Home Life Sciences Size-based Degradation of Therapeutic Proteins - Mechanisms, Modelling and Control
Article Open Access

Size-based Degradation of Therapeutic Proteins - Mechanisms, Modelling and Control

  • Rohit Bansal EMAIL logo , Saurabh Kumar Jha and Niraj Kumar Jha EMAIL logo
Published/Copyright: June 19, 2021

Abstract

Protein therapeutics are in great demand due to their effectiveness towards hard-to-treat diseases. Despite their high demand, these bio-therapeutics are very susceptible to degradation via aggregation, fragmentation, oxidation, and reduction, all of which are very likely to affect the quality and efficacy of the product. Mechanisms and modelling of these degradation (aggregation and fragmentation) pathways is critical for gaining a deeper understanding of stability of these products. This review aims to provide a summary of major developments that have occurred towards unravelling the mechanisms of size-based protein degradation (particularly aggregation and fragmentation), modelling of these size-based degradation pathways, and their control. Major caveats that remain in our understanding and control of size-based protein degradation have also been presented and discussed.

Introduction

Bio-therapeutics have become the fastest growing segment of the pharmaceutical industry, owing to their ability to combat and treat important debilitating life threatening diseases like cancers, autoimmune disorders, and metabolism related diseases [1]. The demand and growth of this product class has been significantly higher than that of the traditional small molecule pharmaceuticals [2]. However, despite the significant growth in bio-therapeutic products, these products are extremely complex and vulnerable to the various degradation pathways, which threaten to impact the safety and efficacy of the bio-therapeutic [3]. As a result, controlling degradation is a major task for the biopharmaceutical industry and the industry spends significant resources on production, transport and storage of these products so as to contain if not completely avoid degradation [4,5].

It is accepted that it is not economically feasible and, in some cases, even technically feasible to make a product that is completely devoid of any kind of impurity (both intrinsic and extrinsic) [4]. The process for the manufacturing of these bio-therapeutic drug products is quite complex, with multiple unit operations used for production, purification, characterization, storage, filling and shipping [6]. At various stages of production, the protein therapeutic may be exposed to different types of stresses such as chemical stress (pH), thermal stress (high temperature, cold temperature, freeze thaw, mechanical stress, and stress due to interaction amongst various types of stresses [7].

Structure and configuration of a protein determines its vulnerability towards the various degradation pathways such as oxidation, deamidation, and hydrolysis [8]. Degradation is influenced by a myriad of factors, which can be categorised as intrinsic and extrinsic. Intrinsic factors represent properties of the protein under consideration. Extrinsic factors include those that represent the environment (pH, temperature, buffer solutions, and changes in excipients) of the protein [9]. Hence, typical strategies for reducing protein degradation aim to either incorporate changes in the protein's structure (protein engineering) to make the protein resistant to aggregation or to control the environmental conditions [3].

To monitor protein size-based degradation, numerous analytical characterization techniques have been used. Some of the commonly used quantitative analytical tools include size exclusion – high performance liquid chromatography (SE-HPLC), dynamic light scattering (DLS), nano particle tracking analysis (NTA), micro flow imaging (MFI), analytical ultra-centrifugation (AUC), coulter counter, multi-angle laser light scattering (MALLS), and turbidimetry [10,11]. On the other hand, commonly used qualitative tools that can provide structural information about the protein include circular dichroism, fluorescence spectroscopy, infrared (IR) spectroscopy, Raman spectroscopy, ultraviolet (UV) visible spectroscopy, light and electron microscopy, fluorescence microscopy, and atomic force microscopy (AFM). Each of the above-mentioned analytical tools works well in a defined size range and hence a combination of analytical tools is required if one intents to monitor size-based degradation from small fragments and dimerization to formation of sub-visible particles and visible precipitates [12,13].

This review aims to provide a summary of major developments that have occurred towards unravelling the mechanisms of protein‘s size-based degradation (particularly aggregation and fragmentation), modelling of these degradation pathways, and control of protein‘s size-based degradation. This review focuses on the work done by the researchers on modelling, mechanism and control of protein‘s size-based degradation. Major caveats that remain in our understanding and control of protein‘s size-based degradation have also been presented and discussed.

Factors affecting size-based protein degradation

One of the major challenges in rationalizing protein‘s size-based degradation is represented by the variety of possible scenarios that can occur, which depend on the specific combination of protein, environmental conditions and exposed stress [14]. Figure 1 illustrates the various intrinsic and extrinsic factors that affect protein‘s size-based degradation. Table 1 summarises the various factors affecting size-based protein degradations.

Figure 1 Factors affecting protein aggregations.
Figure 1

Factors affecting protein aggregations.

Table 1

Summary of the factors affecting size-based protein degradation.

Factors affecting Size-based protein degradation Category (Intrinsic or Extrinsic) Mechanism of size-based degradation Reference(s)
Protein sequence Intrinsic Greater the number of hydrophobic residues in a protein, more it is susceptible to aggregation. [15,16,17]
Protein structure Intrinsic Proteins with high amount of beta sheets are more prone to aggregation as compared to alpha helix rich proteins. [9]
Post translation modifications Extrinsic PTMs like glycosylation affects structure of proteins affecting its susceptibility to aggregation. [22,23]
pH Extrinsic Low pH causes unfolding of the proteins, exposing the hydrophobic residues, increasing the chances of aggregation. [4,6,50]
Temperature Extrinsic High temperature causes protein to unfold and increases the random Brownian motion, leading to more collision between molecules, hence enhancing aggregation. Low temperature causes protein unfolding via cold denaturation. [34,35,36,37]
Ionic strength Extrinsic High ionic strength causes short range electrostatic interactions promoting aggregation. [46,47,48]
Protein concentration Extrinsic High concentration of protein therapeutics can cause crowding effect and protein-protein self-interaction in solution, leading to enhanced aggregation. [51]
Interfacial shear stress Extrinsic At the interfaces, proteins experience stresses such as hydrophobic forces, charge and mechanical stress leading to the formation of visible and sub-visible particles in biologics. [53,54,55,56]

Intrinsic factors include protein structure and sequence. One of the key contributors towards protein aggregation is the number of hydrophobic residues present in a protein sequence, which in turn affects the formation of aggregation prone regions [15,16,17]. Along with the primary sequence of the protein in IgG based mAb molecules, presence of glycosylation and types of glycans in CH2 domains also affects the stability of the complete immunoglobulin protein (Table I) [15,16,17]. Changes in the CH2 domain of mAb molecules affect the tertiary structure of the proteins, thereby affecting both the Fc and Fab regions of the IgG. The secondary structure of a protein also plays a key role in affecting protein stability. Proteins containing high percentage of alpha helices are more stable as compared to proteins primarily containing beta sheets (Table I) [9]. Those amino acids or blocks in a protein which are susceptible to aggregation can affect the overall aggregation propensity if they are exposed to the protein surface [13]. These aggregation prone blocks are generally up to seven amino acids long and in IgGs, 14 such blocks or motifs have been discovered [18].

The effect of sequence in protein stability is becoming more important with the advent of biosimilars. In the case of mAbs, although there might be differences amongst an innovator and its biosimilars in terms of PTMs, manufacturing methods, and product profiles, similar behaviour under forced degradation studies has been reported in case of infliximab innovator product (Remicade®) and its biosimilar (Remsima®) [19]. The authors concluded that this similarity in degradation pattern of these molecules is due to their common primary sequence. Other researchers have observed that mAbs belonging to different subclasses of IgGs, although having identical sequences of variable regions, can still differ in their aggregation susceptibility at low pH [20]. mAbs belonging to IgG1 subclass typically exhibit better stability as compared to IgG2 and IgG4 subclasses, likely due to enhanced flexibility of the hinge region in IgG1 mAbs causing a shielding effect for the hydrophobic residues in the hinge region [20]. In the case of IgG2 and IgG4 subclasses, unfolding of the Fc region at low pH results in exposure of the hydrophobic patches to the outside buffer environment and thereby inducing aggregation. However, if the mAb is glycosylated at these positions, then its aggregation propensity is significantly lowered [21].

Post translation modifications are known to significantly impact the overall stability of a bio-therapeutic [22]. PTMs generally determine the final structure and function of a bio-therapeutic and are very cell and species specific and pose a major challenge for the biopharmaceutical industry during production and characterization [23]. PTMs like glycosylation help in maintaining the structural integrity and activity of protein therapeutics, as has been demonstrated by various techniques used for structural analysis like NMR and X-ray crystallography (Table 1) [22]. Glycans can affect the overall FcγR binding and complement fixation for mAb based therapeutics.

Aggregates generated due to variations in PTMs can sometimes cause immune response and generation of ADAs in the patient which are beyond control and can show enhanced immunogenicity in patient [23]. mAbs have natural capacity to form immune complexes (ICs) that may help them in binding to their target and enhance the response. Aggregated mAbs due to modifications in PTMs can have compromised functions and their ability to recognise their specific targets may get affected (Table 1) [23].

In addition to intrinsic factors, aggregation can be induced by a variety of extrinsic factors which can affect both chemical and physical stability [10,24] (Figure 1 and Table 1). Chemical degradation includes oxidation, photo-irradiation, deamidation, disulphide reduction, and glycation [25,26]. Oxidation from exposure to hydrogen peroxide and metal ions can affect surface hydrophobicity and protein-protein interactions, and decrease the conformational stability of proteins [27,28,29]. Glycation affects protein aggregation by affecting surface charge of the protein [30,31,32].

Thermal exposure is well known to affect protein stability, leading to changes in protein conformations which in turn can modify colloidal interactions (Table 1) [32]. This complex effect often results in non-Arrhenius kinetics. As a consequence, it is challenging to estimate the shelf-life of biologics at low temperature from accelerated studies at higher temperatures [32]. For instance, a recent study showed that aggregation rates measured at 40°C do not correlate with that at 5°C, while the correlation was better between 30°C and 5°C [33]. Moreover, for the same monomer loss, different types of aggregates were formed at different temperatures, indicating changes in the aggregation mechanisms.

In addition to high temperature, exposure to ultralow temperatures can also unfold proteins via so called cold-denaturation (Table 1) [34,35,36], which is important in the context of freeze thaw. The tertiary structure of a protein is greatly affected by freeze thawing, depending on the pH of the solution and additives present. These changes, which can be monitored by fluorescence spectroscopy, could be reversible as well as irreversible [37] and could lead to aggregation of protein molecules caused by alterations in their tertiary structure.

Formulation and buffer conditions such as pH, ionic strength, protein concentration, and buffer types strongly affect protein stability in liquid formulations [38,39,40,41,42]. Exposure to extremely low pH [43,44,45] and changes in ionic strength [46,47,48] can significantly affect protein stability, often in a non-monotonic and complex way [10,43,49] (Table 1, Figure 1). For instance, changes in pH experienced during downstream processing (DSP) of protein therapeutics like in protein A chromatography, viral inactivation, or neutralization mostly leads to increase in aggregation [4,6,50]. The level of aggregation is also dependent on the hold time of the protein under various DSP operations [41]. Researchers have characterized and established an aggregation assay to monitor aggregation due to changes in pH, mostly at pH 2 and then neutralizing it to slightly neutral pH of 6.5 [42]. At low pH there is unfolding of the protein but as the pH is raised, refolding occurs, which is due to intermolecular attraction. But if this intermolecular attraction between molecules further increases beyond refolding, it may result in protein aggregation (Table 1) [42]. Protein concentration influences the stability of protein therapeutics by affecting both aggregation and viscosity due to the increase in protein-protein self-interactions. This problem is a growing worry for the formulation scientists working in the domain of development of subcutaneous administration of these bio-therapeutics (50). Protein aggregation is a complex pathway consisting of multiple microscopic reactions of nucleation and growth. The effect of protein concentration on the rate of these individual steps is described by their corresponding reaction orders, which depend on the types of stresses and buffer conditions. Therefore, changes in protein concentration can have different effects on nucleation and growth reactions depending on the specific environmental conditions. This, in turn, has consequence for the size distribution of the formed aggregates. For instance, one study showed that increase in protein concentration led to an increase in the concentration of smaller particles and a decrease in the concentration of the larger particles when the protein sample was subjected to freeze thaw stress under varying protein concentrations and buffer formulations [51]. In contrast, another study showed that concentration of larger particles increased as compared to smaller particles on increasing protein concentration as the therapeutic was subjected to heat stress at 70ºC [52]. Apart from the buffer conditions, effect of surfaces and interfaces is very critical in affecting the pace of protein therapeutics’ degradation. One of the phenomena that is related to surface is shear stress (Table 1) [53]. Shear stress at the interface is reported to be more aggregation causing as compared to bulk shear stress. These interfacial stresses are generally encountered during drug manufacturing, development and administration to patients. At the interfaces there can be many interactions like solid-liquid interaction, liquid-liquid interaction or vapor-liquid interaction. Due to these interactions, there are chances of formation of soluble as well as sub-visible and visible particles [53]. There is also probability of variation in concentration of therapeutic product because of surface adsorption of the drug product. Changes in the shape and structures of molecules due to interfacial effects are also probable. Conformational changes occur due to mechanical or hydrophobic stresses. Major surface interactions occur during drug product manufacturing and storage processes like filtration, container surfaces, mixing and mechanical processes, freeze-thawing during excipient addition. During storage, main aggregation prone exposures include vial filling operations. Air liquid interfacial stress can be introduced due to headspace of the vial, transportation agitation. Interfacial stress is also possible during drug product administration via exposure to surfaces like IV bags, infusion sets, etc. (Table 1) [53].

Shear stress could also make the proteins susceptible to protein aggregation. It is often found in high concentration mAb formulations. One of such recent studies has shown the effect of shear stresses (20,000 s−1–2,50,000 s−1) on high concentration mAb formulations having concentrations greater than 100 mg/ml [54]. In this study, Bee et al. proposed that shear during production processes not only causes aggregation. Adsorption to solid surfaces, particulate contamination, air-bubble entrainment or pump cavitation has also a significant role in promoting aggregation mAb therapeutics. This also has been reported in other protein therapeutics apart from mAbs, like human serum albumin (HSA) as reported by Duerkop et. al. This group reported checked that whether there are aggregates formation due to formation of hydroxyl radicals caused by cavitation bubble collapse, as reported in some previous literatures. It has been suggested in the conclusion of this study that aggregation occurs due to increase in the surface area because of cavitation bubble growth and not because of hydroxyl radical release or shear stress [55].

Another study by Arosio et al. studies the role of hydrodynamic flow on protein aggregation [56]. In this study they have studies the interfacial stress effect by setting up an experimental design where flow and interface induced aggregation of two IgGs have been investigated. Shear rates of 10 s−1–1000 s−1 have been kept but negligible change in % aggregation has been observed. Rather considerable monomer loss has been observed when the material of the syringe barrels has been changed. This led them to the conclusion that there is significant of solid-liquid interface in promoting aggregation during flow operations. They have suggested a synergistic effect of hydrodynamic flow and interfaces in causing protein aggregation [56].

Mechanisms of protein‘s size-based degradation

Two of the most common degradation pathways of proteins therapeutics include aggregation and fragmentation.

Mechanism of protein aggregation

Proteins can form aggregates following a variety of microscopic mechanisms which include multiple events of nucleation and growth. Nucleation is often promoted by conformational changes or chemical modifications. Heterogeneous nucleation triggered by interactions with surfaces is also commonly observed. All these steps can be reversible or irreversible, although non-native protein aggregation often leads to irreversible aggregates, and differs from other types of reversible protein phase transitions such as liquid-liquid phase separation and protein crystallization. (Figure 2):

  1. Nucleation – Changes in conformation of protein molecules can result in partial exposure of hydrophobic patches or residues [57], which can lead to formation of oligomers or nuclei, the first step towards aggregation [17,57] (Figure 3). These oligomers are typically dimeric and non-native [58,59,60]. The understanding and prediction of nucleation rates remains one of the most challenging aspect in protein aggregation. Nucleation is mediated by a variety of interactions such as electrostatic interactions, hydrophobic interactions, hydrogen bonds, and the van der Waals forces. Net and average parameters describing protein-protein interactions often do not correlate with protein aggregation [61,62,63,64], since interactions that are relevant for protein aggregation can be confined to specific aggregation protein regions. These regions can be located in various locations of protein molecules like the complementary determining region 1 (CDR1) regions, Fab and Fc, and CH2 or CH3 domains [17,65] (Figure 3). The nucleation rate typically increases with protein concentration [62,63], unless unimolecular conformational changes are the rate-limiting step in the process.

  2. Formation of higher order aggregates – Formation of oligomers is often followed by formation of larger insoluble and visible particulates [66]. Aggregates can grow in size either by addition of monomers via chain polymerization or by cluster-cluster interactions [59,60,67,68] (Figures 2 and 3). The relative impact of these steps is controlled by conformational and colloidal stability [43,60,69,70].

Figure 2 Mechanism of protein aggregation.
Figure 2

Mechanism of protein aggregation.

Figure 3 Overview of non-native protein aggregation pathways. APRs, aggregation prone regions.
Figure 3

Overview of non-native protein aggregation pathways. APRs, aggregation prone regions.

Mechanism of protein fragmentation

Fragmentation in bio-therapeutics, especially monoclonal antibodies, is another common degradation pathway [71]. Fragmentation refers to the breakage or disruption of the peptide covalent bond between the amino acid residues of a protein molecule. Fragmentation, either spontaneous or enzymatic, leads to changes in the primary sequence and consequently in higher order structure [72]. Peptide covalent bonds are generally considered strong associations unless they are exposed to harsh environments like drastic pH or extreme temperature [71,73]. Another key factor affecting protein fragmentation is the flexibility of the side chains of the molecule. Fragmentation is typically more frequent in those regions of the molecule which are somewhat flexible in nature or are exposed to the solvent [74] and less frequent in regions that are buried inside the molecule, although the primary sequence of that region is prone to cleavage. Fragmentation could be either due to β-elimination or due to peptide bond hydrolysis, primarily depending on the pH of the protein solution [71]. Presence of certain amino acids (Asp, Gly, Ser, Thr, Cys or Asn) makes these regions more susceptible to fragmentation [71]. Each of these amino acids aids in cleavage of peptide bonds in mAbs, thereby accelerating fragmentation by acting through their side chains [75]. The given amino acids generally act in a wide range of pH (pH 0.3–10) and temperature (25°C – 70°C) [76]. Apart from side chain mediation, fragmentation can also occur due to peptide bond hydrolysis [74]. Many computational and experimental studies have shown the significant role of hydrolysis in peptide bond cleavage.

Metals ions also play a crucial role in initiating or accelerating protein fragmentation. Copper has been reported to accelerate fragmentation of IgG1 at the hinge region. This copper mediated cleavage is pH and mAb specific and has been shown to decelerate upon addition of chelating agents like EDTA [77]. Various IgG molecules behave differently in the presence of cupric ions, depending on the affinity between the protein molecules and the metal ions as well as on the isotype of IgG since each IgG isotype varies in its sequence and conformation of the hinge region [77]. Apart from the copper ions, other metal ions like MgII, MnII, ZnII, FeIII and NiII can also impact protein fragmentation, although to a lesser degree [74,77].

Hydrolysis of the peptide bond (which is one of the mechanisms of fragmentation) in the hinge region of the mAb yields Fc-Fab and Fab fragments [71]. Fragmentation can be observed during protein production in cell culture, purification, storage, and even while circulating in blood. Better understanding of the mechanism behind hydrolytic fragmentation is required for its prevention, including its origin, solvent conditions (pH, ionic strength), temperature and kinetic behaviour [78]. There is a significant difference among the different mAbs with respect to their vulnerability towards fragmentation. Conformational stability and local structural dynamics are also known to influence fragmentation [77]. Fragmentation is usually detected by size exclusion-high performance liquid chromatography (SE-HPLC) [75,76]. Other analytical techniques which can be used for characterizing fragmentation include mass spectrometry (MS), sodium dodecyl sulphate-poly acrylamide gel electrophoresis (SDS-PAGE), and DLS [79].

Modelling of protein‘s size-based degradation

An important step towards controlling protein‘s size-based degradation is understanding its mechanism and this is typically achieved via kinetic and thermodynamic modelling [80]. A powerful strategy involves acquiring time resolved data of the degradation process and comparing this information with simulations based on different theoretical models. This comparison leads to the identification of the aggregation mechanism and to the quantification of the corresponding rate constants of nucleation and growth [80].

Numerous models for aggregation have been proposed in the literature [81,82,83] in different contexts. Indeed, protein aggregation is observed in a variety of biological and biotechnological systems. In some cases, aggregation inside biological systems is undesirable and leads to pathological species associated with a variety of neurodegenerative diseases [10], while in other cases protein self-assembly is functional, as in the case of actin [10]. In the biotechnological context, protein aggregation is undesirable, leading to product loss and to formation of potential immunogenic species [10,84]. Although aggregate types vary significantly with respect to their heterogeneity, the Lumry Eyring model is the most commonly used model for predicting kinetics of protein aggregation [85]. According to this model, the monomer first converts into an unfolded or misfolded species and then forms small oligomers. Small oligomers may undergo conformational changes to form nucleus species which are highly reactive and contribute to formation of larger oligomers. These large oligomers can interact with each other to form either fibrils, amorphous aggregates, or protofilaments. All through this aggregation process, numerous intermediate species which have ring or annular structures are formed and this has been confirmed by researchers using atomic force microscopy (AFM) and electron microscopy (EM) [86,87,88].

The Lumry Eyring model, in combination with population balance equations, has been widely applied to describe the aggregation of a variety of biologics under several stresses.

Recent developments in the domain of protein‘s size-based degradation mechanisms and modelling

A host of analytical techniques are available for the aggregate characterization like SE-HPLC, DLS, SLS, fluorescence spectroscopy, circular dichroism, mass spectroscopy, micro flow imaging (MFI), nanoparticle tracking analysis (NTA) [87]. Each of these characterization techniques gives different information about aggregation. SE-HPLC helps us to separate lower order aggregate species of various sizes and with the help of a molecular ladder or an integrated MALLS, sizes and molecular weights of different aggregates can be easily computed [87]. DLS helps us to calculate average hydrodynamic radius of an aggregated sample. Spectroscopic tools like circular dichroism, FTIR and fluorescence spectroscopy give information about the protein's secondary and tertiary structure. Flow imaging and tracking tools like NTA and MFI compute the particle sizes in a flow medium [87].

On the basis of the data obtained from these characterization tools, sizes of the various particles formed during aggregation are calculated and this size data can be used to deduce various plausible aggregation pathways taken by protein molecules. Researchers have described one such mechanism where it has been mentioned that reversible protein aggregation is a complex phenomenon where the smaller aggregates like dimers and oligomers form rigid and stable complexes which are in an equilibrium state with the monomer and their stability depends on the dissociation constant (Kd) [89]. Here they have compared two molecules – insulin and mAb. In the case of insulin, the Kd value is around 10 μM without zinc. Insulin is generally formulated and administered at a higher dose than this Kd value. This causes insulin to exist as dimer at physiological pH. But at lower pH (pH = 2.0), it exists as a monomer. At protein concentrations around this Kd, both monomers and dimers coexist in equal amounts [89]. But at concentrations below Kd, dimers are more stable than monomers. In the case of mAb therapeutics, Kd values are much higher than insulin. Stable dimers and other oligomers are generally observed at high mAb concentrations (100 mg/ml). Larger aggregates are stable and observable by imaging techniques as compared to dimers and oligomers, which are quite unstable and are in dynamic equilibrium with each other [89]. But one thing to mention here is that aggregates with number of monomeric units up to 100 are considered as molecular aggregates and macroscopic particles are considered only when the particles become very large in dimensions and macroscopic phase separation occurs [89]. Besides reversible aggregation, irreversible aggregation or formation of non-native aggregates is also important where the aggregates are not dissociable upon dilution or changing protein's environmental conditions [89]. These non-native aggregates formed as a result of complete loss of protein's secondary and tertiary structures and formation of strong non-covalent interactions amongst multiple molecular chains. This non-native aggregation includes formation of nucleus misfolded species [89]. This nucleus specie is generally very stable and further growth of non-native aggregates occurs by adding of other monomeric units to this nucleus. The process or step of nucleation is generally slow and rate limiting in non-native aggregation, which is kinetically controlled and depends on the concentration of monomer and various aggregated species present in a solution [89].

Protein stability can be estimated and quantified in terms of a parameter known as Fuchs stability ratio (W) [90,91]. It is calculated on the basis of interaction potential of two particles. This value is highly dependent on changes in surface potential or interactions. It is accurately determined by fitting experimental data to population balance modelling equations. Researchers have described the relation between aggregate reactivity and aggregate size. For the ideal case, the kernel size was determined using this Fuchs stability ratio (W) assuming the aggregating spheres have a uniform reactivity [91]. But in actual situation, reactivity of a protein is highly dependent on confirmation of protein in three-dimensional space and amount of exposure of hydrophobic patches to the external buffer environment. It can be done by taking into consideration all kinds of monomeric and aggregated species like monomer, nucleus species, oligomeric species and aggregated species. The relationship between protein reactivity and coverage of patches which are hydrophobic in nature is inverse because as the aggregation increases, the number of available hydrophobic patches for contributing to aggregation decreases, ultimately affecting the reactivity of aggregates [80,91]. Researchers have used this approach to predict heat induced aggregation in mAbs. Their study confirmed that although two mAbs when treated by same stress condition, behave differently and follow different aggregation pathways and can have high W value. The factors which contribute to this differential behaviour could be due to electrostatic repulsive forces, presence of scattered hydrophobic patches on protein's surface and hydration forces, which lead to stability and differential aggregation propensity of mAbs in solution [52,80,91].

These approaches can be used to model and understand the effect of excipients on individual steps of aggregation [80]. Researchers have used population balance for predicting aggregation kinetics for monoclonal antibody products in acidic conditions [90]. Changes in the radius of gyration and hydrodynamic radius of the mAb samples were predicted as a function of time and used to model aggregation kinetics using Smoluchowski's population balance modelling [90]. In situ light scattering was used to compute these parameters and to derive fractal dimensions, and colloidal stability of protein particles was described in terms of Fuchs stability ratio, a parameter that offers insight into the Arrhenius dependency of the different rate constants of aggregation and also into the mechanism of aggregation. This ratio is hence indicative of whether aggregation is due to monomer-monomer addition or cluster-cluster interaction. The researchers concluded that reactive monomer species are more prone to aggregation as compared to reactive aggregate species because of a decrease in number of hydrophobic patches that are available for conglomeration [90]. Other researchers have reviewed the various theories that have been applied for understanding protein aggregation [91]. They have used coarse grain modelling for computing the colloidal and conformational stability of therapeutic proteins and explained the use of fractal dimension for characterizing irregular aggregate morphologies.

Studying molecular mechanisms and kinetics of protein‘s size-based degradation have been an interesting area to explore and many researchers in the past have demonstrated the same using various computational tools. One such recent investigation examines aggregation kinetics of different classes of mAbs (IgG1 and IgG2) when subjected to thermal stress [80]. SE-HPLC and light scattering techniques were deployed in combination to compute the time-based growth of various aggregate species (monomer, dimer and trimer) along with their molecular weight and hydrodynamic radius. Using experimental data, population balance modelling was performed to evaluate the role of each basic step on aggregation. The main outcome of the study was that subclass type has a significant role to play in deciding the aggregation route even if both the molecules are kept under similar operating conditions. In mAbs, while monomer depletion is rate limited by conformational changes occurring inside the monomer molecule, collision between two monomers is the rate limiting step for further aggregation. Fuchs stability ratio (W) was used to calculate the kinetic rate constants for measuring interaction potentials amongst various protein molecules. Protein aggregation is generally considered as a multi-step process that starts with formation of a non-native unfolding intermediate which is highly susceptible to aggregation. This unfolded protein then accelerates the formation of nucleus species as well as other higher order oligomers. This study reported that protein aggregation depends on both colloidal and conformational stability of protein. Colloidal stability on one hand depends on interactions between the different protein molecules which are prone to aggregation. On the other hand, conformational stability is dependent upon both the rate of protein unfolding (kinetics) as well as the free energy change that occurs during protein unfolding (thermodynamics) [80]. In another work, researchers have focussed on aggregation pattern in concentrated conditions under thermal stress [91]. A kinetic model using population balance equations has been described and significance of colloidal and conformational stability towards overall aggregation rate has been examined [91]. The model accounts for the impact of rate constant, protein-protein interactions, solution viscosity and aggregation compactness under both dilute and concentrated conditions.

Protein‘s size-based degradation mechanisms (both aggregation and fragmentation) have been related to various factors and correlations of these mechanisms with factors have been the research interests of many scientists. Researchers have described the aggregation of IgG1 based mAb therapeutics under various buffer conditions, comparing different factors like temperature, pH, salt concentrations and buffers and finding their correlation using protein aggregation models like Lumry Eyring nucleated polymerization models, extended Lumry Eyring model and Finke Watzky model [6]. In their study, pH was found to be the most important factor that plays a crucial role in aggregation of mAb based therapeutics. Temperature, salt concentration and buffers come next with respect to their effect on protein aggregation kinetics. LENP model provided better fit with the experimental data as compared to the ELE model showing that the aggregation is primarily nucleation dominated and controlling the monomer half-life [6].

Proteins in their native conformations can form multitude of aggregates varied in size and shape, which is highly dependent on external buffer conditions of the sample. Researchers have reviewed the various mechanisms and factors leading to formation of larger aggregates [92]. They have categorized the various supramolecular structures for higher order aggregates which are as follows: amorphous aggregate, oligomers, fibrils, protofibrils, superstructure, amyloid like sephrulites, and protein particulates. In terms of mechanisms for the formation of these structures, transitions between soluble precursors and insoluble aggregates occur due to formation of various inter and intra molecular interactions [92]. These interactions are largely affected by the protein structure, concentration, and the physicochemical conditions of the buffer. The models for these higher order aggregates are largely dependent on the formation of nucleus and growth of these aggregates. Various steps in formation of these structures include micelle formation, protein conformational changes, and joining of various filaments. A number of characterization tools have been listed as used for characterization of these higher order aggregate structures like – FTIR, thioflavin/ANS/tryptophan fluorescence, far-UV CD, DLS, AFM, TEM, and fluorescence microscopy [92,93]. As discussed in the previous section, of the factors affecting protein aggregation, protein concentration is an important factor that affects protein stability. Researchers have related the interactions between protein molecules and aggregation propensity to their dependency on protein concentration [94]. Aggregation rates were measured across a pH range of 5–6.5 in the presence of various excipients like sucrose and NaCl using SE-HPLC, SLS and DLS, and it was observed that at low protein concentrations, small changes in protein conformation and protein-protein interactions greatly impact the protein aggregation kinetics. Various hypotheses have been tested for correlating protein interaction to changes in protein aggregation propensity like thermodynamic changes during protein aggregation, statistical mechanical fluctuation theory, and surface contact probabilities. Amongst these hypotheses, only the hypothesis of surface contact probability has been found to be consistent with experimental results and a semi-quantitative relationship has been established between protein-protein interactions and protein aggregation rates [94].

Recent usage of computational approach for studying protein aggregation has been gaining interest. Researchers have described and reviewed various methods and applications for computational approaches for protein aggregation [95]. They compared atomistic and coarse -grained models for predicting and deeper insights into protein aggregation mechanism. In the atomistic approach, monomeric and small oligomeric molecules were examined for their interactions with dyes, small molecules and peptide inhibitors. Atomistic models provide information related to early stages of aggregation. The unstable states like partially folded structures and aggregation prone structures are very difficult to study experimentally [95]. Simulation based analysis has been proven useful in analysing these unstable states. In coarse-grained model approach, peptide aggregation has been extensively studied. Coarse grained models have been used to study higher order aggregation. Various coarse-grained models have been used like lower resolution models, phenomenological models, Martini force field model, prime model, discontinuous molecular dynamics, and the optimized potential for efficient protein structure prediction (OPEP) protein model [95].

Researchers have described the use of kinetic approach for revealing various mechanisms used by molecular chaperones for prohibiting the aggregation [96]. Aggregates play a crucial role in a myriad of human diseases. The formation of aggregating proteins is strongly affected by molecular chaperones. The authors used kinetics of protein aggregation for shedding light on interactions amongst chaperones and protein molecules. This method helped in identifying protein components which are actively targeted and various steps that are inhibited by these chaperones [96].

Kinetics of protein fragmentation has also been studied in detail recently, showing the significance of various factors affecting mAb fragmentation using kinetic modelling [71]. The fragmentation studied by them is nonenzymatic in nature under high temperature conditions. The fragmentation rate and fragment species formed have been characterized by various analytical tools like SE-HPLC, DLS, SDS-PAGE, LC-MS. Based on the results obtained from kinetic analysis, the authors concluded that at high temperature, mAb gets broken or cleaved at hinge region, resulting in formation of Fab-Fc and Fab/Fc fragments [71].

Control of protein‘s size-based degradation

Protein stabilization is critical for commercial success of a bio-therapeutic. Excipients are known to stabilize protein therapeutic formulations via different mechanisms such as reinforcement of stabilizing forces or weakening of denatured protein state. Some excipients also act by directly binding to the protein molecule [97,98]. Protein excipients could be categorized into various classes such as salts, antioxidants, surfactants, polyols, buffers, amino acids like arginine and lysine, and sugars like sorbitol, sucrose, mannitol and polysaccharides and small peptides [98,99,100]. These excipients are generally used in concentrations ranging from 0.1M to 1M. They act by decreasing surface adsorption and facilitating physiological osmolality. The chief interaction of these excipients is with protein, container surface, and water. Excipients stabilise proteins either by direct or indirect interactions. Direct interactions are generally encountered in dried or lyophilized state [101]. Approaches to reduce aggregation and increase protein stability could include strategies like lyophilisation or spray drying which transform therapeutics into drug conjugates such as antibody drug conjugates (ADCs), and use of inline filters prior to administration.

Excipients have been shown to stabilise protein solutions by weakening of denatured state, reinforcement of protein stabilizing forces, and sometimes also by direct binding of protein to the excipient at various stages of formulation development like drying, lyophilisation, purification [102]. Formulations that are in liquid form as compared to dried or lyophilized are more susceptible to instability due to the higher mobility of the protein molecule, which increases the chances of molecular collisions amongst molecules [103,104]. Lyophilized formulations are generally more stable and offer greater shelf life. In liquid formulations, interactions at liquid air interface or solid-liquid interface are responsible for instability due to adsorption and protein unfolding. But sometimes, lyophilisation cycles can cause instability in protein formulations. Therefore, proper optimization of the lyophilisation process can aid in minimizing lyophilisation induced aggregation [104,105,106].

Antimicrobial agents like benzyl alcohol that are present in the liquid formulations have been known to promote aggregation by enhancing the formation of partially unfolded states. Some solutes may also stabilize the protein by minimizing the surface area which is exposed to the excipient or co-solute [102]. This reduction in exposed surface area can be achieved by either self-association of monomer or adsorption of the monomer to a surface.

A key strategy to control aggregation is via modification of the hot spot regions of proteins that are more susceptible to instability. Hot spot modifications can help in reducing aggregation by modifying the protein surface charge and improving protein solubility. However, major changes in the protein's primary sequence or structure can also significantly impact protein activity in an undesirable fashion [97,98]. Attachment of hydrophilic moieties such as polyethylene glycol (PEG), hydroxyethyl starch (HES), or glycan moieties to the protein sequence may improve protein solubility in aqueous solutions [97,98,99,101].

Since there are many risks associated with changes in the primary structure of the protein w.r.t. immunogenicity and bioactivity, creating a stable formulation is preferred [102]. This entails choice of environmental conditions (pH, conductivity) and use of stabilizers. Monovalent cations are known to be more efficient in stabilising protein formulations as compared to multivalent cations [103]. Stabilizers commonly act by increasing conformational and colloidal stability, shielding the hydrophobic patches or enhancing protein disaggregation. High hydrostatic pressure with or without assistance of high temperature can also be used to dissociate aggregates. Researchers have successfully demonstrated this for green fluorescent protein inclusion bodies and growth hormone aggregates [107]. Inline filters can be also used to filter out large aggregate particles that may be formed due to presence of saline or dextrose as diluents in these formulations [108]. Commonly used surfactants that have been demonstrated to improve stability of protein formulations include polysorbates (PS 80 or PS 20), alkyl saccharides, and amino acid derivatives [9,13,27,109,110]. These stabilizers generally act by minimizing self-interaction between protein molecules and reducing adsorption to surface interfaces.

Recent developments in control of protein‘s size-based degradation

Use of the various types of excipients has been in demand recently for preventing protein therapeutics from aggregating. Here we focus on the important categories of salts, surfactants and sugars. Researchers have defined excipient as an ingredient other than the main protein component in a protein formulation which improves the quality and performance of a drug product [111,112]. As discussed earlier, several classes of excipients are used in protein formulations. Excipients used in biologics include buffering agents, antioxidants, surfactants, salts, sugars, carbohydrates, and amino acids. These excipients help in maintaining protein solubility, osmolarity and osmolality in the solution, thereby enhancing the therapeutic's shelf life. Excipients help proteins in maintaining their structure and conformation in solution. Buffers have a critical role in maintaining pH of the protein solution and providing stability to it. Commonly used buffers are citrate, acetate, phosphate, glycine and histidine. One study has showed that citrate at near neutral pH is favourable for protein formulation due to its lesser crystallization tendency and more collapse temperature. Another study has shown that glycine when added as excipient in phosphate buffer reduce its crystallization tendency, hence enhancing overall formulation stability [113]. Sugars and carbohydrates have shown their ability to provide stability to protein formulations. These are especially used during freeze drying and freeze thawing operations. They operate by various mechanisms like reduced mobility, water replacement mechanism and preferential exclusion. Commonly used sugars are trehalose and sucrose. Amongst these two sugars, trehalose is much preferred to sucrose because of higher glass transition temperature of the former [113,114].

Salts have been used to stabilize protein therapeutic solutions at high concentration by reducing the viscosity of the protein formulations [115]. Use of these salts directly like NaCl, NH4Cl, NaAC, Na2SO4 or salt derivatives of some common amino acids like ArgHCl, HisHCl, LysHCl could help in stabilising high concentration protein formulations. High concentration protein solutions have high viscosity due to increase charge-charge interactions. Salts and salts derivatives of amino acids stabilise the high concentration therapeutic formulations by lowering these charge-charge interactions amongst protein molecules in a solution [115].

Commonly used surfactants in protein therapeutics formulations like PS 80, PS 20, Triton X-100, Pluronic F-68, and caprylic acid have been shown to be effective in providing stability to protein formulations against various types of mechanical, thermal and photo degradations. Surfactants play a critical role in controlling stability of protein therapeutics by affecting their interfacial properties like surface tension and competing for surface adsorption. Examples include polyethylene glycol, poloxamer-188 and PS 20 [112,114]. These surfactants adsorb to the interface, thereby effectively reducing the concentration of soluble aggregates (<100 nm). Insoluble aggregates are controlled due to the formation of cohesive network between adsorbed mAb and surfactant [112]. In this study, various interfacial parameters and properties were used to elucidate mechanism of stability provided by surfactants to protein therapeutics. These mechanisms include surface tension, surface tension gradients for various surfactants, interfacial rheology, and surface tension [112]. In a recent study, several ionic and non-ionic surfactants were evaluated for their ability to provide stability to mAb therapeutics against degradations [114]. It was observed that caprylic acid, an anionic surfactant, has a slightly negative effect on the stability of protein therapeutics against mechanical and light related degradations as compared to other non-ionic surfactants like PS 80 or PS 20. A concentration of at least 0.005% (w/v) is needed to offer stability to mAb based therapeutics [114].

A recent publication explored the stabilizing effects (against both aggregation and conformational perturbation) of different commercially available surfactants used in therapeutic protein formulations [116], including Tween 20, Tween 80, PEG dodecyl ether (Brij35), Chaps, Triton X-100, Pluronic F68 and F127 and SDS. These surfactants were selected based on their diverse physical and chemical properties and stability of two different mAbs was examined using various analytical techniques. The researchers found that non-ionic surfactants are better performers in terms of safety and efficacy as compared to ionic surfactants in the concentration range of 0.02 – 2 mg/ml. The concentration of the surfactant in protein formulation should be below their critical micelle concentration (CMC). However, lower surfactant concentration (<0.002 mg/ml) is generally found to be ineffective in providing stability to protein formulation. Above 2 mg/ml, structural perturbation is likely a phenomenon that is highly dependent on the surfactant type. This proves that enhancing stability of protein formulation via surfactants is highly dependent on their absolute mass concentrations and not on their molecular weights. Hydrophobic surfactants such as tween 80 and triton x 100 offer stable protein formulations at or above concentrations of 2 mg/ml. Ionic surfactants like Chaps are also effective in providing stability to protein formulations when used below 2 mg/ml of concentration. But these have to be used with caution as they can decrease the thermal and long-term protein stability. SDS falls under the category of ionic surfactants should never be used for stabilising protein formulations due to their significant impact on protein structure. Poloxamers like F68 and F127 are also effective in providing stability to protein formulations due to their aggregation preventing and protein stability enhancing role. Stabilising effects of surfactants is also highly dependent on protein concentration and its type [116]. At protein concentration of 50 mg/ml, one of the proteins displayed higher stability in presence tween 20 and tween 80 as compared to the other protein, indicating that surfactant stabilization of proteins is highly protein and concentration dependent. The stabilising property of a surfactant is generally credited to their ability to reduce the structure perturbation caused by surface adsorption to various interfaces and also to their capacity for inhibiting protein-protein interactions by hydrophobic surface binding [116].

To conclude, usage of surfactants in IgG formulations is highly dependent on type and concentration of surfactant and protein and its mechanism of aggregation. There should be a fine balance between the stability enhancing and perturbing forces for achieving the desired outcome of a stable protein formulation [116].

In addition to surfactants, sugars are another common class of excipients. The mechanism of stabilization of different protein therapeutics by various sugars in solid and drying state has been reviewed recently [117]. Sugars interact with protein molecules during drying operations and this helps in preserving the protein in its original native form and lowers the molecule's overall movement during storage. It has been observed that smaller the sugar molecules, better is the interaction between protein molecules and sugar molecules due to less steric interference. Overall stability provided by a sugar or a saccharide is dependent on many factors like protein under study, conditions under which formulation is prepared as well as storage conditions [117].

Excipient crystallization also has a strong impact on the unfolding abilities of protein molecules, ultimately affecting protein aggregation. Researchers have demonstrated the consequences of crystallization of mannitol and trehalose on the stability of IgG1 at different pH conditions by repetitive freeze-thaw cycles [118]. At low pH, the recurring stress of freeze thaw induce aggregation could be lessened by the use of a surfactant. Increment in the number of freeze-thaw cycles, along with decrease in the formulation pH resulted in formation of soluble and insoluble particles [118].

Researchers have shown the use of approach of modelling as well as experimental characterization for exploring the effect of sucrose and polyols (sorbitol, mannitol, glycerol and threitol) on the mAb aggregates formed by heating [119]. The authors concluded that the protein unfolding rate governs the overall monomer depletion kinetics [119].

Apart from conventional excipients, use of non-conventional excipients is also gaining interest in providing stability to protein formulations. One of the recent publications has demonstrated use of gold nanoparticles (AuNPs) for treatment of Alzheimer's disease [120]. These are used as anti-aggregating particles which makes them one of the promising candidates as neuro-medicines for the treatment of Alzheimer's disease [120]. Use of peptides and peptide based dendrons have also been shown to stabilise mAb and therapeutics formulations [100,121]. Researchers have shown the use of small peptide-based ligands complementary to hydrophobic regions of insulin in preventing aggregation and providing stability [90]. A similar work has shown the use of peptide based dendrons for providing stability to mAb based therapeutics. They have shown the use of lysine based dendrons of third generation in effectively preventing aggregation as well as fragmentation [121].

Conclusions

Protein‘s size-based degradation continues to be a potent problem for the biopharmaceutical industry. It is imperative to understand how protein stability, mechanisms and modelling of protein‘s size-based degradation impact quality, safety and efficacy of a bio-therapeutic. Lumry Eyring Model and its variants (LENP and ELE) are the most accepted models for describing aggregation kinetics in bio-therapeutic proteins. Knowledge of the mechanism as well as the various species that are formed in the process can be useful in creating appropriate strategy for monitoring product stability during production and storage. Strategies for mitigating protein‘s size-based degradation include making changes in formulations by using excipients, which can provide stability to protein formulations. It is hoped that the information presented here will be of interest to researchers working on formulation development and stability of bio-therapeutic products.

Abbreviations Used

ADAs -

Anti-Drug Antibodies

ADCs -

Antibody Drug Conjugates

AFM -

Atomic Force Microscopy

ANS -

8-Anilinonaphthalene-1-Sulfonic Acid

AUC -

Analytical Ultra Centrifugation

CD -

Circular Dichroism

CDR -

Complementarity Determining Region

CMC -

Critical Micelle Concentration

DLS -

Dynamic Light Scattering

DSP -

Downstream Processing

EDTA -

Ethylenediaminetetraacetic Acid

ELE -

Extended Lumry Eyring

EM -

Electron Microscopy

Fab -

Fragment Antigen Binding

Fc -

Fragment Crystallizable

FcγR -

Fragment Crystallizable Gamma Receptor

FTIR -

Fourier Transform Infrared Spectroscopy

HES -

Hydroxyethyl Starch

IgG -

Immunoglobulin G

IR -

Infra-Red

LCMS -

Liquid Chromatography Mass Spectrometry

LENP -

Lumry Eyring Nucleated Polymerization

mAbs -

Monoclonal Antibodies

MALLS -

Multi Angle Laser Light Scattering

MFI -

Micro Flow Imaging

MS -

Mass Spectrometry

NMR -

Nuclear Magnetic Resonance

NTA -

Nao Particle Tracking Analysis

OPEP -

Optimized Potential For Efficient Protein Structure Prediction

PEG -

Poly Ethylene Glycol

PS -

Polysorbate

PTM -

Post Translation Modification

SDS-PAGE -

Sodium Dodecyl Sulphate - Poly Acrylamide Gel Electrophoresis

SE-HPLC -

Size Exclusion-High Performance Liquid Chromatography

SLS -

Static Light Scattering

UV -

Ultra Violet

Acknowledgements

The authors would like to express their gratitude to the unknown referees for carefully reading the paper and giving valuable suggestions.

  1. Funding information: Authors state no funding involved.

  2. Conflict of interest: Authors state no conflict of interest.

  3. Data availability statement: Data sharing is not applicable to this article as no datasets were generated or analysed during the current study.

References

[1] Ecker DM, Jones SD, Levine HL. The therapeutic monoclonal antibody market. MAbs. 2015;7(1):9–14.10.4161/19420862.2015.989042Search in Google Scholar PubMed PubMed Central

[2] Otvos L Jr, Wade JD. Current challenges in peptide-based drug discovery. Front Chem. 2014 Aug;2:62–72.10.3389/fchem.2014.00062Search in Google Scholar PubMed PubMed Central

[3] Ratanji KD, Derrick JP, Dearman RJ, Kimber I. Immunogenicity of therapeutic proteins: influence of aggregation. J Immunotoxicol. 2014 Apr–Jun;11(2):99–109.10.3109/1547691X.2013.821564Search in Google Scholar PubMed PubMed Central

[4] Rathore AS, Joshi V, Yadav N. Aggregation of monoclonal antibody products: formation and removal. Biopharm Int. 2013;26(3):40–5.Search in Google Scholar

[5] Arora I, Bansal R, Joshi V, Rathore AS. Aggregation Kinetics for Monoclonal Antibody Products. Int J Chem Eng Appl. 2014;5(5):433–8.10.7763/IJCEA.2014.V5.424Search in Google Scholar

[6] Singla A, Bansal R, Joshi V, Rathore AS. Aggregation kinetics for IgG1-based monoclonal antibody therapeutics. AAPS J. 2016 May;18(3):689–702.10.1208/s12248-016-9887-0Search in Google Scholar PubMed PubMed Central

[7] Moussa EM, Panchal JP, Moorthy BS, Blum JS, Joubert MK, Narhi LO, et al. Immunogenicity of therapeutic protein aggregates. J Pharm Sci. 2016 Feb;105(2):417–30.10.1016/j.xphs.2015.11.002Search in Google Scholar PubMed

[8] Kumar V, Sami N, Kashav T, Islam A, Ahmad F, Hassan MI. Protein aggregation and neurodegenerative diseases: from theory to therapy. Eur J Med Chem. 2016 Nov;124:1105–20.10.1016/j.ejmech.2016.07.054Search in Google Scholar PubMed

[9] Roberts CJ. Therapeutic protein aggregation: mechanisms, design, and control. Trends Biotechnol. 2014 Jul;32(7):372–80.10.1016/j.tibtech.2014.05.005Search in Google Scholar PubMed PubMed Central

[10] Chaturvedi SK, Siddiqi MK, Alam P, Khan RH. Protein misfolding and aggregation: mechanism, factors and detection. Process Biochem. 2016;51(9):1183–92.10.1016/j.procbio.2016.05.015Search in Google Scholar

[11] Carpenter JF, Randolph TW, Jiskoot W, Crommelin DJ, Middaugh CR, Winter G. Potential inaccurate quantitation and sizing of protein aggregates by size exclusion chromatography: essential need to use orthogonal methods to assure the quality of therapeutic protein products. J Pharm Sci. 2010 May;99(5):2200–8.10.1002/jps.21989Search in Google Scholar PubMed

[12] Bansal R, Gupta S, Rathore AS. Analytical platform for monitoring aggregation of monoclonal antibody therapeutics. Pharm Res. 2019 Aug;36(11):152.10.1007/s11095-019-2690-8Search in Google Scholar PubMed

[13] Alam P, Siddiqi K, Chturvedi SK, Khan RH. Protein aggregation: from background to inhibition strategies. Int J Biol Macromol. 2017 Oct;103:208–19.10.1016/j.ijbiomac.2017.05.048Search in Google Scholar PubMed

[14] Kueltzo LA, Wang W, Randolph TW, Carpenter JF. Effects of solution conditions, processing parameters, and container materials on aggregation of a monoclonal antibody during freeze-thawing. J Pharm Sci. 2008 May;97(5):1801–12.10.1002/jps.21110Search in Google Scholar PubMed

[15] Buck PM, Kumar S, Singh SK. On the role of aggregation prone regions in protein evolution, stability, and enzymatic catalysis: insights from diverse analyses. PLOS Comput Biol. 2013;9(10):e1003291–310.10.1371/journal.pcbi.1003291Search in Google Scholar PubMed PubMed Central

[16] Buck PM, Kumar S, Singh SK. Consequences of glycan truncation on Fc structural integrity. MAbs. 2013 Nov–Dec;5(6):904–16.10.4161/mabs.26453Search in Google Scholar PubMed PubMed Central

[17] Meric G, Robinson AS, Roberts CJ. Driving forces for nonnative protein aggregation and approaches to predict aggregation-prone regions. Annu Rev Chem Biomol Eng. 2017 Jun;8(1):139–59.10.1146/annurev-chembioeng-060816-101404Search in Google Scholar PubMed

[18] Courtois F, Schneider CP, Agrawal NJ, Trout BL. Rational design of biobetters with enhanced stability. J Pharm Sci. 2015 Aug;104(8):2433–40.10.1002/jps.24520Search in Google Scholar PubMed

[19] Pisupati K, Benet A, Tian Y, Okbazghi S, Kang J, Ford M, et al. Biosimilarity under stress: a forced degradation study of Remicade® and RemsimaTM. MAbs. 2017. Oct;9(7):1197–209.10.1080/19420862.2017.1347741Search in Google Scholar PubMed PubMed Central

[20] Skamris T, Tian X, Thorolfsson M, Karkov HS, Rasmussen HB, Langkilde AE, et al. Monoclonal antibodies follow distinct aggregation pathways during production-relevant acidic incubation and neutralization. Pharm Res. 2016 Mar;33(3):716–28.10.1007/s11095-015-1821-0Search in Google Scholar PubMed

[21] Nakamura H, Oda-Ueda N, Ueda T, Ohkuri T. Introduction of a glycosylation site in the constant region decreases the aggregation of adalimumab Fab. Biochem Biophys Res Commun. 2018 Sep;503(2):752–6.10.1016/j.bbrc.2018.06.071Search in Google Scholar PubMed

[22] Hmiel LK, Brorson KA, Boyne MT 2nd. Post-translational structural modifications of immunoglobulin G and their effect on biological activity. Anal Bioanal Chem. 2015 Jan;407(1):79–94.10.1007/s00216-014-8108-xSearch in Google Scholar PubMed

[23] Lu Y, Khawli LA, Purushothama S, Theil F-P, Partridge MA. Recent advances in assessing immunogenicity of therapeutic proteins: impact on biotherapeutic development. J Immunol Res. 2016;2016:8141269.10.1155/2016/8141269Search in Google Scholar PubMed PubMed Central

[24] Siddiqi MK, Alam P, Chaturvedi SK, Shahein YE, Khan RH. Mechanisms of protein aggregation and inhibition. Front Biosci (Elite Ed). 2017 Jan;9(1):1–20.10.2741/e781Search in Google Scholar PubMed

[25] Trnková L, Dršata J, Boušová I. Oxidation as an important factor of protein damage: implications for Maillard reaction. J Biosci. 2015 Jun;40(2):419–39.10.1007/s12038-015-9523-7Search in Google Scholar PubMed

[26] Tsuchida D, Yamazaki K, Akashi S. Comprehensive characterization of relationship between higher-order structure and FcRn binding affinity of stress-exposed monoclonal antibodies. Pharm Res. 2016 Apr;33(4):994–1002.10.1007/s11095-015-1845-5Search in Google Scholar PubMed

[27] Fazeli A, Haji-Abdolvahab M, Shojaosadati SA, Schellekens H, Khalifeh K, Moosavi-Movahedi AA, et al. Effect of arginine on pre-nucleus stage of interferon beta-1b aggregation. AAPS PharmSciTech. 2014 Dec;15(6):1619–29.10.1208/s12249-014-0192-xSearch in Google Scholar PubMed PubMed Central

[28] Hada S, Kim NA, Lim DG, Lim JY, Kim KH, Adhikary P, et al. Evaluation of antioxidants in protein formulation against oxidative stress using various biophysical methods. Int J Biol Macromol. 2016 Jan;82:192–200.10.1016/j.ijbiomac.2015.10.048Search in Google Scholar PubMed

[29] Arthur KK, Dinh N, Gabrielson JP. Technical decision making with higher order structure data: utilization of differential scanning calorimetry to elucidate critical protein structural changes resulting from oxidation. J Pharm Sci. 2015 Apr;104(4):1548–54.10.1002/jps.24313Search in Google Scholar PubMed

[30] Adrover M, Mariño L, Sanchis P, Pauwels K, Kraan Y, Lebrun P, et al. Mechanistic insights in glycation-induced protein aggregation. Biomacromolecules. 2014 Sep;15(9):3449–62.10.1021/bm501077jSearch in Google Scholar PubMed

[31] Oliveira LM, Gomes RA, Yang D, Dennison SR, Família C, Lages A, et al. Insights into the molecular mechanism of protein native-like aggregation upon glycation. Biochim Biophys Acta. 2013 Jun;1834(6):1010–22.10.1016/j.bbapap.2012.12.001Search in Google Scholar PubMed

[32] Wang W, Roberts CJ. Non-Arrhenius protein aggregation. AAPS J. 2013 Jul;15(3):840–51.10.1208/s12248-013-9485-3Search in Google Scholar PubMed PubMed Central

[33] Wälchli R, Vermeire PJ, Massant J, Arosio P. Accelerated aggregation studies of monoclonal antibodies: considerations for storage stability. J Pharm Sci. 2020 Jan;109(1):595–602.10.1016/j.xphs.2019.10.048Search in Google Scholar PubMed

[34] Luan B, Shan B, Baiz C, Tokmakoff A, Raleigh DP. Cooperative cold denaturation: the case of the C-terminal domain of ribosomal protein L9. Biochemistry. 2013 Apr;52(14):2402–9.10.1021/bi3016789Search in Google Scholar PubMed

[35] Meliga SC, Farrugia W, Ramsland PA, Falconer RJ. Cold-induced precipitation of a monoclonal IgM: a negative activation enthalpy reaction. J Phys Chem B. 2013 Jan;117(2):490–4.10.1021/jp309109kSearch in Google Scholar PubMed

[36] Rosa M, Lopes C, Melo EP, Singh SK, Geraldes V, Rodrigues MA. Measuring and modeling hemoglobin aggregation below the freezing temperature. J Phys Chem B. 2013 Aug;117(30):8939–46.10.1021/jp4035369Search in Google Scholar PubMed

[37] Liu L, Braun LJ, Wang W, Randolph TW, Carpenter JF. Freezing-induced perturbation of tertiary structure of a monoclonal antibody. J Pharm Sci. 2014 Jul;103(7):1979–86.10.1002/jps.24013Search in Google Scholar PubMed PubMed Central

[38] Chakroun N, Hilton D, Ahmad SS, Platt GW, Dalby PA. Mapping the aggregation kinetics of a therapeutic antibody fragment. Mol Pharm. 2016 Feb;13(2):307–19.10.1021/acs.molpharmaceut.5b00387Search in Google Scholar PubMed

[39] Kim NA, An IB, Lim DG, Lim JY, Lee SY, Shim WS, et al. Effects of pH and buffer concentration on the thermal stability of etanercept using DSC and DLS. Biol Pharm Bull. 2014;37(5):808–16.10.1248/bpb.b13-00926Search in Google Scholar PubMed

[40] Kim NA, Lim DG, Lim JY, Kim KH, Jeong SH. Comprehensive evaluation of etanercept stability in various concentrations with biophysical assessment. Int J Pharm. 2014 Jan;460(1–2):108–18.10.1016/j.ijpharm.2013.11.019Search in Google Scholar PubMed

[41] Joshi V, Shivach T, Kumar V, Yadav N, Rathore A. Avoiding antibody aggregation during processing: establishing hold times. Biotechnol J. 2014 Sep;9(9):1195–205.10.1002/biot.201400052Search in Google Scholar PubMed

[42] Oyetayo OO, Kiefer H. Experimental model system to study pH shift-induced aggregation of monoclonal antibodies under controlled conditions. Pharm Res. 2016 Jun;33(6):1359–69.10.1007/s11095-016-1878-4Search in Google Scholar PubMed

[43] Neergaard MS, Nielsen AD, Parshad H, Van De Weert M. Stability of monoclonal antibodies at high-concentration: head-to-head comparison of the IgG1 and IgG4 subclass. J Pharm Sci. 2014 Jan;103(1):115–27.10.1002/jps.23788Search in Google Scholar PubMed

[44] Zhang L, Yu L, Zhang-Van Enk J, Huang G, Zhang J. Phase Behavior of an Fc-Fusion Protein Reveals Generic Patterns of Ion-Specific Perturbation on Protein-Protein Interactions. J Pharm Sci. 2017 Nov;106(11):3287–92.10.1016/j.xphs.2017.07.011Search in Google Scholar PubMed

[45] Calero-Rubio C, Ghosh R, Saluja A, Roberts CJ. Predicting protein-protein interactions of concentrated antibody solutions using dilute solution data and coarse-grained molecular models. J Pharm Sci. 2018 May;107(5):1269–81.10.1016/j.xphs.2017.12.015Search in Google Scholar PubMed PubMed Central

[46] Wang S, Zhang X, Wu G, Tian Z, Qian F. Optimization of high-concentration endostatin formulation: harmonization of excipients’ contributions on colloidal and conformational stabilities. Int J Pharm. 2017 Sep;530(1–2):173–86.10.1016/j.ijpharm.2017.07.057Search in Google Scholar PubMed

[47] Casaz P, Boucher E, Wollacott R, Pierce BG, Rivera R, Sedic M, et al. Resolving self-association of a therapeutic antibody by formulation optimization and molecular approaches. MAbs. 2014;6(6):1533–9.10.4161/19420862.2014.975658Search in Google Scholar PubMed PubMed Central

[48] Kutsch M, Hortmann P, Herrmann C, Weibels S, Weingärtner H. Dissecting ion-specific from electrostatic salt effects on amyloid fibrillation: A case study of insulin. Biointerphases. 2016 Mar;11(1):019008.10.1116/1.4941008Search in Google Scholar PubMed

[49] Buell AK, Hung P, Salvatella X, Welland ME, Dobson CM, Knowles TP. Electrostatic effects in filamentous protein aggregation. Biophys J. 2013 Mar;104(5):1116–26.10.1016/j.bpj.2013.01.031Search in Google Scholar PubMed PubMed Central

[50] Arosio P, Barolo G, Müller-Späth T, Wu H, Morbidelli M. Aggregation stability of a monoclonal antibody during downstream processing. Pharm Res. 2011 Aug;28(8):1884–94.10.1007/s11095-011-0416-7Search in Google Scholar PubMed

[51] Hauptmann A, Podgoršek K, Kuzman D, Srčič S, Hoelzl G, Loerting T. Impact of buffer, protein concentration and sucrose addition on the aggregation and particle formation during freezing and thawing. Pharm Res. 2018 Mar;35(5):101. Erratum in: Pharm Res. 2018 Jul;36(9):132.10.1007/s11095-018-2378-5Search in Google Scholar PubMed PubMed Central

[52] Nicoud L, Jagielski J, Pfister D, Lazzari S, Massant J, Lattuada M, et al. Kinetics of monoclonal antibody aggregation from dilute toward concentrated conditions. J Phys Chem B. 2016 Apr;120(13):3267–80.10.1021/acs.jpcb.5b11791Search in Google Scholar PubMed

[53] Li J, Krause ME, Chen X, Cheng Y, Dai W, Hill JJ, et al. Interfacial stress in the development of biologics: fundamental understanding, current practice, and future perspective. AAPS J. 2019 Mar;21(3):44.10.1208/s12248-019-0312-3Search in Google Scholar PubMed PubMed Central

[54] Bee JS, Stevenson JL, Mehta B, Svitel J, Pollastrini J, Platz R, et al. Response of a concentrated monoclonal antibody formulation to high shear. Biotechnol Bioeng. 2009 Aug;103(5):936–43.10.1002/bit.22336Search in Google Scholar PubMed PubMed Central

[55] Duerkop M, Berger E, Dürauer A, Jungbauer A. Influence of cavitation and high shear stress on HSA aggregation behavior. Eng Life Sci. 2018 Mar;18(3):169–78.10.1002/elsc.201700079Search in Google Scholar PubMed PubMed Central

[56] Grigolato F, Arosio P. Synergistic effects of flow and interfaces on antibody aggregation. Biotechnol Bioeng. 2020 Feb;117(2):417–28.10.1002/bit.27212Search in Google Scholar PubMed

[57] Abbas SA, Sharma VK, Patapoff TW, Kalonia DS. Characterization of antibody-polyol interactions by static light scattering: implications for physical stability of protein formulations. Int J Pharm. 2013 May;448(2):382–9.10.1016/j.ijpharm.2013.03.058Search in Google Scholar PubMed

[58] Buchanan LE, Dunkelberger EB, Tran HQ, Cheng PN, Chiu CC, Cao P, et al. Mechanism of IAPP amyloid fibril formation involves an intermediate with a transient β-sheet. Proc Natl Acad Sci USA. 2013 Nov;110(48):19285–90.10.1073/pnas.1314481110Search in Google Scholar PubMed PubMed Central

[59] Arosio P, Rima S, Morbidelli M. Aggregation mechanism of an IgG2 and two IgG1 monoclonal antibodies at low pH: from oligomers to larger aggregates. Pharm Res. 2013 Mar;30(3):641–54.10.1007/s11095-012-0885-3Search in Google Scholar PubMed

[60] Kim N, Remmele RL Jr, Liu D, Razinkov VI, Fernandez EJ, Roberts CJ. Aggregation of anti-streptavidin immunoglobulin gamma-1 involves Fab unfolding and competing growth pathways mediated by pH and salt concentration. Biophys Chem. 2013 Feb;172:26–36.10.1016/j.bpc.2012.12.004Search in Google Scholar PubMed

[61] Mossuto MF. Disulfide bonding in neurodegenerative misfolding diseases. Int J Cell Biol. 2013:318319.10.1155/2013/318319Search in Google Scholar PubMed PubMed Central

[62] Esfandiary R, Parupudi A, Casas-Finet J, Gadre D, Sathish H. Mechanism of reversible self-association of a monoclonal antibody: role of electrostatic and hydrophobic interactions. J Pharm Sci. 2015 Feb;104(2):577–86.10.1002/jps.24237Search in Google Scholar PubMed

[63] Yearley EJ, Godfrin PD, Perevozchikova T, Zhang H, Falus P, Porcar L, et al. Observation of small cluster formation in concentrated monoclonal antibody solutions and its implications to solution viscosity. Biophys J. 2014 Apr;106(8):1763–70.10.1016/j.bpj.2014.02.036Search in Google Scholar PubMed PubMed Central

[64] Wälchli R, Fanizzi F, Massant J, Arosio P. Relationship of PEG-induced precipitation with protein-protein interactions and aggregation rates of high concentration mAb formulations at 5°C. Eur J Pharm Biopharm. 2020 Jun;151:53–60.10.1016/j.ejpb.2020.03.011Search in Google Scholar PubMed

[65] Wu H, Kroe-Barrett R, Singh S, Robinson AS, Roberts CJ. Competing aggregation pathways for monoclonal antibodies. FEBS Lett. 2014 Mar;588(6):936–41.10.1016/j.febslet.2014.01.051Search in Google Scholar PubMed

[66] Berg JM, Tymoczko JL, Stryer L. Biochemistry. 5th edition. New York: W H Freeman; 2002.Search in Google Scholar

[67] Barnett GV, Qi W, Amin S, Neil Lewis E, Roberts CJ. Aggregate structure, morphology and the effect of aggregation mechanisms on viscosity at elevated protein concentrations. Biophys Chem. 2015 Dec;207:21–9.10.1016/j.bpc.2015.07.002Search in Google Scholar PubMed

[68] Sahin Z, Demir YK, Kayser V. Global kinetic analysis of seeded BSA aggregation. Eur J Pharm Sci. 2016 Apr;86:115–24.10.1016/j.ejps.2016.03.007Search in Google Scholar PubMed

[69] Kalonia C, Toprani V, Toth R, Wahome N, Gabel I, Middaugh CR, et al. Effects of protein conformation, apparent solubility, and protein—protein interactions on the rates and mechanisms of aggregation for an IgG1Monoclonal antibody. J Phys Chem B. 2016 Jul;120(29):7062–75.10.1021/acs.jpcb.6b03878Search in Google Scholar PubMed

[70] Borzova VA, Markossian KA, Kara DA, Kurganov B. Kinetic regime of dithiothreitol-induced aggregation of bovine serum albumin. Int J Biol Macromol. 2015 Sep;80:130–8.10.1016/j.ijbiomac.2015.06.040Search in Google Scholar PubMed

[71] Ravuluri S, Bansal R, Chhabra N, Rathore AS. Kinetics and Characterization of Non-enzymatic Fragmentation of Monoclonal Antibody Therapeutics. Pharm Res. 2018 May;35(7):142–54.10.1007/s11095-018-2415-4Search in Google Scholar PubMed

[72] Yin S, Pastuskovas CV, Khawli LA, Stults JT. Characterization of therapeutic monoclonal antibodies reveals differences between in vitro and in vivo time-course studies. Pharm Res. 2013 Jan;30(1):167–78.10.1007/s11095-012-0860-zSearch in Google Scholar PubMed

[73] Salinas BA, Sathish HA, Shah AU, Carpenter JF, Randolph TW. Buffer-dependent fragmentation of a humanized full-length monoclonal antibody. J Pharm Sci. 2010 Jul;99(7):2962–74.10.1002/jps.22056Search in Google Scholar PubMed PubMed Central

[74] Fekete S, Gassner AL, Rudaz S, Schappler J, Guillarme D. Analytical strategies for the characterization of therapeutic monoclonal antibodies. Trends Analyt Chem. 2013;42:74–83.10.1016/j.trac.2012.09.012Search in Google Scholar

[75] Zhu ZC, Chen Y, Ackerman MS, Wang B, Wu W, Li B, et al. Investigation of monoclonal antibody fragmentation artifacts in non-reducing SDS-PAGE. J Pharm Biomed Anal. 2013 Sep;83:89–95.10.1016/j.jpba.2013.04.030Search in Google Scholar PubMed

[76] Nowak C, K Cheung J, M Dellatore S, Katiyar A, Bhat R, Sun J, et al. Forced degradation of recombinant monoclonal antibodies: A practical guide. MAbs. 2017 Nov/Dec;9(8):1217–30.10.1080/19420862.2017.1368602Search in Google Scholar PubMed PubMed Central

[77] Glover ZK, Basa L, Moore B, Laurence JS, Sreedhara A. Metal ion interactions with mAbs: Part 1. MAbs. 2015;7(5):901–11.10.1080/19420862.2015.1062193Search in Google Scholar PubMed PubMed Central

[78] Weiss WF 4th, Young TM, Roberts CJ. Principles, approaches, and challenges for predicting protein aggregation rates and shelf life. J Pharm Sci. 2009 Apr;98(4):1246–77.10.1002/jps.21521Search in Google Scholar PubMed

[79] Zheng S, Qiu D, Adams M, Li J, Mantri RV, Gandhi R. Investigating the degradation behaviors of a therapeutic monoclonal antibody associated with pH and buffer species. AAPS PharmSciTech. 2017 Jan;18(1):42–8.10.1208/s12249-015-0403-0Search in Google Scholar PubMed

[80] Nicoud L, Arosio P, Sozo M, Yates A, Norrant E, Morbidelli M. Kinetic analysis of the multistep aggregation mechanism of monoclonal antibodies. J Phys Chem B. 2014 Sep;118(36):10595–606.10.1021/jp505295jSearch in Google Scholar PubMed

[81] Morris AM, Watzky MA, Finke RG. Protein aggregation kinetics, mechanism, and curve-fitting: a review of the literature. Biochim Biophys Acta. 2009 Mar;1794(3):375–97.10.1016/j.bbapap.2008.10.016Search in Google Scholar PubMed

[82] Morris AM, Watzky MA, Agar JN, Finke RG. Fitting neurological protein aggregation kinetic data via a 2-step, minimal/“Ockham's razor” model: the Finke-Watzky mechanism of nucleation followed by autocatalytic surface growth. Biochemistry. 2008 Feb;47(8):2413–27.10.1021/bi701899ySearch in Google Scholar PubMed

[83] Kendrick BS, Cleland JL, Lam X, Nguyen T, Randolph TW, Manning MC, et al. Aggregation of recombinant human interferon gamma: kinetics and structural transitions. J Pharm Sci. 1998 Sep;87(9):1069–76.10.1021/js9801384Search in Google Scholar PubMed

[84] Bansal R, Dash R, Rathore AS. Impact of mAb Aggregation on Its Biological Activity: Rituximab as a Case Study. J Pharm Sci. 2020 Sep;109(9):2684–98.10.1016/j.xphs.2020.05.015Search in Google Scholar PubMed

[85] Lumry R, Eyring H. Conformation changes of proteins. J Phys Chem. 1954;58(2):110–20.10.1021/j150512a005Search in Google Scholar

[86] Hamrang Z, Rattray NJ, Pluen A. Proteins behaving badly: emerging technologies in profiling biopharmaceutical aggregation. Trends Biotechnol. 2013 Aug;31(8):448–58.10.1016/j.tibtech.2013.05.004Search in Google Scholar PubMed

[87] den Engelsman J, Garidel P, Smulders R, Koll H, Smith B, Bassarab S, et al. Strategies for the assessment of protein aggregates in pharmaceutical biotech product development. Pharm Res. 2011 Apr;28(4):920–33.10.1007/s11095-010-0297-1Search in Google Scholar PubMed PubMed Central

[88] Kumar M, Pant A, Bansal R, Pandey A, Gomes J, Khare K, et al. Electron microscopy-based semi-automated characterization of aggregation in monoclonal antibody products. Comput Struct Biotechnol J. 2020 Jun;18:1458–65.10.1016/j.csbj.2020.06.009Search in Google Scholar PubMed PubMed Central

[89] Roberts CJ. Protein aggregation and its impact on product quality. Curr Opin Biotechnol. 2014 Dec;30:211–7.10.1016/j.copbio.2014.08.001Search in Google Scholar PubMed PubMed Central

[90] Arosio P, Rima S, Lattuada M, Morbidelli M. Population balance modeling of antibodies aggregation kinetics. J Phys Chem B. 2012 Jun;116(24):7066–75.10.1021/jp301091nSearch in Google Scholar PubMed

[91] Nicoud L, Owczarz M, Arosio P, Morbidelli M. A multiscale view of therapeutic protein aggregation: a colloid science perspective. Biotechnol J. 2015 Mar;10(3):367–78.10.1002/biot.201400858Search in Google Scholar PubMed

[92] Vetri V, Foderà V. The route to protein aggregate superstructures: particulates and amyloid-like spherulites. FEBS Lett. 2015 Sep;589(19PartA 19 Pt A):2448–63.10.1016/j.febslet.2015.07.006Search in Google Scholar PubMed

[93] Filipe V, Que I, Carpenter JF, Löwik C, Jiskoot W. In vivo fluorescence imaging of IgG1 aggregates after subcutaneous and intravenous injection in mice. Pharm Res. 2014 Jan;31(1):216–27.10.1007/s11095-013-1154-9Search in Google Scholar PubMed

[94] Ghosh R, Calero-Rubio C, Saluja A, Roberts CJ. Relating Protein-Protein Interactions and Aggregation Rates From Low to High Concentrations. J Pharm Sci. 2016 Mar;105(3):1086–96.10.1016/j.xphs.2016.01.004Search in Google Scholar PubMed

[95] Morriss-Andrews A, Shea JE. Computational studies of protein aggregation: methods and applications. Annu Rev Phys Chem. 2015 Apr;66(1):643–66.10.1146/annurev-physchem-040513-103738Search in Google Scholar PubMed

[96] Arosio P, Michaels TC, Linse S, Månsson C, Emanuelsson C, Presto J, et al. Kinetic analysis reveals the diversity of microscopic mechanisms through which molecular chaperones suppress amyloid formation. Nat Commun. 2016 Mar;7(1):10948.10.1038/ncomms10948Search in Google Scholar PubMed PubMed Central

[97] Holm LS, Mcumber A, Rasmussen JE, Obiols-Rabasa M, Thulstrup PW, Kasimova MR, et al. The effect of protein PEGylation on physical stability in liquid formulation. J Pharm Sci. 2014 Oct;103(10):3043–54.10.1002/jps.24094Search in Google Scholar PubMed

[98] Ferebee R, Hakem IF, Koch A, Chen M, Wu Y, Loh D, et al. Light scattering analysis of mono-and multi-PEGylated bovine serum albumin in solution: role of composition on structure and interactions. J Phys Chem B. 2016 May;120(20):4591–9.10.1021/acs.jpcb.6b03097Search in Google Scholar PubMed

[99] Liebner R, Meyer M, Hey T, Winter G, Besheer A. Head to head comparison of the formulation and stability of concentrated solutions of HESylated versus PEGylated anakinra. J Pharm Sci. 2015 Feb;104(2):515–26.10.1002/jps.24253Search in Google Scholar PubMed

[100] Mishra A, Bansal R, Sreenivasan S, Dash R, Joshi S, Singh R, et al. Structure-Based Design of Small Peptide Ligands to Inhibit Early-Stage Protein Aggregation Nucleation. J Chem Inf Model. 2020 Jun;60(6):3304–14.10.1021/acs.jcim.0c00226Search in Google Scholar PubMed

[101] Dave AC, Loveday SM, Anema SG, Jameson GB, Singh H. Glycation as a tool to probe the mechanism of β-lactoglobulin nanofibril self-assembly. J Agric Food Chem. 2014 Apr;62(14):3269–78.10.1021/jf405441gSearch in Google Scholar PubMed

[102] Kemter K, Altrichter J, Derwand R, Kriehuber T, Reinauer E, Scholz M. Amino Acid-Based Advanced Liquid Formulation Development for Highly Concentrated Therapeutic Antibodies Balances Physical and Chemical Stability and Low Viscosity. Biotechnol J. 2018 Jul;13(7):e1700523.10.1002/biot.201700523Search in Google Scholar PubMed

[103] Scholz M, Altrichter J, Kemter K, inventors; Leukocare AG, assignee. Novel stabilisation method for viruses or bacteria. Worldwide patent WO/2013/001034. 2013 Jan 3.Search in Google Scholar

[104] Kemter K. Terminal Radiation Sterilization of Combination Products. Biofunctional Surface Engineering. In: Scholz M., editor. Biofunctional Surface Engineering. New York: Jenny Stanford Publishing; 2014. p. 11–25.10.1201/b15667-3Search in Google Scholar

[105] Cicerone MT, Pikal MJ, Qian KK. Stabilization of proteins in solid form. Adv Drug Deliv Rev. 2015 Oct;93:14–24.10.1016/j.addr.2015.05.006Search in Google Scholar PubMed PubMed Central

[106] Wang B, Cicerone MT, Aso Y, Pikal MJ. The impact of thermal treatment on the stability of freeze-dried amorphous pharmaceuticals: II. Aggregation in an IgG1 fusion protein. J Pharm Sci. 2010 Feb;99(2):683–700.10.1002/jps.21960Search in Google Scholar PubMed

[107] Malavasi NV, Foguel D, Bonafe CF, Braga C, Chura-Chambi RM, Vieira JM, et al. Protein refolding at high pressure: optimization using eGFP as a model. Process Biochem. 2011;46(2):512–8.10.1016/j.procbio.2010.10.002Search in Google Scholar

[108] Pardeshi NN, Qi W, Dahl K, Caplan L, Carpenter JF. Microparticles and nanoparticles delivered in intravenous saline and in an intravenous solution of a therapeutic antibody product. J Pharm Sci. 2017 Feb;106(2):511–20.10.1016/j.xphs.2016.09.028Search in Google Scholar PubMed PubMed Central

[109] Singh SM, Bandi S, Jones DN, Mallela KM. Effect of Polysorbate 20 and Polysorbate 80 on the higher-order structure of a monoclonal antibody and its fab and fc fragments probed using 2D nuclear magnetic resonance spectroscopy. J Pharm Sci. 2017 Dec;106(12):3486–98.10.1016/j.xphs.2017.08.011Search in Google Scholar PubMed PubMed Central

[110] Wang X, An Z, Luo W, Xia N, Zhao Q. Molecular and functional analysis of monoclonal antibodies in support of biologics development. Protein Cell. 2018 Jan;9(1):74–85.10.1007/s13238-017-0447-xSearch in Google Scholar PubMed PubMed Central

[111] Rayaprolu BM, Strawser JJ, Anyarambhatla G. Excipients in parenteral formulations: selection considerations and effective utilization with small molecules and biologics. Drug Dev Ind Pharm. 2018 Oct;44(10):1565–71.10.1080/03639045.2018.1483392Search in Google Scholar PubMed

[112] Kannan A, Shieh IC, Fuller GG. Linking aggregation and interfacial properties in monoclonal antibody-surfactant formulations. J Colloid Interface Sci. 2019 Aug;550:128–38.10.1016/j.jcis.2019.04.060Search in Google Scholar PubMed

[113] Bhugra C, Pikal MJ. Role of thermodynamic, molecular, and kinetic factors in crystallization from the amorphous state. J Pharm Sci. 2008 Apr;97(4):1329–49.10.1002/jps.21138Search in Google Scholar PubMed

[114] Agarkhed M, O’Dell C, Hsieh MC, Zhang J, Goldstein J, Srivastava A. Effect of surfactants on mechanical, thermal, and photostability of a monoclonal antibody. AAPS PharmSciTech. 2018 Jan;19(1):79–92.10.1208/s12249-017-0845-7Search in Google Scholar PubMed

[115] Wang S, Zhang N, Hu T, Dai W, Feng X, Zhang X, et al. Viscosity-lowering effect of amino acids and salts on highly concentrated solutions of two IgG1 monoclonal antibodies. Mol Pharm. 2015 Dec;12(12):4478–87.10.1021/acs.molpharmaceut.5b00643Search in Google Scholar PubMed

[116] Wang S, Wu G, Zhang X, Tian Z, Zhang N, Hu T, et al. Stabilizing two IgG1 monoclonal antibodies by surfactants: balance between aggregation prevention and structure perturbation. Eur J Pharm Biopharm. 2017 May;114:263–77.10.1016/j.ejpb.2017.01.025Search in Google Scholar PubMed

[117] Mensink MA, Frijlink HW, van der Voort Maarschalk K, Hinrichs WL. How sugars protect proteins in the solid state and during drying (review): mechanisms of stabilization in relation to stress conditions. Eur J Pharm Biopharm. 2017 May;114:288–95.10.1016/j.ejpb.2017.01.024Search in Google Scholar PubMed

[118] Horn J, Jena S, Aksan A, Friess W. Freeze/thaw of IGG solutions. Eur J Pharm Biopharm. 2019 Jan;134:185–9.10.1016/j.ejpb.2018.12.001Search in Google Scholar PubMed

[119] Nicoud L, Cohrs N, Arosio P, Norrant E, Morbidelli M. Effect of polyol sugars on the stabilization of monoclonal antibodies. Biophys Chem. 2015 Feb;197:40–6.10.1016/j.bpc.2014.12.003Search in Google Scholar PubMed

[120] Shaikh S, Nazam N, Danish Rizvi SM, Hussain T, Farhana A, Choi I. Anti-amyloid aggregating gold nanoparticles: can they really be translated from bench to bedside for Alzheimer's disease treatment? Curr Protein Pept Sci. 2020;21(12):1184–92.10.2174/1389203721666200226101930Search in Google Scholar PubMed

[121] Bansal R, Dhawan S, Chattopadhyay S, Maurya GP, Haridas V, Rathore AS. Peptide Dendrons as Thermal-Stability Amplifiers for Immunoglobulin G1 Monoclonal Antibody Biotherapeutics. Bioconjug Chem. 2017 Oct;28(10):2549–59.10.1021/acs.bioconjchem.7b00389Search in Google Scholar PubMed

Received: 2021-03-12
Accepted: 2021-05-07
Published Online: 2021-06-19

© 2021 Rohit Bansal et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. In Our Image: The Ethics of CRISPR Genome Editing
  2. Doxorubicin loaded magnetism nanoparticles based on cyclodextrin dendritic-graphene oxide inhibited MCF-7 cell proliferation
  3. The Effects of Oral Liposomal Glutathione and In Vitro Everolimus in Altering the Immune Responses against Mycobacterium bovis BCG Strain in Individuals with Type 2 Diabetes
  4. Myosteatosis in NAFLD patients correlates with plasma Cathepsin D
  5. Polyphenols from Passiflora ligularis Regulate Inflammatory Markers and Weight Gain
  6. Everolimus-induced effector mechanism in macrophages and survivability of Erdman, CDC1551 and HN878 strains of Mycobacterium tuberculosis infection
  7. Pathophysiological Aspects of the Development of Abdominal Aortic Aneurysm with a Special Focus on Mitochondrial Dysfunction and Genetic Associations
  8. Size-based Degradation of Therapeutic Proteins - Mechanisms, Modelling and Control
  9. Incidence and characterisation of Bacillus cereus bacteriophages from Thua Nao, a Thai fermented soybean product
  10. Bcl2 negatively regulates Protective Immune Responses During Mycobacterial Infection
  11. Insulin resistance is positively associated with plasma cathepsin D activity in NAFLD patients
  12. 3D host cell and pathogen-based bioassay development for testing anti-tuberculosis (TB) drug response and modeling immunodeficiency
  13. Universal scanning-free initiation of eukaryote protein translation–a new normal
  14. Characterization of Differentially Expressed miRNAs by CXCL12/SDF-1 in Human Bone Marrow Stromal Cells
  15. Combination therapy using TGF-β1 and STI-571 can induce apoptosis in BCR-ABL oncogene-expressing cells
  16. SIT1 transporter as a potential novel target in treatment of COVID-19
  17. Primary Erlotinib Resistance in a Patient with Non-Small Cell Lung Cancer Carrying Simultaneous Compound EGFR L718A, Q849H, and L858R Mutations
  18. Plant growth-promoting properties of bacterial endophytes isolated from roots of Thymus vulgaris L. and investigate their role as biofertilizers to enhance the essential oil contents
  19. Implication of plant growth-promoting rhizobacteria of Bacillus spp. as biocontrol agents against wilt disease caused by Fusarium oxysporum Schlecht. in Vicia faba L.
  20. Antibodies targeting enzyme inhibition as potential tools for research and drug development
  21. Retraction Note
  22. Retraction of: Alzheimer's and Danish dementia peptides induce cataract and perturb retinal architecture in rats
Downloaded on 21.1.2026 from https://www.degruyterbrill.com/document/doi/10.1515/bmc-2021-0008/html
Scroll to top button