throbber
Microbial Cell Factories
`
`BioMed Central
`
`Open Access
`Review
`Strategies for the recovery of active proteins through refolding of
`bacterial inclusion body proteins
`Luis Felipe Vallejo and Ursula Rinas*
`
`Address: Biochemical Engineering Division, GBF German Research Center for Biotechnology, Mascheroder Weg 1, 38124 Braunschweig, Germany
`
`Email: Luis Felipe Vallejo - lfv@gbf.de; Ursula Rinas* - uri@gbf.de
`* Corresponding author
`
`Published: 02 September 2004
`
`Microbial Cell Factories 2004, 3:11
`
`doi:10.1186/1475-2859-3-11
`
`Received: 29 June 2004
`Accepted: 02 September 2004
`
`This article is available from: http://www.microbialcellfactories.com/content/3/1/11
`
`© 2004 Vallejo and Rinas; licensee BioMed Central Ltd.
`This is an open-access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/2.0),
`which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
`
`Abstract
`Recent advances in generating active proteins through refolding of bacterial inclusion body proteins
`are summarized in conjunction with a short overview on inclusion body isolation and solubilization
`procedures. In particular, the pros and cons of well-established robust refolding techniques such as
`direct dilution as well as less common ones such as diafiltration or chromatographic processes
`including size exclusion chromatography, matrix- or affinity-based techniques and hydrophobic
`interaction chromatography are discussed. Moreover, the effect of physical variables (temperature
`and pressure) as well as the presence of buffer additives on the refolding process is elucidated. In
`particular, the impact of protein stabilizing or destabilizing low- and high-molecular weight additives
`as well as micellar and liposomal systems on protein refolding is illustrated. Also, techniques
`mimicking the principles encountered during in vivo folding such as processes based on natural and
`artificial chaperones and propeptide-assisted protein refolding are presented. Moreover, the
`special requirements for the generation of disulfide bonded proteins and the specific problems and
`solutions, which arise during process integration are discussed. Finally, the different strategies are
`examined regarding their applicability for large-scale production processes or high-throughput
`screening procedures.
`
`Background
`Recombinant DNA technology made available several
`simple techniques for transferring and efficiently express-
`ing desired genes in a foreign cell. Thus, it was thought
`that unlimited and inexpensive sources of otherwise rare
`proteins would become accessible. It soon was observed
`that the host cell had a great influence on the quality and
`quantity of the produced recombinant protein. For exam-
`ple, recombinant protein production in mammalian cells
`yields a biologically active protein with all the required
`posttranslational modifications. However, mammalian
`cell cultivation is characterized by low volumetric yields
`of the recombinant protein, long cultivation times and
`
`requirements for expensive bioreactors and medium com-
`ponents. All these points have a great impact on the pro-
`duction costs. On the other hand, bacterial cultivation
`processes are based on inexpensive media in which fast
`growth and high cell concentrations can be obtained.
`These high cell concentrations combined with higher pro-
`duction rates of the bacterial expression system result in
`higher volumetric productivities. However, the produc-
`tion of recombinant proteins in bacteria such as
`Escherichia coli frequently yields an inactive protein, aggre-
`gated in the form of so-called inclusion bodies.
`
`Page 1 of 12
`(page number not for citation purposes)
`
`Amgen Exhibit 2010
`Lupin, Ltd. v. Amgen Inc.
`IPR2021-00326
`Page 00001
`
`

`

`Microbial Cell Factories 2004, 3:11
`
`http://www.microbialcellfactories.com/content/3/1/11
`
`Though, producing an inactive target protein in the form
`of inclusion bodies is an important drawback, it also has
`several advantages such as the high degree of purity of the
`target protein in the aggregate fraction and the increased
`protection from proteolytic degradation compared to the
`soluble counterpart. Inclusion bodies have long been con-
`sidered completely inert towards in vivo dissolution; only
`recently it was shown that proteins can be resolubilized in
`vivo from inclusion body deposits [1]. Although inclusion
`bodies in general consist of inactive proteins, E. coli can be
`the superior expression system compared to eukaryotic
`expression systems when the activity of the recombinant
`protein can be regained through refolding from the pro-
`duced inclusion bodies. However, one needs to consider
`that the decision to select a specific expression system fre-
`quently is based on more trivial reasons such as staff
`knowledge and available equipment and facilities of the
`producing company/institute.
`
`A good example to demonstrate the diverse routes that
`can be used for recombinant protein production is the
`manufacturing of tissue-type plasminogen activator (tPA).
`This protein enables the dissolution of blood clots and is
`used therapeutically for the treatment of myocardial inf-
`arction, thrombosis, pulmonary embolism, and strokes.
`To assure sufficient tPA for such a widespread application,
`an economic production process is a necessity. From the
`beginning, both the mammalian as well as the microbial
`route were explored for the production of tPA [2]. tPA is a
`fairly large (527 amino acids) monomeric protein con-
`taining 17 disulfide bridges. Because of this complexity,
`tPA was first produced in E. coli in the form of inclusion
`bodies while the mammalian expression system yielded
`an active protein that was secreted into the culture
`medium. More recently, obtaining active tPA through
`secretion into the periplasm of E. coli was attempted [3-5].
`The early unsatisfactory yields have been improved [6,7]
`rendering the E. coli secretion system as a future potential
`alternative route to generate functional tPA. Other recom-
`binant organisms such as yeast [8], fungi [9] or insect cells
`[10] have not yet been considered as industrial producers
`for this protein.
`
`Initially, the recombinant tPA introduced into the market
`was obtained from genetically engineered mammalian
`cells [2]. At that time, generating biologically active tPA
`from E. coli produced material was a process with a poor
`overall yield [2]. Today, the majority of commercial tPA
`(alteplase, Activase®) is still produced using a mammalian
`expression system (Genentech: http://www.gene.com/
`gene/products/information/cardiovascular/activase/). In
`addition, an amino substituted tPA produced by the
`mammalian expression system with increased half-life
`(tenecteplase) was developed. Alternatively, a non-glyco-
`sylated, truncated tPA (reteplase, Retavase®) produced in
`
`E. coli in form of inclusion bodies and afterwards refolded
`to its biologically active form is now on the market (Cen-
`tocor: http://www.retavase.com/) and apparently gains
`market share at the cost of the mammalian-derived prod-
`uct(s) (see Genentech 2004 First Quarter Report).
`
`Thus, continuous research effort focused on developing
`new refolding techniques or improving existing ones by
`including novel refolding aiding agents can make the bac-
`terial inclusion body system an excellent alternative to the
`mammalian expression system or other expression sys-
`tems that can directly generate active proteins with a com-
`plex disulfide bond structure. The foremost aim in
`improving protein refolding from E. coli produced inclu-
`sion bodies is to increase both the allowed protein con-
`centrations during the refolding process and the final
`refolding yield. Recent advances in this area are summa-
`rized in conjunction with a short overview on inclusion
`body isolation and solubilization procedures. Moreover,
`the different techniques are discussed regarding their
`applicability for large-scale production processes or high-
`throughput screening procedures.
`
`Isolation and solubilization of inclusion bodies
`A high degree of purification of the recombinant protein
`can be achieved by inclusion body isolation [for recent
`reviews on various aspects of inclusion body formation
`and renaturation of inclusion body proteins please refer
`also to [11-18]]. Inclusion bodies are in general recovered
`by low speed centrifugation of bacterial cells mechanically
`disrupted either by using ultrasonication for small, French
`press for medium, or high pressure homogenization for
`large scale. Main protein contaminants in the crude inclu-
`sion body fraction are proteins from the cell envelope, the
`outer membrane proteins [19]. These proteins are not
`integral inclusion body contaminants but coprecipitate
`together with other insoluble cell material during inclu-
`sion body recovery. Lysozyme-EDTA treatment before cell
`homogenization facilitates cell disruption. Addition of
`detergents such as Triton X-100 and/or low concentra-
`tions of chaotropic compounds either prior to mechanical
`cell breakage or for washing crude inclusion body prepa-
`rations allow the removal of membrane proteins or other
`nonspecifically adsorbed cell material [11-14].
`
`After their isolation, inclusion bodies are commonly sol-
`ubilized by high concentrations of chaotropic agents such
`as guanidinium hydrochloride or urea. Although expen-
`sive, guanidinium hydrochloride is in general preferred
`due to its superior chaotropic properties. Moreover, urea
`solutions may contain and spontaneously produce
`cyanate [20], which can carbamylate the amino groups of
`the protein [21]. In addition, inclusion body solubiliza-
`tion by urea is pH dependent and optimum pH condi-
`tions must be determined for each protein [22]. There are
`
`Page 2 of 12
`(page number not for citation purposes)
`
`Page 00002
`
`

`

`Microbial Cell Factories 2004, 3:11
`
`http://www.microbialcellfactories.com/content/3/1/11
`
`also reports that inclusion bodies can be solubilized at
`extreme pH in the presence or absence of low concentra-
`tions of denaturants [23-25]. However, extreme pH treat-
`ments can result in irreversible protein modifications such
`as deamidation and alkaline desulfuration of cysteine res-
`idues [26]. Finally, inclusion bodies can be solubilized
`with different types of detergents [27,28], low concentra-
`tions of denaturants [29,30], or even by utilization of the
`aggregation suppressor arginine [29]. Inclusion body pro-
`teins solubilized under these mild conditions can possess
`a native-like secondary structure [28-30], and may even
`reveal some biological activity [29,31]. It has also been
`demonstrated that the utilization of milder solubilization
`conditions can lead to higher final refolding yields com-
`pared to solubilization by high concentrations of gua-
`nidinium hydrochloride or urea [27].
`
`In addition to the solubilizing agent, the presence of low
`molecular weight thiol reagents such as dithiothreitol
`(DTT) or 2-mercaptoethanol is generally required. These
`substances will reduce nonnative inter- and intramolecu-
`lar disulfide bonds possibly formed by air oxidation dur-
`ing cell disruption and will also keep the cysteines in their
`reduced state [14,15]. Optimum conditions for disrup-
`tion of existing disulfide bonds are found at mild alkaline
`pH since the nucleophilic attack on the disulfide bond is
`carried out by the thiolate anion. Residual concentrations
`of reducing substances can negatively affect the refolding
`process, thus, they are frequently removed (e.g. by dialy-
`sis) before starting the refolding procedure. As an alterna-
`tive,
`immobilized
`reducing
`agents
`(e.g. DTT;
`VectraPrime™, Biovectra) could simplify reducing agent
`removal by centrifugation after the solubilization process.
`Finally, the pH must be reduced before the removal of the
`reducing agent from the solution containing the solubi-
`lized protein to prevent the formation of undesired
`disulfide bonds.
`
`Principles of refolding solubilized and unfolded proteins
`Correct refolding versus aggregation
`In general, the methods used for inclusion body solubili-
`zation result in a soluble protein that is devoid of its
`native conformation. This protein must then be trans-
`ferred into conditions that allow the formation of the
`native structure (e.g. low denaturant concentration).
`Moreover, appropriate redox conditions have to be estab-
`lished when the protein contains disulfide bonds in the
`native state. When proper conditions for refolding are
`identified, the refolding process can require a few seconds
`or several days. During this period, the correct refolding
`pathway competes, often in disadvantage, with misfold-
`ing and aggregation of the target protein (Figure 1). Pro-
`tein refolding involves intramolecular interactions and
`follows first order kinetics [32-35]. Protein aggregation,
`however, involves intermolecular interactions and, thus,
`
`Simplified model of correct folding versus misfolding and aggregationFigure 1
`
`
`Simplified model of correct folding versus misfolding and
`aggregation. The correct protein folding pathway (1) often
`competes with misfolding (2) and aggregation (3). Aggrega-
`tion occurs among intermediates with exposed hydrophobic
`patches, which are buried in the correctly folded protein
`(blue lines, hydrophilic solvent-exposed parts of the protein;
`red lines: hydrophobic patches).
`
`is a kinetic process of second or higher order, which is
`favored at high protein concentrations [32-35]. In fact,
`refolding yields commonly decrease with increasing ini-
`tial concentrations of the unfolded protein independent
`of the refolding method applied [35-40].
`
`Aggregates are formed by nonnative intermolecular
`hydrophobic interactions between protein folding inter-
`mediates, which have not yet buried their hydrophobic
`amino acid stretches (Figure 1). When the refolding proc-
`ess is beyond these aggregation-prone intermediates, the
`productive folding pathway is favored and aggregation
`does not occur. Therefore, prevention of hydrophobic
`intermolecular interaction during the first steps of refold-
`ing is crucial to allow successful renaturation at high pro-
`tein concentrations. Only recently a non-empirical
`method for predicting the fate of proteins during the
`refolding process was proposed [41]. It is based on the
`second viral coefficient, which indicates the magnitude of
`protein interaction under certain refolding conditions,
`and thus its tendency to aggregate. However, though
`being soluble in the refolding buffer is essential for a pro-
`tein molecule to refold, it does not ensure that it will fold
`into the native form.
`
`Page 3 of 12
`(page number not for citation purposes)
`
`Page 00003
`
`

`

`Microbial Cell Factories 2004, 3:11
`
`http://www.microbialcellfactories.com/content/3/1/11
`
`Are further purification steps required after solubilization of inclusion
`bodies?
`The recombinant target protein represents in general the
`major fraction of the inclusion body proteins. Therefore,
`refolding attempts can be undertaken directly after solubi-
`lization of the inclusion bodies. Some reports, however,
`claim higher refolding yields when the solubilized inclu-
`sion body proteins are purified prior to the refolding
`attempt [36,39,42,43]. Additional purification has been
`recommended when the protein of interest represents less
`than 2–5% of the total cell protein [26] or less than 2/3 of
`the total inclusion body protein [42]. The type of contam-
`inants can also be crucial for the success of the refolding
`process. For example, typical non-proteinaceous contam-
`inants of inclusion body preparations did not affect
`refolding yields of lysozyme, while proteinaceous con-
`taminants, which have a high tendency towards aggrega-
`tion significantly reduced refolding yields [42]. Further
`purification prior to the refolding attempt does not seem
`to be required, even at low target protein concentrations,
`when the solubilized inclusion body proteins are sub-
`jected to refolding conditions during size exclusion chro-
`matography where refolding and purification can occur
`simultaneously [44]. All pros and cons of any further puri-
`fication step have to be carefully evaluated as they cause
`potential protein loss and additional production costs.
`
`Techniques for protein refolding
`Direct dilution
`The simplest refolding procedure is to dilute the concen-
`trated protein-denaturant solution into a refolding buffer
`that allows the formation of the native structure of the
`protein. Most frequently, the final protein concentration
`after dilution is in the 1–10 µg/ml range in order to favor
`the productive refolding instead of the unproductive
`aggregation pathway. Though ideal at laboratory scale,
`this technique has serious drawbacks during scale-up as
`huge refolding vessels and additional cost-intensive con-
`centration steps are required after renaturation.
`
`A major improvement of this technique was the develop-
`ment of a method where the solubilized, denatured pro-
`tein is added in pulses or continuously into the refolding
`buffer [37,40,45-47]. This technique still keeps the sim-
`plicity of the direct dilution method while considerably
`increasing the final concentration of the refolded protein.
`Prerequisite is an appropriate knowledge of the folding
`kinetics of the target protein. The addition of the concen-
`trated protein-denaturant solution should occur at rates
`slower than the rate-determining folding step of the target
`protein, thereby avoiding the accumulation of aggrega-
`tion-prone folding intermediates [37,46]. For pulse addi-
`tion it has been recommended that 80% of the maximum
`refolding yield should be reached before adding the next
`pulse [14]. Other factors to be considered are the increas-
`
`ing residual concentration of the denaturant with each
`pulse, which should not surpass concentrations that affect
`the refolding of the protein, and the amount of protein
`added per pulse, which should be optimized in batch
`experiments to minimize aggregation [14].
`
`Membrane controlled denaturant removal
`Another technique to transfer the solubilized and
`unfolded protein to conditions allowing the formation of
`the native structure is the utilization of dialysis and diafil-
`tration systems for denaturant removal [e.g. [48-51]]. In
`contrast to the direct dilution method, the change from
`denaturing to native buffer conditions occurs gradually.
`Thus, the protein passes through different regimes of
`denaturant concentrations, where folding intermediates
`that are prone to aggregation may become populated.
`Most often, these techniques cause more aggregation dur-
`ing refolding compared to the direct dilution method [e.g.
`[52]]. Additionally, refolding yields can be negatively
`affected by non-specific adsorption of protein to the
`membrane. However, for some proteins and with the
`appropriate denaturant removal rates, adapted to the
`requirements of the target protein, high refolding yields at
`high protein concentrations can be obtained [50-53]. A
`fairly simple device was recently introduced allowing con-
`tinuous or pulse refolding in a similar way as in the direct
`dilution method [54].
`
`Chromatographic methods for protein refolding
`Protein refolding based on size exclusion chromatography
`Buffer exchange for denaturant removal can also be car-
`ried out by using size exclusion chromatography (SEC).
`Most frequently, the denaturant-protein solution is
`injected into a column previously equilibrated with the
`refolding buffer [44,55-58]. Subsequent elution with the
`refolding buffer results in a refolded protein in the eluate
`fraction with a considerably higher concentration com-
`pared to concentrations that can be reached by the simple
`dilution technique [44,56,58]. Protein refolding may be
`completed in the column or for proteins with slow folding
`kinetics the final folding steps may occur in the eluate
`fraction [44]. Aggregate formation is supposed to be
`reduced either by physical separation of aggregation-
`prone folding intermediates in the porous structures of
`the gel [56] or, more likely, by resolubilization of formed
`aggregates through the delayed running front of the
`denaturant, which gives the solubilized aggregates
`another opportunity to refold [14]. For proteins, which
`exhibit superior refolding yields during gradual denatu-
`rant removal, such as lysozyme, elution during SEC is
`preferably performed by using a decreasing denaturant
`gradient [38,52,59]. In specific cases, the denaturant
`removal can be accompanied with other changes in the
`buffer composition (i.e. pH) for further optimization of
`refolding conditions [52]. An additional advantage of this
`
`Page 4 of 12
`(page number not for citation purposes)
`
`Page 00004
`
`

`

`Microbial Cell Factories 2004, 3:11
`
`http://www.microbialcellfactories.com/content/3/1/11
`
`chromatographic method is the concomitant purification
`of the target protein during the refolding process [44].
`Furthermore, some recent applications have shown the
`feasibility of using SEC for continuous processes of pro-
`tein refolding [60,61]. Also, SEC in combination with the
`use of an annular chromatography system can be coupled
`to an ultrafiltration and recycling unit for reinjection of
`resolubilized aggregates, which may form during the
`refolding process [60].
`
`Some parameters for refolding using SEC are of key
`importance. For example, protein aggregation during
`sample injection can cause low refolding yields [62];
`injecting the sample followed by an additional small vol-
`ume of denaturant solution solves
`this problem
`[44,52,62]. Also, optimum results can only be reached
`when the properties of the chromatographic resin allow
`efficient separation of the renatured target protein from
`different folding intermediates, misfolded protein, and
`aggregates that might form during the refolding process
`[59,63,64]. In general, lower refolding yields are obtained
`by injecting the denatured protein at high concentrations
`[38,58,60,61,64] and/or by elution at high rates
`[52,62,64]. Both conditions result in poor separation
`among different folding intermediates thereby boosting
`protein precipitation.
`
`Matrix-assisted protein refolding
`Attaching the solubilized and unfolded protein to a solid
`support prior to changing from denaturing to native
`buffer conditions is another approach to avoid the
`unwanted intermolecular interaction between aggrega-
`tion-prone folding intermediates. Binding of the solubi-
`lized and unfolded protein to the matrix requires the
`formation of a stable protein-matrix complex withstand-
`ing the presence of chaotropic agents. However, after
`changing to native buffer conditions, the detachment of
`the refolded target protein from the matrix should easily
`be accomplished. Several combinations of binding
`motives and matrices have been employed for binding the
`unfolded protein to the solid support. For example, pro-
`teins with a natural occurring charged patch in the
`unfolded chain, which binds to ion exchange resins
`[59,65-67], or proteins containing artificially engineered
`peptide tags such as the hexahistidine tag, which binds to
`immobilized metal ions [59,68-70], or N- or C-terminal
`hexaarginine tags binding to a polyanionic support [71],
`or protein fusions with denaturant-resistant binding
`domains, such as a glutathione S-transferase fragment,
`which binds to an anion exchange matrix [72] or the cel-
`lulose binding domain of the cellulose degrading mul-
`tienzyme complex of
`the
`thermophilic bacterium
`Clostridium thermocellum, which binds to a cellulose
`matrix [73], have been employed. After binding, the
`matrix-protein complex is brought to refolding conditions
`
`by any of the above-mentioned techniques such as dilu-
`tion [71], dialysis [68,73], or buffer exchange through
`chromatography [59,66,69,70,72]. Finally, the refolded
`protein can be detached from the matrix, e.g. in the case
`hexahistidine-tagged proteins by elution with EDTA [69]
`or imidazole [59,70] or by buffers with high ionic
`strength in the case of proteins bound by ionic interac-
`tions [59,65,66,71,72]. Due to the selective binding,
`matrix-assisted refolding can combine the renaturation of
`the target protein along with its purification from host cell
`protein contaminants [69,70,72].
`
`Refolding using hydrophobic interaction chromatography
`Hydrophobic interaction chromatography (HIC) has also
`been successfully used for protein refolding with concom-
`itant removal of contaminating proteins during the rena-
`turation process [74-78]. Unfolded proteins are applied to
`the column at high salt concentrations and refolded and
`eluted with a decreasing salt gradient. In contrast to the
`above-mentioned chromatographic methods there is no
`requirement for typical refolding aiding agents such as
`arginine during the in-column refolding process. Moreo-
`ver, refolding of the disulfide containing protein proinsu-
`lin was even obtained in the absence of a redox system in
`the mobile phase [76].
`
`It has been proposed that refolding is facilitated during
`HIC because unfolded proteins adsorb at high salt con-
`centrations to the hydrophobic matrix and, thus, are not
`prone to aggregation. Additionally, hydrophobic regions
`of the protein that adsorb to the HIC matrix form micro-
`domains around which native structure elements can
`form. During migration through the column, the protein
`will pass through several steps of adsorption and desorp-
`tion, controlled by the salt concentration and hydropho-
`bicity of the intermediate(s), resulting finally in the
`formation of the native structure [75].
`
`Physical and chemical features improving protein refolding
`yields
`Apart from any of the above-mentioned techniques for
`protein refolding, there are physical and chemical varia-
`bles that have a great impact on the final yield of biologi-
`cally active protein. For example, temperature as well as
`the composition of the refolding buffer are important var-
`iables influencing the final refolding yield.
`
`Physical variables aiding protein refolding
`The most important physical variable influencing the
`refolding yield is the temperature [40,45,50,51]. Temper-
`ature has a dual effect on the refolding process. On one
`side, it influences the speed of folding and on the other it
`influences the propensity towards aggregation of folding
`intermediates with exposed hydrophobic patches. Also,
`there is limited temperature range in which each protein
`
`Page 5 of 12
`(page number not for citation purposes)
`
`Page 00005
`
`

`

`Microbial Cell Factories 2004, 3:11
`
`http://www.microbialcellfactories.com/content/3/1/11
`
`is thermodynamically stable in a given buffer system [79].
`In general, low temperatures support the productive fold-
`ing pathway as hydrophobic aggregation is suppressed.
`However, low temperatures also slow down the folding
`rates, thus increasing the time required for renaturation
`[51]. For refolding attempts of a new protein, 15°C has
`been proposed as a good starting point [14].
`
`Pressure was identified as another important physical var-
`iable affecting protein structure as well as protein refold-
`ing processes [80]. It was shown that high pressure up to
`3 kbar can disrupt oligomeric protein structures [80] and
`can dissolve protein aggregates and inclusion bodies
`[81,82]. The disassembled protein monomers retain
`native-like secondary structure up to 5 kbar [80]. After
`gradual depressurization, they can reach their native state
`even at high protein concentrations, because folding
`intermediates prone to aggregate at atmospheric pressure
`are prevented from aggregation by high pressure [81-83].
`
`Chemicals aiding protein refolding
`Certainly, L-arginine is nowadays the most commonly
`used refolding aiding agent [14]. It impedes aggregate for-
`mation by enhancing the solubility of folding intermedi-
`ates, presumably by shielding hydrophobic regions of
`partially folded chains. In addition, it has been shown
`that numerous other low molecular weight additives such
`as detergents, protein-stabilizing agents such as glycerol or
`even low residual concentrations of denaturants improve
`refolding yields by suppressing aggregation [14]. In addi-
`tion, high-molecular weight additives such as polyethyl-
`ene glycol were used successfully for enhancing protein
`refolding yields [84]. More recently, low-molecular
`weight non-detergent zwitterionic agents such as sulfo-
`betaines, substituted pyridines and pyrroles and acid sub-
`stituted aminocyclohexanes have been employed
`successfully for protein renaturation [40,85-87]. Moreo-
`ver, polymers with temperature-dependent hydrophobic-
`ity were effectively applied for protein refolding at higher
`temperatures [88,89]. The benefit of each of these refold-
`ing aiding agents for a given renaturation system has to be
`elucidated experimentally, as they are not equally advan-
`tageous for all proteins. The mechanisms of interactions
`of these refolding aiding agents with the folding interme-
`diates remain often obscure although it is clear that all
`these substances suppress aggregation in favor of the pro-
`ductive folding pathway [90].
`
`Micelles and liposomes as protein refolding aiding systems
`Detergents [91,92] and phospholipids [93,94], in the
`form of micelles and liposomes [95], respectively, as well
`as mixed micelle systems formed by phospholipids and
`detergents [92,95,96] have shown potential to aid protein
`refolding. Most likely, illegitimate hydrophobic interac-
`tions between folding intermediates are suppressed by
`
`transient nonpolar interactions between the protein and
`the micelle or liposome [91-93]. Additional transient
`polar interactions in mixed micelles are supposed to be
`responsible for higher refolding yields compared to only
`detergent-based micelle systems [92,96]. Moreover, lipo-
`somes linked covalently to chromatographic resins have
`potential to combine renaturation and separation of the
`refolded target protein [93,94].
`
`Reversed micelles, formed when an aqueous detergent
`solution is mixed with an organic solvent, can also facili-
`tate protein refolding by avoiding aggregate formation
`[97]. The denatured protein, once transferred to this solu-
`tion, tries to avoid the organic phase, and, after reaching
`the hydrophilic core of the reversed micelle, can refold as
`a single molecule [97]. Recently, it was demonstrated that
`protein precipitates can be solubilized by direct addition
`into the reversed micellar system allowing refolding with
`high yields at high protein concentrations [98-100]. Yet,
`direct solubilization of inclusion bodies in reversed micel-
`lar systems has not been reported. In addition, recovery of
`refolded protein from these micellar structures is not eas-
`ily accomplished [97,99].
`
`Chemical and biological protein refolding aiding agents
`mimicking in vivo folding conditions
`Natural chaperones
`Chaperones are a group of proteins conserved in all king-
`doms, which play a key role in assisting in vivo protein
`folding and protecting cellular proteins from different
`types of environmental stress by suppressing protein
`aggregation. For example, the major E. coli chaperonin
`GroEL is involved in the in vivo folding of 10% of all newly
`synthesized proteins at normal growing conditions, and
`of 30% under stress conditions [101]. GroEL assists pro-
`tein folding by a first capturing step of aggregation-prone
`folding intermediates [102]. The release of the folding-
`competent form is then accomplished in an ATP-depend-
`ent fashion through the action of the cochaperonin GroES
`[102].
`
`Natural chaperones have also been applied successfully to
`refold various proteins in vitro [103]. However, their rou-
`tine application is limited by their cost, the relatively high
`chaperone concentration required (at least equimolar to
`the target protein) and the need for their removal after the
`refolding procedure [103,104]. Some procedures have
`tried to overcome these limitations by utilizing immobi-
`lized and reusable (mini)-chaperone systems [104-106].
`Nevertheless, chaperone-based refolding processes are not
`robust enough for large-scale processes [14].
`
`Artificial chaperones
`A further development of the detergent-based micellar
`system mimics the two-step mechanism of chaperone-
`
`Page 6 of 12
`(page number not for citation purposes)
`
`Page 00006
`
`

`

`Microbial Cell Factories 2004, 3:11
`
`http://www.microbialcellfactories.com/content/3/1/11
`
`assisted protein folding. The capturing step is performed
`by diluting the denatured protein into a detergent solu-
`tion, which prevents protein aggregation through the for-
`mation of mixed protein-detergent micelles [107-110].
`Aqueous solutions of hydrogel nanoparticles (e.g. self-
`assembly of hydrophobized polysaccharides such as cho-
`lesterol-bearing pullulan) have been also used for the cap-
`turing step [111]. The release of the folding-competent
`protein is subsequently initiated by the addition of cyclo

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


Or .

Accessing this document will incur an additional charge of $.

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

Accept $ Charge
throbber

Still Working On It

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

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

throbber

A few More Minutes ... Still Working

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

Thank you for your continued patience.

This document could not be displayed.

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

Your account does not support viewing this document.

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

Your account does not support viewing this document.

Set your membership status to view this document.

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

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

Become a Member

One Moment Please

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

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

Your document is on its way!

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

Sealed Document

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

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


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket