BBAPAP-39367; No. of pages: 8; 4C: 2, 3 Biochimica et Biophysica Acta xxx (2014) xxx–xxx

Contents lists available at ScienceDirect

Biochimica et Biophysica Acta journal homepage: www.elsevier.com/locate/bbapap

Review

2

Principles and engineering of antibody folding and assembly☆

3

Matthias J. Feige a,⁎, Johannes Buchner b,⁎

4 5

a

6

a r t i c l e

7 8 9 10 11

Article history: Received 9 April 2014 Received in revised form 4 June 2014 Accepted 6 June 2014 Available online xxxx

12 13 14 15

Keywords: Antibody folding Antibody assembly Antibody engineering

a b s t r a c t

R O

i n f o

O

Department of Tumor Cell Biology, St. Jude Children's Research Hospital, Memphis 38105, TN, USA CIPSM at the Department of Chemistry, Technische Universität München, 85748 Garching, Germany

D

P

Antibodies are uniquely suited to serve essential roles in the human immune defense as they combine several specific functions in one hetero-oligomeric protein. Their constant regions activate effector functions and their variable domains provide a stable framework that allows incorporation of highly diverse loop sequences. The combination of non-germline DNA recombination and mutation together with heavy and light chain assembly allows developing variable regions that specifically recognize essentially any antigen they may encounter. However, it also requires tailor-made mechanisms to guarantee that folding and association of antibodies is carefully controlled before the protein is secreted from a plasma cell. Accordingly, the generic immunoglobulin fold β-barrel structure of antibody domains has been fine-tuned during evolution to fit the different requirements. Work over the past decades has identified important aspects of the folding and assembly of antibody domains and chains revealing domain specific variations of a general scheme. The most striking is the folding of an intrinsically disordered antibody domain in the context of its partner domain as the basis for antibody assembly and its control on the molecular level in the cell. These insights have not only allowed a better understanding of the antibody folding process but also provide a wealth of opportunities for rational optimization of antibody molecules. In this review, we summarize current concepts of antibody folding and assembly and discuss how they can be utilized to engineer antibodies with improved performance for different applications. This article is part of a Special Issue entitled: Recent advances in molecular engineering of antibody. © 2014 Published by Elsevier B.V.

C

T

E

b

F

1

E

32 33 37 35 34

R

36

1. Introduction

39

The human immune defense depends on reliable and efficient discrimination of self from non-self. In all higher vertebrates, antibodies play a major role in this task and protect the organism against threats ranging from toxins to microbial pathogens [1]. Antibodies are complex glycoproteins that are secreted in large quantities by plasma cells [2,3]. They are made up of heavy chains (HCs) and light chains (LCs) with the simplest human antibodies being hetero-tetramers that comprise two HCs and two LCs each (Fig. 1). Five antibody classes exist in humans (IgA, G, D, E and M) that differ in their HCs (αHC, γHC, δHC, εHC and μHC) comprising either four (αHC, γHC, δHC) or five (εHC, μHC) immunoglobulin (Ig) domains. Only two classes of light chains (κLC and λLC) are shared between the different human antibody classes. Even though the various antibody classes differ in their biological functions and structural details, they have two important features in common that enable them to defend the organism against most infectious or toxic

46 47 48 49 50 51 52 53

N C O

44 45

U

42 43

R

38

40 41

16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31

☆ This article is part of a Special Issue entitled: Recent advances in molecular engineering of antibody. ⁎ Corresponding authors. E-mail addresses: [email protected] (M.J. Feige), [email protected] (J. Buchner).

challenges it may encounter: variable regions that bind to antigens and constant regions that link the antibody and its bound antigen to the cellular immune defense (Fig. 1). The constant region of antibodies incorporates multiple biological functions. It binds to Fc receptors inducing antibody-dependent cellmediated cytotoxicity (ADCC) and activates the complement system [4]. The constant region furthermore has an important impact on the biological half-life of antibodies due to its binding to the neonatal Fc receptor that induces antibody recycling after cellular uptake into endosomes [5,6]. Variable regions of one HC and one LC each form a composite antigen binding surface (paratope) (Fig. 1). Antigen binding sites of the variable regions are generated by non-germline genetic rearrangements, non-templated base pair additions and hypermutation in a microevolutionary process [7,8] that gives rise to proteins that can bind essentially any target structure with high affinity and specificity. An array of techniques has been developed, from ribosome, phage and yeast display to genetically engineered mice with a human antibody repertoire to obtain antibodies with the desired specificities [9–11]. Recently, this possible repertoire has even been further extended by generating bispecific antibodies [6,12–14]. Together, the highly specific binding of antibodies and their link to the cellular immune response have rendered them the most widespread

http://dx.doi.org/10.1016/j.bbapap.2014.06.004 1570-9639/© 2014 Published by Elsevier B.V.

Please cite this article as: M.J. Feige, J. Buchner, Principles and engineering of antibody folding and assembly, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbapap.2014.06.004

54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76

2

M.J. Feige, J. Buchner / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

Fv

e op ng) rat ndi pa en bi

f ra gm

tig (an

V

en

H

t

VH VL

V

b Fa

f ra

gm

en

t

L

C

C H1

H

1

CL

C L

S

S S S S

S

S

S

CH 2

CH 2

a b f c

ed

IgG1

F

CH3 CH3

Fab

CL

O

Fc fragment

(effector functions)

g

P

R O

Fig. 1. Schematic and structural elements of the IgG1 antibody molecule and the Ig fold. Left panel: The IgG1 antibody consists of two heavy chains (blue) and two light chains (green) that together form two identical antigen binding sites (paratopes), which are part of the Fab fragment (fragment antigen binding). The smallest complete dimeric antigen binding entity is the Fv fragment (fragment variable). Effector functions are mediated via the Fc fragment (fragment crystallizable). The CH2 domains are glycosylated (gray hexagons) and disulfide bridges (S\S) connect the two heavy chains as well as the heavy and light chains. Middle panel: The Fab fragment is connected by an interchain disulfide bond and each of the domains has one buried disulfide bond (yellow, CPK representation). Proline residues within the Fab fragment are highlighted in red. Trans prolines are shown in a stick representation, and cis prolines are shown in a CPK representation. The small helices in the CL domain, which are found in all most antibody constant domains except for e.g. CH1, are colored in orange. Right panel: As shown for the CL domain, the constant domain Ig fold consists of seven strands (a–f), two of which (b and f) are connected by the buried disulfide bond. Prolines and small helices are depicted in the same manner as in the middle panel.

85 86 87 88 89 90

2. Engineering antibodies: stability and folding pathways

92 93

Each of the antibody domains shows the Ig fold, the most widespread protein topology in secretory pathway proteins, with variable domains belonging to the V-set and constant domains to the C-set [23,24]. Ig domains show a Greek key β-barrel topology generally composed of either seven (C-type) or nine (V-type) β-strands [23,24] (Fig. 1). A characteristic of this topology is that β-strands that are adjacent in space are not necessarily adjacent in sequence (Fig. 1). Most of the published studies on antibody optimization have focused on the antigen binding variable regions due to their immediate role in target recognition and the possibility to design smaller binding molecules without constant domains [20,25]. The opposite, however, applies when it comes to studies that have assessed Ig folding pathways in detail. These mostly focused on antibody constant domains or other members of the Ig superfamily [26–29]. Mutational approaches combined with kinetic analyses of a protein folding process allow the structural assessment of protein folding transition states (Phi value analysis) [30–33]. This approach has revealed that Ig domains initiate their folding by a nucleus composed of Ig strands b, c, e and f around which the Ig fold forms [34–36] (Fig. 1). However, antibody domains comprise several structural features that significantly influence this general mechanism with direct implications for protein engineering.

99 100 101 102 103 104 105 106 107 108 109 110 111 112 113

C

N

98

U

96 97

O

91

94 95

D

2.1. Disulfide bonds in antibody folding

E

A defining characteristic of most antibody domains is an intramolecular buried disulfide bridge perpendicular to the two sheets of the Ig fold [37] (Fig. 1). Early on, it has been recognized that this disulfide bond plays a crucial role in the stability of antibody domains in vitro [38–40] and complete antibodies in vivo [41]. Besides, the oligomeric structure of antibodies is stabilized by disulfide bonds connecting LCs and HCs as well as inter-HC disulfide bonds [42,43] (Fig. 1). In the case of the constant domain of antibody LCs (CL domain, Fig. 1), the internal disulfide bridge accelerates folding and inhibits off-pathway misfolding reactions [44,45]. Of note, in antibody constant domains the internal disulfide bridge connects strands b and f (Fig. 1), which are part of the common Ig topology folding nucleus [34–36]. For the variable domains, the intramolecular disulfide bond, also connecting strands b and f, similarly reduces their tendency to aggregate [46]. Together this argues for an intimate link between folding of antibody domains and formation of their disulfide bond which was also found in vivo [47] where disulfide bonds in antibody domains and between antibody chains already form co-translationally [41,42]. Disulfide bond formation is aided by protein disulfide isomerases (PDIs) in vivo [48,49] and in vitro [50] (Fig. 2). The time window for the action of PDI can be extended in the presence of the Hsp70 chaperone BiP [51], a major ER-resident protein involved in the folding, assembly and quality control of antibodies in the cell [52,53] (Fig. 2), by inhibiting premature collapse of the polypeptide chains. Engineering additional disulfide bonds into the Ig fold or between Ig domains is one of the most widespread approaches to stabilize antibody domains. A significant stabilization can be achieved this way [54–58] as described in several recent reviews [59–62]. A particularly well studied example in this regard is Fv fragments, the most simple dimeric binding module derived from antibodies, which comprise only one VL and one VH domain (Fig. 1). Nowadays, Fv fragments are mainly used as single chain Fv fragments (scFv) where the C-terminus of the VL and the N-terminus of the VH domain are connected by a peptide linker [63,64]. Disulfide bonds have been shown to be particularly useful to stabilize scFv fragments either by engineering the antigen binding site [65], or by, more generally applicable, the framework regions to stabilize VL–VH heterodimers [66,67]. Indeed, dissociation of the variable domains is often an initial step preceding aggregation of scFv fragments under challenging conditions and thus disulfide bond engineering can

T

C

83 84

E

81 82

R

79 80

diagnostic and therapeutic biomolecules that can be used for a variety of applications from detecting and marking tumors in vivo, via treating autoimmune diseases to combating cancer cells [6]. Recent developments have given rise to antibodies that are able to pass the blood brain barrier, opening up a whole new field of diseases that could potentially be targeted with tailored antibodies [15,16]. Thus, a vital interest exists in optimizing antibody production, stability and half-life in the organism as well as in inhibiting side reactions like aggregation, which can cause serious side effects including immunogenicity [17–20]. As strategies to engineer specificity have been extensively reviewed elsewhere [9–11, 21,22], in the following, we will focus on recent insights into the determinants of in vitro and in vivo antibody folding and assembly pathways, with an emphasis on how these can be used to design better antibody therapeutics.

R

77 78

Please cite this article as: M.J. Feige, J. Buchner, Principles and engineering of antibody folding and assembly, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbapap.2014.06.004

114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152

N

C

VH

SH SH

HS HS

BiP HS HS

VH

CH1

HS S

HS

CH3

R SH PDIs R S +2H+/2eR SH R S SH SH

H

SH

disulfide bridge formation

C

C H1 HS

BiP HS

SH SH

C H2

V

H

O

VL

C H2

CH2

VH

VL

HC

Cαi-1 C

ω=1

O

80

°

Cαi

PPIases

N C αi

Cαi-1 C

ω=0

°

kon koff

N

O

HC-LC assembly

peptidyl-prolyl isomerization

CL

VL SH structure formation

HS CL

SH

SH

LC

VL CL SH

V

C

S S S C C H1 S H1 S S HS S S HS

CH2

H

C H2

V

VL

L

C

T

assembly-induced folding

V

VH

H

CH 2

E

re ec

H

s

1C L

S

disulfide bridge formation

C

C H1

n

tio

V

L

CL

L

SH SH

SS S S S SS

CH 2

kunfolding

CH 2 unfolding

peptidyl-prolyl isomerization

CH3CH3

HS

L

SS S S S SS

CH3CH3

disulfide bridge formation chaperoness

H1 C

CH 2

E

CH3CH3

V

C

C H1 S

R

S S SH S S

H

L

CL

R

1

V

VH

D

CH3CH3

P

R O

M.J. Feige, J. Buchner / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

Please cite this article as: M.J. Feige, J. Buchner, Principles and engineering of antibody folding and assembly, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbapap.2014.06.004

U

O

Fig. 2. Critical steps in antibody folding, secretion and in vivo stability. IgG antibodies are assembled from two heavy and two light chains in the endoplasmic reticulum before secretion via the Golgi occurs. The molecular chaperone BiP stably binds to the permanently unfolded CH1 domain until displaced by the LC [53,80,103,104] and also transiently to other antibody domains [166,167]. Critical molecular events occurring in each of the steps are highlighted in the boxes. These represent prime targets for antibody engineering and optimization.

F

3

211

2.3. Small helices in antibody domain folding

212 213

Generally, the Ig fold is an all-β structure where flexible loops connect the β-strands that make up the topology. For antibodies, however, by far most constant domains possess two small helices that connect

181 182 183 184 185 186 187 188 189 190 191 192 193 194 195 196 197 198 199 200 201 202 203 204 205 206 207 208 209

214

C

179 180

E

177 178

R

175 176

R

173 174

O

172

C

170 171

N

168 169

U

166 167

3. Engineering antibodies: cellular quality control

227

Like essentially all secreted proteins, antibodies are produced in the endoplasmic reticulum (ER), an organelle dedicated to protein folding [93,94]. Quality control measures exist in the ER to ensure that only properly folded and assembled proteins leave the ER whereas incompletely folded proteins are retained and ultimately transported back to the cytosol to be degraded by the proteasome [95]. Quality control measures are of particular relevance for antibodies due to their intrinsic variability, which might affect folding and their potent effector functions once secreted. Thus, the cell must ensure that antibodies are folded and assembled correctly prior to secretion. Indeed, only completely assembled antibodies leave the ER. Most insights into this process have been obtained for IgM and IgG. In the case of IgM, an unpaired cysteine in the C-terminal tailpiece of the μHC, which forms a disulfide bond upon IgM oligomerization, plays a key role in retaining μHCs in the cell [96]. ERp44, a member of the ER-localized thioredoxin domain containing proteins, plays a major role in this process by engaging proteins that contain free cysteines [97,98]. If these proteins leave the ER prematurely, ERp44 re-shuttles them back to the ER. This is mediated by exposure of the ERp44 ER retention sequence in the more acidic post-ER compartments [97–101]. Another major quality control step for antibodies has been discovered for IgG. The first of the γHC constant domains, CH1, is permanently unfolded and thus remains stably bound to a complex of ER chaperones, the most prominent of it being BiP (heavy chain binding protein) [48,52,53,80,102] (Fig. 2). Light chains are able to displace BiP and induce folding of the CH1 domain [80,103, 104]. This reaction is limited by prolyl-isomerization, catalyzed by the PPIase CypB and rendered irreversible by the formation of the HC-LC interchain disulfide bond [80,103,104] (Fig. 2). Thus, LC-induced folding of the CH1 domain is a critical step in antibody quality control that sets apart unassembled HCs from assembled ones and might be amenable to optimization, by either improving the initial HC-LC assembly step, the assembly-dependent folding reaction itself or by manipulating the levels or composition of the chaperone machinery involved in this process [48,105,106] (Fig. 2). Alternatively, recapitulating a similar environment as in the ER in simpler organisms is an actively pursued direction [107] even though achieving authentic antibody glycosylation, important for many biological antibody functions [108,109], presents a major obstacle. It is noteworthy that although the details of antibody assembly pathways vary for different antibody classes, i.e. formation of the HC-dimer might precede assembly with LCs or, alternatively, HCLC heterodimers form first [110,111], LC-induced folding of the CH1 domain seems to be a generic feature and thus a particularly attractive target for optimization. A variation of this theme is found in pre-B cells, which express membrane-bound HCs but LC rearrangement has not yet taken place. These cells express the so-called surrogate light chain (SLC), a protein similar to LCs but made up of two non-covalently linked polypeptides, VpreB and λ5, that interact via a β-strand exchange mechanism [112–117]. HCs and SLC pair to form the pre-B cell receptor (pre-BCR) [112–116]. Even though detailed insights are still missing, the split nature of the SLC seems to be uniquely suited to test on the one hand for correct VH rearrangement and on the other hand for CH1

228

F

210

Among the 20 natural amino acids, proline is the only one for which a trans conformation of the peptide bond preceding the proline residue is only slightly more energetically favorable than the generally disfavored cis conformation [72] (Fig. 2). Thus, in the absence of structural constraints, e.g. in the unfolded state of a protein, prolines undergo cis/ trans isomerization. In the native state, however, generally only one isomer is compatible with the native structure often giving rise to slow folding reactions caused by peptidyl–prolyl isomerization reactions [73–75]. Antibody domains are rich in prolines. These make up ca. 5–10% of the primary sequence and are generally found in turns connecting the β-strands (Fig. 1). Thus, very often antibody domain folding is complicated by peptidyl–prolyl isomerization [28,76]. On the other hand, proline residues allow the formation of tight turns connecting the strands of Ig domains, which might be important in increasing protease resistance in the extracellular space, and prolines have been suggested to reduce the aggregation tendency of Ig domains [77], maybe due to their low propensity to form β-strands. Peptidyl– prolyl isomerization reactions are slow, taking seconds to minutes to complete at room temperature [75]. Thus, the cell has developed a class of enzymes, the peptidyl–prolyl isomerases (PPIases) to accelerate these reactions [78,79] (Fig. 2). For antibodies, it was shown that among those PPIases, the ER-localized PPIase cyclophilin B (CypB) plays an important role [48,80]. Its inhibition leads to reduced IgG secretion [81] and a general deficiency in protein quality control of the cell [82]. In terms of antibody engineering, the role of proline residues has not been assessed in detail yet. Cis-proline residues can be generally expected to be essential for the formation of the native antibody domain structure and a conserved cis-proline residue is found between strands b and c of most antibody constant domains [37,83]. In fact, mutation of this residue to an alanine was used to trap a partially folded antibody state [84]. Exceptions exist, however, and of note, the crystal structure of the Ig domains from shark IgNAR (immunoglobulin new antigen receptor) antibodies has revealed the absence of cis-proline residues in this otherwise conserved position between strands b and c for two particularly stable IgNAR domains [83]. Thus, the replacement of these prolines and their immediate surrounding amino acids by optimized loops might be a viable strategy to increase domain stability and simplify the folding reaction to optimize protein behavior [85]. Selectively replacing transproline residues might be an alternative option for optimization of antibody domain folding pathways as cis to trans isomerization can have a significant effect on antibody domain folding [86]. Additionally, optimizing residues in sequence proximity to proline residues might be feasible as these residues influence the cis/trans equilibrium of the peptidyl\prolyl bond and the catalysis by PPIases [87–89]. This might be used to favor either a trans or a cis peptidyl\prolyl bond, dependent on the native state of the respective residue under scrutiny, or improve catalysis of the isomerization reaction in vivo.

160 161

O

164 165

159

R O

2.2. Proline residues in antibody folding

157 158

215 216

P

163

155 156

strands a and b and e and f, respectively (Fig. 1). These helices are highly conserved in evolution [37,83]. Recently, they have been shown to form early on in the folding of the CL domain and reduce the tendency of this domain to misfold [84]. The second helix often contains a hydrophobic or aromatic amino acid that is properly positioned to participate in the hydrophobic core of the antibody domain [83,84]. Since the formation of α-helices is relatively well understood [90,91] optimization of these small helices in antibody domain folding might be an attractive way of increasing the folding efficiency and decreasing the misfolding tendency of antibody domains (Fig. 2) or Ig domains in general. This has already been shown to be feasible for β2-microglobulin [84], a member of the Ig superfamily with a tendency to form amyloids [29,92].

T

162

be complemented by framework engineering [68]. Insights into the structural details of antibody domain folding/unfolding pathways can help to identify particularly well-suited positions for introducing disulfide bonds and thus may even further increase the positive impact of engineered disulfide bonds [69–71]. In particular, results from the experimental analysis of the native state dynamics of antibody domains and their unfolding pathways might be useful in this context. This is of particular interest since often aggregation following unfolding renders the unfolding process irreversible, constituting a major problem in therapy [20] (Fig. 2).

D

153 154

M.J. Feige, J. Buchner / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

E

4

Please cite this article as: M.J. Feige, J. Buchner, Principles and engineering of antibody folding and assembly, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbapap.2014.06.004

217 218 219 220 221 222 223 224 225 226

229 230 231 232 233 234 235 236 237 238 239 240 241 242 243 244 245 246 247 248 249 250 251 252 253 254 255 256 257 258 259 260 261 262 263 264 265 266 267 268 269 270 271 272 273 274 275 276 277 278

M.J. Feige, J. Buchner / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

295 296 297 298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 336 337 338 339 340 341 342

C

293 294

E

5. Summary and perspectives

378

The seminal work by Goto and Hamaguchi determined crucial elements in the structure formation of antibody domains including the identification of proline cis/trans isomerization as the basis of the slow folding phase of antibodies and the role of the buried disulfide bridge in folding and stability of antibody domains [39,40,44,76]. These experiments were possible before the advent of recombinant protein expression as some patients secreted large amounts of light chains, the so-called Bence–Jones proteins. Also, further diseases exist that involve antibody deposits, such as the light chain deposition disease (LCDD). LCDD is a severe illness during which large amounts of antibody fibrils accumulate in different organs [158,159]. Some point mutations in the VL domain seem to predispose it for fibril formation [160]. Thus, antibodies, among the most important components in the defense mechanisms of our immune system, can have detrimental effects when things go awry. Work in recent years revealed the built in mechanisms to combat these processes and control the quality of each antibody molecule [28]. Antibodies are also special because they represent a rare case where insights into folding and assembly have been gained both from in vitro and in vivo studies. Antibodies were an early model system to demonstrate co-translational folding and disulfide bond formation [41,42] and our understanding of the chaperone assistance of structure formation, oxidation and assembly of proteins in the ER has significantly been shaped by studies on the biogenesis of antibodies which provided many insights into the members, organization and functions of the ER chaperone machinery [48,52,53,98,103,104,161–163]. Engineering of antibodies is important both for basic science aiming at understanding the foundations of protein structure, especially of the

379

F

291 292

R

289 290

R

287 288

N C O

285 286

U

283 284

O

Simpler and more stable antibodies are of particular interest for diagnostics and therapy. In a case of convergent evolution [121], nature has developed simpler than human antibodies at least twice. Camelids (e.g. camels and llamas) and cartilaginous fish (e.g. sharks, skates and rays) each possess antibodies naturally devoid of light chains, which are HC-only homodimers named HC-only antibodies (HCAbs) in camelids and IgNAR in sharks [122,123]. Normally, HC-LC assembly is controlled via assembly of the LC CL domain with the otherwise unfolded HC CH1 domain [53,80,103,104]. Accordingly, HCAbs in camelids and cartilaginous fish both are devoid of a CH1 domain thus allowing secretion of HC dimers. In camelids, the CH1 domain is still present in the genomic HCAb sequence yet absent in the HCAb transcript due to a mutation in a necessary splice site [124]. In cartilaginous fish the origin of a HC devoid of a CH1 domain is less clear [121]. That nature has developed HCAbs multiple times argues for distinct evolutionary advantages of HCAbs. Some of those are of particular interest for antibody engineering. HCAbs can be readily raised against many different antigens and due to the antigen-binding function encoded in a single domain, the underlying nucleotide sequences are relatively easy to obtain [125–127]. Their small size qualifies the variable domains of HCAbs as very attractive minimal antigen binding units by themselves, called nanobodies [125], some of which have been shown to be possible effective therapeutics tolerated by the immune system [128,129]. Of particular relevance in terms of antibody folding and assembly, the variable domains of HCAbs have undergone a long evolution towards being stable entities that function as monomers. Indeed, variable domains from HCAbs derived from both camelids and cartilaginous fish have more hydrophilic residues in parts of their framework region that would normally interact with the variable domain of the LC [130–135]. However, whereas the camelid monomeric variable domain VHH (variable domain of the heavy chain of HCAbs) has clearly developed from a conventional heavy chain variable domain [122,130,131], the shark IgNAR variable domain VNAR shows characteristics of both constant and variable antibody domains as well as T cell receptor variable domains and it might have evolved from a cell surface receptor [132,133]. Structural insights into naturally occurring monomeric variable domains have been successfully used to guide the design of monomeric human VH domains [125,136–138]. A particular attractive feature of camelid VHH domains is their ability to often refold after thermal denaturation, conditions where other mammalian dimeric variable domains tend to aggregate [139–141], even though exceptions to this general rule could be established [142–144]. This beneficial behavior qualifies VHH domains for tasks where classical variable domains would be too unstable, e.g. at elevated temperatures, or to use them as templates to optimize the biophysical properties of human VH domains by comprehensively assessing the impact of different mutations rather than by simply copying structural elements from VHH domains [145]. As an example, whereas in VHH and VNAR domains one of the antigen binding loops, the CDR3 loop (complementarity determining region 3), often contributes to shielding hydrophobic remnants of the former dimerization interface [130–133], this structural feature is not necessary to generate stable variable domains [145], and thus is not a restricting factor for the binding repertoire of monomeric variable domains. Alternatively, particularly stable human domains might act as a template for optimization, even though their CDRs seem to play an important role in stability thus maybe restricting the available repertoires of stable domains [146,147]. Besides their beneficial biophysical properties, their small size and their easy production in simple organisms [125,148], single chain variable domains from camelids and cartilaginous fish also have functional characteristics, which set them apart from dimeric variable domains

R O

282

P

4. Engineering antibodies: biomimetic approaches

343 344

D

281

and might explain why they have evolved [121]: Their extended CDR3 loop is often stabilized by aromatic residues and in particular additional disulfide bridges, which confer stability and increase antigen binding affinity [131–133,135,149,150]. It can thus extend away from the variable domain framework to allow recognition of buried, cryptic antigens that are not easily accessible to conventional dimeric antibodies [131,132, 151]. These cryptic antigens also include enzyme active sites and pathogen coat proteins [149], which cannot easily be mutated by a pathogen to escape immune detection. That indeed recognition of cryptic antigens is an evolutionary advantage is further underlined by the recent structural insights into bovine antibodies that can encompass extremely long CDR3 regions which form knob like structures stabilized by disulfide bonds and are present on a long stalk extruding from the VH region [152]. Extended and stabilized CDR regions, which have been developed multiple times in nature in different formats [130–133,152,153], might thus be interesting engineering options for hard to target antigens. Apart from providing insights into how simpler, monomeric variable regions can be obtained, shark IgNAR antibodies might also hold a clue for how to further stabilize human antibodies for therapy and diagnostics. Sharks enrich urea in their blood to inhibit the osmotic loss of water in a marine environment [154]. Urea is a well-known protein denaturant. Accordingly, shark antibodies are believed to be particularly stable [155,156], even though sharks also counteract the effect of urea by small osmolytes like Trimethylamine-N-oxid [157]. In a recent study it could indeed be shown that some constant domains derived from IgNAR are particularly stable against thermal and chemical denaturation and the crystal structures of those IgNAR constant domains revealed structural elements that contribute to their high stability: a salt bridge between the second of the small antibody helices and the loop connecting strands c and d as well as a slightly extended hydrophobic core [83]. Transferring these elements to a human CL domain stabilized this protein and led to increased antibody secretion from mammalian cells [83], likely since a better-behaved CL domain provides an optimized template for the CH1 domain to fold on in antibody assembly control [48,52,53,80,102].

E

folding as prerequisites for further B cell development mediated by preBCR signaling [118–120].

T

279 280

5

Please cite this article as: M.J. Feige, J. Buchner, Principles and engineering of antibody folding and assembly, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbapap.2014.06.004

345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377

380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406

431 432

Funding of MJF by the Leopoldina — National Academy of Sciences, grant number LPDS 2009-32 and of JB by the Deutsche Forschungsgemeinschaft (DFG) is gratefully acknowledged. We thank Julia Behnke for the helpful comments on the manuscript.

433

References

T

C

[1] M.F. Flajnik, M. Kasahara, Origin and evolution of the adaptive immune system: genetic events and selective pressures, Nat. Rev. Genet. 11 (2010) 47–59. [2] S. Cenci, R. Sitia, Managing and exploiting stress in the antibody factory, FEBS Lett. 581 (2007) 3652–3657. [3] J.W. Brewer, L.M. Hendershot, Building an antibody factory: a job for the unfolded protein response, Nat. Immunol. 6 (2005) 23–29. [4] M. Raghavan, P.J. Bjorkman, Fc receptors and their interactions with immunoglobulins, Annu. Rev. Cell Dev. Biol. 12 (1996) 181–220. [5] F.W. Brambell, W.A. Hemmings, I.G. Morris, A Theoretical model of gammaglobulin catabolism, Nature 203 (1964) 1352–1354. [6] N.A. Buss, S.J. Henderson, M. McFarlane, J.M. Shenton, L. de Haan, Monoclonal antibody therapeutics: history and future, Curr. Opin. Pharmacol. 12 (2012) 615–622. [7] J.M. Di Noia, M.S. Neuberger, Molecular mechanisms of antibody somatic hypermutation, Annu. Rev. Biochem. 76 (2007) 1–22. [8] S.D. Fugmann, A.I. Lee, P.E. Shockett, I.J. Villey, D.G. Schatz, The RAG proteins and V(D)J recombination: complexes, ends, and transposition, Annu. Rev. Immunol. 18 (2000) 495–527. [9] E.T. Boder, M. Raeeszadeh-Sarmazdeh, J.V. Price, Engineering antibodies by yeast display, Arch. Biochem. Biophys. 526 (2012) 99–106. [10] A. Pluckthun, Ribosome display: a perspective, Methods Mol. Biol. 805 (2012) 3–28. [11] A.L. Nelson, E. Dhimolea, J.M. Reichert, Development trends for human monoclonal antibody therapeutics, Nat. Rev. Drug Discov. 9 (2010) 767–774. [12] S.M. Lewis, X. Wu, A. Pustilnik, A. Sereno, F. Huang, H.L. Rick, G. Guntas, A. LeaverFay, E.M. Smith, C. Ho, C. Hansen-Estruch, A.K. Chamberlain, S.M. Truhlar, E.M. Conner, S. Atwell, B. Kuhlman, S.J. Demarest, Generation of bispecific IgG antibodies by structure-based design of an orthogonal Fab interface, Nat. Biotechnol. 32 (2014) 191–198. [13] W. Schaefer, J.T. Regula, M. Bahner, J. Schanzer, R. Croasdale, H. Durr, C. Gassner, G. Georges, H. Kettenberger, S. Imhof-Jung, M. Schwaiger, K.G. Stubenrauch, C. Sustmann, M. Thomas, W. Scheuer, C. Klein, Immunoglobulin domain crossover as a generic approach for the production of bispecific IgG antibodies, Proc. Natl. Acad. Sci. U. S. A. 108 (2011) 11187–11192. [14] S. Fenn, C.B. Schiller, J.J. Griese, H. Duerr, S. Imhof-Jung, C. Gassner, J. Moelleken, J.T. Regula, W. Schaefer, M. Thomas, C. Klein, K.P. Hopfner, H. Kettenberger, Crystal structure of an anti-Ang2 CrossFab demonstrates complete structural and functional integrity of the variable domain, PLoS ONE 8 (2013) e61953. [15] A.M. Chacko, C. Li, D.A. Pryma, S. Brem, G. Coukos, V. Muzykantov, Targeted delivery of antibody-based therapeutic and imaging agents to CNS tumors: crossing the blood–brain barrier divide, Expert Opin. Drug Deliv. 10 (2013) 907–926. [16] B. Cheng, H. Gong, H. Xiao, R.B. Petersen, L. Zheng, K. Huang, Inhibiting toxic aggregation of amyloidogenic proteins: a therapeutic strategy for protein misfolding diseases, Biochim. Biophys. Acta 1830 (2013) 4860–4871.

E

434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476

R

424 425

R

422 423

O

420 421

C

418 419

N

416 417

U

414 415

F

429 430 Q2

413

O

Acknowledgments

411 412

R O

428

409 410

[17] D. Lowe, K. Dudgeon, R. Rouet, P. Schofield, L. Jermutus, D. Christ, Aggregation, stability, and formulation of human antibody therapeutics, Adv. Protein Chem. Struct. Biol. 84 (2011) 41–61. [18] F. Ducancel, B.H. Muller, Molecular engineering of antibodies for therapeutic and diagnostic purposes, MAbs 4 (2012) 445–457. [19] I.R. Correia, Stability of IgG isotypes in serum, MAbs 2 (2010) 221–232. [20] J.M. Perchiacca, P.M. Tessier, Engineering aggregation-resistant antibodies, Annu. Rev. Chem. Biomol. Eng. 3 (2012) 263–286. [21] A. Doerner, L. Rhiel, S. Zielonka, H. Kolmar, Therapeutic antibody engineering by high efficiency cell screening, FEBS Lett. 588 (2014) 278–287. [22] G. Winter, A.D. Griffiths, R.E. Hawkins, H.R. Hoogenboom, Making antibodies by phage display technology, Annu. Rev. Immunol. 12 (1994) 433–455. [23] P. Bork, L. Holm, C. Sander, The immunoglobulin fold. Structural classification, sequence patterns and common core, J. Mol. Biol. 242 (1994) 309–320. [24] A.F. Williams, A.N. Barclay, The immunoglobulin superfamily—domains for cell surface recognition, Annu. Rev. Immunol. 6 (1988) 381–405. [25] S. Ewert, A. Honegger, A. Pluckthun, Stability improvement of antibodies for extracellular and intracellular applications: CDR grafting to stable frameworks and structure-based framework engineering, Methods 34 (2004) 184–199. [26] J.H. Han, S. Batey, A.A. Nickson, S.A. Teichmann, J. Clarke, The folding and evolution of multidomain proteins, Nat. Rev. Mol. Cell Biol. 8 (2007) 319–330. [27] A.P. Capaldi, S.E. Radford, Kinetic studies of beta-sheet protein folding, Curr. Opin. Struct. Biol. 8 (1998) 86–92. [28] M.J. Feige, L.M. Hendershot, J. Buchner, How antibodies fold, Trends Biochem. Sci. 35 (2010) 189–198. [29] Y. Goto, H. Yagi, K. Yamaguchi, E. Chatani, T. Ban, Structure, formation and propagation of amyloid fibrils, Curr. Pharm. Des. 14 (2008) 3205–3218. [30] A.R. Fersht, A. Matouschek, L. Serrano, The folding of an enzyme. I. Theory of protein engineering analysis of stability and pathway of protein folding, J. Mol. Biol. 224 (1992) 771–782. [31] L. Serrano, A. Matouschek, A.R. Fersht, The folding of an enzyme. III. Structure of the transition state for unfolding of barnase analysed by a protein engineering procedure, J. Mol. Biol. 224 (1992) 805–818. [32] A. Matouschek, L. Serrano, A.R. Fersht, The folding of an enzyme. IV. Structure of an intermediate in the refolding of barnase analysed by a protein engineering procedure, J. Mol. Biol. 224 (1992) 819–835. [33] L. Serrano, J.T. Kellis Jr., P. Cann, A. Matouschek, A.R. Fersht, The folding of an enzyme. II. Substructure of barnase and the contribution of different interactions to protein stability, J. Mol. Biol. 224 (1992) 783–804. [34] S.B. Fowler, J. Clarke, Mapping the folding pathway of an immunoglobulin domain: structural detail from Phi value analysis and movement of the transition state, Structure 9 (2001) 355–366. [35] E. Cota, A. Steward, S.B. Fowler, J. Clarke, The folding nucleus of a fibronectin type III domain is composed of core residues of the immunoglobulin-like fold, J. Mol. Biol. 305 (2001) 1185–1194. [36] S.J. Hamill, A. Steward, J. Clarke, The folding of an immunoglobulin-like Greek key protein is defined by a common-core nucleus and regions constrained by topology, J. Mol. Biol. 297 (2000) 165–178. [37] R. Huber, J. Deisenhofer, P.M. Colman, M. Matsushima, W. Palm, Crystallographic structure studies of an IgG molecule and an Fc fragment, Nature 264 (1976) 415–420. [38] H. Kikuchi, Y. Goto, K. Hamaguchi, Reduction of the buried intrachain disulfide bond of the constant fragment of the immunoglobulin light chain: global unfolding under physiological conditions, Biochemistry 25 (1986) 2009–2013. [39] Y. Goto, K. Hamaguchi, The role of the intrachain disulfide bond in the conformation and stability of the constant fragment of the immunoglobulin light chain, J. Biochem. 86 (1979) 1433–1441. [40] Y. Goto, K. Hamaguchi, Formation of the intrachain disulfide bond in the constant fragment of the immunoglobulin light chain, J. Mol. Biol. 146 (1981) 321–340. [41] L.W. Bergman, W.M. Kuehl, Formation of an intrachain disulfide bond on nascent immunoglobulin light chains, J. Biol. Chem. 254 (1979) 8869–8876. [42] L.W. Bergman, W.M. Kuehl, Formation of intermolecular disulfide bonds on nascent immunoglobulin polypeptides, J. Biol. Chem. 254 (1979) 5690–5694. [43] G.M. Edelman, B.A. Cunningham, W.E. Gall, P.D. Gottlieb, U. Rutishauser, M.J. Waxdal, The covalent structure of an entire gammaG immunoglobulin molecule, Proc. Natl. Acad. Sci. U. S. A. 63 (1969) 78–85. [44] Y. Goto, K. Hamaguchi, Unfolding and refolding of the reduced constant fragment of the immunoglobulin light chain. Kinetic role of the intrachain disulfide bond, J. Mol. Biol. 156 (1982) 911–926. [45] M.J. Feige, F. Hagn, J. Esser, H. Kessler, J. Buchner, Influence of the internal disulfide bridge on the folding pathway of the CL antibody domain, J. Mol. Biol. 365 (2007) 1232–1244. [46] R. Glockshuber, T. Schmidt, A. Pluckthun, The disulfide bonds in antibody variable domains: effects on stability, folding in vitro, and functional expression in Escherichia coli, Biochemistry 31 (1992) 1270–1279. [47] M.H. Skowronek, L.M. Hendershot, I.G. Haas, The variable domain of nonassembled Ig light chains determines both their half-life and binding to the chaperone BiP, Proc. Natl. Acad. Sci. U. S. A. 95 (1998) 1574–1578. [48] L. Meunier, Y.K. Usherwood, K.T. Chung, L.M. Hendershot, A subset of chaperones and folding enzymes form multiprotein complexes in endoplasmic reticulum to bind nascent proteins, Mol. Biol. Cell 13 (2002) 4456–4469. [49] R.A. Roth, S.B. Pierce, In vivo cross-linking of protein disulfide isomerase to immunoglobulins, Biochemistry 26 (1987) 4179–4182. [50] H. Lilie, S. McLaughlin, R. Freedman, J. Buchner, Influence of protein disulfide isomerase (PDI) on antibody folding in vitro, J. Biol. Chem. 269 (1994) 14290–14296. [51] M. Mayer, U. Kies, R. Kammermeier, J. Buchner, BiP and PDI cooperate in the oxidative folding of antibodies in vitro, J. Biol. Chem. 275 (2000) 29421–29425.

E

426 427

immunoglobulin fold, but also in the context of diseases based on antibodies and in particular for facilitating the use of therapeutic antibodies. Efforts in this regard are rather recent and it is envisioned that important discoveries are still ahead of us. Of particular relevance will be a more detailed understanding of antibody dissociation and unfolding pathways as antibody aggregation poses a major problem in therapy [20,164]. Methodological advancements to understand aggregation [164] and antibody optimization to resist aggregation [20,165] are recent developments of direct relevance to antibody therapy. These, however, will have to be complemented by kinetic and mechanistic insight into antibody unfolding pathways to allow rational design of aggregation-resistant antibody molecules. A deeper understanding of the rooting of the antibody fold in evolution contributed to the identification of structural features that influence basic aspects such as stability. The large database of antibody sequences available today (abys data base) is a further asset to identify points of intervention. Based on the increasing insight into the fundamentals of antibody structure and folding, progress in the areas described above is envisioned for the near future with profound implications for antibodies as agents to prevent and treat disease that can make use of the potent immune defense mechanisms already present in humans.

P

407 408

M.J. Feige, J. Buchner / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

D

6

Please cite this article as: M.J. Feige, J. Buchner, Principles and engineering of antibody folding and assembly, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbapap.2014.06.004

477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562

M.J. Feige, J. Buchner / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

N C O

R

R

E

C

D

P

R O

O

F

[84] M.J. Feige, S. Groscurth, M. Marcinowski, Z.T. Yew, V. Truffault, E. Paci, H. Kessler, J. Buchner, The structure of a folding intermediate provides insight into differences in immunoglobulin amyloidogenicity, Proc. Natl. Acad. Sci. U. S. A. 105 (2008) 13373–13378. [85] R.P. Jakob, B.K. Zierer, U. Weininger, S.D. Hofmann, S.H. Lorenz, J. Balbach, H. Dobbek, F.X. Schmid, Elimination of a cis-proline-containing loop and turn optimization stabilizes a protein and accelerates its folding, J. Mol. Biol. 399 (2010) 331–346. [86] M.J. Feige, S. Walter, J. Buchner, Folding mechanism of the CH2 antibody domain, J. Mol. Biol. 344 (2004) 107–118. [87] G. Zoldak, T. Aumuller, C. Lucke, J. Hritz, C. Oostenbrink, G. Fischer, F.X. Schmid, A library of fluorescent peptides for exploring the substrate specificities of prolyl isomerases, Biochemistry 48 (2009) 10423–10436. [88] H.Y. Meng, K.M. Thomas, A.E. Lee, N.J. Zondlo, Effects of i and i + 3 residue identity on cis-trans isomerism of the aromatic(i + 1)-prolyl(i + 2) amide bond: implications for type VI beta-turn formation, Biopolymers 84 (2006) 192–204. [89] W.J. Wu, D.P. Raleigh, Local control of peptide conformation: stabilization of cis proline peptide bonds by aromatic proline interactions, Biopolymers 45 (1998) 381–394. [90] R. Aurora, T.P. Creamer, R. Srinivasan, G.D. Rose, Local interactions in protein folding: lessons from the alpha-helix, J. Biol. Chem. 272 (1997) 1413–1416. [91] V. Munoz, L. Serrano, Helix design, prediction and stability, Curr. Opin. Biotechnol. 6 (1995) 382–386. [92] T. Eichner, S.E. Radford, Understanding the complex mechanisms of beta2microglobulin amyloid assembly, FEBS J. 278 (2011) 3868–3883. [93] T. Gidalevitz, F. Stevens, Y. Argon, Orchestration of secretory protein folding by ER chaperones, Biochim. Biophys. Acta 1833 (2013) 2410–2424. [94] I. Braakman, N.J. Bulleid, Protein folding and modification in the mammalian endoplasmic reticulum, Annu. Rev. Biochem. 80 (2011) 71–99. [95] S.S. Vembar, J.L. Brodsky, One step at a time: endoplasmic reticulum-associated degradation, Nat. Rev. Mol. Cell Biol. 9 (2008) 944–957. [96] R. Sitia, M. Neuberger, C. Alberini, P. Bet, A. Fra, C. Valetti, G. Williams, C. Milstein, Developmental regulation of IgM secretion: the role of the carboxy-terminal cysteine, Cell 60 (1990) 781–790. [97] T. Anelli, M. Alessio, A. Bachi, L. Bergamelli, G. Bertoli, S. Camerini, A. Mezghrani, E. Ruffato, T. Simmen, R. Sitia, Thiol-mediated protein retention in the endoplasmic reticulum: the role of ERp44, EMBO J. 22 (2003) 5015–5022. [98] T. Anelli, M. Alessio, A. Mezghrani, T. Simmen, F. Talamo, A. Bachi, R. Sitia, ERp44, a novel endoplasmic reticulum folding assistant of the thioredoxin family, EMBO J. 21 (2002) 835–844. [99] L. Wang, S. Vavassori, S. Li, H. Ke, T. Anelli, M. Degano, R. Ronzoni, R. Sitia, F. Sun, C.C. Wang, Crystal structure of human ERp44 shows a dynamic functional modulation by its carboxy-terminal tail, EMBO Rep. 9 (2008) 642–647. [100] S. Vavassori, M. Cortini, S. Masui, S. Sannino, T. Anelli, I.R. Caserta, C. Fagioli, M.F. Mossuto, A. Fornili, E. van Anken, M. Degano, K. Inaba, R. Sitia, A pH-regulated quality control cycle for surveillance of secretory protein assembly, Mol. Cell 50 (2013) 783–792. [101] L.M. Hendershot, M.J. Feige, J. Buchner, Acidification activates ERp44—a molecular litmus test for protein assembly, Mol. Cell 50 (2013) 779–781. [102] M. Vanhove, Y.K. Usherwood, L.M. Hendershot, Unassembled Ig heavy chains do not cycle from BiP in vivo but require light chains to trigger their release, Immunity 15 (2001) 105–114. [103] L.M. Hendershot, Immunoglobulin heavy chain and binding protein complexes are dissociated in vivo by light chain addition, J. Cell Biol. 111 (1990) 829–837. [104] Y.K. Lee, J.W. Brewer, R. Hellman, L.M. Hendershot, BiP and immunoglobulin light chain cooperate to control the folding of heavy chain and ensure the fidelity of immunoglobulin assembly, Mol. Biol. Cell 10 (1999) 2209–2219. [105] J. Behnke, L.M. Hendershot, The large hsp70 grp170 binds to unfolded protein substrates in vivo with a regulation distinct from conventional hsp70s, J. Biol. Chem. 289 (2014) 2899–2907. [106] Y. Shen, L.M. Hendershot, ERdj3, a stress-inducible endoplasmic reticulum DnaJ homologue, serves as a cofactor for BiP's interactions with unfolded substrates, Mol. Biol. Cell 16 (2005) 40–50. [107] D. Groff, S. Armstrong, P.J. Rivers, J. Zhang, J. Yang, E. Green, J. Rozzelle, S. Liang, J.D. Kittle, A.R. Steiner, R. Baliga, C.D. Thanos, T.J. Hallam, A.K. Sato, A.Y. Yam, Engineering toward a bacterial “endoplasmic reticulum” for the rapid expression of immunoglobulin proteins, MAbs 6 (2014). [108] J.N. Arnold, M.R. Wormald, R.B. Sim, P.M. Rudd, R.A. Dwek, The impact of glycosylation on the biological function and structure of human immunoglobulins, Annu. Rev. Immunol. 25 (2007) 21–50. [109] M. Dalziel, M. Crispin, C.N. Scanlan, N. Zitzmann, R.A. Dwek, Emerging principles for the therapeutic exploitation of glycosylation, Science 343 (2014) 1235681. [110] R. Baumal, M. Potter, M.D. Scharff, Synthesis, assembly, and secretion of gamma globulin by mouse myeloma cells. 3. Assembly of the three subclasses of IgG, J. Exp. Med. 134 (1971) 1316–1334. [111] J. Buxbaum, M.D. Scharff, The synthesis, assembly, and secretion of gamma globulin by mouse myeloma cells. VI. Assembly of IgM proteins, J. Exp. Med. 138 (1973) 278–288. [112] H. Karasuyama, A. Kudo, F. Melchers, The proteins encoded by the VpreB and lambda 5 pre-B cell-specific genes can associate with each other and with mu heavy chain, J. Exp. Med. 172 (1990) 969–972. [113] A. Kudo, F. Melchers, A second gene, VpreB in the lambda 5 locus of the mouse, which appears to be selectively expressed in pre-B lymphocytes, EMBO J. 6 (1987) 2267–2272. [114] S. Pillai, D. Baltimore, Formation of disulphide-linked mu 2 omega 2 tetramers in pre-B cells by the 18 K omega-immunoglobulin light chain, Nature 329 (1987) 172–174.

E

T

[52] I.G. Haas, M. Wabl, Immunoglobulin heavy chain binding protein, Nature 306 (1983) 387–389. [53] L. Hendershot, D. Bole, G. Kohler, J.F. Kearney, Assembly and secretion of heavy chains that do not associate posttranslationally with immunoglobulin heavy chain-binding protein, J. Cell Biol. 104 (1987) 761–767. [54] G. Wozniak-Knopp, F. Ruker, A C-terminal interdomain disulfide bond significantly stabilizes the Fc fragment of IgG, Arch. Biochem. Biophys. 526 (2012) 181–187. [55] G. Wozniak-Knopp, J. Stadlmann, F. Ruker, Stabilisation of the Fc fragment of human IgG1 by engineered intradomain disulfide bonds, PLoS ONE 7 (2012) e30083. [56] R. Gong, B.K. Vu, Y. Feng, D.A. Prieto, M.A. Dyba, J.D. Walsh, P. Prabakaran, T.D. Veenstra, S.G. Tarasov, R. Ishima, D.S. Dimitrov, Engineered human antibody constant domains with increased stability, J. Biol. Chem. 284 (2009) 14203–14210. [57] D.Y. Kim, R. To, H. Kandalaft, W. Ding, H. van Faassen, Y. Luo, J.D. Schrag, N. StAmant, M. Hefford, T. Hirama, J.F. Kelly, R. MacKenzie, J. Tanha, Antibody light chain variable domains and their biophysically improved versions for human immunotherapy, MAbs 6 (2014) 219–235. [58] T. Ying, W. Chen, Y. Feng, Y. Wang, R. Gong, D.S. Dimitrov, Engineered soluble monomeric IgG1 CH3 domain: generation, mechanisms of function, and implications for design of biological therapeutics, J. Biol. Chem. 288 (2013) 25154–25164. [59] H. Liu, K. May, Disulfide bond structures of IgG molecules: structural variations, chemical modifications and possible impacts to stability and biological function, MAbs 4 (2012) 17–23. [60] A. Honegger, Engineering antibodies for stability and efficient folding, Handb. Exp. Pharmacol. (2008) 47–68. [61] Y. Hagihara, D. Saerens, Improvement of single domain antibody stability by disulfide bond introduction, Methods Mol. Biol. 911 (2012) 399–416. [62] E.E. Weatherill, K.L. Cain, S.P. Heywood, J.E. Compson, J.T. Heads, R. Adams, D.P. Humphreys, Towards a universal disulphide stabilised single chain Fv format: importance of interchain disulphide bond location and vL-vH orientation, Protein Eng. Des. Sel. 25 (2012) 321–329. [63] J.S. Huston, D. Levinson, M. Mudgett-Hunter, M.S. Tai, J. Novotny, M.N. Margolies, R.J. Ridge, R.E. Bruccoleri, E. Haber, R. Crea, et al., Protein engineering of antibody binding sites: recovery of specific activity in an anti-digoxin single-chain Fv analogue produced in Escherichia coli, Proc. Natl. Acad. Sci. U. S. A. 85 (1988) 5879–5883. [64] R.E. Bird, K.D. Hardman, J.W. Jacobson, S. Johnson, B.M. Kaufman, S.M. Lee, T. Lee, S.H. Pope, G.S. Riordan, M. Whitlow, Single-chain antigen-binding proteins, Science 242 (1988) 423–426. [65] R. Glockshuber, M. Malia, I. Pfitzinger, A. Pluckthun, A comparison of strategies to stabilize immunoglobulin Fv-fragments, Biochemistry 29 (1990) 1362–1367. [66] N.M. Young, C.R. MacKenzie, S.A. Narang, R.P. Oomen, J.E. Baenziger, Thermal stabilization of a single-chain Fv antibody fragment by introduction of a disulphide bond, FEBS Lett. 377 (1995) 135–139. [67] U. Brinkmann, Y. Reiter, S.H. Jung, B. Lee, I. Pastan, A recombinant immunotoxin containing a disulfide-stabilized Fv fragment, Proc. Natl. Acad. Sci. U. S. A. 90 (1993) 7538–7542. [68] A. Worn, A. Pluckthun, Stability engineering of antibody single-chain Fv fragments, J. Mol. Biol. 305 (2001) 989–1010. [69] T. Wang, Y. Duan, Probing the stability-limiting regions of an antibody single-chain variable fragment: a molecular dynamics simulation study, Protein Eng. Des. Sel. 24 (2011) 649–657. [70] D. Kuroda, H. Shirai, M.P. Jacobson, H. Nakamura, Computer-aided antibody design, Protein Eng. Des. Sel. 25 (2012) 507–521. [71] W. Hoyer, K. Ramm, A. Pluckthun, A kinetic trap is an intrinsic feature in the folding pathway of single-chain Fv fragments, Biophys. Chem. 96 (2002) 273–284. [72] R.L. Baldwin, The search for folding intermediates and the mechanism of protein folding, Annu. Rev. Biophys. 37 (2008) 1–21. [73] J.F. Brandts, H.R. Halvorson, M. Brennan, Consideration of the possibility that the slow step in protein denaturation reactions is due to cis-trans isomerism of proline residues, Biochemistry 14 (1975) 4953–4963. [74] J.F. Brandts, M. Brennan, L. Lung-Nan, Unfolding and refolding occur much faster for a proline-free proteins than for most proline-containing proteins, Proc. Natl. Acad. Sci. U. S. A. 74 (1977) 4178–4181. [75] T. Kiefhaber, Protein folding kinetics, Methods Mol. Biol. 40 (1995) 313–341. [76] Y. Goto, K. Hamaguchi, Unfolding and refolding of the constant fragment of the immunoglobulin light chain, J. Mol. Biol. 156 (1982) 891–910. [77] A. Steward, S. Adhya, J. Clarke, Sequence conservation in Ig-like domains: the role of highly conserved proline residues in the fibronectin type III superfamily, J. Mol. Biol. 318 (2002) 935–940. [78] G. Fischer, T. Tradler, T. Zarnt, The mode of action of peptidyl prolyl cis/trans isomerases in vivo: binding vs. catalysis, FEBS Lett. 426 (1998) 17–20. [79] F.X. Schmid, L.M. Mayr, M. Mucke, E.R. Schonbrunner, Prolyl isomerases: role in protein folding, Adv. Protein Chem. 44 (1993) 25–66. [80] M.J. Feige, S. Groscurth, M. Marcinowski, Y. Shimizu, H. Kessler, L.M. Hendershot, J. Buchner, An unfolded CH1 domain controls the assembly and secretion of IgG antibodies, Mol. Cell 34 (2009) 569–579. [81] J. Lee, T.G. Choi, J. Ha, S.S. Kim, Cyclosporine A suppresses immunoglobulin G biosynthesis via inhibition of cyclophilin B in murine hybridomas and B cells, Int. Immunopharmacol. 12 (2012) 42–49. [82] R. Bernasconi, T. Solda, C. Galli, T. Pertel, J. Luban, M. Molinari, Cyclosporine Asensitive, cyclophilin B-dependent endoplasmic reticulum-associated degradation, PLoS ONE 5 (2010). [83] M.J. Feige, M.A. Grawert, M. Marcinowski, J. Hennig, J. Behnke, D. Auslander, E.M. Herold, J. Peschek, C.D. Castro, M. Flajnik, L.M. Hendershot, M. Sattler, M. Groll, J. Buchner, The structural analysis of shark IgNAR antibodies reveals evolutionary principles of immunoglobulins, Proc. Natl. Acad. Sci. U. S. A. (2014).

U

563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 Q3

7

Please cite this article as: M.J. Feige, J. Buchner, Principles and engineering of antibody folding and assembly, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbapap.2014.06.004

649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734

D

P

R O

O

F

[141] J.M. Perez, J.G. Renisio, J.J. Prompers, C.J. van Platerink, C. Cambillau, H. Darbon, L.G. Frenken, Thermal unfolding of a llama antibody fragment: a two-state reversible process, Biochemistry 40 (2001) 74–83. [142] D. Christ, K. Famm, G. Winter, Repertoires of aggregation-resistant human antibody domains, Protein Eng. Des. Sel. 20 (2007) 413–416. [143] L. Jespers, O. Schon, K. Famm, G. Winter, Aggregation-resistant domain antibodies selected on phage by heat denaturation, Nat. Biotechnol. 22 (2004) 1161–1165. [144] K. Famm, L. Hansen, D. Christ, G. Winter, Thermodynamically stable aggregationresistant antibody domains through directed evolution, J. Mol. Biol. 376 (2008) 926–931. [145] P.A. Barthelemy, H. Raab, B.A. Appleton, C.J. Bond, P. Wu, C. Wiesmann, S.S. Sidhu, Comprehensive analysis of the factors contributing to the stability and solubility of autonomous human VH domains, J. Biol. Chem. 283 (2008) 3639–3654. [146] M. Dumoulin, K. Conrath, A. Van Meirhaeghe, F. Meersman, K. Heremans, L.G. Frenken, S. Muyldermans, L. Wyns, A. Matagne, Single-domain antibody fragments with high conformational stability, Protein Sci. 11 (2002) 500–515. [147] L. Jespers, O. Schon, L.C. James, D. Veprintsev, G. Winter, Crystal structure of HEL4, a soluble, refoldable human V(H) single domain with a germ-line scaffold, J. Mol. Biol. 337 (2004) 893–903. [148] S.D. Nuttall, R.B. Walsh, Display scaffolds: protein engineering for novel therapeutics, Curr. Opin. Pharmacol. 8 (2008) 609–615. [149] M. Lauwereys, M. Arbabi Ghahroudi, A. Desmyter, J. Kinne, W. Holzer, E. De Genst, L. Wyns, S. Muyldermans, Potent enzyme inhibitors derived from dromedary heavy-chain antibodies, EMBO J. 17 (1998) 3512–3520. [150] J. Govaert, M. Pellis, N. Deschacht, C. Vincke, K. Conrath, S. Muyldermans, D. Saerens, Dual beneficial effect of interloop disulfide bond for single domain antibody fragments, J. Biol. Chem. 287 (2012) 1970–1979. [151] E. De Genst, K. Silence, K. Decanniere, K. Conrath, R. Loris, J. Kinne, S. Muyldermans, L. Wyns, Molecular basis for the preferential cleft recognition by dromedary heavychain antibodies, Proc. Natl. Acad. Sci. U. S. A. 103 (2006) 4586–4591. [152] F. Wang, D.C. Ekiert, I. Ahmad, W. Yu, Y. Zhang, O. Bazirgan, A. Torkamani, T. Raudsepp, W. Mwangi, M.F. Criscitiello, I.A. Wilson, P.G. Schultz, V.V. Smider, Reshaping antibody diversity, Cell 153 (2013) 1379–1393. [153] J. Johansson, M. Aveskogh, B. Munday, L. Hellman, Heavy chain V region diversity in the duck-billed platypus (Ornithorhynchus anatinus): long and highly variable complementarity-determining region 3 compensates for limited germline diversity, J. Immunol. 168 (2002) 5155–5162. [154] H. Dooley, M.F. Flajnik, Antibody repertoire development in cartilaginous fish, Dev. Comp. Immunol. 30 (2006) 43–56. [155] D. Frommel, G.W. Litman, J. Finstad, R.A. Good, The evolution of the immune response. XI. The immunoglobulins of the horned shark, Heterodontus francisci: purification, characterization and structural requirement for antibody activity, J. Immunol. 106 (1971) 1234–1243. [156] D. Saerens, G.H. Ghassabeh, S. Muyldermans, Single-domain antibodies as building blocks for novel therapeutics, Curr. Opin. Pharmacol. 8 (2008) 600–608. [157] P.H. Yancey, G.N. Somero, Counteraction of urea destabilization of protein structure by methylamine osmoregulatory compounds of elasmobranch fishes, Biochem. J. 183 (1979) 317–323. [158] L.M. Blancas-Mejia, M. Ramirez-Alvarado, Systemic amyloidoses, Annu. Rev. Biochem. 82 (2013) 745–774. [159] J. Buxbaum, G. Gallo, Nonamyloidotic monoclonal immunoglobulin deposition disease. Light-chain, heavy-chain, and light- and heavy-chain deposition diseases, Hematol. Oncol. Clin. North Am. 13 (1999) 1235–1248. [160] M.R. Hurle, L.R. Helms, L. Li, W. Chan, R. Wetzel, A role for destabilizing amino acid replacements in light-chain amyloidosis, Proc. Natl. Acad. Sci. U. S. A. 91 (1994) 5446–5450. [161] Y. Shimizu, L. Meunier, L.M. Hendershot, pERp1 is significantly up-regulated during plasma cell differentiation and contributes to the oxidative folding of immunoglobulin, Proc. Natl. Acad. Sci. U. S. A. 106 (2009) 17013–17018. [162] E. van Anken, F. Pena, N. Hafkemeijer, C. Christis, E.P. Romijn, U. Grauschopf, V.M. Oorschot, T. Pertel, S. Engels, A. Ora, V. Lastun, R. Glockshuber, J. Klumperman, A.J. Heck, J. Luban, I. Braakman, Efficient IgM assembly and secretion require the plasma cell induced endoplasmic reticulum protein pERp1, Proc. Natl. Acad. Sci. U. S. A. 106 (2009) 17019–17024. [163] K. Leitzgen, M.R. Knittler, I.G. Haas, Assembly of immunoglobulin light chains as a prerequisite for secretion. A model for oligomerization-dependent subunit folding, J. Biol. Chem. 272 (1997) 3117–3123. [164] Z. Hamrang, N.J. Rattray, A. Pluen, Proteins behaving badly: emerging technologies in profiling biopharmaceutical aggregation, Trends Biotechnol. 31 (2013) 448–458. [165] R. Rouet, D. Lowe, D. Christ, Stability engineering of the human antibody repertoire, FEBS Lett. 588 (2014) 269–277. [166] M.R. Knittler, I.G. Haas, Interaction of BiP with newly synthesized immunoglobulin light chain molecules: cycles of sequential binding and release, EMBO J. 11 (1992) 1573–1581. [167] R. Hellman, M. Vanhove, A. Lejeune, F.J. Stevens, L.M. Hendershot, The in vivo association of BiP with newly synthesized proteins is dependent on the rate and stability of folding and not simply on the presence of sequences that can bind to BiP, J. Cell Biol. 144 (1999) 21–30.

N

C

O

R

R

E

C

T

[115] N. Sakaguchi, F. Melchers, Lambda 5, a new light-chain-related locus selectively expressed in pre-B lymphocytes, Nature 324 (1986) 579–582. [116] T. Tsubata, M. Reth, The products of pre-B cell-specific genes (lambda 5 and VpreB) and the immunoglobulin mu chain form a complex that is transported onto the cell surface, J. Exp. Med. 172 (1990) 973–976. [117] Y. Minegishi, L.M. Hendershot, M.E. Conley, Novel mechanisms control the folding and assembly of lambda5/14.1 and VpreB to produce an intact surrogate light chain, Proc. Natl. Acad. Sci. U. S. A. 96 (1999) 3041–3046. [118] A.J. Bankovich, S. Raunser, Z.S. Juo, T. Walz, M.M. Davis, K.C. Garcia, Structural insight into pre-B cell receptor function, Science 316 (2007) 291–294. [119] C. Vettermann, K. Herrmann, C. Albert, E. Roth, M.R. Bosl, H.M. Jack, A unique role for the lambda5 nonimmunoglobulin tail in early B lymphocyte development, J. Immunol. 181 (2008) 3232–3242. [120] S. Herzog, M. Reth, H. Jumaa, Regulation of B-cell proliferation and differentiation by pre-B-cell receptor signalling, Nat. Rev. Immunol. 9 (2009) 195–205. [121] M.F. Flajnik, N. Deschacht, S. Muyldermans, A case of convergence: why did a simple alternative to canonical antibodies arise in sharks and camels? PLoS Biol. 9 (2011) e1001120. [122] C. Hamers-Casterman, T. Atarhouch, S. Muyldermans, G. Robinson, C. Hamers, E.B. Songa, N. Bendahman, R. Hamers, Naturally occurring antibodies devoid of light chains, Nature 363 (1993) 446–448. [123] A.S. Greenberg, D. Avila, M. Hughes, A. Hughes, E.C. McKinney, M.F. Flajnik, A new antigen receptor gene family that undergoes rearrangement and extensive somatic diversification in sharks, Nature 374 (1995) 168–173. [124] V.K. Nguyen, R. Hamers, L. Wyns, S. Muyldermans, Loss of splice consensus signal is responsible for the removal of the entire C(H)1 domain of the functional camel IGG2A heavy-chain antibodies, Mol. Immunol. 36 (1999) 515–524. [125] S. Muyldermans, Nanobodies: natural single-domain antibodies, Annu. Rev. Biochem. 82 (2013) 775–797. [126] M.R. Muller, R. O'Dwyer, M. Kovaleva, F. Rudkin, H. Dooley, C.J. Barelle, Generation and isolation of target-specific single-domain antibodies from shark immune repertoires, Methods Mol. Biol. 907 (2012) 177–194. [127] S. Zielonka, N. Weber, S. Becker, A. Doerner, A. Christmann, C. Christmann, C. Uth, J. Fritz, E. Schafer, B. Steinmann, M. Empting, P. Ockelmann, M. Lierz, H. Kolmar, Shark Attack: High affinity binding proteins derived from shark vNAR domains by stepwise in vitro affinity maturation, J. Biotechnol. (2014). [128] T.N. Baral, S. Magez, B. Stijlemans, K. Conrath, B. Vanhollebeke, E. Pays, S. Muyldermans, P. De Baetselier, Experimental therapy of African trypanosomiasis with a nanobody-conjugated human trypanolytic factor, Nat. Med. 12 (2006) 580–584. [129] V. Cortez-Retamozo, N. Backmann, P.D. Senter, U. Wernery, P. De Baetselier, S. Muyldermans, H. Revets, Efficient cancer therapy with a nanobody-based conjugate, Cancer Res. 64 (2004) 2853–2857. [130] S. Spinelli, L. Frenken, D. Bourgeois, L. de Ron, W. Bos, T. Verrips, C. Anguille, C. Cambillau, M. Tegoni, The crystal structure of a llama heavy chain variable domain, Nat. Struct. Biol. 3 (1996) 752–757. [131] A. Desmyter, T.R. Transue, M.A. Ghahroudi, M.H. Thi, F. Poortmans, R. Hamers, S. Muyldermans, L. Wyns, Crystal structure of a camel single-domain VH antibody fragment in complex with lysozyme, Nat. Struct. Biol. 3 (1996) 803–811. [132] R.L. Stanfield, H. Dooley, M.F. Flajnik, I.A. Wilson, Crystal structure of a shark singledomain antibody V region in complex with lysozyme, Science 305 (2004) 1770–1773. [133] V.A. Streltsov, J.N. Varghese, J.A. Carmichael, R.A. Irving, P.J. Hudson, S.D. Nuttall, Structural evidence for evolution of shark Ig new antigen receptor variable domain antibodies from a cell-surface receptor, Proc. Natl. Acad. Sci. U. S. A. 101 (2004) 12444–12449. [134] M.M. Harmsen, R.C. Ruuls, I.J. Nijman, T.A. Niewold, L.G. Frenken, B. de Geus, Llama heavy-chain V regions consist of at least four distinct subfamilies revealing novel sequence features, Mol. Immunol. 37 (2000) 579–590. [135] V.K. Nguyen, R. Hamers, L. Wyns, S. Muyldermans, Camel heavy-chain antibodies: diverse germline V(H)H and specific mechanisms enlarge the antigen-binding repertoire, EMBO J. 19 (2000) 921–930. [136] O.V. Kovalenko, A. Olland, N. Piche-Nicholas, A. Godbole, D. King, K. Svenson, V. Calabro, M.R. Muller, C.J. Barelle, W. Somers, D.S. Gill, L. Mosyak, L. Tchistiakova, Atypical antigen recognition mode of a shark immunoglobulin new antigen receptor (IgNAR) variable domain characterized by humanization and structural analysis, J. Biol. Chem. 288 (2013) 17408–17419. [137] J. Davies, L. Riechmann, ‘Camelising’ human antibody fragments: NMR studies on VH domains, FEBS Lett. 339 (1994) 285–290. [138] J. Davies, L. Riechmann, Single antibody domains as small recognition units: design and in vitro antigen selection of camelized, human VH domains with improved protein stability, Protein Eng. 9 (1996) 531–537. [139] S. Ewert, C. Cambillau, K. Conrath, A. Pluckthun, Biophysical properties of camelid V(HH) domains compared to those of human V(H)3 domains, Biochemistry 41 (2002) 3628–3636. [140] R.H. van der Linden, L.G. Frenken, B. de Geus, M.M. Harmsen, R.C. Ruuls, W. Stok, L. de Ron, S. Wilson, P. Davis, C.T. Verrips, Comparison of physical chemical properties of llama VHH antibody fragments and mouse monoclonal antibodies, Biochim. Biophys. Acta 1431 (1999) 37–46.

U

735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768 769 770 Q4 771 772 773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792 793 794 795 796 797 798 799 800 801 802 803 804 805 806 807 808 809 810 811 812 813 892

M.J. Feige, J. Buchner / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

E

8

Please cite this article as: M.J. Feige, J. Buchner, Principles and engineering of antibody folding and assembly, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbapap.2014.06.004

814 815 816 817 818 819 820 821 822 823 824 825 826 827 828 829 830 831 832 833 834 835 836 837 838 839 840 841 842 843 844 845 846 847 848 849 850 851 852 853 854 855 856 857 858 859 860 861 862 863 864 865 866 867 868 869 870 871 872 873 874 875 876 877 878 879 880 881 882 883 884 885 886 887 888 889 890 891

Principles and engineering of antibody folding and assembly.

Antibodies are uniquely suited to serve essential roles in the human immune defense as they combine several specific functions in one hetero-oligomeri...
2MB Sizes 2 Downloads 3 Views