Skip to main content
Interface Focus logoLink to Interface Focus
. 2017 Oct 20;7(6):20170030. doi: 10.1098/rsfs.2017.0030

Factors affecting the physical stability (aggregation) of peptide therapeutics

Karolina L Zapadka 1, Frederik J Becher 1, A L Gomes dos Santos 2,, Sophie E Jackson 1,
PMCID: PMC5665799  PMID: 29147559

Abstract

The number of biological therapeutic agents in the clinic and development pipeline has increased dramatically over the last decade and the number will undoubtedly continue to increase in the coming years. Despite this fact, there are considerable challenges in the development, production and formulation of such biologics particularly with respect to their physical stabilities. There are many cases where self-association to form either amorphous aggregates or highly structured fibrillar species limits their use. Here, we review the numerous factors that influence the physical stability of peptides including both intrinsic and external factors, wherever possible illustrating these with examples that are of therapeutic interest. The effects of sequence, concentration, pH, net charge, excipients, chemical degradation and modification, surfaces and interfaces, and impurities are all discussed. In addition, the effects of physical parameters such as pressure, temperature, agitation and lyophilization are described. We provide an overview of the structures of aggregates formed, as well as our current knowledge of the mechanisms for their formation.

Keywords: physical stabilities, therapeutics, peptides

1. Introduction

The aggregation of peptides is one of the most common and troubling processes encountered in almost all phases of biological drug development (table 1) [32]. Aggregation can take several different forms and the term is used to describe a number of different processes during which peptide molecules associate into larger species consisting of multiple polypeptide chains (figure 1). Aggregates can be amorphous [33] or highly structured, e.g. amyloid fibrils [34], and can form in solution or on surfaces due to adsorption [3537]. They can arise as a result of the non-covalent association of polypeptide chains, or from covalent linkage of chains. In some cases, aggregation is reversible while in others it is effectively irreversible. In either case, it reduces the physical stability of the peptide in question, not only leading to a loss in activity but also other critical problems such as toxicity and immunogenicity [3840].

Table 1.

Some examples of peptide and protein pharmaceuticals that form different types of aggregate.

peptide or protein length (residues) conditions type of aggregate references
GLP-1(7–37) 31 pH 7.5–8.2 fibril [1]
GLP-1(7–36)NH2 30 pH 7.4 fibril [2]
glucagon 29 pH 2.0 fibril [35]
glucagon 29 pH 8.5–9.7 fibril [6,7]
salmon calcitonin 32 neutral pH fibril [811]
amylin 37 under different conditions fibril [1216]
insulin 51 (31 chain A, 20 chain B) pH 2.0 fibril [1719]
pH 1.1–3.2 fibril [2023]
pH 1.5–7.5 fibril [24]
insulin 51 (31 chain A, 20 chain B) pH 1.75–2.0
conc < 5 mg ml−1
conc > 5 mg ml−1

spherulite
fibril
[25]
insulin analogues 51 (31 chain A, 20 chain B) high temperature and mechanical agitation precipitation [26]
human parathyroid hormone 84 fibril [27]
human interferon a2b chemical degradation-oxidation covalent aggregate [28,29]
Proleukin +SDS formation of micellar-like structures containing approx. 27 molecules [30]
Met-G-CSF pH 6.9, 37°C precipitation [31]

Figure 1.

Figure 1.

Amorphous and amyloid: atomic force microscopy images of aggregates of GLP-1 formed under different conditions. (a) Amorphous aggregates of 300 µM GLP-1 in water formed after incubation at room temperature for one week. (b) Fibrils of 25 µM GLP-1 in 25 mM sodium citrate at pH 3.5 formed after 100 h of incubation at 37°C. (c) Fibrils of 150 µM GLP-1 in sodium phosphate buffer at pH 7.5 formed after incubation for 120 h at 37°C. (d) Oligomers formed after dilution of GLP-1 fibrils in ddH2O. Fibrils of 150 μM GLP-1 were first formed in sodium phosphate buffer at pH 7.5 formed after 200 h incubation at 37°C then diluted 100 fold in ddH2O prior to imaging. (e) Fibrils of 150 µM GLP-1 in 25 mM Tris at pH 8.5 formed after incubation for 144 h at 37°C. (f) Fibrils of 150 µM GLP-1 in 25 mM Tris, pH 8.5 formed after incubation for 144 h at 37°C, and then diluted 10-fold in ddH2O prior to imaging. The scale bars represent 1 µm. (Online version in colour.)

In recent years, numerous and extensive research programmes have been undertaken and significant progress been made increasing our understanding of the mechanisms by which peptide and protein aggregation occurs and many of the factors that influence the reaction. However, much of the research has focused on peptides and proteins where aggregation in vivo is associated with disease states, particularly neurodegenerative diseases [41,42]. A considerable amount of the research on disease-related peptide and protein aggregation can be applied to the aggregation of peptide-based therapeutic agents currently being developed by the pharmaceutical industry. In general, however, the aggregation pathways of peptides currently coming through drug pipelines have not been characterized in depth.

In this review, we discuss the factors that are known to affect the aggregation of peptides rather than proteins. We focus here mainly on peptides that do not adopt a stable tertiary structure in the absence of their receptor/target proteins. The main differences between peptides, usually relatively unstructured as monomers, versus proteins, which usually form highly structured monomers sometimes within a larger oligomeric state/complex, is that for the latter any factor known to affect the thermodynamic or kinetic stability of the native structure of the monomer and/or complex also affects the aggregation propensity of the protein. There is extensive literature covering this area and for interested readers the following reviews are recommended: influence of thermodynamic and kinetic stability of native states [4345], the effect of oligomer formation [4649], the binding of ligands [50,51]. Within this review, wherever possible, we use therapeutically relevant peptides as examples, but we also refer to peptides such as Aβ or intrinsically disordered proteins, such as α-synuclein and tau, whose aggregation is associated with diseases which have been extensively studied. In the occasional case, where there are no suitable peptide-based examples, we use a highly structured protein to illustrate the point.

First, we describe the types of aggregation processes that have been observed and the structure and properties of the different aggregates, as well as some of the fundamental insights that have been gained on the mechanisms of aggregation. Then, we discuss in more detail the factors known to affect aggregation, and how these factors might influence the aggregation of peptide therapeutics. Finally, at the end of the review, we briefly discuss strategies for preventing or minimizing aggregation and also highlight those areas where research has been lacking and many questions remain.

1.1. Aggregation: the formation of highly structured amyloid fibrils

Many peptides and proteins, under the right conditions, have been shown to self-assemble into highly structured amyloid fibrils (figure 2) [34]. These fibrils contain a cross-β-sheet structure comprising tightly interacting intermolecular β-sheets [52]. Each sheet is formed from many molecules of the peptide each molecule adopting a very similar and highly regular β-strand conformation (figure 2). The sheets are stabilized by backbone hydrogen bonds that result in a spacing between the β-strands of 4.8 Å and are parallel to the fibril axis [34,52]. A pair of β-sheets is known as a protofilament. Two or more protofilaments interdigitate and intertwine with a spacing of 10–11 Å [5355] to form a mature fibril. These fibrils are stabilized by the hydrophobic effect, van der Waals interactions, π–π stacking and hydrogen bonds. A number of different cross-β-structures can form and are classified according to whether their β-strands are parallel or anti-parallel, and whether the packing of the β-sheets occurs face-to-face or face-to-back, and whether the β-sheets are organized parallel or anti-parallel to each other (figure 2). For an excellent review on the different structures that have been observed and further details of the packing of the side chains within the different structures see Riek & Eisenberg [34].

Figure 2.

Figure 2.

Structure of amyloid-like fibrils. (a) Structure of the mature fibril formed by Aβ (pdb code: 3ZPK). (b) Two views of the atomic resolution β-structure within the amyloid fibrils formed by the GYMLGS peptide from human prion protein (pdb code: 3NHC). The top view shows the juxtaposition of side chains looking down the main axis of the fibril. The bottom view shows the anti-parallel nature of the β-sheet formed and packing of side chains from adjacent β-strands. (c) Two views of the atomic resolution showing the cross β-structure within amyloid fibrils formed by the AIIGLM peptide from Aβ (pdb code: 2Y3J). The top view shows the stacking of side chains of the same type from the same residue looking down the main axis of the fibril while the bottom view shows the parallel arrangements of the β-strands in the structure. See [34] for an excellent review on atomic resolution structures of amyloid fibrils. Figures were made using PyMol. (Online version in colour.)

It is interesting to note that, although many amyloid fibrils have been shown to be associated with disease states and are therefore termed pathological fibrils, nature has exploited this type of structure and there are now a considerable number of functional fibrils [56].

1.2. Aggregation to form other types of structures

1.2.1. Higher-order fibrillar structures

Higher-order structures comprised of amyloid-like fibrils have been observed for a number of proteins. Insulin and β-lactoglobulin have both been shown to form spherulites, spherical structures consisting of radiating fibrils from a central core [57]. In addition, a number of peptides that self-assemble into fibrillar structures have also been shown to form hydrogels [58]. These types of higher-order fibrillar structures are beyond the scope of this review and will not be discussed further.

1.2.2. Amorphous aggregates

Amorphous aggregates, also known as disordered aggregates, are unstructured and there are no ordered intermolecular interactions [33]. They are often granular in appearance when imaged by atomic force microscopy, figure 1, and can vary in size and solubility. When insoluble they form precipitates which present a major challenge in biotechnology.

1.3. Mechanisms of aggregation

Aggregation is a complex phenomenon and can occur from many different states (figure 3). In many cases, aggregation occurs from an unfolded or largely unstructured state, i.e. in the case of many peptides and intrinsically disordered proteins. For amyloid fibril formation, the monomeric peptides first self-associate to form oligomeric structures which go on to form a critical nucleus (figure 4). Elongation of the nucleus is rapid relative to its formation giving rise to the sigmoidal aggregation kinetics often observed (figure 5a). For structured proteins, aggregation can also occur from partially structured folding intermediates or, as is the case for a number of monoclonal antibodies, self-association of the native state can take place which gives rise to inactive oligomers [59,60].

Figure 3.

Figure 3.

General aggregation processes from relatively unstructured peptides. Peptides can aggregate to form amorphous (left) or highly structured amyloid-like fibrils (right). This process can be induced by a solid–water (surface), or water–air interface or occur in solution. Many intrinsic and extrinsic factors affect the conformational equilibria (β-strand shown as a green arrow, α-helix shown as a red cylinder), including chemical degradation/modifications (shown as a purple circle). Fragmentation can also lead to peptides with a high propensity to aggregate. (Online version in colour.)

Figure 4.

Figure 4.

Mechanisms of amyloid fibril formation from monomeric peptides. (a) Shows a classic nucleation–polymerization mechanism where a nucleus (shown by three light blue circles in a rectangular box) which may be thermodynamically unfavourable and/or slow to form from monomer (red circle) is followed by a relatively rapid elongation step where additional monomers add to the end of the nucleus or fibril. In some cases, this is followed by secondary nucleation steps—nucleation on the surface of an existing fibril to form an additional nucleus (top) which is then followed by further elongation steps. Alternatively, a fibril can fragment to create two shorter fibrils thus increasing the number of nuclei and therefore elongation/growth sites (bottom). Fragmentation is followed by further elongation steps. In this mechanism, a number of on-pathway oligomers may be formed prior to the nucleus. In this case, classic kinetics where the lag time and t1/2 decrease with increasing monomer concentration are usually observed (figure 5b–d). (b) Shows an alternative mechanism in which the on-pathway nucleation–polymerization (shown by the light blue species) is in competition with the formation of off-pathway oligomeric species (green circles). If there is a unimolecular step such as a conformational change on the direct pathway to fibrils, then unusual aggregation kinetics can be observed (figure 5eg). (Online version in colour.)

Figure 5.

Figure 5.

Aggregation kinetics: formation of amyloid-like fibrils. (a) Typical sigmoidal kinetics frequently observed in the aggregation of peptides and proteins into amyloid-like fibrils. Aggregation is usually probed by the increase in fluorescence of the dye thioflavin T (ThT) on binding to fibrils containing regular β-sheet structure. The curve shows the key kinetic parameters that can be obtained from a fit of the aggregation data to a simple sigmoidal function. Three kinetic parameters are obtained from the fit of the data to a simple sigmoidal function: t1/2, the time at which the ThT fluorescence is 50% of its final value (red dashed line), the apparent growth rate which corresponds to the gradient of the steepest part of the rapid growth phase (blue dotted line), and tlag, the lag time at which point sufficient nuclei are present in solution for elongation and growth of fibrils to become rapid. (b) Typical kinetics for a reaction following a simple nucleation–polymerization mechanism with only on-pathway oligomeric intermediate states. This figure illustrates how the t1/2 and tlag decrease with increasing peptide/protein concentration (dark purple, low concentrations, to light blue, high concentrations, of peptide; the increasing peptide concentration is also shown by the grey arrow). The data are simulated using KinTek using Model 1 in [1]. (c,d) Kinetic parameters for the aggregation of GLP-1 into amyloid fibrils at pH 8.2 taken from [1] under conditions where the peptide follows a simple nucleation–polymerization mechanism with on-pathway oligomers only. Dependence of (c) t1/2 and (d) tlag on peptide concentration. (e) Simulated data to show how the aggregation kinetics change when off-pathway oligomeric intermediates are populated as well as on-pathway intermediate species and where there is a unimolecular step on the pathway to fibril formation (dark purple, low concentrations, to light blue, high concentrations, of peptide; the increasing peptide concentration is also shown by the grey arrow). KinTek is used, see Model 3 in [1]. (f,g) Kinetic parameters for the aggregation of GLP-1 into amyloid fibrils at pH 7.5 taken from [1] under conditions where the populates on and off-pathway oligomers. Dependence of (f) t1/2 and (g) tlag on peptide concentration. (Online version in colour.)

2. Factors affecting the physical stability of peptides

2.1. Peptide concentration

One of the most important factors influencing the physical stability of peptides is peptide concentration. This is long-established through numerous kinetic studies on the formation of amyloid fibrils [6165], amorphous aggregates [66,67] and adsorption on surfaces [68]. With respect to the formation of amyloid-like fibrils, extensive kinetic studies have been performed over wide ranges of peptide concentrations and, in the majority of cases, the kinetics have been shown to be consistent with a nucleation–polymerization mechanism (figure 4) [69]. In this case, the monomeric starting material is converted into oligomers, and ultimately a species that can act as a nucleus for rapid fibril growth. Addition of usually monomers to a nucleus or growing fibrils occurs rapidly to form protofibrils and mature fibrils. The essential features of such systems are that the kinetics show a lag phase, a rapid elongation or growth phase and a plateau. During the lag phase, a number of oligomeric species, including nucleating species, and short amyloid fibrils form until a point is reached at which the growth of fibrils is rapid (figure 4). The formation of fibrils then slows as a plateau is reached which represents either the point where monomers are completely depleted or an equilibrium point at which mature fibrils are in equilibrium with starting monomers. The sigmoidal kinetic traces can be fitted to obtain key kinetic parameters such as lag time t1/2 and the growth rate (figure 5a). How these kinetic parameters depend upon peptide concentration has been established for many systems, including Aβ [70], α-synuclein [71], tau [72], glucagon-like peptide-1 (GLP-1) [1] and other peptides. The majority of peptides and conditions show kinetics typical of nucleation–polymerization mechanisms, and in these cases, generally both t1/2 and tlag decrease with increasing peptide concentration as shown in figure 5bd. The growth rate has been shown to decrease or increase, or be independent of peptide concentration, depending upon the peptide/conditions and rate-limiting step. Cases where there is an increase in growth rate with peptide concentration point towards association of monomer and fibril being rate limiting to the growth. In contrast, if the growth rate is independent of peptide concentration this can indicate some other step such as a conformational change of the monomeric unit is rate limiting. The growth rate is also dependent upon secondary nucleation steps, if they exist, such as fragmentation of the fibrils to form small fibrils with a net increase in reactive fibril ends on which growth can occur, or secondary nucleation sites on the surface of a fibril (figure 4a) [62].

This highlights the issues of physical stability of peptides, which may not be a significant issue at low peptide concentrations, but which can become a serious problem at high concentrations such as those frequently used in therapeutics.

In stark contrast to the systems and kinetics described above, there are an increasing number of peptides and conditions under which very different kinetics have been observed. In these cases, increasing concentrations of peptide leads to a decrease in t1/2 and tlag [1,73,74] (figure 5eg). This unusual kinetic behaviour has been attributed to the formation of a different form of oligomeric species where the oligomers formed are off-pathway, and either cannot directly be converted into on-pathway oligomers or nucleating species, or where the interconversion is extremely slow. One such system is GLP-1, which is a particularly interesting system as in this case there is a switch in the kinetic behaviour with pH [1]. At pH > 8.0, GLP-1 follows the more common nucleation–polymerization kinetics with t1/2 and tlag decreasing with increasing peptide concentration (figure 5bd); however, at pH 7.5, the kinetics reverse such that both lag time and t1/2 decrease with increasing peptide concentration indicating off-pathway oligomers have formed [1]. At an intermediate pH, 8.0, there is little dependence on peptide concentration as there is a delicate balance between on- and off-pathway species [1]. Although off-pathway oligomers have been observed now for a number of peptides and proteins [1,73,74] these systems have not been studied in the detail that peptides and proteins showing simpler nucleation–polymerization kinetics have been and therefore not much is known about the nature of off-pathway oligomers, for example, their size, structure and cellular toxicity.

2.2. Amino acid sequence

How the amino acid sequence of a peptide affects it intrinsic propensity to aggregate, to form either amorphous or amyloid-like fibrils is now very well understood, such that there are numerous prediction programmes available [33,75]. Factors such as hydrophobicity, charge state, β-sheet forming propensity along with other properties are all known to contribute [33,7577]. A series of five or more residues which all have a high intrinsic propensity form what is known as an aggregation-prone region, APR [75,7881]. In some cases, these methods have been used to predict the effect of an amino acid substitution on aggregation propensity, for example, the effect of pathogenic mutations on the aggregation of Aβ in vitro and in vivo [75].

A number of approaches have been developed to predict aggregation propensity some of which are primarily based on our knowledge of the intrinsic propensity of amino acids to aggregate and form fibrils. As our understanding of the factors that affect aggregation in terms of amino acid sequence has improved, methods have also been developed that not only take into account APRs in peptide sequence but also other factors such as the presence of gatekeeper residues that help prevent aggregation [8284], and the flexibility and solvent accessibility of the peptide chain [85]. Other algorithms also take into account the propensity of amino acids to form other forms of stable structure such as α-helices, as aggregation is always in competition with the formation of other conformational states [86].

Since high-resolution structures of amyloid fibrils have become available from X-ray crystallographic and NMR studies [34], structure-based methods for predicting APRs have also been developed [33,75]. In these cases, APRs which are compatible with the formation of the cross β-structure seen in the core of amyloid fibrils are identified by threading peptide and protein sequences through templates of different amyloid structures scoring each resultant structure in terms of packing density [87], pairing energies for side chains that pack against each other in the structure [8890] and other metrics [91,92]. In another case, a combination of the two approaches has been used and a position-specific scoring matrix developed taking into account both intrinsic propensities and also structure [93].

For peptides which do not form stable tertiary or quaternary structures, these approaches work well and predictions are generally reasonably accurate. In addition, while many algorithms which are not structure-based predict both the propensity for a peptide to form either an amorphous or highly structured aggregate, the structure-based methods are largely accurate only for predicting the formation of amyloid-like fibrillar structures.

Given these prediction methods, summarized in table 2, it is relatively straightforward to identify APRs in largely unstructured peptides and to rationally design amino acid substitutions to reduce aggregation propensity. Of course, such changes may also affect activity. At the moment, there are no programs available to predict the effects of incorporating non-natural amino acids nor the effects of chemical modification of a peptide on its aggregation propensity.

Table 2.

Methods for predicting aggregation-prone regions and aggregation propensity in peptides.

name description of method references
Zygreggator APRs and gatekeepers, protein flexibility and solvent accessibility [85]
TANGO statistical thermodynamics approach. Identifies APRs but takes into account competition β-aggregated state and other conformations [86]
SALSA uses β-propensity for each amino acid and a sliding average to identify APRs [81]
AGGRESCAN uses aggregation propensity for each amino acid [94,95]
FoldAmyloid packing density [87]
3D profile method structure-based method uses amyloid fibril structure as a template [91]
Pre-AMYL structure-based method uses amyloid fibril structure as a template [92]
PASTA pairing energies calculated for pairs of residues within an amyloid structure. Sequences are scored by energies to identify APRs [88,89,96]
BETASCAN β-strand pairing energies using amyloid core structure [90]

2.3. pH and net charge

Electrostatic interactions are known to play major roles in the self-association of peptides to form all form of aggregates [97102]. Essentially, for systems carrying an overall net charge, unfavourable electrostatic repulsion needs to be overcome in order for self-association/aggregation to take place. Generally, the higher the net charge the slower the aggregation and the lower the net charge the higher the propensity to aggregate [97102]. As such, solution conditions such as pH, ionic strength, nature of cations and anions in solution and the presence of polyelectrolytes can all have a dramatic effect on both the rate and extent of aggregation. One of the best characterized systems in terms of the effects of net charge, ionic strength, and counter ions on aggregation is the 37-residue positively charged peptide IAPP which has no acidic groups [103]. A comprehensive and rigorous study has shown that there are multiple effects on aggregation depending upon the pH [103]. At pH 8.0 there is a strong dependence on ionic strength due to Debye screening and increasing ionic strength accelerates aggregation by up to a factor of ten. In this case, specific ion effects were also observed, the nature of the cation being relatively unimportant but the rate depending crucially on the identity of the anion (factors of four were observed with different anions). The effects were observed both on the lag phase and the apparent growth phase. At pH 5.5, where IAPP has a higher overall net charge, the aggregation kinetics were considerably slower than at pH 8.0, but there were similar effects of ionic strength [103]. For lysozyme and acylphosphatase, studies have established that polyelectrolytes can also have significant effects. In these cases, a number of different polyelectrolytes were found to bind to and stabilize mature fibrils [104].

Although it is easy to establish some general rules of thumb with regards to the effects of pH, net charge, ionic strength, ions and polyelectrolytes, these rules should be viewed with caution. There are an increasing number of examples where these rules are not followed due to the complexity of aggregation reactions. For example, around the isoelectric point of peptides and proteins, where net charges are essentially zero, then monodisperse and quasi-amorphous aggregates have been observed instead of amyloid-like fibrils [57]. Thus, aggregation kinetics can be very different from those of amyloid-like fibril formation which is often seen at pH values away from the isoelectric point. In addition, other changes in mechanism have been found at different values of pH and net charge. For example, as already discussed in §2.1, the aggregation kinetics of GLP-1 into amyloid-like fibrils is highly sensitive to pH [1]. In this case, the presence of off-pathway species slows the rate of aggregation considerably particularly at high peptide concentrations. Thus, although the net charge on the protein at pH 7.5 is lower than at pH > 8.0, the rate of aggregation is considerably slower, contrary to what would be expected [1].

2.4. Chemical degradation

Chemical degradation pathways in biological pharmaceuticals have been reviewed many times [105,106]. A number of chemical degradation pathways have been characterized including: deamidation and isomerization [107], oxidation [106,108], hydrolysis, disulfide bond breakage and formation, succinimidation [109], deglycosylation, Maillard reaction, racemization and β-elimination and many of these have been shown to increase the aggregation of biologics [109]. This is in effect because chemical degradation/modification often changes the physical properties such as hydrophobicity, secondary (relevant for relatively unstructured peptides) and/or tertiary structure, and the thermodynamic and/or kinetic barriers to unfold (relevant only for highly structured proteins and not discussed further here) [105,109].

2.4.1. The effect of deamidation on the physical stability of peptides

Deamidation is a chemical reaction in which the amide group in Asn or Gln side chains is hydrolysed at pH < 3 or undergoes intramolecular cyclization at pH > 6 to form the corresponding carboxylate: Asp or Glu [105], where the isomerization of Asp to form iso-Asp involves backbone-to-side chain rearrangement. The deamidation of Asn at pH < 4 is via direct hydrolysis of the side chain [110] leading to a single product, Asp. However, in basic conditions (pH > 6), deamidation generates two products: Asp and iso-Asp [105,109] and it can also lead to peptide and protein fragmentation [111].

Deamidation of Gln to Glu can also take place but is rarely observed, unless the pH < 3. It is well known that Asn is more prone to deamidation than Gln [112115]. Moreover, it has been reported that the deamidation of Asn is strongly dependent on primary structure with the adjacent amino acid on the C-terminal side having the greatest influence on the rate. In peptides, an Asn-Gly sequence is particularly prone to deamidation and is considered a ‘hot spot’ of reactivity. Asp degradation is similar to Asn degradation resulting in the formation of iso-Asp and racemic products. Asp residues can also undergo backbone hydrolysis.

There is limited literature on the link between deamidation and physical stability; however, deamidation may alter the local structure of the peptide. Deamidation to Asp results in a change of charge (−1) but Asp is isosteric with Asn. In contrast, iso-Asp, which contains a methylene group in its peptide backbone is likely to be more disruptive to local structure. In some cases, deamidation has been shown to accelerate amyloid formation and alters amyloid fibril structure, e.g. amylin20–29 [116] and amylin [117]. Moreover, deamidation of Asn76 in human γS-crystallin promotes dimer formation [118] and deamidation of lens protein, βA3-crystallin destabilizes and triggers aggregation of that protein [119].

2.4.2. The effect of oxidation on the physical stability of peptides

It is well known that some amino acid side chains can be oxidatively modified during any stage of peptide production, purification [120], formulation [121], storage [105,106,122]. This includes aromatic amino acids His, Trp and Tyr and sulfur-containing side chains Met and Cys [105,109,111]. Moreover, complicated oxidation reactions can modify the primary sequence of a protein and may cause aggregation [123].

There is rather limited literature on the link between oxidation and the physical stability of peptides, such that the effects are still poorly understood. However, oxidation of the side chain of Met has been discussed in a few papers. Interestingly, studies on the peptide Aβ1–40 have shown that oxidation of the methionine residue reduces lag-times of fibril formation but extends lag-times for Aβ1–42 [124]. It has been shown, for both peptides, that Met35 oxidation affects fibril morphology and length [124]. Studies on methionine oxidation also in α-synuclein at physiological pH show that oxidation inhibits fibril formation significantly [125].

2.4.3. The effect of hydrolysis on the physical stability of peptides and proteins

Most peptide bonds are stable except those in -X-Asp-Y- sequence which is at least 100 times more labile than other peptide bonds in dilute acid. The Asp residue undergoes backbone hydrolysis of peptide bonds both N- and C-terminal to the residue as discussed earlier. In addition, the side chain of Trp is known to undergo hydrolysis and the primary degradation product is called kynurenine [105,109]. N-terminal cyclization reactions are closely related to hydrolytic reactions, both degradation pathways involving nucleophilic attack [105]. They have been described as occurring by two general pathways: diketopiperazine (DKP) formation and pyroglutamic acid (pGlu) formation [105]. An example of a peptide forming DKP is the undecapeptide substance P [126]. Pyroglutamic acid formation follows a mechanism similar to that for DKP formation described earlier [105] and forms a cyclized, N-terminal structure which is has been observed in some peptide hormones [127]. Unfortunately, there is very little literature describing how these hydrolysis reactions in peptides affect the physical stability of the system apart from in a very few cases. For example, it has been shown that pGlu-Aβ peptides contribute to the formation of amorphous and fibrillar deposits in Alzheimer's disease, familial British dementia, and familial Danish dementia [128].

2.4.4. The effect of β-elimination and racemization on the physical stability of peptides and proteins

β-Elimination reactions typically require high pH and elevated temperatures [105,111], and thus there are few reports of β-elimination in proteins of pharmaceutical relevance, except for β-elimination at Cys being observed in insulin in the solid state [129]. The racemization reaction is very similar to β-elimination and has been observed in Asp residues in crystallin as been described in a number of studies [130]; however, the impact of racemization on subsequent physical stability is not well known. It has been suggested that isomerization and/or racemization of Asp at position 1 or 7 is associated with the pathological role of Aβ1–42 peptide [131]. On the other hand, racemization and/or isomerization of Asp23 in the same system is thought not to be related to pathogenesis, but to be a consequence of chemical reactions during long-term deposition of fibrils [131].

2.5. Chemical modifications

Chemical modifications of different types are increasingly used to optimize the chemical/physical and biological properties of therapeutic peptides. This includes the incorporation of non-natural amino acids [132], engineering of disulfide bonds [133137] and lipidation or acylation of the peptide [138,139].

Lipidation/acylation of peptides is becoming increasingly important in the field of therapeutic peptides [140,141]. Inspired by endogenous lipidation processes, it has been shown to improve many properties of a peptide, particularly its ability to bind to human serum albumin increasing the half-life of the peptide in vivo dramatically and also enhancing its ability to permeate body tissues [140,141]. Current therapeutics which use this approach include liraglutide [142] (a lipidated form of GLP-1), and insulin detemir, a lipidated conjugate of insulin [143]. Recently, GlaxoSmithKline have developed this idea further and created a GLP1-albumin fusion to achieve similar effects [144,145].

Although considerable research has been undertaken on how this class of chemical modification affects the activity of the peptide in vitro and in vivo, rather little is known about how these modifications affect the physical stability of the system. One recent biophysical study on liraglutide by Wang and co-workers showed that this lipidated peptide can form oligomers in solution [146]. In this case, a pH-sensitive equilibrium between octamers and dodecamers was observed in the vicinity of pH 7 [146]. In another case, acylation of the anti-microbial peptide novicidin resulted in an increased tendency to form α-helical structure when a C12 acyl group was attached. The concentration dependence of this conformational change was attributed to the increased tendency of the peptide to form micelles as shown by NMR [147]. Although not peptides with any therapeutic potential, the Prive group has published extensively on the properties of lipopeptide detergents (LPDs) which have been developed for structural studies of membrane proteins [148]. LPDs, which consist of an α–helical peptide with alkyl chains at either end of the helix, have been shown to self-assemble into cylindrical micelles that have a densely packed hydrophobic core [148].

In contrast to the chemical modifications used by the pharmaceuticals industry to optimize the properties of therapeutic peptides, where little is known about the effect of the modifications on physical stability, a great deal has been reported for peptides whose aggregation is associated with disease states. In these cases, the modifications investigated are largely post-translational modifications that occur in vivo and which, in many cases, are thought to affect the disease. For example, the effects of many post-translational modifications on the aggregation of Aβ peptides have been studied. An N-terminal truncation and pyroglutamate modification at residue 3 of Aβ peptide has been shown to have a higher propensity for oligomerization and aggregation than full-length Aβ [149,150]. In addition, phosphorylation of Aβ has been shown to increase oligomer formation [151153], while citrullination of Aβ decreases the rate of fibril formation but may increase the amount of toxic oligomers [154]. Similarly, nitration has also been shown to increase the amount of oligomers of Aβ in solution [155]. Backbone modification of Aβ at Gly29 also shows decreased fibril formation and increased oligomer formation [156].

The aggregation and fibril formation of α-synuclein is associated with another neurodegenerative disease, Parkinson's disease. In this case, there is also a large body of data that establishes that numerous post-translational modifications in vivo affect the propensity of the protein to aggregate and thus potentially the disease state. For example, N-acetylation has been shown to decrease fibrillation [157,158], post-translational modifications due to oxidative stress such as HNE conjugation, nitration and oxidation have been shown to reduce fibril formation but increase oligomer formation [159,160], and oxidation of Met5 also decreases fibril formation but increases the concentration of oligomers [161].

The intracellular protein tau, associated with Alzheimer's disease, is also known to undergo a large number of post-translational modifications in vivo including phosphorylation, ubiquitination, nitration, truncation, prolyl isomerization, glycosylation and glycation. Many of these are thought to affect neurofibrillar tangle formation, the aggregated form of tau in cells [162]. A few, but not many, of these post-translational modifications on tau aggregation have been studied in detail. For example, cysteine guanylation of tau has been shown to inhibit fibril formation with a corresponding increase in the population of oligomers [162].

The effects of a large number of other modifications on physical stability have also been reported including different glycosylations [163] and fatty acid conjugation [164].

2.6. Surfaces and interfaces

Surface-induced aggregation, often called ‘surface adsorption’, is a physical degradation process whereby there is accumulation and adhesion of peptide molecules to a surface, without penetration of the surface. During this process, peptide molecules change their physical state, i.e. they are no longer in solution [165168]. There are different mechanisms of surface adsorption of biomolecules and some of the major factors driving adsorption include intra-molecular forces, hydrophobicity [169171], and ionic or electrostatic interactions [168,172].

Extensive research has been conducted on the surface adsorption of human insulin [171,173177]. It is well known that insulin aggregates to form fibres in the presence of hydrophobic surfaces [170,171,175,176] and especially in a peristaltic pump system [177]. In this case, the potential formation of aggregation-prone intermediates is very problematic as this may catalyse the aggregation of other molecules in bulk solution. Other studies have established that the aggregation of Aβ1–40 into fibrils is also influenced by surfaces [178182]. The surface adsorption was studied for other systems such as GLP-1 [172], GLP-2 [183], acylated glucagon-like peptide-2 [183,184], glucagon [185], poly(ethylene glycol) (PEG)-glucagon [186], and α-synuclein [187].

It has been demonstrated that any interface (between solid, liquid or air) will affect the ability of water molecules to form a dynamic hydrogen-bonding network as in bulk water, and therefore air–water and ice–water interfaces are described in the literature as hydrophobic interfaces, because these interfaces do not allow hydrogen bonding [167]. Adsorption of peptides through their hydrophobic side chains to such hydrophobic interfaces can thus ‘hide’ the hydrophobic surface from the aqueous bulk. Some research shows that, for example, Aβ1–40 [179,188] and insulin both aggregate at air–water interfaces [189].

Many studies have confirmed that the rate of peptide and protein adsorption is normally governed by diffusion, and is thus a function of concentration [109]. The amount of surface-adsorbed species slowly increases over time, usually due to further peptide aggregation at the site [178]. Of course, there may also be slow structural rearrangements of the adsorbed peptide [178]. The adsorbed layer often changes its properties over time; this is often associated with a change from a reversible to an irreversible process [168]. The final morphologies of the aggregates formed can be sensitive to the surface chemistry [172,174,178].

It is well known that after sufficient time most biomolecules will adsorb to solid surfaces so strongly that they will not (or only very slowly) desorb from the attached surface back into solution [190]. However, it is possible to some degree to control this process by adding surface active agents which can lead to desorption. The degree of adsorption depends on multiple factors, including: concentration, pH, excipients and temperature [167,168].

2.7. Excipients

Excipients of many types have been employed to try and reduce aggregation in numerous systems including salts, surfactants, osmolytes, preservatives, chelators and antioxidants, specific ligands, sugars and carbohydrates. Below is a brief description of some of the excipients that have been used and their effects on aggregation.

2.7.1. Buffering agents

As already discussed, pH strongly influences peptide stability and the potential for biomolecular aggregation. The most commonly used buffers in pharmaceutical development are acetate, citrate, histidine, phosphate, Tris and glycine [105,109,191]. A complete description of the effects of buffer on peptide aggregation is beyond the scope of this review; however, interested readers are directed towards the recent excellent review by Zbacnik and co-workers [192].

2.7.2. Salts

Standard physiological salts have been used as tonicity modifiers in peptide formulations [109,191]. Salts have complex effects on the physical stability of biomolecules affecting both conformational and colloidal stability. Their effects frequently vary according to the surface charge on the peptide or protein, and the overall effect of a salt on physical stability is a balance of different and multiple mechanisms by which salt interacts with water and biomolecules. Salts can influence physical stability by altering the properties of the peptide–solvent system (Hofmeister effects) and by screening electrostatic interactions (Debye–Hückel effects). Many different studies on the dependence of amyloid formation on ionic strength over the past 10 years have shown that ions can influence both the kinetics of aggregation and the structure of fibrils formed. For example, islet amyloid polypeptide (37-residue) [103], human glucagon (29-residue) [193], and Aβ peptide [100,102,194].

2.7.3. Surfactants

Surface-active agents are amphiphilic molecules that tend to orient such that the exposure of the hydrophobic part of the molecule to the aqueous solution is minimized. Non-ionic surfactants such as Tween20 and Tween80 are often added to peptide solutions during pharmaceutical development to prevent aggregation or adsorption during purification, filtration, shaking and transportation [191,195198]. However, the chemical stability of Tween20 and Tween80 in pharmaceutical formulations is very important. Unfortunately both surfactants are known to undergo oxidation and cleavage at their ethylene oxide subunits, as well as hydrolysis of the fatty acid ester bond [199202]. Thus, their use can be complicated as these chemical processes can potentially promote precipitation in some peptide formulations. The use of Tween80 is also far from straightforward as it has been reported that it can have a dual effect on physical stability. For example, Tween80 in the formulation of the protein IL-2 mutein inhibited shaking-induced aggregation, but also had a dramatic effect on the oxidation and aggregation of the protein during storage of the liquid formulation [203,204]. To counteract this issue, methionine and tryptophan are used in formulation development to prevent Tween from oxidative degradation [205]. Readers interested in the interactions between surfactants and peptides are directed towards two excellent recent reviews by Otzen [206] and Khan and co-workers [196].

2.7.4. Amino acids

A number of free amino acids have been used to stabilize proteins and reduce aggregation. Arginine, histidine, lysine, glycine and aspartic acid have been found to reduce aggregation for a number of biomolecules [191].

2.7.5. Osmolytes

Osmolytes such as sucrose, trehalose, sorbitol and glycine are commonly used in formulation development [191]. Sucrose has been shown to inhibit IL-1ra dimer formation [109]. However, their use is not always beneficial and it has been reported that some osmolytes accelerate the rate of aggregation, for example, of glucagon [207].

2.7.6. Antioxidants and chelators

Peptide oxidation, which was described earlier in this review, is a major cause of chemical instability and also sometimes linked to physical stability. Amino acids such as methionine, cysteine, histidine, tyrosine and tryptophan in peptides are susceptible to oxidation under some conditions encountered during pharmaceutical development. Therefore, a number of antioxidants are used as excipients including ascorbic acid [191]. It has also been reported that sodium thiosulfate, methionine, catalase or platinum, and the chelating agents EDTA and DTPA are also effective in reducing the oxidation of biologics [191,208210].

2.7.7. Preservatives

Preservatives are used as antimicrobial agents and are very common in liquid formulations [191,211]. Their role is to prevent bacterial growth during storage [191]. The most commonly used preservatives in pharmaceutical development are m-cresol, phenol and benzyl alcohol. However, it has been reported that preservatives can sometimes cause protein aggregation. For example, benzyl alcohol induces aggregation of recombinant human interleukin-1 receptor [212], and m-cresol induces aggregation of cytochrome c [213]. Moreover, destabilizing effects of phenol and m-cresol were reported for insulin lispro and insulin aspart [214]. Little is known about their effects on the physical stability of peptides.

2.7.8. Polymers and proteins

A number of polymers have been shown to stabilize drug products including PEGs. Other polymers that are used as excipients include dextrans, heparin, gelatins type A, poly-l-glutamic acid, poly-l-lysine, poly-Asp and poly-Gl [105,109,191]. For example, it has been reported that different PEGs stabilize the proteins lysozyme and bovine serum albumin [109]. Human serum albumin is also commonly used as an excipient to inhibit protein/peptide adsorption onto surfaces during product development [109]. However, adding protein-based excipients adds more complexity to pharmaceutical formulations [191].

Recently, some excellent reviews on biologics formulation [215,216] and excipients, including protein–excipients interactions [191,216,217] have been published.

2.8. Impurities

The results of in vitro aggregation studies are often controversial and difficult to reproduce by others. One of the reasons for this problem is the variable amounts of intrinsic impurities in peptide and protein preparations that can affect aggregation rates. For example, the aggregation kinetics of synthetic Aβ has been reported to vary not only batch-to-batch but also with storage and solubilization conditions [218]. In line with these findings, kinetics studies revealed that the impurities left in the synthetic peptides of glucagon decreased fibrillation rates dramatically compared to that of pharmaceutical-grade glucagon [193,219,220].

Numerous sorts of impurities may be encountered in synthetic peptides, which either originate from the raw materials, the manufacturing process, formed by degradation during the manufacturing process or during storage. Some examples of potential synthetic peptide process-related impurities are deletions, truncations, diastereoisomers, substitution (Leu/Ile) and disulfide modification. The most common peptide degradation impurities are deamidation of Gln/Asn/C-terminus, methionine oxidation, acetylation of amino functions and disulfide modification (see §2.4). Non-peptide process-related impurities can be present. Some examples are residual solvents and heavy metals, traces of reagents used in the process, such as trifluoroacetic acid (TFA), acetate and hydrochloride salts, inorganic fluoride compounds from protective groups such as t-butyloxycarbonyl- (Boc-). Unfortunately, there are few reports about the impact of degradation impurities and counter ion salts on the kinetics of peptide aggregation and there is a lack of information about the influence of other common impurities in synthetic peptides on either the mechanism or the kinetics of aggregation.

2.8.1. Peptide-related impurities

The addition of 1% oxidized glucagon (Met27) has shown to greatly delay the rate of fibril formation. The authors suggest that glucagon analogues can interfere with fibril formation by either competing with or blocking unmodified glucagon adsorption at the ends of fibrils [3,4]. Another study shows that in the absence of impurities, amylin 20–29 does not spontaneously aggregate and is not amyloidogenic. However, this peptide can spontaneously deamidate, and the presence of less than 5% of deamidation impurities leads to the formation of aggregates that have the hallmarks of amyloid. In addition, small amounts of deamidated material can induce amyloid formation of the purified peptide. These results have fundamental implications for the definition of an amyloidogenic sequence and for the standards of purity of peptides and proteins used for aggregation studies [58,221227].

2.8.2. Non-peptide-related impurities

The majority of synthetic peptides are still manufactured by solid-phase procedures. Due to the cleavage and standardized purification steps, the resulting product usually contains TFA salts. The presence of TFA in the peptidic material affects its physico-chemical properties. Several investigations have documented the influence of TFA on peptide structure. For example, characterization of synthetic Aβ peptide has shown that there is a strong influence of TFA counter ions on the physico-chemical properties of the peptide [228]. In particular, it has been shown that TFA can change the secondary structure, solubility, and aggregation propensity of synthetic Aβ peptides, and even affect the ultimate size, flexibility, and geometry of fibrils formed [229231]. Fezoui and co-workers showed that solvation of Aβ synthetic peptide with sodium hydroxide followed by lyophilization produced stocks with superior solubility and fibrillogenesis characteristics [232,233]. Relative to trifluoroacetate or hydrochloric acid Aβ peptide counter ion salts, yields of low molecular weight (monomers and/or dimers) were improved significantly after pre-treatment with sodium hydroxide [232,233].

2.9. Temperature/pressure/agitation/lyophilization

Numerous other external factors including temperature, pressure, agitation and lyophilization/freeze drying can affect the aggregation of peptides, in many cases, in ways difficult to predict.

2.9.1. Pressure

High hydrostatic pressure (HHP) has been used in a number of studies [234]. It is now well established that HHP pushes the conformational equilibria towards ensembles of structures that occupy smaller volumes, it acting on cavities that are excluded from solvent that can be found in the hydrophobic cores of natively folded proteins and aggregates including fibrils [234]. HHP has been used to dissociate multimers in oligomeric structures and, in some cases, dissociate aggregates [234,235]. The degree to which HHP acts to dissociate aggregated forms of peptides depends upon the structures that are formed and also their inherent thermodynamic stability. For example, if fibrils have cavities within them (cavity defects as they are called) then often low pressures are sufficient to dissociate the aggregated form; for example, early aggregates and protofibrils are generally more readily dissociated by HHP than mature fibrils [236]. For a peptide from transthyretin that can form fibrils, TTR105–115, 220 MPa dissociates aggregates and protofibrils while mature fibrils remain stable up to 1.3 GPa [236]. For α-synuclein, pathogenic variant fibrils are less resistant to HHP than those of the wild-type protein, hence more toxic oligomers are present [234]. The dissociation of aggregates and fibrils by HHP is reversible and can be enhanced by the addition of cosolvents such as glycerol, methanol and dimethylsulfoxide [236].

2.9.2. Temperature

Aggregation kinetics frequently show non-Arrhenius behaviour even over short temperature ranges [237]. A consequence of this is that it is challenging to extrapolate the results of aggregation kinetics commonly measured at higher temperatures to lower temperatures.

However, cold denaturation of amyloid fibrils has been observed [238,239], and low temperatures have been combined with high pressures to dissociate aggregates and fibrils [236].

2.9.3. Agitation

Agitation, in the form of a magnetic stirring bar or by use of a shaker table, is routinely used in studies of peptide and protein aggregation. It has been employed in order to accelerate otherwise slow aggregation kinetics and also used to improve the reproducibility of experimental results [240]. Despite the frequent use of agitation little is known about the shear forces that are produced and the effects on different steps in the aggregation reaction (figure 4). In addition, as stirring/shaking is often different in different studies, it is challenging to undertake a comparative analysis on systems studied in different research groups [241].

2.9.4. Lyophilization/freeze drying

The stabilization of peptides in a solid form is a very common approach to increase both their chemical and physical stability [242]. Lyophilization and freeze drying are frequently used to prepare peptides in a solid state. However, for a number of systems studies have shown these processes can result in conformational changes leading to increased aggregation after reconstitution [243].

3. Summary

The physical stability of peptide- and protein-based therapeutics remains a huge challenge for the pharmaceutical industry. Table 1 provides just a few examples of the different types of aggregation processes that can occur that lead to low physical stability; however, the number of potential therapeutic agents known to aggregate in some form is very large indeed [30,32,123,196,215]. In contrast to peptides that aggregate and that are associated with disease states, the relative propensity of therapeutic peptides to aggregate may be low. However, the fact that these peptides are frequently used at high concentrations and are required to have long-term physical stability under these conditions makes the problem particularly acute. In addition, therapeutic peptides have to withstand a number of processes during production, formulation and use, that increase their propensity to aggregate and which exacerbate the problem.

Fundamentally, the basic thermodynamic and kinetic parameters that underpin the aggregation of peptides whether they be disease-associated or therapeutic must be the same. In many cases, the intrinsic and extrinsic factors that affect the physical stability of peptides, such as amino acid sequence, pH etc., will be common to both classes. In other cases, factors will differ. For example, the physical stability of disease-associated peptides depends upon many factors in vivo such as the presence of molecular chaperones, proteostasis and membrane surfaces which are not relevant to therapeutic peptides. In contrast, therapeutic peptides are exposed to a range of surfaces and processes, such as high pressures and freeze-drying, that are not relevant to in vivo systems. Despite this, much knowledge is directly transferable between the different classes from studying the fundamentals of aggregation in vitro in either system, particularly from a mechanistic perspective. It might even be that strategies for increasing the physical stability of therapeutic peptides may come from a deeper understanding of how nature deals with the problem.

Given the large number of different factors that can affect the aggregation of peptides, and which are outlined in this review, it is no great surprise that this remains a challenging area in the development and use of peptide-based therapeutic agents. In some cases, great strides have been made in recent years in our understanding of some of the factors, particularly the sequence of the peptide. In other cases, such as impurities, there is almost no knowledge on which and how these species affect aggregation although it is known to be a significant issue. In other cases, there is a large amount of empirical data available, for example on how post-translation modifications affect oligomer and fibril formation in disease-related peptides. Even in this case though, there is a fundamental lack of understanding of how the modifications affect each step on the aggregation pathway. Without such, it is impossible to rationally predict effects. In addition, there are an increasing number of chemical modifications employed by the pharmaceutical industry to enhance the properties of therapeutic peptides, whose effects on physical stability are not known.

The number of factors that affect aggregation reactions to some degree reflects the intrinsic complexity of the reactions and very subtle changes in the physico-chemical properties of the system can result in large changes in aggregation propensity and rates. There is clearly a significant amount of research that is still required in order for scientists to fully understand the complex energy landscape of aggregation and how this is affected by numerous variables. As well as empirical studies, more fundamental research is needed for this major problem to be fully addressed.

Acknowledgements

We are grateful to many scientists in the protein and peptide aggregation field for interesting discussions.

Data accessibility

This article has no additional data.

Authors' contributions

All authors co-wrote the manuscript.

Competing interests

We declare we have no competing interests.

Funding

K.L.Z. was funded by MedImmune. F.J.B. is funded by MedImmune and Peterhouse, Cambridge. A.L.G.d.S. is an employee of MedImmune.

References

  • 1.Zapadka KL, Becher FJ, Uddin S, Varley PG, Bishop S, Gomes dos Santos AL, Jackson SE. 2016. A pH-induced switch in human glucagon-like peptide-1 aggregation kinetics. J. Am. Chem. Soc. 138, 16 259–16 265. ( 10.1021/jacs.6b05025) [DOI] [PubMed] [Google Scholar]
  • 2.Poon S, Birkett N, Fowler S, Luisi B, Dobson C, Zurdo J. 2009. Amyloidogenicity and aggregate cytotoxicity of human glucagon-like peptide-1 (hGLP-1). Protein Pept. Lett. 16, 1548–1556. ( 10.2174/092986609789839232) [DOI] [PubMed] [Google Scholar]
  • 3.Pedersen JS. 2010. The nature of amyloid-like glucagon fibrils. J. Diabetes Sci. Technol. 4, 1357–1367. ( 10.1177/193229681000400609) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Pedersen JS, Andersen CB, Otzen DE. 2010. Amyloid structure—one but not the same: the many levels of fibrillar polymorphism. FEBS J. 277, 4591–4601. ( 10.1111/j.1742-4658.2010.07888.x) [DOI] [PubMed] [Google Scholar]
  • 5.De Jong KL, Incledon B, Yip CM, DeFelippis MR. 2006. Amyloid fibrils of glucagon characterized by high-resolution atomic force microscopy. Biophys. J. 91, 1905–1914. ( 10.1529/biophysj.105.077438) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Caputo N, et al. 2014. Biochemical stabilization of glucagon at alkaline pH. Diabetes Technol. Ther. 16, 747–758. ( 10.1089/dia.2014.0047) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Caputo N, Castle JR, Bergstrom CP, Carroll JM, Bakhtiani PA, Jackson MA, Roberts CT, David LL, Ward WK. 2013. Mechanisms of glucagon degradation at alkaline pH. Peptides 45, 40–47. ( 10.1016/j.peptides.2013.04.005) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Gilchrist PJ, Bradshaw JP. 1993. Amyloid formation by salmon calcitonin. Biochim. Biophys. Acta 1182, 111–114. ( 10.1016/0925-4439(93)90160-3) [DOI] [PubMed] [Google Scholar]
  • 9.Gaudiano MC, Colone M, Bombelli C, Chistolini P, Valvo L, Diociaiuti M. 2005. Early stages of salmon calcitonin aggregation: effect induced by ageing and oxidation processes in water and in the presence of model membranes. Biochim. Biophys. Acta 1750, 134–145. ( 10.1016/j.bbapap.2005.04.008) [DOI] [PubMed] [Google Scholar]
  • 10.Diociaiuti M, Macchia G, Paradisi S, Frank C, Camerini S, Chistolini P, Gaudiano MC, Petrucci TC, Malchiodi-Albedi F. 2014. Native metastable prefibrillar oligomers are the most neurotoxic species among amyloid aggregates. Biochim. Biophys. Acta 1842, 1622–1629. ( 10.1016/j.bbadis.2014.06.006) [DOI] [PubMed] [Google Scholar]
  • 11.Diociaiuti M, Gaudiano MC, Malchiodi-Albedi F. 2011. The slowly aggregating salmon calcitonin: a useful tool for the study of the amyloid oligomers structure and activity. Int. J. Mol. Sci. 12, 9277–9295. ( 10.3390/ijms12129277) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Sumner Makin O, Serpell LC. 2004. Structural characterisation of islet amyloid polypeptide fibrils. J. Mol. Biol. 335, 1279–1288. ( 10.1016/j.jmb.2003.11.048) [DOI] [PubMed] [Google Scholar]
  • 13.Luca S, Yau W-M, Leapman R, Tycko R. 2007. Peptide conformation and supramolecular organization in amylin fibrils: constraints from solid-state NMR. Biochemistry 46, 13 505–13 522. ( 10.1021/bi701427q) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Bedrood S, Li Y, Isas JM, Hegde BG, Baxa U, Haworth IS, Langen R. 2012. Fibril structure of human islet amyloid polypeptide. J. Biol. Chem. 287, 5235–5241. ( 10.1074/jbc.M111.327817) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Jha S, Snell JM, Sheftic SR, Patil SM, Daniels SB, Kolling FW, Alexandrescu AT. 2014. pH dependence of amylin fibrillization. Biochemistry 53, 300–310. ( 10.1021/bi401164k) [DOI] [PubMed] [Google Scholar]
  • 16.Weirich F, Gremer L, Mirecka EA, Schiefer S, Hoyer W, Heise H, van der Wel P. 2016. Structural characterization of fibrils from recombinant human islet amyloid polypeptide by solid-state NMR: the central FGAILS segment is part of the beta-sheet core. PLoS ONE 11, e0161243 ( 10.1371/journal.pone.0161243) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Jimenez JL, Nettleton EJ, Bouchard M, Robinson CV, Dobson CM, Saibil HR. 2002. The protofilament structure of insulin amyloid fibrils. Proc. Natl Acad. Sci. USA 99, 9196–9201. ( 10.1073/pnas.142459399) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Hua QX, Weiss MA. 2004. Mechanism of insulin fibrillation: the structure of insulin under amyloidogenic conditions resembles a protein-folding intermediate. J. Biol. Chem. 279, 21 449–21 460. ( 10.1074/jbc.M314141200) [DOI] [PubMed] [Google Scholar]
  • 19.Smith JF, Knowles TPJ, Dobson CM, MacPhee CE, Welland ME. 2006. Characterization of the nanoscale properties of individual amyloid fibrils. Proc. Natl Acad. Sci. USA 103, 15 806–15 811. ( 10.1073/pnas.0604035103) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Kurouski D, Deckert-Gaudig T, Deckert V, Lednev IK. 2012. Structure and composition of insulin fibril surfaces probed by TERS. J. Am. Chem. Soc. 134, 13 323–13 329. ( 10.1021/ja303263y) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Kurouski D, Dukor RK, Lu X, Nafie LA, Lednev IK. 2012. Normal and reversed supramolecular chirality of insulin fibrils probed by vibrational circular dichroism at the protofilament level of fibril structure. Biophys. J. 103, 522–531. ( 10.1016/j.bpj.2012.04.042) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Kurouski D, Dukor RK, Lu X, Nafie LA, Lednev IK. 2012. Spontaneous inter-conversion of insulin fibril chirality. Chem. Commun. 48, 2837–2839. ( 10.1039/c2cc16895b) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Kurouski D, Washington J, Ozbil M, Prabhakar R, Shekhtman A, Lednev IK, Colombo G. 2012. Disulfide bridges remain intact while native insulin converts into amyloid fibrils. PLoS ONE 7, e36989 ( 10.1371/journal.pone.0036989) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Yoshihara H, Saito J, Tanabe A, Amada T, Asakura T, Kitagawa K, Asada S. 2016. Characterization of novel insulin fibrils that show strong cytotoxicity under physiological pH. J. Pharm. Sci. 105, 1419–1426. ( 10.1016/j.xphs.2016.01.025) [DOI] [PubMed] [Google Scholar]
  • 25.Smith MI, Foderà V, Sharp JS, Roberts CJ, Donald AM. 2012. Factors affecting the formation of insulin amyloid spherulites. Colloids Surf. B Biointerfaces 89, 216–222. ( 10.1016/j.colsurfb.2011.09.018) [DOI] [PubMed] [Google Scholar]
  • 26.Kerr D, Wizemann E, Senstius J, Zacho M, Ampudia-Blasco FJ. 2013. Stability and performance of rapid-acting insulin analogs used for continuous subcutaneous insulin infusion: a systematic review. J. Diabetes Sci. Technol. 7, 1595–1606. ( 10.1177/193229681300700620) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Gopalswamy M, et al. 2015. Structural characterization of amyloid fibrils from the human parathyroid hormone. Biochim. Biophys. Acta 1854, 249–257. ( 10.1016/j.bbapap.2014.12.020) [DOI] [PubMed] [Google Scholar]
  • 28.Hermeling S, Aranha L, Damen JMA, Slijper M, Schellekens H, Crommelin DJA, Jiskoot W. 2005. Structural characterization and immunogenicity in wild-type and immune tolerant mice of degraded recombinant human interferon α2b. Pharm. Res. 22, 1997–2006. ( 10.1007/s11095-005-8177-9) [DOI] [PubMed] [Google Scholar]
  • 29.Torosantucci R, Sharov VS, van Beers M, Brinks V, Schöneich C, Jiskoot W. 2013. Identification of oxidation sites and covalent cross-links in metal catalyzed oxidized interferon β-1a: potential implications for protein aggregation and immunogenicity. Mol. Pharm. 10, 2311–2322. ( 10.1021/mp300665u) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Lipiainen T, Peltoniemi M, Sarkhel S, Yrjönen T, Vuorela H, Urtti A, Juppo A. 2015. Formulation and stability of cytokine therapeutics. J. Pharm. Sci. 104, 307–326. ( 10.1002/jps.24243) [DOI] [PubMed] [Google Scholar]
  • 31.Krishnan S, Chi EY, Webb JN, Chang BS, Shan D, Goldenberg M, Manning MC, Randolph TW, Carpenter JF. 2002. Aggregation of granulocyte colony stimulating factor under physiological conditions: characterization and thermodynamic inhibition. Biochemistry 41, 6422–6431. ( 10.1021/bi012006m) [DOI] [PubMed] [Google Scholar]
  • 32.Wang W. 2015. Advanced protein formulations. Protein Sci. 24, 1031–1039. ( 10.1002/pro.2684) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Trainor K, Broom A, Meiering EM. 2017. Exploring the relationships between protein sequence, structure and solubility. Curr. Opin. Struct. Biol. 42, 136–146. ( 10.1016/j.sbi.2017.01.004) [DOI] [PubMed] [Google Scholar]
  • 34.Riek R, Eisenberg DS. 2016. The activities of amyloids from a structural perspective. Nature 539, 227–235. ( 10.1038/nature20416) [DOI] [PubMed] [Google Scholar]
  • 35.Ozboyaci M, Kokh DB, Corni S, Wade RC. 2016. Modeling and simulation of protein–surface interactions: achievements and challenges. Q. Rev. Biophys. 49, e4 ( 10.1017/S0033583515000256) [DOI] [PubMed] [Google Scholar]
  • 36.Kabiri M, Unsworth LD. 2014. Application of isothermal titration calorimetry for characterizing thermodynamic parameters of biomolecular interactions: peptide self-assembly and protein adsorption case studies. Biomacromolecules 15, 3463–3473. ( 10.1021/bm5004515) [DOI] [PubMed] [Google Scholar]
  • 37.Wei Q, Becherer T, Angioletti-Uberti S, Dzubiella J, Wischke C, Neffe AT, Lendlein A, Ballauff M, Haag R. 2014. Protein interactions with polymer coatings and biomaterials. Angew. Chem. Int. Ed. Engl. 53, 8004–8031. ( 10.1002/anie.201400546) [DOI] [PubMed] [Google Scholar]
  • 38.Wang W, Singh SK, Li N, Toler MR, King KR, Nema S. 2012. Immunogenicity of protein aggregates—concerns and realities. Int. J. Pharm. 431, 1–11. ( 10.1016/j.ijpharm.2012.04.040) [DOI] [PubMed] [Google Scholar]
  • 39.DeWitte RS. 2006. Avoiding physicochemical artefacts in early ADME-Tox experiments. Drug Discov. Today 11, 855–859. ( 10.1016/j.drudis.2006.07.012) [DOI] [PubMed] [Google Scholar]
  • 40.Moussa EM, Panchal JP, Moorthy BS, Blum JS, Joubert MK, Narhi LO, Topp EM. 2016. Immunogenicity of therapeutic protein aggregates. J. Pharm. Sci. 105, 417–430. ( 10.1016/j.xphs.2015.11.002) [DOI] [PubMed] [Google Scholar]
  • 41.Knowles TP, Vendruscolo M, Dobson CM. 2014. The amyloid state and its association with protein misfolding diseases. Nat. Rev. Mol. Cell Biol. 15, 384–396. ( 10.1038/nrm3810) [DOI] [PubMed] [Google Scholar]
  • 42.Eisenberg D, Jucker M. 2012. The amyloid state of proteins in human diseases. Cell 148, 1188–1203. ( 10.1016/j.cell.2012.02.022) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Buell AK, et al. 2011. Population of nonnative states of lysozyme variants drives amyloid fibril formation. J. Am. Chem. Soc. 133, 7737–7743. ( 10.1021/ja109620d) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Sekijima Y, Kelly JW, Ikeda S. 2008. Pathogenesis of and therapeutic strategies to ameliorate the transthyretin amyloidoses. Curr. Pharm. Des 14, 3219–3230. ( 10.2174/138161208786404155) [DOI] [PubMed] [Google Scholar]
  • 45.Pauwels K, et al. 2012. Structural basis for increased toxicity of pathological Aβ42:Aβ40 ratios in Alzheimer disease. J. Biol. Chem. 287, 5650–5660. ( 10.1074/jbc.M111.264473) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Foss TR, Wiseman RL, Kelly JW. 2005. The pathway by which the tetrameric protein transthyretin dissociates. Biochemistry 44, 15 525–15 533. ( 10.1021/bi051608t) [DOI] [PubMed] [Google Scholar]
  • 47.Jiang X, Buxbaum JN, Kelly JW. 2001. The V122I cardiomyopathy variant of transthyretin increases the velocity of rate-limiting tetramer dissociation, resulting in accelerated amyloidosis. Proc. Natl Acad. Sci. USA 98, 14 943–14 948. ( 10.1073/pnas.261419998) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Nettleton EJ, Sunde M, Lai Z, Kelly JW, Dobson CM, Robinson CV. 1998. Protein subunit interactions and structural integrity of amyloidogenic transthyretins: evidence from electrospray mass spectrometry. J. Mol. Biol. 281, 553–564. ( 10.1006/jmbi.1998.1937) [DOI] [PubMed] [Google Scholar]
  • 49.McCutchen SL, Colon W, Kelly JW. 1993. Transthyretin mutation Leu-55-Pro significantly alters tetramer stability and increases amyloidogenicity. Biochemistry 32, 12 119–12 127. ( 10.1021/bi00096a024) [DOI] [PubMed] [Google Scholar]
  • 50.Trivella DB, Bleicher L, Palmieri LC, Wiggers HJ, Montanari CA, Kelly JW, Lima LMTR, Foguel D, Polikarpov I. 2010. Conformational differences between the wild type and V30M mutant transthyretin modulate its binding to genistein: implications to tetramer stability and ligand-binding. J. Struct. Biol. 170, 522–531. ( 10.1016/j.jsb.2010.03.002) [DOI] [PubMed] [Google Scholar]
  • 51.Wiseman RL, Johnson SM, Kelker MS, Foss T, Wilson IA, Kelly JW. 2005. Kinetic stabilization of an oligomeric protein by a single ligand binding event. J. Am. Chem. Soc. 127, 5540–5551. ( 10.1021/ja042929f) [DOI] [PubMed] [Google Scholar]
  • 52.Jahn TR, Makin OS, Morris KL, Marshall KE, Tian P, Sikorski P, Serpell LC. 2010. The common architecture of cross-beta amyloid. J. Mol. Biol. 395, 717–727. ( 10.1016/j.jmb.2009.09.039) [DOI] [PubMed] [Google Scholar]
  • 53.Serpell LC, Sunde M, Benson MD, Tennent GA, Pepys MB, Fraser PE. 2000. The protofilament substructure of amyloid fibrils. J. Mol. Biol. 300, 1033–1039. ( 10.1006/jmbi.2000.3908) [DOI] [PubMed] [Google Scholar]
  • 54.Serpell LC. 2000. Alzheimer's amyloid fibrils: structure and assembly. Biochim. Biophys. Acta 1502, 16–30. ( 10.1016/S0925-4439(00)00029-6) [DOI] [PubMed] [Google Scholar]
  • 55.Lashuel HA, LaBrenz SR, Woo L, Serpell LC, Kelly JW. 2000. Protofilaments, filaments, ribbons, and fibrils from peptidomimetic self-assembly: implications for amyloid fibril formation and materials science. J. Am. Chem. Soc. 122, 5262–5277. ( 10.1021/ja9937831) [DOI] [PubMed] [Google Scholar]
  • 56.Knowles TP, Buehler MJ. 2011. Nanomechanics of functional and pathological amyloid materials. Nat. Nanotechnol. 6, 469–479. ( 10.1038/nnano.2011.102) [DOI] [PubMed] [Google Scholar]
  • 57.Krebs MR, Domike KR, Donald AM. 2009. Protein aggregation: more than just fibrils. Biochem. Soc. Trans. 37, 682–686. ( 10.1042/BST0370682) [DOI] [PubMed] [Google Scholar]
  • 58.Bowerman CJ, Nilsson BL. 2012. Self-assembly of amphipathic β-sheet peptides: insights and applications. Biopolymers 98, 169–184. ( 10.1002/bip.22058) [DOI] [PubMed] [Google Scholar]
  • 59.Sarangapani PS, et al. 2016. Both reversible self-association and structural changes underpin molecular viscoelasticity of mAb solutions. J. Pharm. Sci. 105, 3496–3506. ( 10.1016/j.xphs.2016.08.020) [DOI] [PubMed] [Google Scholar]
  • 60.Geoghegan JC, Fleming R, Damschroder M, Bishop SM, Sathish HA, Esfandiary R. 2016. Mitigation of reversible self-association and viscosity in a human IgG1 monoclonal antibody by rational, structure-guided Fv engineering. mAbs 8, 941–950. ( 10.1080/19420862.2016.1171444) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Arosio P, Knowles TP, Linse S. 2015. On the lag phase in amyloid fibril formation. Phys. Chem. Chem. Phys. 17, 7606–7618. ( 10.1039/C4CP05563B) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Meisl G, Kirkegaard JB, Arosio P, Michaels TCT, Vendruscolo M, Dobson CM, Linse S, Knowles TPJ. 2016. Molecular mechanisms of protein aggregation from global fitting of kinetic models. Nat. Protoc. 11, 252–272. ( 10.1038/nprot.2016.010) [DOI] [PubMed] [Google Scholar]
  • 63.Buell AK, Dobson CM, Knowles TP. 2014. The physical chemistry of the amyloid phenomenon: thermodynamics and kinetics of filamentous protein aggregation. Essays Biochem. 56, 11–39. ( 10.1042/bse0560011) [DOI] [PubMed] [Google Scholar]
  • 64.Gillam JE, MacPhee CE. 2013. Modelling amyloid fibril formation kinetics: mechanisms of nucleation and growth. J. Phys. Condens. Matter 25, 373101 ( 10.1088/0953-8984/25/37/373101) [DOI] [PubMed] [Google Scholar]
  • 65.Hall D, Kardos J, Edskes H, Carver JA, Goto Y. 2015. A multi-pathway perspective on protein aggregation: implications for control of the rate and extent of amyloid formation. FEBS Lett. 589, 672–679. ( 10.1016/j.febslet.2015.01.032) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Borzova VA, Markossian KA, Kurganov BI. 2014. Relationship between the initial rate of protein aggregation and the lag period for amorphous aggregation. Int. J. Biol. Macromol. 68, 144–150. ( 10.1016/j.ijbiomac.2014.04.046) [DOI] [PubMed] [Google Scholar]
  • 67.Borgia MB, Nickson AA, Clarke J, Hounslow MJ. 2013. A mechanistic model for amorphous protein aggregation of immunoglobulin-like domains. J. Am. Chem. Soc. 135, 6456–6464. ( 10.1021/ja308852b) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Morriss-Andrews A, Shea JE. 2012. Kinetic pathways to peptide aggregation on surfaces: the effects of beta-sheet propensity and surface attraction. J. Chem. Phys. 136, 065103 ( 10.1063/1.3682986) [DOI] [PubMed] [Google Scholar]
  • 69.Brorsson AC, Kumita JR, MacLeod I, Bolognesi B, Speretta E, Luheshi LM, Knowles TP, Dobson CM, Crowther DC. 2010. Methods and models in neurodegenerative and systemic protein aggregation diseases. Front. Biosci. (Landmark Ed) 15, 373–396. [DOI] [PubMed] [Google Scholar]
  • 70.Meisl G, Yang X, Frohm B, Knowles TPJ, Linse S. 2016. Quantitative analysis of intrinsic and extrinsic factors in the aggregation mechanism of Alzheimer-associated Aβ-peptide. Sci. Rep. 6, 18728 ( 10.1038/srep18728) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Iljina M, et al. 2016. Kinetic model of the aggregation of α-synuclein provides insights into prion-like spreading. Proc. Natl Acad. Sci. USA 113, E1206–E1215. ( 10.1073/pnas.1524128113) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Shammas SL, et al. 2015. A mechanistic model of tau amyloid aggregation based on direct observation of oligomers. Nat. Commun. 6, 7025 ( 10.1038/ncomms8025) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Deva T, Lorenzen N, Vad BS, Petersen SV, Thørgersen I, Enghild JJ, Kristensen T, Otzen DE. 2013. Off-pathway aggregation can inhibit fibrillation at high protein concentrations. Biochim. Biophys. Acta 1834, 677–687. ( 10.1016/j.bbapap.2012.12.020) [DOI] [PubMed] [Google Scholar]
  • 74.Powers ET, Powers DL. 2008. Mechanisms of protein fibril formation: nucleated polymerization with competing off-pathway aggregation. Biophys. J. 94, 379–391. ( 10.1529/biophysj.107.117168) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.De Baets G, Schymkowitz J, Rousseau F. 2014. Predicting aggregation-prone sequences in proteins. Essays Biochem. 56, 41–52. ( 10.1042/bse0560041) [DOI] [PubMed] [Google Scholar]
  • 76.Bolognesi B, Tartaglia GG. 2013. Physicochemical principles of protein aggregation. Prog. Mol. Biol. Transl. Sci. 117, 53–72. ( 10.1016/B978-0-12-386931-9.00003-9) [DOI] [PubMed] [Google Scholar]
  • 77.Chiti F, Stefani M, Taddei N, Ramponi G, Dobson CM. 2003. Rationalization of the effects of mutations on peptide and protein aggregation rates. Nature 424, 805–808. ( 10.1038/nature01891) [DOI] [PubMed] [Google Scholar]
  • 78.Agrawal NJ, Kumar S, Wang X, Helk B, Singh SK, Trout BL. 2011. Aggregation in protein-based biotherapeutics: computational studies and tools to identify aggregation-prone regions. J. Pharm. Sci. 100, 5081–5095. ( 10.1002/jps.22705) [DOI] [PubMed] [Google Scholar]
  • 79.Pawar AP, DuBay KF, Zurdo J, Chiti F, Vendruscolo M, Dobson CM. 2005. Prediction of ‘aggregation-prone’ and ‘aggregation-susceptible’ regions in proteins associated with neurodegenerative diseases. J. Mol. Biol. 350, 379–392. ( 10.1016/j.jmb.2005.04.016) [DOI] [PubMed] [Google Scholar]
  • 80.DuBay KF, Pawar AP, Chiti F, Zurdo J, Dobson CM, Vendruscolo M. 2004. Prediction of the absolute aggregation rates of amyloidogenic polypeptide chains. J. Mol. Biol. 341, 1317–1326. ( 10.1016/j.jmb.2004.06.043) [DOI] [PubMed] [Google Scholar]
  • 81.Zibaee S, Makin OS, Goedert M, Serpell LC. 2007. A simple algorithm locates beta-strands in the amyloid fibril core of alpha-synuclein, Aβ, and tau using the amino acid sequence alone. Protein Sci. 16, 906–918. ( 10.1110/ps.062624507) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Beerten J, Jonckheere W, Rudyak S, Xu J, Wilkinson H, De Smet F, Schymkowitz J, Rousseau F. 2012. Aggregation gatekeepers modulate protein homeostasis of aggregating sequences and affect bacterial fitness. Protein Eng. Des. Sel. 25, 357–366. ( 10.1093/protein/gzs031) [DOI] [PubMed] [Google Scholar]
  • 83.De Baets G, Van Durme J, Rousseau F, Schymkowitz J. 2014. A genome-wide sequence-structure analysis suggests aggregation gatekeepers constitute an evolutionary constrained functional class. J. Mol. Biol. 426, 2405–2412. ( 10.1016/j.jmb.2014.04.007) [DOI] [PubMed] [Google Scholar]
  • 84.Reumers J, Maurer-Stroh S, Schymkowitz J, Rousseau F. 2009. Protein sequences encode safeguards against aggregation. Hum. Mutat. 30, 431–437. ( 10.1002/humu.20905) [DOI] [PubMed] [Google Scholar]
  • 85.Tartaglia GG, Pawar AP, Campioni S, Dobson CM, Chiti F, Vendruscolo M. 2008. Prediction of aggregation-prone regions in structured proteins. J. Mol. Biol. 380, 425–436. ( 10.1016/j.jmb.2008.05.013) [DOI] [PubMed] [Google Scholar]
  • 86.Fernandez-Escamilla AM, Rousseau F, Schymkowitz J, Serrano L. 2004. Prediction of sequence-dependent and mutational effects on the aggregation of peptides and proteins. Nat. Biotechnol. 22, 1302–1306. ( 10.1038/nbt1012) [DOI] [PubMed] [Google Scholar]
  • 87.Galzitskaya OV, Garbuzynskiy SO, Lobanov MY. 2006. Prediction of amyloidogenic and disordered regions in protein chains. PLoS Comput. Biol. 2, e177 ( 10.1371/journal.pcbi.0020177) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Bui JM, Cavalli A, Gsponer J. 2008. Identification of aggregation-prone elements by using interaction-energy matrices. Angew. Chem. Int. Ed. Engl. 47, 7267–7269. ( 10.1002/anie.200802345) [DOI] [PubMed] [Google Scholar]
  • 89.Trovato A, Seno F, Tosatto SC. 2007. The PASTA server for protein aggregation prediction. Protein Eng. Des. Sel. 20, 521–523. ( 10.1093/protein/gzm042) [DOI] [PubMed] [Google Scholar]
  • 90.Bryan AW Jr, Menke M, Cowen LJ, Lindquist SL, Berger B, Dunbrack R. 2009. BETASCAN: probable beta-amyloids identified by pairwise probabilistic analysis. PLoS Comput. Biol. 5, e1000333 ( 10.1371/journal.pcbi.1000333) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Thompson MJ, Sievers SA, Karanicolas J, Ivanova MI, Baker D, Eisenberg D. 2006. The 3D profile method for identifying fibril-forming segments of proteins. Proc. Natl Acad. Sci. USA 103, 4074–4078. ( 10.1073/pnas.0511295103) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Zhang Z, Chen H, Lai L. 2007. Identification of amyloid fibril-forming segments based on structure and residue-based statistical potential. Bioinformatics 23, 2218–2225. ( 10.1093/bioinformatics/btm325) [DOI] [PubMed] [Google Scholar]
  • 93.Maurer-Stroh S, et al. 2010. Exploring the sequence determinants of amyloid structure using position-specific scoring matrices. Nat. Methods 7, 237–242. ( 10.1038/nmeth.1432) [DOI] [PubMed] [Google Scholar]
  • 94.Conchillo-Sole O, de Groot NS, Avilés FX, Vendrell J, Daura X, Ventura S. 2007. AGGRESCAN: a server for the prediction and evaluation of ‘hot spots’ of aggregation in polypeptides. BMC Bioinformatics 8, 65 ( 10.1186/1471-2105-8-65) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Sanchez de Groot N, Pallarés I, Avilés FX, Vendrell J, Ventura S. 2005. Prediction of ‘hot spots’ of aggregation in disease-linked polypeptides. BMC Struct. Biol. 5, 18 ( 10.1186/1472-6807-5-18) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Walsh I, Seno F, Tosatto SCE, Trovato A. 2014. PASTA 2.0: an improved server for protein aggregation prediction. Nucleic Acids Res. 42, W301–W307. ( 10.1093/nar/gku399) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Marshall KE, Morris KL., Charlton D, O'Reilly N, Lewis L, Walden H, Serpell LC. 2011. Hydrophobic, aromatic, and electrostatic interactions play a central role in amyloid fibril formation and stability. Biochemistry 50, 2061–2071. ( 10.1021/bi101936c) [DOI] [PubMed] [Google Scholar]
  • 98.Kamihira M, Oshiro Y, Tuzi S, Nosaka AY, Saitô H, Naito A. 2003. Effect of electrostatic interaction on fibril formation of human calcitonin as studied by high resolution solid state 13C NMR. J. Biol. Chem. 278, 2859–2865. ( 10.1074/jbc.M205285200) [DOI] [PubMed] [Google Scholar]
  • 99.Yun S, Urbanc B, Cruz L, Bitan G, Teplow DB, Stanley HE. 2007. Role of electrostatic interactions in amyloid β-protein (Aβ) oligomer formation: a discrete molecular dynamics study. Biophys. J. 92, 4064–4077. ( 10.1529/biophysj.106.097766) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Raman B, Chatani E, Kihara M, Ban T, Sakai M, Hasegawa K, Naiki H, Rao CM, Goto Y. 2005. Critical balance of electrostatic and hydrophobic interactions is required for β2-microglobulin amyloid fibril growth and stability. Biochemistry 44, 1288–1299. ( 10.1021/bi048029t) [DOI] [PubMed] [Google Scholar]
  • 101.Jain S, Udgaonkar JB. 2010. Salt-induced modulation of the pathway of amyloid fibril formation by the mouse prion protein. Biochemistry 49, 7615–7624. ( 10.1021/bi100745j) [DOI] [PubMed] [Google Scholar]
  • 102.Klement K, Wieligmann K, Meinhardt J, Hortschansky P, Richter W, Fändrich M. 2007. Effect of different salt ions on the propensity of aggregation and on the structure of Alzheimer's aβ(1–40) amyloid fibrils. J. Mol. Biol. 373, 1321–1333. ( 10.1016/j.jmb.2007.08.068) [DOI] [PubMed] [Google Scholar]
  • 103.Marek PJ, Patsalo V, Green DF, Raleigh DP. 2012. Ionic strength effects on amyloid formation by amylin are a complicated interplay among Debye screening, ion selectivity, and Hofmeister effects. Biochemistry 51, 8478–8490. ( 10.1021/bi300574r) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Calamai M, Kumita JR, Mifsud J, Parrini C, Ramazzotti M, Ramponi G, Taddei N, Chiti F, Dobson CM. 2006. Nature and significance of the interactions between amyloid fibrils and biological polyelectrolytes. Biochemistry 45, 12 806–12 815. ( 10.1021/bi0610653) [DOI] [PubMed] [Google Scholar]
  • 105.Manning MC, Chou DK, Murphy BM, Payne RW, Katayama DS. 2010. Stability of protein pharmaceuticals: an update. Pharm. Res. 27, 544–575. ( 10.1007/s11095-009-0045-6) [DOI] [PubMed] [Google Scholar]
  • 106.Torosantucci R, Schoneich C, Jiskoot W. 2014. Oxidation of therapeutic proteins and peptides: structural and biological consequences. Pharm. Res. 31, 541–553. ( 10.1007/s11095-013-1199-9) [DOI] [PubMed] [Google Scholar]
  • 107.Wakankar AA, Borchardt RT. 2006. Formulation considerations for proteins susceptible to asparagine deamidation and aspartate isomerization. J. Pharm. Sci. 95, 2321–2336. ( 10.1002/jps.20740) [DOI] [PubMed] [Google Scholar]
  • 108.Ji JA, Zhang B, Cheng W, Wang YJ. 2009. Methionine, tryptophan, and histidine oxidation in a model protein, PTH: mechanisms and stabilization. J. Pharm. Sci. 98, 4485–4500. ( 10.1002/jps.21746) [DOI] [PubMed] [Google Scholar]
  • 109.Wang W. 1999. Instability, stabilization, and formulation of liquid protein pharmaceuticals. Int. J. Pharm. 185, 129–188. ( 10.1016/S0378-5173(99)00152-0) [DOI] [PubMed] [Google Scholar]
  • 110.Bhatt NP, Patel K, Borchardt RT. 1990. Chemical pathways of peptide degradation. I. Deamidation of adrenocorticotropic hormone. Pharm. Res. 7, 593–599. ( 10.1023/A:1015862026539) [DOI] [PubMed] [Google Scholar]
  • 111.Manning MC, Patel K, Borchardt RT. 1989. Stability of protein pharmaceuticals. Pharm. Res. 6, 903–918. ( 10.1023/A:1015929109894) [DOI] [PubMed] [Google Scholar]
  • 112.Joshi AB, Kirsch LE. 2002. The relative rates of glutamine and asparagine deamidation in glucagon fragment 22–29 under acidic conditions. J. Pharm. Sci. 91, 2331–2345. ( 10.1002/jps.10213) [DOI] [PubMed] [Google Scholar]
  • 113.Joshi AB, Rus E, Kirsch LE. 2000. The degradation pathways of glucagon in acidic solutions. Int. J. Pharm. 203, 115–125. ( 10.1016/S0378-5173(00)00438-5) [DOI] [PubMed] [Google Scholar]
  • 114.Joshi AB, Kirsch LE. 2004. The estimation of glutaminyl deamidation and aspartyl cleavage rates in glucagon. Int. J. Pharm. 273, 213–219. ( 10.1016/j.ijpharm.2004.01.006) [DOI] [PubMed] [Google Scholar]
  • 115.Joshi AB, Sawai M, Kearney WR, Kirsch LE. 2005. Studies on the mechanism of aspartic acid cleavage and glutamine deamidation in the acidic degradation of glucagon. J. Pharm. Sci. 94, 1912–1927. ( 10.1002/jps.20405) [DOI] [PubMed] [Google Scholar]
  • 116.Nilsson MR, Driscoll M, Raleigh DP. 2002. Low levels of asparagine deamidation can have a dramatic effect on aggregation of amyloidogenic peptides: implications for the study of amyloid formation. Protein Sci. 11, 342–349. ( 10.1110/ps.48702) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Dunkelberger EB, Buchanan LE, Marek P, Cao P, Raleigh DP, Zanni MT. 2012. Deamidation accelerates amyloid formation and alters amylin fiber structure. J. Am. Chem. Soc. 134, 12 658–12 667. ( 10.1021/ja3039486) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Ray NJ, Hall D, Carver JA. 2016. Deamidation of N76 in human γS-crystallin promotes dimer formation. Biochim. Biophys. Acta 1860, 315–324. ( 10.1016/j.bbagen.2015.08.015) [DOI] [PubMed] [Google Scholar]
  • 119.Takata T, Oxford JT, Demeler B, Lampi KJ. 2008. Deamidation destabilizes and triggers aggregation of a lens protein, βA3-crystallin. Protein Sci. 17, 1565–1575. ( 10.1110/ps.035410.108) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Mohammadian-Mosaabadi J, Naderi-Manesh H, Maghsoudi N, Khalilzadeh R, Shojaosadati SA, Ebrahimi M. 2005. Effect of oxidative stress on the production of recombinant human interferon-γ in Escherichia coli. Biotechnol. Appl. Biochem. 41, 37–42. ( 10.1042/BA20030230) [DOI] [PubMed] [Google Scholar]
  • 121.Kerwin BA. 2008. Polysorbates 20 and 80 used in the formulation of protein biotherapeutics: structure and degradation pathways. J. Pharm. Sci. 97, 2924–2935. ( 10.1002/jps.21190) [DOI] [PubMed] [Google Scholar]
  • 122.Takenawa T, Yokota A, Oda M, Takahashi H, Iwakura M. 2009. Protein oxidation during long storage: identification of the oxidation sites in dihydrofolate reductase from Escherichia coli through LC-MS and fragment studies. J. Biochem. 145, 517–523. ( 10.1093/jb/mvp003) [DOI] [PubMed] [Google Scholar]
  • 123.Wang W. 2005. Protein aggregation and its inhibition in biopharmaceutics. Int. J. Pharm. 289, 1–30. ( 10.1016/j.ijpharm.2004.11.014) [DOI] [PubMed] [Google Scholar]
  • 124.Gu M, Viles JH. 2016. Methionine oxidation reduces lag-times for amyloid-beta(1–40) fiber formation but generates highly fragmented fibers. Biochim. Biophys. Acta 1864, 1260–1269. ( 10.1016/j.bbapap.2016.04.009) [DOI] [PubMed] [Google Scholar]
  • 125.Glaser CB, et al. 1992. Oxidation of a specific methionine in thrombomodulin by activated neutrophil products blocks cofactor activity. A potential rapid mechanism for modulation of coagulation. J. Clin. Invest. 90, 2565–2573. ( 10.1172/JCI116151) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Kertscher U, Bienert M, Krause E, Sepetov NF, Mehlis B. 1993. Spontaneous chemical degradation of substance P in the solid phase and in solution. Int. J. Pept. Protein Res. 41, 207–211. ( 10.1111/j.1399-3011.1993.tb00328.x) [DOI] [PubMed] [Google Scholar]
  • 127.Schilling S, Wasternack C, Demuth HU. 2008. Glutaminyl cyclases from animals and plants: a case of functionally convergent protein evolution. Biol. Chem. 389, 983–991. ( 10.1515/BC.2008.111) [DOI] [PubMed] [Google Scholar]
  • 128.Schlenzig D, Manhart S, Cinar Y, Kleinschmidt M, Hause G, Willbold D, Funke SA, Schilling S, Demuth H-U. 2009. Pyroglutamate formation influences solubility and amyloidogenicity of amyloid peptides. Biochemistry 48, 7072–7078. ( 10.1021/bi900818a) [DOI] [PubMed] [Google Scholar]
  • 129.Costantino HR, Langer R, Klibanov AM. 1994. Moisture-induced aggregation of lyophilized insulin. Pharm. Res. 11, 21–29. ( 10.1023/A:1018981208076) [DOI] [PubMed] [Google Scholar]
  • 130.Fujii N, Momose Y, Harada K. 1996. Kinetic study of racemization of aspartyl residues in model peptides of alpha A-crystallin. Int. J. Pept. Protein Res. 48, 118–122. ( 10.1111/j.1399-3011.1996.tb00821.x) [DOI] [PubMed] [Google Scholar]
  • 131.Murakami K, Uno M, Masuda Y, Shimizu T, Shirasawa T, Irie K. 2008. Isomerization and/or racemization at Asp23 of Aβ42 do not increase its aggregative ability, neurotoxicity, and radical productivity in vitro. Biochem. Biophys. Res. Commun. 366, 745–751. ( 10.1016/j.bbrc.2007.12.009) [DOI] [PubMed] [Google Scholar]
  • 132.Qvit N, Rubin SJS, Urban TJ, Mochly-Rosen D, Gross ER. 2017. Peptidomimetic therapeutics: scientific approaches and opportunities. Drug Discov. Today 22, 454–462. ( 10.1016/j.drudis.2016.11.003) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Furman JL, Chiu M, Hunter MJ. 2015. Early engineering approaches to improve peptide developability and manufacturability. AAPS J. 17, 111–120. ( 10.1208/s12248-014-9681-9) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Hagihara Y, Saerens D. 2014. Engineering disulfide bonds within an antibody. Biochim. Biophys. Acta 1844, 2016–2023. ( 10.1016/j.bbapap.2014.07.005) [DOI] [PubMed] [Google Scholar]
  • 135.Nick Pace C, Scholtz JM, Grimsley GR. 2014. Forces stabilizing proteins. FEBS Lett. 588, 2177–2184. ( 10.1016/j.febslet.2014.05.006) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Vinther TN, Kjeldsen TB, Jensen KJ, Hubálek F. 2015. The road to the first, fully active and more stable human insulin variant with an additional disulfide bond. J. Pept. Sci. 21, 797–806. ( 10.1002/psc.2822) [DOI] [PubMed] [Google Scholar]
  • 137.Li Y, Zheng X, Tang L, Xu W, Gong M. 2011. GLP-1 analogs containing disulfide bond exhibited prolonged half-life in vivo than GLP-1. Peptides 32, 1303–1312. ( 10.1016/j.peptides.2011.04.001) [DOI] [PubMed] [Google Scholar]
  • 138.Avadisian M, Gunning PT. 2013. Extolling the benefits of molecular therapeutic lipidation. Mol. Biosyst. 9, 2179–2188. ( 10.1039/c3mb70147f) [DOI] [PubMed] [Google Scholar]
  • 139.Aicart-Ramos C, Valero RA, Rodriguez-Crespo I. 2011. Protein palmitoylation and subcellular trafficking. Biochim. Biophys. Acta 1808, 2981–2994. ( 10.1016/j.bbamem.2011.07.009) [DOI] [PubMed] [Google Scholar]
  • 140.Li Y, Wang Y, Wei Q, Zheng X, Tang L, Kong D, Gong M. 2015. Variant fatty acid-like molecules: conjugation, novel approaches for extending the stability of therapeutic peptides. Sci. Rep. 5, 18039 ( 10.1038/srep18039) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Ward BP, Ottaway NL, Perez-Tilve D, Ma D, Gelfanov VM, Tschöp MH, DiMarchi RD. 2013. Peptide lipidation stabilizes structure to enhance biological function. Mol Metab. 2, 468–479. ( 10.1016/j.molmet.2013.08.008) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Meece J. 2009. Pharmacokinetics and pharmacodynamics of liraglutide, a long-acting, potent glucagon-like peptide-1 analog. Pharmacotherapy 29, 33s–42s. ( 10.1592/phco.29.pt2.33S) [DOI] [PubMed] [Google Scholar]
  • 143.Gilroy CA, Luginbuhl KM, Chilkoti A. 2016. Controlled release of biologics for the treatment of type 2 diabetes. J. Control. Release 240, 151–164. ( 10.1016/j.jconrel.2015.12.002) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Bush MA, Matthews JE, De Boever EH, Dobbins RL, Hodge RJ, Walker SE, Holland MC, Gutierrez M. 2009. Safety, tolerability, pharmacodynamics and pharmacokinetics of albiglutide, a long-acting glucagon-like peptide-1 mimetic, in healthy subjects. Diabetes Obes. Metab. 11, 498–505. ( 10.1111/j.1463-1326.2008.00992.x) [DOI] [PubMed] [Google Scholar]
  • 145.Matthews JE, Stewart MW, De Boever EH, Dobbins RL, Hodge RJ, Walker SE, Holland MC, Bush MA. 2008. Pharmacodynamics, pharmacokinetics, safety, and tolerability of albiglutide, a long-acting glucagon-like peptide-1 mimetic, in patients with type 2 diabetes. J. Clin. Endocrinol. Metab. 93, 4810–4817. ( 10.1210/jc.2008-1518) [DOI] [PubMed] [Google Scholar]
  • 146.Wang Y, Lomakin A, Kanai S, Alex R, Benedek GB. 2015. Transformation of oligomers of lipidated peptide induced by change in pH. Mol. Pharm. 12, 411–419. ( 10.1021/mp500519s) [DOI] [PubMed] [Google Scholar]
  • 147.Vad B, et al. 2010. Divorcing folding from function: how acylation affects the membrane-perturbing properties of an antimicrobial peptide. Biochim. Biophys. Acta 1804, 806–820. ( 10.1016/j.bbapap.2009.12.006) [DOI] [PubMed] [Google Scholar]
  • 148.Prive GG. 2009. Lipopeptide detergents for membrane protein studies. Curr. Opin. Struct. Biol. 19, 379–385. ( 10.1016/j.sbi.2009.07.008) [DOI] [PubMed] [Google Scholar]
  • 149.Gunn AP, Masters CL, Cherny RA. 2010. Pyroglutamate-Abeta: role in the natural history of Alzheimer's disease. Int. J. Biochem. Cell Biol. 42, 1915–1918. ( 10.1016/j.biocel.2010.08.015) [DOI] [PubMed] [Google Scholar]
  • 150.Goldblatt G, Matos JO, Gornto J, Tatulian SA. 2015. Isotope-edited FTIR reveals distinct aggregation and structural behaviors of unmodified and pyroglutamylated amyloid β peptides. Phys. Chem. Chem. Phys. 17, 32 149–32 160. ( 10.1039/C5CP03343H) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Thal DR, Walter J, Saido TC, Fändrich M. 2015. Neuropathology and biochemistry of Aβ and its aggregates in Alzheimer's disease. Acta Neuropathol. 129, 167–182. ( 10.1007/s00401-014-1375-y) [DOI] [PubMed] [Google Scholar]
  • 152.Rezaei-Ghaleh N, Amininasab M, Kumar S, Walter J, Zweckstetter M. 2016. Phosphorylation modifies the molecular stability of beta-amyloid deposits. Nat. Commun. 7, 11359 ( 10.1038/ncomms11359) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Rezaei-Ghaleh N, Kumar S, Walter J, Zweckstetter M. 2016. Phosphorylation interferes with maturation of amyloid-beta fibrillar structure in the N terminus. J. Biol. Chem. 291, 16 059–16 067. ( 10.1074/jbc.M116.728956) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Osaki D, Hiramatsu H. 2016. Citrullination and deamidation affect aggregation properties of amyloid beta-proteins. Amyloid 23, 234–241. ( 10.1080/13506129.2016.1240076) [DOI] [PubMed] [Google Scholar]
  • 155.Guivernau B, et al. 2016. Amyloid-beta peptide nitrotyrosination stabilizes oligomers and enhances NMDAR-mediated toxicity. J. Neurosci. 36, 11 693–11 703. ( 10.1523/JNEUROSCI.1081-16.2016) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Johnson EC, Lanning JD, Meredith SC. 2016. Peptide backbone modification in the bend region of amyloid-beta inhibits fibrillogenesis but not oligomer formation. J. Pept. Sci. 22, 368–373. ( 10.1002/psc.2879) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Bartels T, Kim NC, Luth ES, Selkoe DJ, Pastore A. 2014. N-alpha-acetylation of alpha-synuclein increases its helical folding propensity, GM1 binding specificity and resistance to aggregation. PLoS ONE 9, e103727 ( 10.1371/journal.pone.0103727) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Iyer A, Roeters SJ, Schilderink N, Hommersom B, Heeren RMA, Woutersen S, Claessens MMAE, Subramaniam V. 2016. The impact of N-terminal acetylation of alpha-synuclein on phospholipid membrane binding and fibril structure. J. Biol. Chem. 291, 21 110–21 122. ( 10.1074/jbc.M116.726612) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Xiang W, et al. 2013. Oxidative stress-induced posttranslational modifications of alpha-synuclein: specific modification of α-synuclein by 4-hydroxy-2-nonenal increases dopaminergic toxicity. Mol. Cell. Neurosci. 54, 71–83. ( 10.1016/j.mcn.2013.01.004) [DOI] [PubMed] [Google Scholar]
  • 160.Bae EJ, Ho D-H, Park E, Jung JW, Cho K, Hong JH, Lee H-J, Kim KP, Lee S-J. 2013. Lipid peroxidation product 4-hydroxy-2-nonenal promotes seeding-capable oligomer formation and cell-to-cell transfer of alpha-synuclein. Antioxid. Redox Signal. 18, 770–783. ( 10.1089/ars.2011.4429) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Carmo-Goncalves P, Pinheiro AS, Romão L, Cortines J, Follmer C. 2014. UV-induced selective oxidation of Met5 to Met-sulfoxide leads to the formation of neurotoxic fibril-incompetent alpha-synuclein oligomers. Amyloid 21, 163–174. ( 10.3109/13506129.2014.912208) [DOI] [PubMed] [Google Scholar]
  • 162.Chen F, David D, Ferrari A, Gotz J. 2004. Posttranslational modifications of tau—role in human tauopathies and modeling in transgenic animals. Curr. Drug Targets 5, 503–515. ( 10.2174/1389450043345236) [DOI] [PubMed] [Google Scholar]
  • 163.Lin C, et al. 2014. Comparison of the anti-amyloidogenic effect of O-mannosylation, O-galactosylation, and O-GalNAc glycosylation. Carbohydr. Res. 387, 46–53. ( 10.1016/j.carres.2014.01.025) [DOI] [PubMed] [Google Scholar]
  • 164.Rawat A, Nagaraj R. 2013. Covalently attached fatty acyl chains alter the aggregation behavior of an amyloidogenic peptide derived from human β2-microglobulin. J. Pept. Sci. 19, 770–783. ( 10.1002/psc.2575) [DOI] [PubMed] [Google Scholar]
  • 165.Nygren H, Stenberg M. 1989. Immunochemistry at interfaces. Immunology 66, 321–327. [PMC free article] [PubMed] [Google Scholar]
  • 166.Stenberg M, Nygren H. 1988. Kinetics of antigen-antibody reactions at solid-liquid interfaces. J. Immunol. Methods 113, 3–15. ( 10.1016/0022-1759(88)90376-6) [DOI] [PubMed] [Google Scholar]
  • 167.Wang W, Nema S, Teagarden D. 2010. Protein aggregation—pathways and influencing factors. Int. J. Pharm. 390, 89–99. ( 10.1016/j.ijpharm.2010.02.025) [DOI] [PubMed] [Google Scholar]
  • 168.Rabe M, Verdes D, Seeger S. 2011. Understanding protein adsorption phenomena at solid surfaces. Adv. Colloid Interface Sci. 162, 87–106. ( 10.1016/j.cis.2010.12.007) [DOI] [PubMed] [Google Scholar]
  • 169.Nygren H. 1995. Logarithmic growth in surface adsorption. Adv. Colloid Interface Sci. 62, 137–159. ( 10.1016/0001-8686(95)00277-W) [DOI] [PubMed] [Google Scholar]
  • 170.Nault L, Vendrely C, Bréchet Y, Bruckert F, Weidenhaupt M. 2013. Peptides that form beta-sheets on hydrophobic surfaces accelerate surface-induced insulin amyloidal aggregation. FEBS Lett. 587, 1281–1286. ( 10.1016/j.febslet.2012.11.036) [DOI] [PubMed] [Google Scholar]
  • 171.Nault L, Guo P, Jain B, Bréchet Y, Bruckert F, Weidenhaupt M. 2013. Human insulin adsorption kinetics, conformational changes and amyloidal aggregate formation on hydrophobic surfaces. Acta Biomater. 9, 5070–5079. ( 10.1016/j.actbio.2012.09.025) [DOI] [PubMed] [Google Scholar]
  • 172.Huotari A, Xu W, Mönkäre J, Kovalainen M, Herzig K-H, Lehto V-P, Järvinen K. 2013. Effect of surface chemistry of porous silicon microparticles on glucagon-like peptide-1 (GLP-1) loading, release and biological activity. Int. J. Pharm. 454, 67–73. ( 10.1016/j.ijpharm.2013.06.063) [DOI] [PubMed] [Google Scholar]
  • 173.Mauri S, Volk M, Byard S, Berchtold H, Arnolds H. 2015. Stabilization of insulin by adsorption on a hydrophobic silane self-assembled monolayer. Langmuir 31, 8892–8900. ( 10.1021/acs.langmuir.5b01477) [DOI] [PubMed] [Google Scholar]
  • 174.Pandey LM, Le Denmat S, Delabouglise D, Bruckert F, Pattanayek SK, Weidenhaupt M. 2012. Surface chemistry at the nanometer scale influences insulin aggregation. Colloids Surf. B Biointerfaces 100, 69–76. ( 10.1016/j.colsurfb.2012.05.022) [DOI] [PubMed] [Google Scholar]
  • 175.Feingold V, Jenkins AB, Kraegen EW. 1984. Effect of contact material on vibration-induced insulin aggregation. Diabetologia 27, 373–378. ( 10.1007/BF00304853) [DOI] [PubMed] [Google Scholar]
  • 176.Sefton MV, Antonacci GM. 1984. Adsorption isotherms of insulin onto various materials. Diabetes 33, 674–680. ( 10.2337/diab.33.7.674) [DOI] [PubMed] [Google Scholar]
  • 177.Lougheed WD, Woulfe-Flanagan H, Clement JR, Albisser AM. 1980. Insulin aggregation in artificial delivery systems. Diabetologia 19, 1–9. ( 10.1007/BF00258302) [DOI] [PubMed] [Google Scholar]
  • 178.Moores B, Drolle E, Attwood SJ, Simons J, Leonenko Z. 2011. Effect of surfaces on amyloid fibril formation. PLoS ONE 6, e25954 ( 10.1371/journal.pone.0025954) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Chi EY, Frey SL, Winans A, Lam KLH, Kjaer K, Majewski J, Lee KYC. 2010. Amyloid-β fibrillogenesis seeded by interface-induced peptide misfolding and self-assembly. Biophys. J. 98, 2299–2308. ( 10.1016/j.bpj.2010.01.056) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 180.Giacomelli CE, Norde W. 2005. Conformational changes of the amyloid β-peptide (1–40) adsorbed on solid surfaces. Macromol. Biosci. 5, 401–407. ( 10.1002/mabi.200400189) [DOI] [PubMed] [Google Scholar]
  • 181.Giacomelli CE, Norde W. 2003. Influence of hydrophobic Teflon particles on the structure of amyloid beta-peptide. Biomacromolecules 4, 1719–1726. ( 10.1021/bm034151g) [DOI] [PubMed] [Google Scholar]
  • 182.Kowalewski T, Holtzman DM. 1999. In situ atomic force microscopy study of Alzheimer's β-amyloid peptide on different substrates: new insights into mechanism of β-sheet formation. Proc. Natl Acad. Sci. USA 96, 3688–3693. ( 10.1073/pnas.96.7.3688) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Pinholt C, Kapp SJ, Bukrinsky JT, Hostrup S, Frokjaer S, Norde W, Jorgensen L. 2013. Influence of acylation on the adsorption of GLP-2 to hydrophobic surfaces. Int. J. Pharm. 440, 63–71. ( 10.1016/j.ijpharm.2012.01.040) [DOI] [PubMed] [Google Scholar]
  • 184.Trier S, Linderoth L, Bjerregaard S, Andresen TL, Rahbek UL, Castanho MARB. 2014. Acylation of glucagon-like peptide-2: interaction with lipid membranes and in vitro intestinal permeability. PLoS ONE 9, e109939 ( 10.1371/journal.pone.0109939) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Lu DR. 1993. Glucagon adsorption on polymer surfaces with alpha-helical and extended beta-strand conformations: a computational approach. J. Biomater. Sci. Polym. Ed. 4, 323–335. ( 10.1163/156856293X00014) [DOI] [PubMed] [Google Scholar]
  • 186.Pinholt C, Bukrinsky JT, Hostrup S, Frokjaer S, Norde W, Jorgensen L. 2011. Influence of PEGylation with linear and branched PEG chains on the adsorption of glucagon to hydrophobic surfaces. Eur. J. Pharm. Biopharm. 77, 139–147. ( 10.1016/j.ejpb.2010.11.001) [DOI] [PubMed] [Google Scholar]
  • 187.Rabe M, Soragni A, Reynolds NP, Verdes D, Liverani E, Riek R, Seeger S. 2013. On-surface aggregation of alpha-synuclein at nanomolar concentrations results in two distinct growth mechanisms. ACS Chem. Neurosci. 4, 408–417. ( 10.1021/cn3001312) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Schladitz C, Vieira EP, Hermel H, Möhwald H. 1999. Amyloid-β-sheet formation at the air-water interface. Biophys. J. 77, 3305–3310. ( 10.1016/S0006-3495(99)77161-4) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 189.Mauri S, Weidner T, Arnolds H. 2014. The structure of insulin at the air/water interface: monomers or dimers? Phys. Chem. Chem. Phys. 16, 26 722–26 724. ( 10.1039/C4CP04926H) [DOI] [PubMed] [Google Scholar]
  • 190.Hlady VV, Buijs J. 1996. Protein adsorption on solid surfaces. Curr. Opin. Biotechnol. 7, 72–77. ( 10.1016/S0958-1669(96)80098-X) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Kamerzell TJ, Esfandiary R, Joshi SB, Middaugh CR, Volkin DB. 2011. Protein-excipient interactions: mechanisms and biophysical characterization applied to protein formulation development. Adv. Drug Deliv. Rev. 63, 1118–1159. ( 10.1016/j.addr.2011.07.006) [DOI] [PubMed] [Google Scholar]
  • 192.Zbacnik TJ, et al. 2017. Role of buffers in protein formulations. J. Pharm. Sci. 106, 713–733. ( 10.1016/j.xphs.2016.11.014) [DOI] [PubMed] [Google Scholar]
  • 193.Pedersen JS, Dikov D, Otzen DE. 2006. N- and C-terminal hydrophobic patches are involved in fibrillation of glucagon. Biochemistry 45, 14 503–14 512. ( 10.1021/bi061228n) [DOI] [PubMed] [Google Scholar]
  • 194.Abelein A, Jarvet J, Barth A, Gräslund A, Danielsson J. 2016. Ionic strength modulation of the free energy landscape of Aβ40 peptide fibril formation. J. Am. Chem. Soc. 138, 6893–6902. ( 10.1021/jacs.6b04511) [DOI] [PubMed] [Google Scholar]
  • 195.Bee JS, Randolph TW, Carpenter JF, Bishop SM, Dimitrova MN. 2011. Effects of surfaces and leachables on the stability of biopharmaceuticals. J. Pharm. Sci. 100, 4158–4170. ( 10.1002/jps.22597) [DOI] [PubMed] [Google Scholar]
  • 196.Khan TA, Mahler HC, Kishore RS. 2015. Key interactions of surfactants in therapeutic protein formulations: a review. Eur. J. Pharm. Biopharm. 97, 60–67. ( 10.1016/j.ejpb.2015.09.016) [DOI] [PubMed] [Google Scholar]
  • 197.Lee HJ, McAuley A, Schilke KF, McGuire J. 2011. Molecular origins of surfactant-mediated stabilization of protein drugs. Adv. Drug Deliv. Rev. 63, 1160–1171. ( 10.1016/j.addr.2011.06.015) [DOI] [PubMed] [Google Scholar]
  • 198.Frokjaer S, Otzen DE. 2005. Protein drug stability: a formulation challenge. Nat. Rev. Drug Discov. 4, 298–306. ( 10.1038/nrd1695) [DOI] [PubMed] [Google Scholar]
  • 199.Nidhi K, Indrajeet S, Khushboo M, Gauri K, Sen DJ. 2011. Hydrotropy: a promising tool for solubility enhancement: a review. Int. J. Drug Dev. Res. 3, 26–33. [Google Scholar]
  • 200.Kishore RS, Kiese S, Fischer S, Pappenberger A, Grauschopf U, Mahler H-C. 2011. The degradation of polysorbates 20 and 80 and its potential impact on the stability of biotherapeutics. Pharm. Res. 28, 1194–1210. ( 10.1007/s11095-011-0385-x) [DOI] [PubMed] [Google Scholar]
  • 201.Kishore RS, Pappenberger A, Dauphin IB, Ross A, Buergi B, Staempfli A, Mahler H-C. 2011. Degradation of polysorbates 20 and 80: studies on thermal autoxidation and hydrolysis. J. Pharm. Sci. 100, 721–731. ( 10.1002/jps.22290) [DOI] [PubMed] [Google Scholar]
  • 202.Tomlinson A, Demeule B, Lin B, Yadav S. 2015. Polysorbate 20 degradation in biopharmaceutical formulations: quantification of free fatty acids, characterization of particulates, and insights into the degradation mechanism. Mol. Pharm. 12, 3805–3815. ( 10.1021/acs.molpharmaceut.5b00311) [DOI] [PubMed] [Google Scholar]
  • 203.Wang W, Wang YJ, Wang DQ. 2008. Dual effects of Tween 80 on protein stability. Int. J. Pharm. 347, 31–38. ( 10.1016/j.ijpharm.2007.06.042) [DOI] [PubMed] [Google Scholar]
  • 204.Ha E, Wang W, Wang YJ. 2002. Peroxide formation in polysorbate 80 and protein stability. J. Pharm. Sci. 91, 2252–2264. ( 10.1002/jps.10216) [DOI] [PubMed] [Google Scholar]
  • 205.Maggio ET. 2010. Use of excipients to control aggregation in peptide and protein formulations. Development 1, 40–49. [Google Scholar]
  • 206.Otzen D. 2011. Protein–surfactant interactions: a tale of many states. Biochim. Biophys. Acta 1814, 562–591. ( 10.1016/j.bbapap.2011.03.003) [DOI] [PubMed] [Google Scholar]
  • 207.Macchi F, Eisenkolb M, Kiefer H, Otzen DE. 2012. The effect of osmolytes on protein fibrillation. Int. J. Mol. Sci. 13, 3801–3819. ( 10.3390/ijms13033801) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 208.Lam XM, Yang JY, Cleland JL. 1997. Antioxidants for prevention of methionine oxidation in recombinant monoclonal antibody HER2. J. Pharm. Sci. 86, 1250–1255. ( 10.1021/js970143s) [DOI] [PubMed] [Google Scholar]
  • 209.Tsai PK, et al. 1993. Formulation design of acidic fibroblast growth factor. Pharm. Res. 10, 649–659. ( 10.1023/A:1018939228201) [DOI] [PubMed] [Google Scholar]
  • 210.Zhou S, Sturm E, Teagarden DL, Schöneich C, Kolhe P, Lewis LM, Muralidhara BK, Singh SK. 2010. Comparative evaluation of disodium edetate and diethylenetriaminepentaacetic acid as iron chelators to prevent metal-catalyzed destabilization of a therapeutic monoclonal antibody. J. Pharm. Sci. 99, 4239–4250. ( 10.1002/jps.22141) [DOI] [PubMed] [Google Scholar]
  • 211.Meyer BK, Ni A, Hu B, Shi L. 2007. Antimicrobial preservative use in parenteral products: past and present. J. Pharm. Sci. 96, 3155–3167. ( 10.1002/jps.20976) [DOI] [PubMed] [Google Scholar]
  • 212.Zhang Y, Roy S, Jones LS, Krishnan S, Kerwin BA, Chang BS, Manning MC, Randolph TW, Carpenter JF. 2004. Mechanism for benzyl alcohol-induced aggregation of recombinant human interleukin-1 receptor antagonist in aqueous solution. J. Pharm. Sci. 93, 3076–3089. ( 10.1002/jps.20219) [DOI] [PubMed] [Google Scholar]
  • 213.Singh SM, Hutchings RL, Mallela KM. 2011. Mechanisms of m-cresol-induced protein aggregation studied using a model protein cytochrome c. J. Pharm. Sci. 100, 1679–1689. ( 10.1002/jps.22426) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 214.Teska BM, Alarcón J, Pettis RJ, Randolph TW, Carpenter JF. 2014. Effects of phenol and meta-cresol depletion on insulin analog stability at physiological temperature. J. Pharm. Sci. 103, 2255–2267. ( 10.1002/jps.24039) [DOI] [PubMed] [Google Scholar]
  • 215.Bye JW, Platts L, Falconer RJ. 2014. Biopharmaceutical liquid formulation: a review of the science of protein stability and solubility in aqueous environments. Biotechnol. Lett. 36, 869–875. ( 10.1007/s10529-013-1445-6) [DOI] [PubMed] [Google Scholar]
  • 216.Ohtake S, Kita Y, Arakawa T. 2011. Interactions of formulation excipients with proteins in solution and in the dried state. Adv. Drug Deliv. Rev. 63, 1053–1073. ( 10.1016/j.addr.2011.06.011) [DOI] [PubMed] [Google Scholar]
  • 217.Lee JC. 2000. Biopharmaceutical formulation. Curr. Opin. Biotechnol. 11, 81–84. ( 10.1016/S0958-1669(99)00058-0) [DOI] [PubMed] [Google Scholar]
  • 218.Finder VH, Vodopivec I, Nitsch RM, Glockshuber R. 2010. The recombinant amyloid-β peptide Aβ1–42 aggregates faster and is more neurotoxic than synthetic Aβ1–42. J. Mol. Biol. 396, 9–18. ( 10.1016/j.jmb.2009.12.016) [DOI] [PubMed] [Google Scholar]
  • 219.Pedersen JS, Dikov D, Flink JL, Hjuler HA, Christiansen G, Otzen DE. 2006. The changing face of glucagon fibrillation: structural polymorphism and conformational imprinting. J. Mol. Biol. 355, 501–523. ( 10.1016/j.jmb.2005.09.100) [DOI] [PubMed] [Google Scholar]
  • 220.Pedersen JS, Flink JM, Dikov D, Otzen DE. 2006. Sulfates dramatically stabilize a salt-dependent type of glucagon fibrils. Biophys. J. 90, 4181–4194. ( 10.1529/biophysj.105.070912) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 221.Doran TM, Anderson EA, Latchney SE, Opanashuk LA, Nilsson BL. 2012. Turn nucleation perturbs amyloid beta self-assembly and cytotoxicity. J. Mol. Biol. 421, 315–328. ( 10.1016/j.jmb.2012.01.055) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222.Doran TM, Kamens AJ, Byrnes NK, Nilsson BL. 2012. Role of amino acid hydrophobicity, aromaticity, and molecular volume on IAPP(20–29) amyloid self-assembly. Proteins 80, 1053–1065. ( 10.1002/prot.24007) [DOI] [PubMed] [Google Scholar]
  • 223.Goransson AL, Nilsson KPR, Kågedal K, Brorsson A-C. 2012. Identification of distinct physiochemical properties of toxic prefibrillar species formed by Abeta peptide variants. Biochem. Biophys. Res. Commun. 420, 895–900. ( 10.1016/j.bbrc.2012.03.097) [DOI] [PubMed] [Google Scholar]
  • 224.Ito M, Johansson J, Strömberg R, Nilsson L, Fraternali F. 2012. Effects of ligands on unfolding of the amyloid β-peptide central helix: mechanistic insights from molecular dynamics simulations. PLoS ONE 7, e30510 ( 10.1371/journal.pone.0030510) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225.Psonka-Antonczyk KM, et al. 2012. Nanoscopic and photonic ultrastructural characterization of two distinct insulin amyloid states. Int. J. Mol. Sci. 13, 1461–1480. ( 10.3390/ijms13021461) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 226.Swanekamp RJ, DiMaio JTM, Bowerman CJ, Nilsson BL. 2012. Coassembly of enantiomeric amphipathic peptides into amyloid-inspired rippled β-sheet fibrils. J. Am. Chem. Soc. 134, 5556–5559. ( 10.1021/ja301642c) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 227.Zako T, Sakono M, Kobayashi T, Sörgjerd K, Nilsson KPR, Hammarström P, Lindgren M, Maeda M. 2012. Cell interaction study of amyloid by using luminescent conjugated polythiophene: implication that amyloid cytotoxicity is correlated with prolonged cellular binding. ChemBioChem 13, 358–363. ( 10.1002/cbic.201100467) [DOI] [PubMed] [Google Scholar]
  • 228.Shen CL, Fitzgerald MC, Murphy RM. 1994. Effect of acid predissolution on fibril size and fibril flexibility of synthetic beta-amyloid peptide. Biophys. J. 67, 1238–1246. ( 10.1016/S0006-3495(94)80593-4) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 229.Hilbich C, Kisters-Woike B, Reed J, Masters CL, Beyreuther K. 1991. Human and rodent sequence analogs of Alzheimer's amyloid β A4 share similar properties and can be solubilized in buffers of pH 7.4. Eur. J. Biochem. 201, 61–69. ( 10.1111/j.1432-1033.1991.tb16256.x) [DOI] [PubMed] [Google Scholar]
  • 230.Hilbich C, Kisters-Woike B, Reed J, Masters CL, Beyreuther K. 1991. Aggregation and secondary structure of synthetic amyloid β A4 peptides of Alzheimer's disease. J. Mol. Biol. 218, 149–163. ( 10.1016/0022-2836(91)90881-6) [DOI] [PubMed] [Google Scholar]
  • 231.Burdick D, Soreghan B, Kwon M, Kosmoski J, Knauer M, Henschen A, Yates J, Cotman C, Glabe C. 1992. Assembly and aggregation properties of synthetic Alzheimer's A4/β amyloid peptide analogs. J. Biol. Chem. 267, 546–554. [PubMed] [Google Scholar]
  • 232.Fezoui Y, et al. 2000. An improved method of preparing the amyloid beta-protein for fibrillogenesis and neurotoxicity experiments. Amyloid 7, 166–178. ( 10.3109/13506120009146831) [DOI] [PubMed] [Google Scholar]
  • 233.Fezoui Y, Teplow DB, Hartley DM, Walsh DM, Selkoe DJ, Osterhout JJ. 2000. A de novo designed helix-turn-helix peptide forms nontoxic amyloid fibrils. Nat. Struct. Biol. 7, 1095–1099. ( 10.1038/81937) [DOI] [PubMed] [Google Scholar]
  • 234.Cordeiro Y, Foguel D, Silva JL. 2013. Pressure-temperature folding landscape in proteins involved in neurodegenerative diseases and cancer. Biophys. Chem. 183, 9–18. ( 10.1016/j.bpc.2013.06.002) [DOI] [PubMed] [Google Scholar]
  • 235.Seefeldt MB, Rosendahl M, Cleland J, Hesterberg L. 2009. Application of high hydrostatic pressure to dissociate aggregates and refold proteins. Curr. Pharm. Biotechnol. 10, 447–455. ( 10.2174/138920109788488914) [DOI] [PubMed] [Google Scholar]
  • 236.Liu R, Su R, Liang M, Huang R, Wang M, Qi W, He Z. 2012. Physicochemical strategies for inhibition of amyloid fibril formation: an overview of recent advances. Curr. Med. Chem. 19, 4157–4174. ( 10.2174/092986712802430018) [DOI] [PubMed] [Google Scholar]
  • 237.Wang W, Roberts CJ. 2013. Non-Arrhenius protein aggregation. AAPS J. 15, 840–851. ( 10.1208/s12248-013-9485-3) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 238.Mishra R, Winter R. 2008. Cold- and pressure-induced dissociation of protein aggregates and amyloid fibrils. Angew. Chem. Int. Ed. Engl. 47, 6518–6521. ( 10.1002/anie.200802027) [DOI] [PubMed] [Google Scholar]
  • 239.Kim HY, Cho M-K, Riedel D, Fernandez CO, Zweckstetter M. 2008. Dissociation of amyloid fibrils of alpha-synuclein in supercooled water. Angew. Chem. Int. Ed. Engl. 47, 5046–5048. ( 10.1002/anie.200800342) [DOI] [PubMed] [Google Scholar]
  • 240.Giehm L, Lorenzen N, Otzen DE. 2011. Assays for α-synuclein aggregation. Methods 53, 295–305. ( 10.1016/j.ymeth.2010.12.008) [DOI] [PubMed] [Google Scholar]
  • 241.Batzli KM, Love BJ. 2015. Agitation of amyloid proteins to speed aggregation measured by ThT fluorescence: a call for standardization. Mater. Sci. Eng. C Mater. Biol. Appl. 48, 359–364. ( 10.1016/j.msec.2014.09.015) [DOI] [PubMed] [Google Scholar]
  • 242.Cicerone MT, Pikal MJ, Qian KK. 2015. Stabilization of proteins in solid form. Adv. Drug Deliv. Rev. 93, 14–24. ( 10.1016/j.addr.2015.05.006) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 243.Roy I, Gupta MN. 2004. Freeze-drying of proteins: some emerging concerns. Biotechnol. Appl. Biochem. 39, 165–177. ( 10.1042/BA20030133) [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

This article has no additional data.


Articles from Interface Focus are provided here courtesy of The Royal Society

RESOURCES