to involve
`either unfolded or native states. Inclusion body formation
`and other aggregates formed during protein folding have
`been assumed to arise from hydrophobic aggregation of
`the unfolded or denatured states, whereas amyloid fibrils
`and other extracellular aggregates have been assumed to
`arise from native-like conformations in a process analo-
`gous to the polymerization of hemoglobin S [8]. Recent
`observations suggest that aggregation is much more likely
`to arise from specific partially folded intermediates,
`however. An important consequence of this is that aggre-
`gation will be favored by factors and conditions that favor
`population of these intermediates, and hence it is the
`properties of these intermediates that are important in
`determining whether aggregation occurs. Furthermore,
`the characteristics and properties of the intermediates may
`be significantly different from those of the native (and
`unfolded) conformation.
`
`Several observations indicate that transient aggregation
`occurring during in vitro protein refolding may be mis-
`taken for a transient intermediate [3]. Direct evidence for
`the transient association of partially folded intermediates
`during refolding has been obtained in small-angle X-ray
`scattering experiments of apomyoglobin [11,12], carbonic
`anhydrase and phosphoglycerate kinase [13]. The experi-
`ments show the rapid (milliseconds or less) formation of
`associated states that become monomeric on a slow time-
`scale (typically seconds to minutes or longer). Another
`approach indicative of transient aggregation is that involv-
`ing changes in the rate constants for refolding as a function
`
`MYLAN INST. EXHIBIT 1096 PAGE 1
`
`MYLAN INST. EXHIBIT 1096 PAGE 1
`
`

`

`folding aggregates and amorphous deposits); how the envi-
`ronmental conditions affect the rate and the amount of
`aggregation; and how the aggregation may be prevented.
`
`It seems likely from an evolutionary perspective that pro-
`teins have evolved to avoid sequences that result in a
`strong propensity to aggregate. It is also interesting to con-
`sider that many short peptide sequences containing several
`hydrophobic residues and a high tendency for β-sheet
`formation probably have a strong disposition to form
`aggregates and/or amyloid fibrils. It is only the flanking
`sequences, which are either quite polar and therefore
`increase the solubility limit or are sufficiently bulky to
`sterically prevent the required interactions, that result in
`the lack of aggregation and/or amyloid formation.
`
`Problems due to protein aggregation
`Protein deposition diseases
`Several dozen protein deposition diseases are known.
`The most familiar include the amyloid diseases (amyloid-
`oses), such as Alzheimer’s disease, and the transmissible
`spongiform encephalopathies (TSEs; prion diseases such
`as bovine spongiform encephalopathy, BSE, or Mad Cow
`disease and Creutzfeldt–Jacob disease, CJD, in humans).
`In both amyloid and prion diseases the aggregated pro-
`tein is usually in the form of ordered fibrils. Amyloid fibril
`formation has been observed to arise from both peptides
`and proteins. Several protein deposition diseases involve
`non-ordered protein deposits; some examples are inclu-
`sion body myositis, light-chain deposition disease and
`cataracts. Many thousands of people die each year from
`protein deposition diseases [14]. New diseases are added
`to this list every year, one of the latest being Huntington’s
`disease [15].
`
`Inclusion bodies
`Inclusion body formation is very common when proteins
`are overexpressed. This may facilitate their potential
`purification because inclusion bodies are usually highly
`homogeneous. The problem is that renaturation is fre-
`quently difficult, as a result of aggregation. Several tech-
`niques have been developed to help overcome the
`common problem of their re-aggregation during renatura-
`tion [16] and some are discussed later. Inclusion bodies
`and related insoluble non-ordered protein aggregates are
`also found in certain diseases.
`
`R10 Folding & Design Vol 3 No 1
`
`Figure 1
`
`Absorbance
`
`1700
`
`1680
`
`1640
`1660
`Wavenumber (cm–1)
`
`1620
`
`1600
`
`Native IL-2
`Ammonium sulfate precipitated IL-2
`Aggregated IL-2
`
`Folding & Design
`
`Precipitates of native protein, in this case interleukin-2 (IL-2) formed by
`‘salting out’ with ammonium sulfate, retain the native conformation. The
`figure shows the second derivatives of the FTIR spectra of the amide I
`region of native, ammonium sulfate precipitated IL-2 (dashed line) and,
`for comparison, aggregated (inclusion body) IL-2 (dotted line). IL-2 is
`an all-α protein, as indicated by the dominant band at 1654 cm–1 for
`the native conformation.
`
`of protein concentration [3]. This study indicates that tran-
`sient aggregates can easily be mistaken for structured
`monomers and could be a general problem in time-resolved
`folding studies. Because aggregation is sensitive to protein
`concentration, monitoring the kinetics as a function of
`concentration should reveal potential aggregation artifacts.
`
`Key questions relating to aggregation, many of which are
`not yet fully answered, include the following: the nature
`of the species responsible for aggregation: the detailed
`mechanism that leads to aggregation and the underlying
`kinetics scheme; the structure of the aggregates; the speci-
`ficity of the intermolecular interaction (e.g. are the aggre-
`gates homogenous?); why aggregation (even of the same
`protein) sometimes leads to ordered aggregates (amyloid)
`and sometimes to disordered aggregates (inclusion bodies,
`
`Protein drugs
`Protein aggregation is also a problem in a number of other
`aspects of biotechnology; for example, during storage or
`delivery of protein drugs. There are several reports that
`protein aggregation can occur during lyophilization of pro-
`teins or during their subsequent rehydration, depending
`on the conditions (e.g. the water content of the system is
`critical [17–22]). Because in some cases it appears that the
`dehydration of proteins, which occurs in lyophilization,
`
`MYLAN INST. EXHIBIT 1096 PAGE 2
`
`MYLAN INST. EXHIBIT 1096 PAGE 2
`
`

`

`results in denaturation [23,24], it is probably the ensuing
`rehydration that leads to aggregation, due to the forma-
`tion of partially folded intermediates during the refolding
`(see below).
`
`Figure 2
`
`Native
`
`Review Protein aggregation Fink R11
`
`Ordered
`aggregates
`(amyloid fibrils)
`
`Amorphous
`aggregates
`(e.g. inclusion bodies)
`
`Folding & Design
`
`I
`
`Unfolded
`
`Mechanisms of aggregation
`One of the earliest and most prescient studies of protein
`aggregation was that of Goldberg and coworkers [25] on
`the enzyme tryptophanase, which revealed an intermedi-
`ate at moderate denaturant concentration that aggregated.
`Evidence for the potential specificity of aggregation was
`also observed in that the addition of other folded proteins
`did not affect the amount of aggregated tryptophanase.
`More recently, the idea that partially folded intermediates
`might be responsible for aggregation has been championed
`by King and coworkers [4,26,27] and Wetzel [6,8,10,28].
`Recent reports supporting the involvement of partially
`folded intermediates in the aggregation of several proteins
`suggests the generality of the phenomenon [29–31].
`
`Hypothesis
`Substantial data support the following model for the for-
`mation and structure of protein aggregates, in which spe-
`cific intermolecular interactions between hydrophobic
`surfaces of structural subunits in partially folded interme-
`diates are responsible for the aggregation. Key features of
`the model (Figure 2) are as follows. Protein folding
`involves intermediates, each consisting of an ensemble of
`closely related substates (the various substates of a given
`intermediate will be characterized by having common sec-
`ondary structure and most likely a common core of rela-
`tively native-like structure, with the remainder of the
`polypeptide chain disordered or in unstable structural
`units). The native state is formed by the sequential inter-
`action of substructural units, the building blocks (typically
`subdomains), which may be stable or metastable on their
`own, but are stabilized by the interactions with other such
`building blocks [32]. Formation of the native state
`involves the intramolecular interaction of the hydrophobic
`faces of structural subunits (Figure 3a). Specificity will
`arise from a variety of features, of which the geometric
`shape and extent of the hydrophobic patches, the con-
`straints of the polypeptide chain, and the presence of
`other structural subunits are probably the most important.
`Aggregation occurs when these hydrophobic surfaces
`interact in an intermolecular manner (Figure 3b). Thus,
`the initial stages of aggregation are quite specific in the
`sense that they involve the interaction of specific surface
`elements of the structural subunits of one molecule with
`‘matching’ hydrophobic surface areas of structural sub-
`units of a neighboring molecule. Three-dimensional prop-
`agation of this process leads to large aggregates. Initially,
`the aggregates (e.g. dimers and tetramers) will be soluble,
`but eventually their size will exceed the solubility limit.
`The fact that they may still have significant solvent-
`exposed hydrophobic surfaces would also minimize their
`
`The basic model for protein aggregation. The circled I represents a key
`partially folded intermediate, which can be populated either in the
`folding direction (from unfolded) or from the native state. The
`intermediate I has a strong propensity to aggregate, leading to either
`ordered or disordered (amorphous) aggregates. It is possible that
`there is very local order even in apparently (macroscopically)
`amorphous aggregates. Intermediate I will normally be an intermediate
`on both the in vivo and in vitro folding pathways.
`
`solubility. The intermediates are more prone to aggregate
`than the unfolded state because in the unfolded state the
`hydrophobic sidechains are scattered relatively randomly
`in many small hydrophobic regions, whereas in the par-
`tially folded intermediates there will be large patches of
`contiguous surface hydrophobicity, which will have a
`much stronger propensity for aggregation (these are the
`surfaces that ‘normally’ interact in an intramolecular
`manner to form the native conformation). The role of
`domain (or subdomain) swapping [33] in aggregation is
`unclear, but it is certainly possible that in some cases it
`may be an important factor.
`
`Aggregation often appears to be irreversible, but this is
`usually a reflection of the very slow rates of disaggregation
`and the fact that the equilibrium lies far in favor of the
`aggregate rather than its soluble monomeric form. Under
`certain conditions, aggregates, including in vivo amyloid
`deposits, can be reversed [34,35]. In practice, however,
`once insoluble aggregates form, the process is effectively
`irreversible under native-like conditions.
`
`A number of observations suggest that specific inter-
`actions are involved in protein aggregation. Among the
`clearest evidence for such specificity is the work of
`Brems and coworkers [36–40] with bovine growth hor-
`mone (bGH), which showed that a peptide fragment
`of bGH could inhibit aggregation and that mutations,
`which increased the hydrophobicity of the domain inter-
`face, increased the propensity for aggregation. Another
`strong indication of specificity in aggregation is the
`homogeneity of inclusion bodies [41], which are normally
`highly homogeneous once any contaminating material has
`
`MYLAN INST. EXHIBIT 1096 PAGE 3
`
`MYLAN INST. EXHIBIT 1096 PAGE 3
`
`

`

`R12 Folding & Design Vol 3 No 1
`
`Figure 3
`
`(a)
`
`Unfolded
`
`Intermediates
`
`(b)
`
`Unfolded
`
`Proposed mechanisms for aggregation
`involving structural building blocks
`(subdomains). See text for more details.
`(a) The normal folding situation, in which
`hydrophobic surfaces (black) of the structural
`units interact in an intramolecular manner to
`form the native conformation. Each identifiable
`intermediate along the pathway will consist of
`an ensemble of substates, having in common
`the compact structural unit(s), with the
`remainder of the chain in a disordered or
`unstable structured form. (b) The situation in
`which aggregation occurs by intermolecular
`interaction of the structural building blocks,
`again via interaction of complementary
`hydrophobic surfaces. The common
`intermediate in folding and aggregation can
`be seen.
`
`Native
`
`Aggregates
`
`Folding & Design
`
`been removed [41]. Similar arguments apply to amyloid
`fibril formation.
`
`Factors favoring aggregation
`Circumstances that lead to the population of partially
`folded intermediates, especially if their concentration is
`high, are thus likely to lead to aggregation; these circum-
`stances include mutations that lead to differential destabi-
`lization of the native state relative to the partially folded
`intermediate, as well as environmental conditions. Conse-
`quently, the major factors that determine whether a
`protein will aggregate, and the extent and rate of the
`aggregation, are: the protein amino acid sequence, the
`pH, the temperature and ionic strength, the concentration
`of the protein, the presence of cosolutes (e.g. denaturants
`such as urea, other chaotropes or kosmotropes — includ-
`ing osmolytes — and ligands that interact selectively with
`
`either the native or non-native conformations of the pro-
`tein or the aggregated form), and the presence (or absence)
`of various molecular chaperones. In fact, a major role of
`chaperones is to prevent the potential aggregation of newly
`synthesized proteins, or those denatured through stress.
`
`In most instances of aggregation, especially in vivo, there
`is a kinetic competition between aggregation and other
`processes, such as folding (Figure 4). The environmental
`conditions and the protein concentration significantly affect
`the degree and rate of intermolecular association. The
`protein concentration enters the equation because aggre-
`gation minimally involves a second-order kinetic process
`(although the growth of amyloid fibrils, and other aggre-
`gates, may appear first order). In the context of physiologi-
`cal aggregation, the potential role of post-translational
`processing may be critical. It is likely that in many cases
`
`MYLAN INST. EXHIBIT 1096 PAGE 4
`
`MYLAN INST. EXHIBIT 1096 PAGE 4
`
`

`

`Figure 4
`
`Degradation (proteolysis)
`
`Aggregates
`
`Partially
`folded
`intermediate(s)
`
`Native
`protein
`
`Nascent
`polypeptide
`
`Binding to
`chaperones
`
`Folding & Design
`
`Aggregation usually involves kinetic competition. During in vivo protein
`synthesis, for example, partially folded intermediates are divided
`between pathways leading to spontaneous folding and the native state,
`aggregation, binding to chaperones and proteolytic degradation.
`
`when aggregation occurs from a solution of the native
`protein it is the partially folded intermediates in equilib-
`rium with the native state that are the immediate precursors
`of the aggregates.
`
`The hypothesis that aggregation arises from partially
`folded intermediates explains the apparent lack of correla-
`tion between protein stability and aggregation that is some-
`times observed [42]. Thus, if the decreased stability of the
`native state of a mutant form also destabilizes the partially
`folded intermediate that is responsible for aggregation then
`a correlation may be observed. On the other hand, if desta-
`bilization of the native conformation increases the popula-
`tion of a partially folded intermediate that aggregates then
`increased aggregation will be observed and there will be no
`apparent correlation with the native-state stability.
`
`The propensity for a given protein to aggregate, either in
`vivo or in vitro, may well be determined in part by the
`lifetime of partially folded intermediates. Longer lived
`intermediates are more likely to lead to aggregation for
`two reasons: first, there is a greater chance of interaction
`with another such partially folded intermediate, and
`second, in the in vivo situation, the molecular chaper-
`ones involved in preventing aggregation by sequestering
`the partially folded intermediate may become saturated,
`and thus there will not be enough free chaperones avail-
`able to bind to additional newly synthesized protein.
`Several observations indicate that in vivo and in vitro
`aggregation during folding give rise to similar aggregates,
`suggesting that it is likely that a common partially folded
`
`Review Protein aggregation Fink R13
`
`intermediate is responsible for both types of aggregates
`[26,30,43,44].
`
`In vivo aggregates: inclusion bodies
`It is noteworthy that inclusion body formation is found
`not only in prokaryote and eukaryote cells, but also for
`both heterologous and homologous overexpression. This
`emphasizes that it is the overexpression itself that is
`responsible for the aggregation. For example, in vivo
`aggregation of β-lactamase is only observed when the rate
`of expression exceeds 2.5% of the total protein synthesis
`rate [45,46]. This mirrors the observations with in vitro
`refolding systems, which show that aggregation increases
`as the protein concentration increases [44]. Despite the
`fundamental and practical importance of inclusion bodies,
`rather little is known about their structures and mecha-
`nisms of formation. No correlation has been found
`between inclusion body formation in recombinant pro-
`teins and a wide variety of factors, including size and
`hydrophobicity, although some correlation with average
`charge and fraction of turn-forming residues has been
`observed [47]. Inclusion bodies are frequently refractory
`to renaturation. To obtain functional protein requires
`denaturation and solubilization, often with disulfide
`reducing agents, and subsequent renaturation. It is fre-
`quently observed that the only way to renature significant
`amounts of soluble material is to use extremely low pro-
`tein concentrations to avoid aggregation. As a result, inclu-
`sion bodies are a major problem in biotechnology and the
`development of protein drugs [48].
`
`Although many inclusion bodies are refractile in phase
`contrast microscopy, some large insoluble aggregates in
`Escherichia coli are not refractile, and have been called
`‘floccule-type’ inclusion bodies [49]. At least in some
`cases, these appear to be less tightly packed, and more
`easily solubilized than classical inclusion bodies. It is pos-
`sible that these aggregates arise from native protein of
`very limited solubility, rather than partially folded inter-
`mediates. We have noticed that there is a range in the
`denaturant solubility of in vivo insoluble proteins: classi-
`cal inclusion bodies are relatively resistant to solubiliza-
`tion, whereas some insoluble material is much more
`readily dissolved in denaturant and is clearly morphologi-
`cally different (S. Seshadri and A.L.F., unpublished
`observations). Morphological differences have also been
`observed between cytoplasmic and periplasmic inclusion
`bodies, reflected in differences in denaturant solubility
`and protease resistance [50].
`
`Several factors have been suggested to lead to inclusion
`body formation, including: the high local concentration
`of protein; a reducing environment in the cytoplasm; lack
`of post-translational modifications; improper interactions
`with chaperones and other enzymes involved in in vivo
`folding; and intermolecular crosslinking via disulfides
`
`MYLAN INST. EXHIBIT 1096 PAGE 5
`
`MYLAN INST. EXHIBIT 1096 PAGE 5
`
`

`

`R14 Folding & Design Vol 3 No 1
`
`(although inclusion bodies can form from proteins
`lacking cysteine). Three possible mechanisms for inclu-
`sion body formation have been proposed: aggregation of
`native protein of limited solubility; aggregation of the
`unfolded state; and aggregation of partially folded inter-
`mediate states. Until recently, there was no evidence for
`any regular structure in inclusion bodies and the prevail-
`ing view was that inclusion body formation was an aggre-
`gation process mediated by non-specific interactions in
`the unfolded state [10].
`
`In a series of elegant studies on the tailspike endorhamno-
`sidase of phage P22, King and coworkers [43,51–55] have
`shown that partially folded intermediates are responsible
`for the aggregation of tailspike protein as well as inclusion
`body formation in this unique system. Their results are
`consistent with a model in which aggregate formation,
`both in vivo and in vitro, occurs by the specific association
`of a partially folded intermediate that associates preferen-
`tially in an intermolecular fashion to form aggregates,
`rather than intramolecular association leading to the native
`conformation. A series of multimeric partially folded inter-
`mediates, representing early stages of the aggregation
`pathway for the P22 tailspike protein, have been trapped
`in the cold and isolated by nondenaturing polyacrylamide
`gel electrophoresis [26,27]. It should be noted that the P22
`tailspike protein system is an unusual one and not typical
`of most globular proteins, although similar ladders of mul-
`timers were observed with the P22 coat protein and car-
`bonic anhydrase II [27]. Monoclonal antibodies against
`tailspike chains that discriminate between folding inter-
`mediates and native states were used to distinguish pro-
`ductive folding intermediates and off-pathway aggregation
`intermediates. The aggregation intermediates displayed
`epitopes in common with productive folding intermedi-
`ates, but were not recognized by antibodies against native
`epitopes [55]. Because many other protein expression
`systems also show increased inclusion body formation as
`the temperature is raised, it is likely that ‘thermolabile’
`intermediates are common precursors to inclusion bodies
`at elevated temperatures.
`
`There are several protein systems in which it has been
`shown that point mutations may dramatically affect the
`amount of aggregate (inclusion body) formation; these
`include the P22 tailspike protein [53], interferon-γ [56],
`colicin A [57], Che Y [58], and interleukin-1β (IL-1β)
`[56,59]. For example, with the human IL-1β expression
`system in E. coli, the amount of inclusion body ranges
`from close to 0% for the native to close to 100% for the
`Leu10→Thr and Lys97→Val mutants [60]. No strong
`correlations were observed between the extent of inclu-
`sion body formation and either thermodynamic or thermal
`stability. This and other investigations suggest that the
`tendency of at least some proteins to form inclusion
`bodies is related to the stability or solubility of folding
`
`intermediates rather than native states [42]. The in vivo
`effects of sequence and growth temperature of IL-1β, as
`manifested by inclusion body formation, were quantita-
`tively reproduced in an in vitro system, indicating not
`only that inclusion body formation is based on the intrin-
`sic properties of the protein sequence, but also that inter-
`actions with chaperones or other cellular factors were not
`significant [59].
`
`Consistent with the proposed model, point mutations in
`the hydrophobic face of the immunoglobulin light chain,
`in which polar residues were introduced, significantly
`decreased the fraction of inclusion body formed in an
`E. coli expression system [61]. The potentially strong
`dependence of inclusion body formation on point muta-
`tions may also be a reflection of specificity in aggregation.
`
`We have used attenuated total reflectance (ATR) FTIR
`to examine the structure in various aggregated protein
`forms, including inclusion bodies [30]. These included
`all-β, all-α and α+β proteins. Some common features
`were observed. First, both the inclusion bodies and the
`folding aggregates exhibited substantial secondary struc-
`ture, typically 50–70% of that of the native conformation.
`We interpret this to mean that the inclusion bodies arose
`from the association of partially folded intermediates con-
`taining substantial native-like structure. Second, the
`structure of a given protein in inclusion bodies, refolding
`aggregates, and thermal aggregates (formed by heating a
`concentrated protein solution to a temperature just below
`the beginning of the thermal denaturation transition) is
`the same (Figure 5). Thus, we conclude that a given
`protein will have one partially folded intermediate that is
`particularly prone to aggregate, and consequently most
`aggregates of that protein will arise from that intermediate
`and thus have similar structures.
`
`Furthermore, in all cases, even for all-β proteins, signifi-
`cant new β structure, compared to that in the native confor-
`mation, was observed; typically this amounted to ~20–25%
`of the total secondary structure. We take this to indicate
`that the intermolecular interactions leading to the aggrega-
`tion involve β-sheet-like interactions. The exact nature of
`the intermolecular interactions is unknown, and could be
`different in different aggregates; for example, the number
`of strands in the β sheet may vary. What is clear, however,
`is that aggregation usually leads to an increase in the
`amount of secondary structure with IR components in the
`1623–1637 cm–1 region, which corresponds to the region
`where β-sheet structure is observed. Interestingly, the
`amount of secondary structure in the inclusion body varies
`from one protein to another, as does the amount of disor-
`dered structure. Figure 5 shows the spectra of the native
`inclusion body and folding aggregate forms of IL-2, illus-
`trating the dramatic increase in β-sheet content compared
`to the native conformation.
`
`MYLAN INST. EXHIBIT 1096 PAGE 6
`
`MYLAN INST. EXHIBIT 1096 PAGE 6
`
`

`

`Review Protein aggregation Fink R15
`
`the inclusion bodies were similar regardless of whether
`they were localized in the cytoplasm or periplasm.
`
`In vivo aggregates: amyloid
`A number of human diseases involve the deposition of pro-
`tein aggregates; a subset of these, known as amyloidoses,
`are lethal diseases involving the extracellular deposition of
`amyloid fibrils and plaques. Amyloid plaque is usually
`defined by three features: characteristic birefringent stain-
`ing using Congo Red, fibrous morphology observed by
`electron microscopy, and a distinctive X-ray fiber diffrac-
`tion pattern similar to that of certain insect silks, consistent
`with high β-sheet content (the so-called cross-β structure).
`Amyloid fibrils are typically 7–12 nm in diameter and can
`be dissociated by high concentrations of denaturant.
`
`The molecular mechanisms leading to pathological amy-
`loid formation are not well understood. In most forms of
`amyloidosis the formation of amyloid appears to be caused
`by a combination of factors, including high concentrations
`of protein, proteolysis, mutations, and other unknown
`factors. Certain other molecules are frequently found in
`association with amyloid plaques, namely serum amyloid
`P component [63], proteoglycans [64], apolipoprotein E
`[65] and metal ions. The role of these molecules, if they
`have one, in amyloid deposition is unclear at present.
`There have been reports that these and other ligands may
`increase or decrease aggregation (e.g. [66,67]).
`
`Evidence to support partially folded intermediates as
`amyloid fibril precursors has been found in at least two
`systems. Transthyretin (TTR) is involved in two amyloid
`diseases, familial amyloidotic neuropathy and senile sys-
`temic amyloidosis. A partially folded monomeric interme-
`diate, populated at low pH, has been implicated in in vitro
`amyloid formation from transthyretin [68,69]. Recently,
`two naturally occurring variants of human lysozyme have
`been found to be amyloidogenic [70]. Both mutants are
`significantly less stable than the wild-type protein and
`have populated partially folded intermediates at 37°C that
`are assumed to be the precursors of amyloid. The ther-
`mally aggregated mutants showed an increased level of
`β structure compared to the native conformation.
`
`Some mutations leading to amyloid-fibril formation (in
`transthyretin [71,72], lysozyme [70] and immunoglobulin
`variable light, VL, chain domains [73]) are also observed to
`result in decreased stability of the native state. Presumably,
`in such cases there is differential destabilization of the
`native conformation compared to the aggregating intermedi-
`ate conformation that results in population of the aggregat-
`ing conformation. It has been shown, in an in vitro system,
`that point mutations may convert a non-amyloidogenic pro-
`tein into an amyloidogenic one [29,73]. In Alzheimer’s
`disease, the Aβ40 peptide is significantly less amyloidogenic
`than the Aβ42 peptide [74]. The intermediacy of partially
`
`Figure 5
`
`Absorbance
`

`

`
`1700
`
`1680
`
`1640
`1660
`Wavenumber (cm–1)
`
`1620
`
`1600
`
`Inclusion bodies
`Folding aggregates
`Native protein
`
`Folding & Design
`
`Aggregation results in a substantial increase in β structure compared
`to the native conformation. This figure shows the second derivatives of
`the FTIR spectra of the amide I region of IL-2. The aggregated forms,
`inclusion bodies and folding aggregates, show 25% more β structure
`than the native state. The spectra of the two aggregated forms are very
`similar, indicating similar underlying secondary structure.
`
`Our results, along with those of Georgiou and coworkers
`[44,62] on β-lactamase (see below), suggest that the gener-
`ation of insoluble aggregates during refolding from denat-
`urant can be considered to be an in vitro model system for
`inclusion body formation.
`
`The secondary structure of β-lactamase (an α+β protein)
`inclusion bodies in E. c

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket