Tissue transglutaminase in Alzheimer\'s disease: involvement in pathogenesis and its potential as a therapeutic target

Share Embed


Descripción

1

Journal of Alzheimer’s Disease xx (20xx) x–xx DOI 10.3233/JAD-132492 IOS Press

3

4

oo f

Tissue Transglutaminase in Alzheimer’s Disease: Involvement in Pathogenesis and its Potential as a Therapeutic Target

Pr

2

Review

or

1

5

Micha M.M. Wilhelmusa,∗ , Mieke de Jagera , Erik N.T.P. Bakkerb and Benjamin Drukarcha

6

a Department

8

of Anatomy and Neurosciences, VU University Medical Center, Neuroscience Campus Amsterdam, Amsterdam, The Netherlands b Department of Biomedical Engineering and Physics, Academic Medical Center, Amsterdam, The Netherlands

uth

7

13 14 15 16 17 18 19 20 21

cte

12

rre

11

Abstract. Protein misfolding and the formation of stable insoluble protein complexes by self-interacting proteins, in particular amyloid-␤ and tau protein, play a central role in the pathogenesis of Alzheimer’s disease (AD). Unfortunately, the underlying mechanisms that trigger the misfolding of self-interacting proteins that eventually results in formation of neurotoxic dimers, oligomers, and aggregates remain unclear. Elucidation of the driving forces of protein complex formation in AD is of crucial importance for the development of disease-modifying therapies. Tissue transglutaminase (tTG) is a calcium-dependent enzyme that induces the formation of covalent ␧-(␥-glutamyl)lysine isopeptide bonds, which results in both intra- and intermolecular protein cross-links. These tTG-catalyzed intermolecular cross-links induce stable, rigid, and insoluble protein complexes, whereas intramolecular cross-links change the conformation of proteins. Inhibition of tTG-catalyzed cross-linking counteracts the formation of protein aggregates, as observed in disease-models of other protein misfolding diseases, in particular Parkinson’s and Huntington’s diseases. Although data of tTG activity in AD models is limited, there is compelling evidence from both in vitro and postmortem human brain tissue of AD patients that point toward a crucial role for tTG in the pathogenesis of AD. Here, we review these data on the role of tTG in the initiation and development of protein aggregates in AD, and discuss the possibility to use inhibitors of the cross-linking activity of tTG as a new therapeutic approach for AD.

co

9 10

dA

Accepted 23 February 2014

Keywords: Alzheimer’s disease, amyloid-␤, cerebral amyloid angiopathy, crosslinking, neurofibrillary tangles, senile plaques, therapy, tissue transglutaminase

24

INTRODUCTION

25 26 27 28

Un

23

22

Alzheimer’s disease (AD) is a devastating neurodegenerative disease and the most common form of dementia. Although a great leap has been made in identifying genes involved in early-onset AD and ∗ Correspondence

to: Micha M.M. Wilhelmus, Department of Anatomy and Neurosciences, VU University Medical Center, Neuroscience Campus Amsterdam, Van der Boechorststraat 7, 1081 BT Amsterdam, The Netherlands. Tel.: +31 20 4448103; Fax: +31 20 4448100; E-mail: [email protected].

unravelling the biological mechanisms of AD, only limited progress has been made in the development of truly disease-modifying treatments. It goes without saying that novel therapeutic targets and agents are of crucial importance for the AD field. Almost all current trials are based on the removal and/or prevention of accumulation of neurotoxic protein multimers and aggregates in AD. Unfortunately, however, the mechanisms that drive the formation of these multimers and aggregates remain largely unknown. Senile plaques (SPs), neurofibrillary tangles (NFTs), and cerebral

ISSN 1387-2877/14/$27.50 © 2014 – IOS Press and the authors. All rights reserved

29 30 31 32 33 34 35 36 37 38 39

48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91

oo f

47

Pr

46

or

45

treatment with the competitive tTG-inhibitor, cystamine, inhibition of ␣-synuclein aggregate formation was observed [9]. In cell models of another proteininclusion neurodegenerative disease, i.e., Huntington’s disease, similar effects were observed upon tTG inhibition [10, 11]. In addition, in Huntington’s disease mouse models, cystamine treatment resulted in improved motor function and increased survival compared to non-treated mice [12, 13]. Together, these data strongly support the notion that tTG cross-linking activity is directly linked to protein oligomerization and aggregation, and that inhibition of tTG activity might be a potential therapeutic target to counteract the formation of neurotoxic oligomers and aggregates in these neurodegenerative diseases. In addition to the findings in Parkinson’s and Huntington’s diseases, the potential importance of cross-links induced by tTG in AD pathogenesis was already noted in the early nineteen eighties [14]. Since then, evidence for a role of tTG in the pathogenesis of AD has been mounting [15, 16]. Surprisingly, however, despite the growing body of evidence that supports the idea that tTG plays a key role in protein aggregation in AD, unlike Parkinson’s and Huntington’s diseases, studies on the effects of inhibition of tTG in AD models, either cellular or animal, are very limited. This review intends to provide an overview of the current knowledge on the role of tTG in the initiation and development of neurotoxic oligomers and aggregates in AD, and discusses the possibility to use inhibitors of tTG cross-linking activity as therapeutic agents.

dA

44

cte

43

rre

42

amyloid angiopathy (CAA) [1] are the main pathological brain lesions defining AD. Both SPs and CAA primarily comprise extracellular deposition of aggregated amyloid-␤ (A␤) protein, a 4 kDa proteolytic cleavage product of the amyloid-␤ protein precursor. In SPs, neurotoxic A␤ oligomers and aggregates are formed and deposited in the brain parenchyma, whereas in CAA, A␤ accumulates in the outer portion of the media of brain vessels, leading to degeneration of both cerebrovascular cells and ultimately neurons. NFTs are intraneuronal inclusions in the cytoplasm of the neuron composed of paired helical filaments. These filaments consist of hyperphosphorylated bundles of the microtubule-binding protein tau [2], which is involved in microtubule stabilization. For reasons not well-understood, the tau protein is hyperphosphorylated in neurons in AD, resulting in abnormal self-polymerization and aggregation of tau, leading to disturbances in microtubule dynamics [3, 4]. In this manner, hyperphosphorylated tau aggregation leads to destabilization and disruption of cytoskeleton integrity and thus, cell function and viability [5]. Despite the well-characterized pathology underlying AD, current symptomatic therapies and recent trials designed to counteract and/or remove aggregate formation and deposition in AD patients have met with little success. Tissue transglutaminase (tTG) is a member of the transglutaminase protein family (EC 2.3.2.13) of which there are nine encoded in the human genome. tTG is a calcium-dependent enzyme that catalyzes several reactions, including the formation of (␥-glutamyl)polyamine bonds [6], the deamidation of protein substrates [7], and the formation of covalent ␧-(␥-glutamyl)lysine isopeptide bonds, which result in both intra- and intermolecular protein cross-links (Fig. 1). tTG-catalyzed intermolecular cross-links induce stable, rigid, and insoluble protein complexes [8], whereas intramolecular cross-links change the conformation of proteins (Fig. 1). It is the formation of stable protein complexes which are induced by self-interacting proteins that play a central role in the pathogenesis of a number of neurodegenerative diseases. Thus, inhibition of tTG-catalyzed cross-linking of proteins involved in neurodegenerative diseases might be a potential therapeutic strategy to counteract the disease process. Interestingly, in Parkinson’s disease models, in which the ␣-synuclein protein aggregates and forms neurotoxic inclusions, inhibition of tTG showed a direct inhibitory effect on the formation of ␣-synuclein aggregates [9]. Conversely, in cells transfected with both tTG and ␣-synuclein, enzymatic tTG activity induced aggregation of ␣-synuclein. After

co

41

Un

40

M.M.M. Wilhelmus et al. / Tissue Transglutaminase in Alzheimer’s Disease

uth

2

CURRENT AND EMERGING THERAPIES FOR AD The major current therapeutic strategy for AD is centered around the “cholinergic hypothesis” described in 1982 [17], which resulted in the introduction of several cholinesterase inhibitors, i.e., donepezil, rivastigmine, and galantamine, for use in the clinic today. Treatment with cholinesterase inhibitors results in a small but significant improvement in cognitive function and overall clinical condition. In addition to cholinesterase inhibitors, the N-methyl-D-aspartate (NMDA) blocker memantine is used as a therapeutic agent in AD. The underlying basis for this treatment is the alleged excitotoxicity of the neurotransmitter glutamate in AD, a common process underlying neuronal cell death in neurodegenerative diseases [18]. Exocitotoxicity results from overactivation of the NMDA receptor, leading to excessive intracellular influx of calcium

92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122

123 124

125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140

3

M.M.M. Wilhelmus et al. / Tissue Transglutaminase in Alzheimer’s Disease

K

+

K

K

+ tTG

+

Pr

Q

NH3

oo f

Q

+ calcium

Q

or

K

NH3

K

uth

aminoacids glutamine (Q)

+

lysine (K) tTG-catalyzed cross-link

145 146 147 148 149 150 151 152 153 154 155 156 157 158 159 160 161 162 163 164 165 166 167 168 169

Phase III as they caused clinical worsening and other safety concerns [27, 28], or are still in earlier phases of development [29].

cte

144

rre

143

and subsequent neuronal cell death [19]. Preclinical studies have shown that blockade of the NMDA receptor by memantine inhibits A␤-induced neurotoxicity [20–22]. Clinical studies show that memantine reduces the incidence of clinical worsening in key symptomatic domains in moderate to severe AD [23]. Thus, although all above-described therapeutic strategies have some positive symptomatic effects on AD patients, they do not seem to modify the course of the disease and/or alter the rate of progression of neurodegeneration in AD. For this purpose, new therapeutic strategies are required that modify the processes underlying the disease. Within this context, an important category of emerging therapies for AD are the immuno-based therapies that are designed to remove A␤ from the brain, by either passive immunization, administration of intravenous immunoglobulins, or active immunization schemes. Although these therapeutic strategies appeared promising in preclinical studies and early stage clinical trials, so far, none of the antibody-based therapies where able to meet the primary endpoints of improvement in cognition and global functioning in large scale clinical studies [24–26]. Apart from therapies targeting the removal of A␤ from the brain are those that focus on decreasing A␤ production. These therapies are designed to restrain the generation of A␤ by inhibiting ␤- and ␥-secretases that cleave A␤ from its precursor. Unfortunately, however, trials using secretase inhibitors have been prematurely terminated in

co

142

Un

141

dA

Fig. 1. Inter- and intramolecular cross-linking of peptides induced by tissue transglutaminase (tTG) cross-linking activity. The main activity of tTG is to catalyze a calcium-dependent acyl transfer reaction between the ␥-carboxamide group of a polypeptide-bound glutamine and the ␧-amino group of a polypeptide bound lysine residue to form an ␧-(␥-glutamyl)lysine isopeptide bond, also known as a cross-link. This characteristic reaction of tTG results in a posttranslational modification through either the formation of an intermolecular cross-link between two polypeptide chains, or an intramolecular cross-link within a single peptide chain.

THE AMYLOID CASCADE: TARGET FOR THERAPEUTIC INTERVENTION The amyloid cascade hypothesis for AD combines histopathological and genetic information on AD and posits that the deposition of A␤ peptide in the brain parenchyma initiates a sequence of events, including the formation of NFTs, that eventually leads to AD dementia. In the last few years, however, it has become clear that the rate of A␤ accumulation and deposition as SPs has no linear relationship to the development of dementia or other forms of cognitive impairment, in both humans and AD mouse models [30, 31]. Moreover, evidence is mounting that soluble A␤ oligomers, ranging from dimers up to mid range-molecular weight oligomers (e.g., A␤*56), account for the neurotoxicity observed in AD [32, 33]. This notion is supported by findings of McLean and coworkers, who demonstrated the presence of ∼8 and ∼12 kDa soluble SDS-stable A␤ dimers and trimers in AD brain tissue [34]. Elevated levels of SDS-stable A␤ dimers are also observed in AD mouse models and A␤ dimers isolated from AD brain tissue are capable of impairing long-term potentiation [21, 35, 36]. Overall, it is becoming evident that small A␤ oligomers are not only the likely neuro-

170 171 172

173 174

175 176 177 178 179 180 181 182 183 184 185 186 187 188 189 190 191 192 193 194 195 196

205 206

207

208

209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 236 237 238 239 240 241 242 243 244 245

TISSUE TRANSGLUTAMINASE Structure and function of tTG The secondary structure of tTG is composed of four domains (Fig. 2), the domains 1, 3, and 4 are folded in ␤-structures and domain 2 presents a ␣-helical structure [37]. The 13 tryptophan residues of the protein are all located in domains 1 and 2. The structural information concerning tTG and the domain-specific intrinsic spectroscopic probes are useful in studies of protein unfolding by chemical and thermal denaturation [38]. The transamidating activity of the enzyme is located in domain 2 and is regulated by the catalytic site in which Cys277 , His335 and Asp358 play an essential role (Fig. 2). Especially the highly reactive Cys277 in the catalytic core is of importance as it forms thioesters with peptidylglutamine moieties of the protein substrates. For instance, the in vivo toxicity of acrylamide toward the central nervous system is probably caused by the reaction of tTG with acrylamide, which results in inactivation of the enzyme [39]. Therefore, the high reactivity of active-site Cys277 has been used to develop irreversible inhibitors of the enzyme, as discussed later. Upon calcium binding, a conformational change in the protein is induced, creating an open access to the active site, and thus inducing the transamidating activity. The calcium-induced transamidating activity is counteracted by binding of guanosine-5′ -triphosphate to the enzyme. Guanosine-5′ -triphosphate binds at Lys173 and is finally hydrolyzed in a process which also involves serine171 , leading to a reversible, guanosine5′ -triphosphatase-dependent regulatory mechanism. Although the guanosine-5′ -triphosphate pocket is located at a long distance from the Cys277 , the spatial location is very close [40]. Therefore, guanosine-5′ triphosphate-binding can be considered as an allosteric inhibition of the enzyme. Another potential regulatory effect of tTG activity might occur by the interaction of tTG with phospholipids and following nitrosylation of cysteine residues by nitric oxide donors. Binding of a

oo f

204

Pr

203

or

202

relatively minor membrane phospholipid component, lysophosphatidylcholine, to tTG increases the sensitivity of the enzyme to calcium, so that tTG acquires appreciable activity at near physiological levels of calcium [41]. Thus, this interaction broadens the cellular conditions under which tTG may be active. The nitrosylation of the enzyme by nitric oxide releasing agents induces another regulatory effect of tTG [42]. Nitrosylation results in a marked inhibition of activity and an increased sensitivity to the inhibitory effects of guanosine-5′ -triphosphate. Different sets of cysteine residues are nitrated in the absence or presence of calcium, but this modification only takes place in the presence of calcium and is apparently relevant in regulating the transamidating activity. As discussed, tTG is activated by the binding of calcium. This binding induces the acylation of the active site cysteine residue (Cys277 ) by a proteinbound glutamine residue, resulting in the liberation of ammonia and the formation of a thioester intermediate between tTG and the glutamine bearing protein substrate. Accordingly, the thioester intermediate is attacked by a nucleophilic primary amine, either a small molecule amine such as putrescine or the ␧amino group of a protein-bound lysine residues [43]. This results in the formation of a relatively stable isopeptide bond. This reaction can induce a bridge, cross-link, between a lysine donor residue of one protein with an acceptor glutamine residue of another protein, creating a cross-link between two proteins. In contrast, if both a lysine donor and a glutamine acceptor are present within a protein, this cross-link induces an intramolecular bridge [15]. Despite the lack of a leader sequence, tTG is transported to the cell surface in a controlled manner [44, 45]. Nitric oxide appears to play a role herein, at least in several vascular cell types [46]. Interestingly, tTG interacts at the cell surface with low-density lipoprotein receptor-related protein 1, which is a requirement for its endocytosis [47]. As also A␤ is transported by low-density lipoprotein receptor-related protein 1 in cerebrovascular cells [48], a potential interaction between A␤ and tTG can therefore be speculated upon. Work from Zemskov et al. suggests that tTG is removed from the cell surface via endosomes [47]. This constitutive endocytosis of tTG from the cell surface depends on plasma membrane cholesterol, dynamin-2, and involves both clathrin-coated pits and lipid rafts. In the extracellular milieu, a high calcium environment, tTG interacts with a number of extracellular proteins, such as fibronectin [49, 50], collagen [51], laminin [52], vitronectin and fibrinogen [53, 54]. By cross-linking these

dA

201

cte

200

rre

199

toxic species that directly lead to neurodegeneration in AD and in AD models, but that these stable oligomers also form the molecular building blocks for fibrillar A␤ in vivo. Thus, as stable A␤ oligomers appear to be of key importance in the amyloid cascade, the mechanisms leading to their formation may offer novel and attractive drug targets in AD. It is therefore an attractive hypothesis that disease-specific post-translational modification of monomeric A␤ by tTG might underlie the formation of SDS-stable neurotoxic A␤ oligomers.

co

198

Un

197

M.M.M. Wilhelmus et al. / Tissue Transglutaminase in Alzheimer’s Disease

uth

4

246 247 248 249 250 251 252 253 254 255 256 257 258 259 260 261 262 263 264 265 266 267 268 269 270 271 272 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 293 294 295 296 297

5

cte

dA

uth

or

Pr

oo f

M.M.M. Wilhelmus et al. / Tissue Transglutaminase in Alzheimer’s Disease

co

rre

Fig. 2. Structure of tissue transglutaminase (tTG). The secondary structure of tTG is composed of four domains, the domains 1, 3, and 4 are folded in ␤-structures and domain 2 presents an ␣-helical structure. The N-terminal ␤-sandwich strand contains the fibronectin-binding domain. Domain 2 contains the guanosine-5′ -triphosphate-binding region, Ser171 and Lys173 and the calcium-binding region composed of Ser449 , Glu451 and Glu452 . The transamidating activity of the enzyme is located in domain 2 and is regulated by the catalytic site in which Cys277 , His335 and Asp358 play an essential role. Upon calcium binding, a conformational change in the protein is induced, creating an open access to the active site, and thus inducing the transamidating activity. During activation of the protein, the interactions between domains 2, 3, and 4 breakdown, creating a possibility for calcium to bind at the main binding site located domain 2. In addition, the active site becomes accessible upon calcium binding. As a consequence of the calcium binding, the ␣-helix (H4) unfolds resulting in perturbing of the structure of neighboring loop 455-478, which connects domains 2 and 3, and also the spatial location of domains 3 and 4, which move from each other and from domain 2. Guanosine-5′ -triphosphate binds at Lys173 and is finally hydrolyzed in a process which also involves serine171 , leading to a reversible, Guanosine-5′ -triphosphatase-dependent regulatory mechanism.

309

tTG expression in the human brain

299 300 301 302 303 304 305 306 307

310 311

Un

308

matrix proteins, tTG does not only increase the proteolytic, chemical and mechanical resistance of these proteins, but is also thought to facilitate cell adhesion and motility, although cross-linking is just one of the activities of the enzyme that plays a role in cell adhesion [55, 56]. In this manner, the cross-linking function of tTG has been implicated in the maintenance of tissue integrity following damage. In fact, subsequent to tissue damage, both tTG levels and activity increase, and the enzyme might even be released into the matrix to stabilize the matrix via cross-linking.

298

tTG is abundantly expressed in the human brain and is present in neurons in several brain areas includ-

ing the amygdala, corpus callosum, cerebellum, and frontal cortex [57]. More recent work of our group demonstrated that tTG is also expressed in astrocytes and microglial cells in the white matter [58]. In addition, we detected tTG in brain capillaries and other parenchymal vessels. Apart from the presence of tTG itself, we also detected the presence of TG-catalyzed cross-links in glial cells of both white and grey matter, and weak immunoreactivity of the anti-cross-linking antibody in brain vessels and in corpora amylacea [58, 59]. In cell lines and cultured primary brain-derived cells, tTG expression is mainly observed in the cytosolic compartment, but also in cell membranes and extracellular fractions. In addition, in cultured neuroblastoma cells, tTG has also been detected in the

312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327

336 337 338 339 340 341 342 343 344 345 346

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

TTG IN AD

oo f

335

Pr

334

or

333

these data using postmortem brain tissue point toward a role for tTG in SPs formation and/or stabilization [58] (Fig. 3). Next to the compelling in vivo evidence that tTG is associated with SPs in AD, in vitro data provide direct proof that tTG is able to affect the A␤ aggregation cascade. In early studies, both wild-type A␤1-40 and A␤1-40 with the Dutch mutation (Glu22 to Gln) were found to be good substrates for tTG cross-linking [74]. Moreover, tTG-catalyzed cross-linking resulted in the formation of A␤ oligomers [74] (Figs. 3 and 4). Ho and coworkers demonstrated that the A␤PP is a substrate for tTG-catalyzed cross-linking resulting in A␤PP dimers and multimers [75], and Schmid and colleagues showed that tTG induces intramolecular cross-links in A␤ [76]. Intriguingly, tTG-catalyzed intermolecular cross-linking of A␤ induces the formation of stable A␤ oligomers that are resistant towards A␤ degrading enzymes, in particular insulin degrading enzyme and neprilysin [77]. In addition, tTG-catalyzed crosslinking of A␤ lowers its oligomerization threshold for self aggregation, suggesting that tTG is capable of driving the aggregation process of A␤ at physiological A␤ levels (Fig. 3). Furthermore, these tTG-mediated A␤ oligomers and protofibrils are toxic in that they inhibit long term potentiation in the CA1 region of the hippocampus [77]. Earlier work on the effects of tTGcatalyzed cross-linking of A␤ had shown that when tTG activity is blocked in cultured neuroblastoma cells, A␤-induced cell death is reduced, whereas induction of tTG enhances A␤-driven neurotoxicity [78]. Moreover, A␤1-42 treatment of monocytes induces tTG expression in vitro [79]. Together, these data indicate that tTG is a likely candidate responsible for initiating the aggregation process of A␤ in AD brains by the formation of stable A␤ dimers, oligomers, and protofibrils (Fig. 4).

dA

332

tTG in SPs

There is strong evidence that tTG is linked to the pathology of AD. Both tTG levels and TG activity are elevated in the cortex of AD brains, compared to control patients [69]. Significantly elevated tTG levels have been reported in the cerebrospinal fluid of AD patients compared to controls [70], suggesting the potential to use tTG protein levels in the cerebrospinal fluid as a diagnostic marker for AD. Moreover, the level of ␧-(␥-glutamyl)lysine isopeptides was significantly elevated in the cerebrospinal fluid of AD patients [71]. In line with these data are the findings of Wang and colleagues, who described a correlation between accumulation of insoluble proteins containing isopeptide bonds in the gray matter with cognitive impairment in AD patients [72]. In addition to these biochemical data, immunohistochemical studies on postmortem tissue sections have shown that tTG enzyme is present in the A␤ core of SPs [73]. A more recent study of our own group confirmed these data and additionally showed colocalization of the tTG enzyme with both diffuse and classic SPs [58] (Fig. 3). The presence of tTG in diffuse SPs suggests an early involvement of tTG in SP formation, as diffuse SPs might be precursors of the classic SPs [1] (Fig. 3). More interestingly, we found the presence of TG-catalyzed cross-links in both diffuse and classic SPs [58], indicating that tTG is not only present but also catalytically active within these lesions. Together,

cte

331

rre

330

nucleus [60]. Interestingly, we recently observed that tTG is associated with the endoplasmic reticulum (ER) in SH-SY5Y neuroblastoma cells, and is upregulated and activated at the ER under ER stress conditions induced by the Parkinson’s disease-mimetic 1-methyl4-phenylpyridinium [61]. These data obtained in cultures cells, were in line with observations in dopaminergic neurons in postmortem Parkinson’s disease brain [62]. tTG is suggested to play an important role in the development of the nervous system, since its expression level and activity changes significantly during development. For example, during the brain growth spurt, tTG activity levels increase dramatically, suggesting that tTG is involved in maturation of the brain [63]. In addition, tTG has been demonstrated to be essential for neuritic outgrowth, although its specific role in the formation of neurites in neurons remains to be elucidated [64–68].

co

329

Un

328

M.M.M. Wilhelmus et al. / Tissue Transglutaminase in Alzheimer’s Disease

uth

6

tTG in NFTs The first in vivo evidence that tTG is involved in NFT formation by cross-linking of tau into helical filaments was provided by Selkoe and coworkers in the early eighties [14, 80, 81]. Other groups subsequently found that the tau protein is indeed a substrate for tTG-catalyzed cross-linking, and that tTG-mediated cross-linking induces SDS-stable tau filament formation in a calcium-dependent manner [82–85] (Fig. 3). Both phosphorylated tau, which accumulates in NFTs, as well as non-phosphorylated tau are substrates for tTG [86]. Furthermore, tTG is able to polyaminate the tau protein, resulting in an increased resistance of

377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413

414

415 416 417 418 419 420 421 422 423 424 425 426

7

cte

dA

uth

or

Pr

oo f

M.M.M. Wilhelmus et al. / Tissue Transglutaminase in Alzheimer’s Disease

Un

co

rre

Fig. 3. Tissue transglutaminase (tTG) and its catalyzed cross-linking in the formation Alzheimer’s disease (AD) pathological hallmarks. tTG-catalyzed cross-linking of amyloid-␤ (A␤) induces the formation of stable A␤ complexes in vitro, suggesting an early involvement of tTG cross-linking activity in AD pathology lesion formation. Studies on postmortem brain tissue of AD patients confirm this notion as the tTG enzyme and its cross-linking products are already found in diffuse senile plaques (SPs), the precursors of classic SPs [58]. These diffuse SPs, in contrast to classic SPs, do not contain ␤-pleated sheet A␤ fibrils, but already show the presence of tTG and its cross-linking products. In line with the observations in SP formation are the findings in cerebral amyloid angiopathy (CAA). In early stages of CAA, characterized by partial A␤ deposition in the cerebral vessel wall (CVW), elevated anti-tTG antibody immunoreactivity has been demonstrated compared to non-affected vessels [95]. However, in contrast to diffuse and classic SP, in which tTG and its cross-linking products colocalize with the deposited A␤, tTG (green) and its cross-links are present in halos surrounding the A␤ deposition (red) in late stage CAA [58, 95]. These halos are also characterized by the presence of extracellular matrix (ECM) [95], suggesting that in late stage CAA the deposited A␤ is surrounded by a cross-linked ECM network [16, 95]. Apart from the A␤ lesions in AD, tTG and its cross-linking activity are also observed during neurofibrillary tangles (NFT) formation. tTG is known to cross-link tau into anti-parallel dimers [88] that are suggested to precede filament formation. The presence of tTG-induced filaments and tTG cross-links in pretangles, a precursor stage of NFTs, suggests an early involvement in NFT formation. Finally, both tTG and its cross-links are also observed in mature NFTs in AD. Immunohistochemical staining using antibodies directed against either A␤, hyperphosphorylated tau, or tTG on postmortem AD brain tissue was performed as described in previous reports of our group [58, 62, 90, 95].

427 428 429 430 431 432 433 434

tau to proteolytic degradation by calpain leading to higher levels of non-degradable tau within the neuron [87]. Of importance for the notion of an early involvement of tTG in NFTs formation is the finding of Murthy and colleagues, who demonstrated that tTG cross-links tau into an “anti-parallel dimer” that might act as a seed for further tau aggregation [88] (Fig. 3). Apart from these in vitro findings, both

we and others have observed tTG immunoreactivity and tTG-catalyzed cross-links in NFTs in AD brains [58, 82, 89–91] (Fig. 3). Interestingly, these intraneuronal tTG-catalyzed cross-links were found to increase during aging in the brain, albeit to a lesser extent as compared to AD brains [91, 92]. Purification of the helical filaments of AD brain tissue showed that the tau filaments contained TG-catalyzed cross-links

435 436 437 438 439 440 441 442

M.M.M. Wilhelmus et al. / Tissue Transglutaminase in Alzheimer’s Disease

or

Pr

oo f

8

447 448 449 450 451 452 453

454 455

456 457 458 459 460 461 462 463 464 465 466 467 468 469

cte

446

rre

445

[83, 93]. Interestingly, cross-linked tau filaments are found before NFTs are detected, indeed pointing toward an early role of tTG-catalyzed cross-linking of tau in the formation of NFTs [83, 91] (Fig. 3). Although cellular data on the role of tTG-mediated cross-linking is lacking, a transgenic mouse model of tau pathology, the P301L transgenic mouse, showed that the accumulating tau filaments are positive for TG-catalyzed cross-links and that the level of these crosslinks is significantly elevated compared to control mice [94].

co

444

tTG in CAA and other cerebrovascular-related changes in AD The first evidence that tTG is present in CAA was provided by us in 2009, showing that both tTG and TG-catalyzed cross-links were present in halos surrounding the deposited A␤ in CAA-affected vessels [58] (Fig. 3). Surprisingly, in contrast to SPs, no actual spatial colocalization was found between the deposited A␤ and both tTG and TG-catalyzed crosslinks. However, both tTG and TG-catalyzed cross-links were detected in the endothelial cells at the luminal side and in perivascular cells at the abluminal side of the deposited A␤ [58] (Fig. 3). Following up on this study, we set out to further investigate the association of tTG and its activity with CAA, and found that in an early phase of CAA formation, elevated levels of

Un

443

dA

uth

Fig. 4. Intermolecular tissue transglutaminase (tTG) cross-links initiate the protein aggregation cascade. The aggregation cascade of amyloid-␤ (A␤) starts with an initiation process that induces the formation of dimers, followed by the formation of oligomers, protofibrils, and eventually mature fibrils. tTG is capable of inducing intermolecular cross-links between A␤ peptides. For the formation of a tTG-catalyzed intermolecular cross-link between two A␤ peptides the glutamine at position 15 (Q15 ) of one A␤ peptide and the lysine at position 16 (K16 ) of another A␤ peptide are of key importance. This intermolecular cross-link generates stable A␤ dimers that initiate the A␤ cascade [76, 77]. In addition, tTG-catalyzed cross-linking might also act as a driving force to accelerate further aggregation from dimer into oligomers and protofibrils [9]. Altough tTG and its cross-links are also detected in the pathological lesions containing mature ␤-pleated sheet fibrils, so far, data demonstrating a role for tTG-catalyzed cross-linking in the formation of mature A␤ fibrils formation is lacking.

tTG immunoreactivity do colocalize with the deposited A␤ in CAA [95] (Fig. 3). These findings are in line with the observations made in diffuse SPs, and suggest that tTG is likely to play an early role in both SP and CAA formation. In addition, in CAA, tTG and its cross-linking products colocalize with the extracellular matrix (ECM) proteins and tTG-substrates laminin and fibronectin [95], pointing towards tTG-catalyzed modification of the ECM in CAA (Fig. 3). Although the above described data support the idea that tTG activity might drive AD pathology, the presence tTG and/or its activity with these lesions might also be a consequence of lesion formation and associated factors that induce tTG levels and/or activity, e.g., reactive oxygen species and transforming growth factor-beta [96]. In addition, evidence based merely on the use of a single antibody directed against the ␧(␥-glutamyl)lysine bond needs to be interpreted with caution as immunoblots incubated with these antibodies do not solely detect ␧-(␥-glutamyl)lysine bond [97]. However, together, both the in vitro and the human postmortem AD brain tissue data provide compelling evidence that tTG activity not only colocalizes with the pathological hallmarks of AD but also induces stable and toxic oligomers of A␤ and tau, which are both crucial in AD pathogenesis. Therefore, blockade of tTG-catalyzed cross-linking activity in AD patients using tTG inhibitors may offer a new therapeutic approach for AD.

470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497

M.M.M. Wilhelmus et al. / Tissue Transglutaminase in Alzheimer’s Disease

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

oo f

Pr

505

or

504

uth

503

dA

502

The use of both competitive amine inhibitors and reversible inhibitors as therapeutic agents is questionable, since the inhibitory status of tTG is strongly dependent upon the local concentration of these inhibitors, which is related to their stability and breakdown rate in vivo. In addition, inhibition of tTG by guanosine-5′ -triphosphate is limited to conditions of suboptimal activation by calcium ions [105] which involves modulation of calcium binding affinity, interference with conformational changes induced by calcium, local alterations in domain flexibility, and increased strength of interaction between domain 4 and domains 1 and 2. The third class of inhibitors, the irreversible inhibitors, appear more promising. Irreversible inhibitors act via blockade of enzyme activity by covalently modifying the catalytic site of the enzyme and thereby prevent substrate binding. This modification entails targeting of the active site cysteine using functional groups that form stable chemical bonds after reacting. A well-studied irreversible inhibitor is iodoacetamide, which acts as an inhibitor of tTG by inducing the tTG active site thiol to form a stable thioether bond with iodoacetamide [106, 107]. Structurally, iodoacetamide is the simplest irreversible inhibitor of tTG, and therefore might be expected to be useful in therapeutic settings. However, the small structure of iodoacetamide prevents it from having a strong interaction with the enzyme, and although it is highly reactive, iodoacetamide is fairly non-selective in its inhibitory action toward active site cysteines. Thus, being such a non-selective inhibitor, iodoacetamide is likely to interact with nucleophilic residues other than the active site cysteine within tTG, causing unwanted side effects and toxicity [107]. Another class of irreversible inhibitors are the 3halo-4,5-dihydroisoxazoles, whose structure is based on acivicin that alike iodoacetamide acts as an inhibitor of active site cysteines [108]. These compounds are structurally typified by a 3-bromo-4,5dihydroisoxazole warhead group bound to a natural or synthetic amino acid via a peptide bond typically followed by a hydrophobic, aromatic moiety, which increases binding specificity to the tTG active site [109, 110]. The asymmetric nature of the interaction between dihydroisoxazoles and the tTG active site was established after purification of the R and S enantiomers around the C-5 carbon in the dihydroisoxazole ring, demonstrating that only the S enantiomer is able to irreversibly inhibit the enzyme [110, 111]. Their inhibitory activities involve the displacement of the active site thiol and forming a stable iminothioether. Although

cte

501

rre

500

Based upon their mechanisms of inhibition, currently available tTG inhibitors can be divided into three classes: competitive amine inhibitors, reversible inhibitors, and irreversible inhibitors, respectively. The first class are the competitive amine inhibitors that are the best characterized inhibitors of tTG, since they are widely accessible, stable, and nontoxic [98, 99]. The primary amine bound to an aliphatic unbranched carbon chain of around 4-5 saturated carbon atoms is the characteristic structure of this class of inhibitors, although shorter amines such as hydroxylamine and methylamine are also tTG substrates. Amine inhibitors function by competing with other natural amine substrates for tTG in the transamidation reaction. tTG is therefore still active, yet the isodipeptide cross-link is now formed between the natural glutamine substrate and the competitive amine inhibitor rather than between the natural glutamine substrate and natural amine substrates. Wellknown competitive amine inhibitors are putrescine, cystamine, spermidine, histamine, and cadaverine analogues like monodansylcadaverine [43]. Cystamine is probably the most extensively studied inhibitor of this type, although it should not be considered as a specific inhibitor for tTG when applied in biological systems, since cystamine has also been shown to inhibit the thiol-dependent protease caspase-3 and causes increased production of the anti-oxidant glutathione inside cells [100]. The second class of inhibitors is the reversible inhibitors of tTG, which prevent tTG activity by inhibiting substrate access to the active site without covalently modifying the enzyme. Both GTP and GDP are reversible inhibitors, but also guanosine-5′ -triphosphate analogues such as GTP␥S and ␤␥-methyleneguanosine 5′ -triphosphate demonstrate this reversible inhibitory effect on tTG [101]. As described earlier, guanosine-5′ -triphosphate binds tTG and prevents binding of calcium to its binding site on tTG in an allosteric fashion. The divalent metal ion Zn2+ is also capable of reversibly inhibiting tTG activity by competing with Ca2+ for its binding site in the protein [43, 102]. Recently, a new class of reversible inhibitors for tTG has been discovered. These inhibitors are characterized by having a thieno[2,3-d]pyrimidin-4-one acylhydrazide backbone [103, 104]. They exhibit slow-binding kinetics, and kinetic analysis suggests that they possibly compete with guanosine-5′ -triphosphate, since they share the same tTG-binding site.

co

499

TTG INHIBITORS

Un

498

9

549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600

609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652

oo f

608

Pr

607

or

606

3-halo-4,5-dihydro-isoxazole reactive group, demonstrated to be a potent tTG inhibitor in vitro and in vivo [118] and was found to disrupt fibronectin assembly in the extracellular matrix of gliobastomas, which resulted in enhanced apoptosis of these tumors in vitro [125]. Similar observations were made in vivo using a murine orthotopic brain tumor model [125]. In addition, inhibition of tTG activity by KCC009 influenced the ability of dendritic cells to mature upon LPS stimulation [126]. The use of KCC009 as a therapeutic agent in humans is strengthened by the fact that it is well tolerated in pharmacologically effective doses, at least in rodents, and that it has a short serum half-life, indicating a fast distribution over organs and tissues [109]. Recently, in vivo administration of another irreversible site-directed tTG inhibitor ERW1041E in a murine pulmonary hypertension model resulted in a significant reduction in right ventricular systolic pressure [127]. CONCLUSIONS, CRITICAL NOTES, AND FUTURE PERSPECTIVES

dA

605

cte

604

rre

603

this type of inhibitors seemed promising, as they have good bioavailability and low toxicity, their low solubility in buffers at physiological pH caused difficulties in using them as therapeutic agents. Other irreversible inhibitors were designed using natural substrates of tTG as a backbone. For instance, the tTG substrate carbobenzyloxy-Lglutaminylglycine (Cbz-gln-gly) was used to produce a family of tTG inhibitors having different electrophilic functional groups by substituting the reactive moieties in the Gln position. Functional groups such as the chloroacetamides [112], maleimides [113], 1,2,4thiadiazoles [114, 115], epoxides and ␣,␤-unsaturated amides [106] were used for this purpose. Both the Cbz-gln(epoxide) and the Cbz-gln(␣,␤-unsaturated)gly were found to be the most potent inhibitors of tTG in these studies, albeit so far only in vitro. An alternative approach used the high affinity binding of gluten peptides to tTG [116]. The sequence of these gluten peptides served as a blueprint to design several peptidomimetic inhibitors [117]. The tTG target glutamine in these gluten peptides was replaced by, for instance, 6-diazo-5-oxo-norleucine (DON), which demonstrated a potent irreversible inhibitory effect. In fact, the DON containing peptide was found to be 5 orders of magnitude more potent than the acivicin analogue. More recently, our group confirmed the potency and efficacy of the DON-based inhibitors Boc-DONGln-Ile-Val-OMe and Z-DON-Val-Pro-Leu-OMe to inhibit tTG activity both in vitro and in a cellular assay [118]. Although information is still limited, several irreversible inhibitors of tTG have already demonstrated to be promising as therapeutic agents in human diseases. An effective class of inhibitors of FXIIIa, the 2-[(2-oxopropyl)thio]imidazolium derivatives, have been used as inhibitors of tTG in vivo [119]. In fact, the 2-[(2-oxopropyl)thio]imidazolium inhibitor L682777 prevented gluten peptide deamidation, which resulted in reduced T-cell activation in celiac sprue mouse models [120]. Although this group of compounds are fairly good inhibitors of TG2 [119], and are effective in numerous biological settings [117, 120, 121], including mouse models of cardiovascular disease [122–124], the potential side-effects caused by inhibition of the blood coagulating factor XIIIa, makes the 2-[(2oxopropyl)thio]imidazolium inhibitors inappropriate for use in humans. Newer classes of peptidomimetic selective and irreversible tTG inhibitors have begun to be evaluated. Interestingly, the selective and irreversible tTG-inhibitor KCC009, which contains a

co

602

Un

601

M.M.M. Wilhelmus et al. / Tissue Transglutaminase in Alzheimer’s Disease

uth

10

From a pharmacological perspective, testing of tTG inhibitors in animal models of neurodegenerative diseases is a promising approach. For instance, the tTG inhibitor cystamine is known to delay the onset of neurological symptoms associated with Huntington’s disease when applied in the R6/2 Huntington’s disease mouse model. Moreover, the tTG-inhibitors cystamine, monodansylcadaverine, L-682777, KCC009, and recently ERW1041E [127], have been successfully used to inhibit tTG activity in other human disease models [128]. These data suggest that effective inhibition of tTG activity in vivo is feasible. However, studies on the effects of specific tTG inhibitors in dedicated models of AD are still lacking but will be indispensable for moving onwards in evaluating the therapeutic potential of tTG inhibition. Until recently, the insoluble mature fibrils of A␤ were considered to be the toxic species in the aggregation cascade resulting in neurodegeneration. However, smaller soluble oligomers are now thought to be the most neurotoxic species, as their presence in the brain is directly linked to neurotoxicity and cognitive decline [36, 129]. Since tTG-catalyzed cross-linking predominantly results in the formation of soluble A␤ multimers, and not the formation of insoluble mature ␤-pleated sheet fibrils, tTG inhibition might indeed prove to be a particularly attractive therapeutic target (Fig. 4). Within this context it is of interest that, according to the A␤ cascade, the accumulation of toxic

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

M.M.M. Wilhelmus et al. / Tissue Transglutaminase in Alzheimer’s Disease

711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748

751 752

This work was supported by a grant from The Brain Foundation of the Netherlands (number F2010(1)-06

oo f

REFERENCES [1] [2]

[3]

Pr

710

Selkoe DJ (1991) The molecular pathology of Alzheimer’s disease. Neuron 6, 487-498. Lee VM, Balin BJ, Otvos L Jr, Trojanowski JQ (1991) A68: A major subunit of paired helical filaments and derivatized forms of normal Tau. Science 251, 675-678. Biernat J, Mandelkow EM, Schroter C, Lichtenberg-Kraag B, Steiner B, Berling B, Meyer H, Mercken M, Vandermeeren A, Goedert M, et al. (1992) The switch of tau protein to an Alzheimer-like state includes the phosphorylation of two serine-proline motifs upstream of the microtubule binding region. EMBO J 11, 1593-1597. Trojanowski JQ, Lee VM (1995) Phosphorylation of paired helical filament tau in Alzheimer’s disease neurofibrillary lesions: Focusing on phosphatases. FASEB J 9, 1570-1576. Kowall NW, Kosik KS (1987) Axonal disruption and aberrant localization of tau protein characterize the neuropil pathology of Alzheimer’s disease. Ann Neurol 22, 639-643. Greenberg CS, Birckbichler PJ, Rice RH (1991) Transglutaminases: Multifunctional cross-linking enzymes that stabilize tissues. FASEB J 5, 3071-3077. Lorand L, Graham RM (2003) Transglutaminases: Crosslinking enzymes with pleiotropic functions. Nat Rev Mol Cell Biol 4, 140-156. Folk JE, Finlayson JS (1977) The epsilon-(gammaglutamyl)lysine crosslink and the catalytic role of transglutaminases. Adv Protein Chem 31, 1-133. Junn E, Ronchetti RD, Quezado MM, Kim SY, Mouradian MM (2003) Tissue transglutaminase-induced aggregation of alpha-synuclein: Implications for Lewy body formation in Parkinson’s disease and dementia with Lewy bodies. Proc Natl Acad Sci U S A 100, 2047-2052. Igarashi S, Koide R, Shimohata T, Yamada M, Hayashi Y, Takano H, Date H, Oyake M, Sato T, Sato A, Egawa S, Ikeuchi T, Tanaka H, Nakano R, Tanaka K, Hozumi I, Inuzuka T, Takahashi H, Tsuji S (1998) Suppression of aggregate formation and apoptosis by transglutaminase inhibitors in cells expressing truncated DRPLA protein with an expanded polyglutamine stretch. Nat Genet 18, 111-117. Kahlem P, Green H, Djian P (1998) Transglutaminase action imitates Huntington’s disease: Selective polymerization of Huntingtin containing expanded polyglutamine. Mol Cell 1, 595-601. Dedeoglu A, Kubilus JK, Jeitner TM, Matson SA, Bogdanov M, Kowall NW, Matson WR, Cooper AJ, Ratan RR, Beal MF, Hersch SM, Ferrante RJ (2002) Therapeutic effects of cystamine in a murine model of Huntington’s disease. J Neurosci 22, 8942-8950. Karpuj MV, Becher MW, Steinman L (2002) Evidence for a role for transglutaminase in Huntington’s disease and the potential therapeutic implications. Neurochem Int 40, 31-36. Selkoe DJ (2002) Introducing transglutaminase into the study of Alzheimer’s disease. A personal look back. Neurochem Int 40, 13-16. Wilhelmus MM, van Dam AM, Drukarch B (2008) Tissue transglutaminase: A novel pharmacological target in

or

709

rre

708

[4]

[5]

[6]

[7]

[8]

[9]

[10]

co

707

Un

706

uth

ACKNOWLEDGMENTS

705

to MMM.W.) and a grant from the ‘Internationale Stichting Alzheimer Onderzoek’ (number 11501, to MMMW, BD and ENTPB). Authors’ disclosures available online (http://www.jalz.com/disclosures/view.php?id=2175).

dA

750

704

cte

749

A␤ species precedes the formation of hyperphosphorylated tau resulting in NFTs. One of the proposed actions of toxic intermediates of A␤ on neurons is that they stimulate tau hyperphosphorylation and aggregation, and increase intracellular calcium levels [130]. As phosphorylation of tau alone is not sufficient to cause NFTs, it is an appealing hypothesis that the A␤induced dysregulation of calcium homeostasis results in an intracellular increase in tTG activity, ultimately resulting in production of stable tau aggregates and filaments [82]. Clearly, besides elucidation of tTG’s involvement in the A␤ cascade, further studies are also needed to determine the contribution of tTG activity to NFT formation, and investigate whether inhibition of tTG-catalzyed cross-linking of tau could prevent tau aggregation in AD. Since tTG activity is widely spread throughout the human body and is involved in many biological processes, tTG inhibition might be expected to result in severe side-effects. However, tTG knockout mice do not appear lethal and show no major pathological features [131, 132]. Furthermore, the above-described amine and reversible inhibitors demonstrate low toxicity. Thus, contrary to expectations, specific inhibition of tTG does not appear to cause dramatic side-effects in vivo. Nevertheless, if tTG inhibitors are considered for use in AD, they will have to be administered in a chronic fashion and studies on the long-term use of these inhibitors in animal or humans have not been carried out thus far. Within this context, another important consideration for the use of tTG inhibitors in AD is their ability to cross the blood-brain barrier. Currently, no information regarding their brain penetration is available. Although in theory it is likely that small amine-based inhibitors are capable of crossing the blood-brain barrier, other properties for good bloodbrain barrier transport, i.e., the lack of charged groups and lipophilicty, have not been taken into account in current tTG inhibitor drug design. However, encouraging for future drug design of specific and selective tTG inhibitors able to cross the BBB is the extensive knowledge already present regarding the structure of tTG, as described above. This knowledge will guide the development of optimized tTG inhibitors which will enable investigation of the full therapeutic potential of tTG blockade in the brain, first in AD models and eventually in AD patients.

703

11

[11]

[12]

[13]

[14]

[15]

753 754 755 756 757

758

759 760 761 762 763 764 765 766 767 768 769 770 771 772 773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792 793 794 795 796 797 798 799 800 801 802 803 804 805 806 807 808 809 810 811 812 813

[17]

820 821 822

[18]

823 824 825

[19]

826 827 828

[20]

829 830 831 832

[21]

833 834 835 836 837 838

[22]

839 840 841 842 843

[23]

844 845 846 847

[24]

848 849 850 851 852 853 854 855 856

[25]

co

857 858 859 860 861 862 863 864

[26]

865 866 867 868 869 870 871

[27]

872 873 874 875 876 877 878

[28]

[29]

[30]

[31]

oo f

819

Pr

818

or

817

Andreasson U, Blennow K, Soares H, Albright C, Feldman HH, Berman RM (2012) Safety and tolerability of the gamma-secretase inhibitor avagacestat in a phase 2 study of mild to moderate Alzheimer disease. Arch Neurol 69, 1430-1440. Mullane K, Williams M (2013) Alzheimer’s therapeutics: Continued clinical failures question the validity of the amyloid hypothesis-but what lies beyond? Biochem Pharmacol 85, 289-305. Games D, Adams D, Alessandrini R, Barbour R, Berthelette P, Blackwell C, Carr T, Clemens J, Donaldson T, Gillespie F, et al. (1995) Alzheimer-type neuropathology in transgenic mice overexpressing V717F beta-amyloid precursor protein. Nature 373, 523-527. Price JL, McKeel DW Jr, Buckles VD, Roe CM, Xiong C, Grundman M, Hansen LA, Petersen RC, Parisi JE, Dickson DW, Smith CD, Davis DG, Schmitt FA, Markesbery WR, Kaye J, Kurlan R, Hulette C, Kurland BF, Higdon R, Kukull W, Morris JC (2009) Neuropathology of nondemented aging: Presumptive evidence for preclinical Alzheimer disease. Neurobiol Aging 30, 1026-1036. Glabe CG (2006) Common mechanisms of amyloid oligomer pathogenesis in degenerative disease. Neurobiol Aging 27, 570-575. Walsh DM, Klyubin I, Fadeeva JV, Cullen WK, Anwyl R, Wolfe MS, Rowan MJ, Selkoe DJ (2002) Naturally secreted oligomers of amyloid beta protein potently inhibit hippocampal long-term potentiation in vivo. Nature 416, 535-539. McLean CA, Cherny RA, Fraser FW, Fuller SJ, Smith MJ, Beyreuther K, Bush AI, Masters CL (1999) Soluble pool of Abeta amyloid as a determinant of severity of neurodegeneration in Alzheimer’s disease. Ann Neurol 46, 860-866. Kawarabayashi T, Shoji M, Younkin LH, Wen-Lang L, Dickson DW, Murakami T, Matsubara E, Abe K, Ashe KH, Younkin SG (2004) Dimeric amyloid beta protein rapidly accumulates in lipid rafts followed by apolipoprotein E and phosphorylated tau accumulation in the Tg2576 mouse model of Alzheimer’s disease. J Neurosci 24, 3801-3809. Mc Donald JM, Savva GM, Brayne C, Welzel AT, Forster G, Shankar GM, Selkoe DJ, Ince PG, Walsh DM (2010) The presence of sodium dodecyl sulphate-stable Abeta dimers is strongly associated with Alzheimer-type dementia. Brain 133, 1328-1341. Casadio R, Polverini E, Mariani P, Spinozzi F, Carsughi F, Fontana A, Polverino de LP, Matteucci G, Bergamini CM (1999) The structural basis for the regulation of tissue transglutaminase by calcium ions. Eur J Biochem 262, 672-679. Bergamini CM, Dean M, Matteucci G, Hanau S, Tanfani F, Ferrari C, Boggian M, Scatturin A (1999) Conformational stability of human erythrocyte transglutaminase. Patterns of thermal unfolding at acid and alkaline pH. Eur J Biochem 266, 575-582. Bergamini CM, Signorini M (1990) In vivo inactivation of transglutaminase during the acute acrylamide toxic syndrome in the rat. Experientia 46, 278-281. Iismaa SE, Wu MJ, Nanda N, Church WB, Graham RM (2000) GTP binding and signaling by Gh/transglutaminase II involves distinct residues in a unique GTP-binding pocket. J Biol Chem 275, 18259-18265. Lai TS, Bielawska A, Peoples KA, Hannun YA, Greenberg CS (1997) Sphingosylphosphocholine reduces the calcium ion requirement for activating tissue transglutaminase. J Biol Chem 272, 16295-16300.

uth

[16]

Un

816

preventing toxic protein aggregation in neurodegenerative diseases. Eur J Pharmacol 585, 464-472. Wilhelmus MM, de JM, Drukarch B (2012) Tissue transglutaminase: A novel therapeutic target in cerebral amyloid angiopathy. Neurodegener Dis 10, 317-319. Bartus RT, Dean RL, III, Beer B, Lippa AS (1982) The cholinergic hypothesis of geriatric memory dysfunction. Science 217, 408-414. Hynd MR, Scott HL, Dodd PR (2004) Glutamate-mediated excitotoxicity and neurodegeneration in Alzheimer’s disease. Neurochem Int 45, 583-595. Lipton SA (2006) Paradigm shift in neuroprotection by NMDA receptor blockade: Memantine and beyond. Nat Rev Drug Discov 5, 160-170. Hu NW, Klyubin I, Anwyl R, Rowan MJ (2009) GluN2B subunit-containing NMDA receptor antagonists prevent Abeta-mediated synaptic plasticity disruption in vivo. Proc Natl Acad Sci U S A 106, 20504-20509. Shankar GM, Li S, Mehta TH, Garcia-Munoz A, Shepardson NE, Smith I, Brett FM, Farrell MA, Rowan MJ, Lemere CA, Regan CM, Walsh DM, Sabatini BL, Selkoe DJ (2008) Amyloid-beta protein dimers isolated directly from Alzheimer’s brains impair synaptic plasticity and memory. Nat Med 14, 837-842. Um JW, Nygaard HB, Heiss JK, Kostylev MA, Stagi M, Vortmeyer A, Wisniewski T, Gunther EC, Strittmatter SM (2012) Alzheimer amyloid-beta oligomer bound to postsynaptic prion protein activates Fyn to impair neurons. Nat Neurosci 15, 1227-1235. Wilkinson D, Wirth Y, Goebel C (2013) Memantine in patients with moderate to severe Alzheimer’s disease: Metaanalyses using realistic definitions of response. Dement Geriatr Cogn Disord 37, 71-85. Vellas B, Carrillo MC, Sampaio C, Brashear HR, Siemers E, Hampel H, Schneider LS, Weiner M, Doody R, Khachaturian Z, Cedarbaum J, Grundman M, Broich K, Giacobini E, Dubois B, Sperling R, Wilcock GK, Fox N, Scheltens P, Touchon J, Hendrix S, Andrieu S, Aisen P (2013) Designing drug trials for Alzheimer’s disease: What we have learned from the release of the phase III antibody trials: A report from the EU/US/CTAD Task Force. Alzheimers Dement 9, 438-444. Dodel R, Rominger A, Bartenstein P, Barkhof F, Blennow K, Forster S, Winter Y, Bach JP, Popp J, Alferink J, Wiltfang J, Buerger K, Otto M, Antuono P, Jacoby M, Richter R, Stevens J, Melamed I, Goldstein J, Haag S, Wietek S, Farlow M, Jessen F (2013) Intravenous immunoglobulin for treatment of mild-to-moderate Alzheimer’s disease: A phase 2, randomised, double-blind, placebo-controlled, dose-finding trial. Lancet Neurol 12, 233-243. Winblad B, Andreasen N, Minthon L, Floesser A, Imbert G, Dumortier T, Maguire RP, Blennow K, Lundmark J, Staufenbiel M, Orgogozo JM, Graf A (2012) Safety, tolerability, and antibody response of active Abeta immunotherapy with CAD106 in patients with Alzheimer’s disease: Randomised, double-blind, placebo-controlled, first-in-human study. Lancet Neurol 11, 597-604. Doody RS, Raman R, Farlow M, Iwatsubo T, Vellas B, Joffe S, Kieburtz K, He F, Sun X, Thomas RG, Aisen PS, Siemers E, Sethuraman G, Mohs R (2013) A phase 3 trial of semagacestat for treatment of Alzheimer’s disease. N Engl J Med 369, 341-350. Coric V, van Dyck CH, Salloway S, Andreasen N, Brody M, Richter RW, Soininen H, Thein S, Shiovitz T, Pilcher G, Colby S, Rollin L, Dockens R, Pachai C, Portelius E,

[32]

[33]

dA

815

cte

814

M.M.M. Wilhelmus et al. / Tissue Transglutaminase in Alzheimer’s Disease

rre

12

[34]

[35]

[36]

[37]

[38]

[39]

[40]

[41]

879 880 881 882 883 884 885 886 887 888 889 890 891 892 893 894 895 896 897 898 899 900 901 902 903 904 905 906 907 908 909 910 911 912 913 914 915 916 917 918 919 920 921 922 923 924 925 926 927 928 929 930 931 932 933 934 935 936 937 938 939 940 941 942 943

M.M.M. Wilhelmus et al. / Tissue Transglutaminase in Alzheimer’s Disease

951 952 953 954 955

[45]

956 957 958 959

[46]

960 961 962 963

[47]

964 965 966 967

[48]

968 969 970 971 972

[49]

973 974 975 976 977

[50]

978 979

[51]

980 981 982 983

[52]

984 985 986

[53]

987 988 989

[54]

990 991 992

[55]

993 994 995

[56]

996 997 998 999 1000 1001

[57]

1002 1003 1004 1005 1006 1007 1008

[58]

[60]

[61]

[62]

[63]

oo f

[44]

Pr

950

or

949

uth

[43]

colocalize with the pathological lesions in Alzheimer’s disease brain. Brain Pathol 19, 612-622. Wilhelmus MM, Verhaar R, Bol JG, van Dam AM, Hoozemans JJ, Rozemuller AJ, Drukarch B (2011) Novel role of transglutaminase 1 in corpora amylacea formation? Neurobiol Aging 32, 845-856. Lesort M, Attanavanich K, Zhang J, Johnson GV (1998) Distinct nuclear localization and activity of tissue transglutaminase. J Biol Chem 273, 11991-11994. Verhaar R, Drukarch B, Bol JG, Jongenelen CA, Musters RJ, Wilhelmus MM (2012) Increase in endoplasmic reticulumassociated tissue transglutaminase and enzymatic activation in a cellular model of Parkinson’s disease. Neurobiol Dis 45, 839-850. Wilhelmus MM, Verhaar R, Andringa G, Bol JG, Cras P, Shan L, Hoozemans JJ, Drukarch B (2011) Presence of tissue transglutaminase in granular endoplasmic reticulum is characteristic of melanized neurons in Parkinson’s disease brain. Brain Pathol 21, 130-139. Bailey CD, Johnson GV (2004) Developmental regulation of tissue transglutaminase in the mouse forebrain. J Neurochem 91, 1369-1379. Maccioni RB, Seeds NW (1986) Transglutaminase and neuronal differentiation. Mol Cell Biochem 69, 161-168. Mahoney SA, Wilkinson M, Smith S, Haynes LW (2000) Stabilization of neurites in cerebellar granule cells by transglutaminase activity: Identification of midkine and galectin-3 as substrates. Neuroscience 101, 141-155. Singh US, Pan J, Kao YL, Joshi S, Young KL, Baker KM (2003) Tissue transglutaminase mediates activation of RhoA and MAP kinase pathways during retinoic acid-induced neuronal differentiation of SH-SY5Y cells. J Biol Chem 278, 391-399. Tucholski J, Lesort M, Johnson GV (2001) Tissue transglutaminase is essential for neurite outgrowth in human neuroblastoma SH-SY5Y cells. Neuroscience 102, 481-491. Zhang J, Lesort M, Guttmann RP, Johnson GV (1998) Modulation of the in situ activity of tissue transglutaminase by calcium and GTP. J Biol Chem 273, 2288-2295. Johnson GV, Cox TM, Lockhart JP, Zinnerman MD, Miller ML, Powers RE (1997) Transglutaminase activity is increased in Alzheimer’s disease brain. Brain Res 751, 323-329. Bonelli RM, Aschoff A, Niederwieser G, Heuberger C, Jirikowski G (2002) Cerebrospinal fluid tissue transglutaminase as a biochemical marker for Alzheimer’s disease. Neurobiol Dis 11, 106-110. Nemes Z, Fesus L, Egerhazi A, Keszthelyi A, Degrell IM (2001) N(epsilon)(gamma-glutamyl)lysine in cerebrospinal fluid marks Alzheimer type and vascular dementia. Neurobiol Aging 22, 403-406. Wang DS, Uchikado H, Bennett DA, Schneider JA, Mufson EJ, Wu J, Dickson DW (2008) Cognitive performance correlates with cortical isopeptide immunoreactivity as well as Alzheimer type pathology. J Alzheimers Dis 13, 53-66. Zhang W, Johnson BR, Suri DE, Martinez J, Bjornsson TD (1998) Immunohistochemical demonstration of tissue transglutaminase in amyloid plaques. Acta Neuropathol 96, 395-400. Dudek SM, Johnson GV (1994) Transglutaminase facilitates the formation of polymers of the beta-amyloid peptide. Brain Res 651, 129-133. Ho GJ, Gregory EJ, Smirnova IV, Zoubine MN, Festoff BW (1994) Cross-linking of beta-amyloid protein

[64]

[65]

dA

948

[59]

cte

947

rre

946

Lai TS, Hausladen A, Slaughter TF, Eu JP, Stamler JS, Greenberg CS (2001) Calcium regulates S-nitrosylation, denitrosylation, and activity of tissue transglutaminase. Biochemistry 40, 4904-4910. Lorand L, Conrad SM (1984) Transglutaminases. Mol Cell Biochem 58, 9-35. Gaudry CA, Verderio E, Aeschlimann D, Cox A, Smith C, Griffin M (1999) Cell surface localization of tissue transglutaminase is dependent on a fibronectin-binding site in its N-terminal beta-sandwich domain. J Biol Chem 274, 30707-30714. Upchurch HF, Conway E, Patterson MK Jr, Birckbichler PJ, Maxwell MD (1987) Cellular transglutaminase has affinity for extracellular matrix. In Vitro Cell Dev Biol 23, 795-800. Jandu SK, Webb AK, Pak A, Sevinc B, Nyhan D, Belkin AM, Flavahan NA, Berkowitz DE, Santhanam L (2013) Nitric oxide regulates tissue transglutaminase localization and function in the vasculature. Amino Acids 44, 261-269. Zemskov EA, Mikhailenko I, Strickland DK, Belkin AM (2007) Cell-surface transglutaminase undergoes internalization and lysosomal degradation: An essential role for LRP1. J Cell Sci 120, 3188-3199. Wilhelmus MM, Otte-Holler I, van Triel JJ, Veerhuis R, Maat-Schieman ML, Bu G, de Waal RM, Verbeek MM (2007) Lipoprotein receptor-related protein-1 mediates amyloid-beta-mediated cell death of cerebrovascular cells. Am J Pathol 171, 1989-1999. Jones RA, Nicholas B, Mian S, Davies PJ, Griffin M (1997) Reduced expression of tissue transglutaminase in a human endothelial cell line leads to changes in cell spreading, cell adhesion and reduced polymerisation of fibronectin. J Cell Sci 110(Pt 19), 2461-2472. Mosher DF (1984) Cross-linking of fibronectin to collagenous proteins. Mol Cell Biochem 58, 63-68. Kleman JP, Aeschlimann D, Paulsson M, van der RM (1995) Transglutaminase-catalyzed cross-linking of fibrils of collagen V/XI in A204 rhabdomyosarcoma cells. Biochemistry 34, 13768-13775. Aeschlimann D, Thomazy V (2000) Protein crosslinking in assembly and remodelling of extracellular matrices: The role of transglutaminases. Connect Tissue Res 41, 1-27. Martinez J, Rich E, Barsigian C (1989) Transglutaminasemediated cross-linking of fibrinogen by human umbilical vein endothelial cells. J Biol Chem 264, 20502-20508. Sane DC, Moser TL, Pippen AM, Parker CJ, Achyuthan KE, Greenberg CS (1988) Vitronectin is a substrate for transglutaminases. Biochem Biophys Res Commun 157, 115-120. Akimov SS, Belkin AM (2001) Cell surface tissue transglutaminase is involved in adhesion and migration of monocytic cells on fibronectin. Blood 98, 1567-1576. Balklava Z, Verderio E, Collighan R, Gross S, Adams J, Griffin M (2002) Analysis of tissue transglutaminase function in the migration of Swiss 3T3 fibroblasts: The active-state conformation of the enzyme does not affect cell motility but is important for its secretion. J Biol Chem 277, 16567-16575. Kim SY, Grant P, Lee JH, Pant HC, Steinert PM (1999) Differential expression of multiple transglutaminases in human brain. Increased expression and cross-linking by transglutaminases 1 and 2 in Alzheimer’s disease. J Biol Chem 274, 30715-30721. Wilhelmus MM, Grunberg SC, Bol JG, van Dam AM, Hoozemans JJ, Rozemuller AJ, Drukarch B (2009) Transglutaminases and transglutaminase-catalyzed cross-links

co

[42]

945

Un

944

[66]

[67]

[68]

[69]

[70]

[71]

[72]

[73]

[74]

[75]

13

1009 1010 1011 1012 1013 1014 1015 1016 1017 1018 1019 1020 1021 1022 1023 1024 1025 1026 1027 1028 1029 1030 1031 1032 1033 1034 1035 1036 1037 1038 1039 1040 1041 1042 1043 1044 1045 1046 1047 1048 1049 1050 1051 1052 1053 1054 1055 1056 1057 1058 1059 1060 1061 1062 1063 1064 1065 1066 1067 1068 1069 1070 1071 1072 1073

1081 1082 1083

[77]

1084 1085 1086 1087 1088

[78]

1089 1090 1091

[79]

1092 1093 1094 1095

[80]

1096 1097 1098 1099

[81]

1100 1101 1102 1103

[82]

1104 1105 1106 1107

[83]

1108 1109 1110

[84]

1111 1112 1113

[85]

1114 1115 1116

[86]

1117 1118 1119

[87]

1120 1121 1122 1123

[88]

1124 1125 1126 1127

[89]

1128 1129 1130 1131

[90]

1132 1133 1134 1135 1136 1137 1138

[91]

[94]

[95]

[96]

oo f

1080

[93]

Pr

1079

or

1078

Nemes Z, Devreese B, Steinert PM, Van BJ, Fesus L (2004) Cross-linking of ubiquitin, HSP27, parkin, and alpha-synuclein by gamma-glutamyl-epsilon-lysine bonds in Alzheimer’s neurofibrillary tangles. FASEB J 18, 11351137. Zemaitaitis MO, Kim SY, Halverson RA, Troncoso JC, Lee JM, Muma NA (2003) Transglutaminase activity, protein, and mRNA expression are increased in progressive supranuclear palsy. J Neuropathol Exp Neurol 62, 173-184. Halverson RA, Lewis J, Frausto S, Hutton M, Muma NA (2005) Tau protein is cross-linked by transglutaminase in P301L tau transgenic mice. J Neurosci 25, 1226-1233. de JM, van der WB, Schul E, Bol JG, van Duinen SG, Drukarch B, Wilhelmus MM (2013) Tissue transglutaminase colocalizes with extracellular matrix proteins in cerebral amyloid angiopathy. Neurobiol Aging 34, 11591169. Lee ZW, Kwon SM, Kim SW, Yi SJ, Kim YM, Ha KS (2003) Activation of in situ tissue transglutaminase by intracellular reactive oxygen species. Biochem Biophys Res Commun 305, 633-640. Johnson GV, LeShoure R Jr (2004) Immunoblot analysis reveals that isopeptide antibodies do not specifically recognize the epsilon-(gamma-glutamyl)lysine bonds formed by transglutaminase activity. J Neurosci Methods 134, 151-158. Karpuj MV, Becher MW, Springer JE, Chabas D, Youssef S, Pedotti R, Mitchell D, Steinman L (2002) Prolonged survival and decreased abnormal movements in transgenic model of Huntington disease, with administration of the transglutaminase inhibitor cystamine. Nat Med 8, 143-149. Verderio E, Nicholas B, Gross S, Griffin M (1998) Regulated expression of tissue transglutaminase in Swiss 3T3 fibroblasts: Effects on the processing of fibronectin, cell attachment, and cell death. Exp Cell Res 239, 119-138. Lesort M, Lee M, Tucholski J, Johnson GV (2003) Cystamine inhibits caspase activity. Implications for the treatment of polyglutamine disorders. J Biol Chem 278, 3825-3830. Lai TS, Slaughter TF, Peoples KA, Hettasch JM, Greenberg CS (1998) Regulation of human tissue transglutaminase function by magnesium-nucleotide complexes. Identification of distinct binding sites for Mg-GTP and Mg-ATP. J Biol Chem 273, 1776-1781. Aeschlimann D, Paulsson M (1994) Transglutaminases: Protein cross-linking enzymes in tissues and body fluids. Thromb Haemost 71, 402-415. Case A, Stein RL (2007) Kinetic analysis of the interaction of tissue transglutaminase with a nonpeptidic slow-binding inhibitor. Biochemistry 46, 1106-1115. Duval E, Case A, Stein RL, Cuny GD (2005) Structureactivity relationship study of novel tissue transglutaminase inhibitors. Bioorg Med Chem Lett 15, 1885-1889. Bergamini CM (1988) GTP modulates calcium binding and cation-induced conformational changes in erythrocyte transglutaminase. FEBS Lett 239, 255-258. de MP, Marrano C, Keillor JW (2002) Synthesis of dipeptide-bound epoxides and alpha,beta-unsaturated amides as potential irreversible transglutaminase inhibitors. Bioorg Med Chem 10, 355-360. Folk JE, Cole PW (1966) Identification of a functional cysteine essential for the activity of guinea pig liver transglutaminase. J Biol Chem 241, 3238-3240. Tso JY, Bower SG, Zalkin H (1980) Mechanism of inactivation of glutamine amidotransferases by the antitumor drug L-(alpha S, 5S)-alpha-amino-3-chloro-4,5-dihydro-

uth

1077

[92]

[97]

[98]

dA

[76]

cte

1076

precursor catalyzed by tissue transglutaminase. FEBS Lett 349, 151-154. Schmid AW, Condemi E, Tuchscherer G, Chiappe D, Mutter M, Vogel H, Moniatte M, Tsybin YO (2011) Tissue transglutaminase-mediated glutamine deamidation of beta-amyloid peptide increases peptide solubility, whereas enzymatic cross-linking and peptide fragmentation may serve as molecular triggers for rapid peptide aggregation. J Biol Chem 286, 12172-12188. Hartley DM, Zhao C, Speier AC, Woodard GA, Li S, Li Z, Walz T (2008) Transglutaminase induces protofibril-like amyloid beta-protein assemblies that are protease-resistant and inhibit long-term potentiation. J Biol Chem 283, 1679016800. Wakshlag JJ, Antonyak MA, Boehm JE, Boehm K, Cerione RA (2006) Effects of tissue transglutaminase on betaamyloid1-42-induced apoptosis. Protein J 25, 83-94. Curro M, Ferlazzo N, Condello S, Caccamo D, Ientile R (2010) Transglutaminase 2 silencing reduced the betaamyloid-effects on the activation of human THP-1 cells. Amino Acids 39, 1427-1433. Selkoe DJ, Abraham C, Ihara Y (1982) Brain transglutaminase: In vitro crosslinking of human neurofilament proteins into insoluble polymers. Proc Natl Acad Sci U S A 79, 6070-6074. Selkoe DJ, Ihara Y, Salazar FJ (1982) Alzheimer’s disease: Insolubility of partially purified paired helical filaments in sodium dodecyl sulfate and urea. Science 215, 1243-1245. Appelt DM, Kopen GC, Boyne LJ, Balin BJ (1996) Localization of transglutaminase in hippocampal neurons: Implications for Alzheimer’s disease. J Histochem Cytochem 44, 1421-1427. Norlund MA, Lee JM, Zainelli GM, Muma NA (1999) Elevated transglutaminase-induced bonds in PHF tau in Alzheimer’s disease. Brain Res 851, 154-163. Miller CC, Anderton BH (1986) Transglutaminase and the neuronal cytoskeleton in Alzheimer’s disease. J Neurochem 46, 1912-1922. Lesort M, Tucholski J, Miller ML, Johnson GV (2000) Tissue transglutaminase: A possible role in neurodegenerative diseases. Prog Neurobiol 61, 439-463. Dudek SM, Johnson GV (1993) Transglutaminase catalyzes the formation of sodium dodecyl sulfate-insoluble, Alz-50reactive polymers of tau. J Neurochem 61, 1159-1162. Tucholski J, Kuret J, Johnson GV (1999) Tau is modified by tissue transglutaminase in situ: Possible functional and metabolic effects of polyamination. J Neurochem 73, 18711880. Murthy SN, Wilson JH, Lukas TJ, Kuret J, Lorand L (1998) Cross-linking sites of the human tau protein, probed by reactions with human transglutaminase. J Neurochem 71, 2607-2614. Citron BA, Suo Z, SantaCruz K, Davies PJ, Qin F, Festoff BW (2002) Protein crosslinking, tissue transglutaminase, alternative splicing and neurodegeneration. Neurochem Int 40, 69-78. Wilhelmus MM, de JM, Rozemuller AJ, Breve J, Bol JG, Eckert RL, Drukarch B (2012) Transglutaminase 1 and its regulator tazarotene-induced gene 3 localize to neuronal tau inclusions in tauopathies. J Pathol 226, 132-142. Singer SM, Zainelli GM, Norlund MA, Lee JM, Muma NA (2002) Transglutaminase bonds in neurofibrillary tangles and paired helical filament tau early in Alzheimer’s disease. Neurochem Int 40, 17-30.

rre

1075

co

1074

M.M.M. Wilhelmus et al. / Tissue Transglutaminase in Alzheimer’s Disease

Un

14

[99]

[100]

[101]

[102]

[103]

[104]

[105]

[106]

[107]

[108]

1139 1140 1141 1142 1143 1144 1145 1146 1147 1148 1149 1150 1151 1152 1153 1154 1155 1156 1157 1158 1159 1160 1161 1162 1163 1164 1165 1166 1167 1168 1169 1170 1171 1172 1173 1174 1175 1176 1177 1178 1179 1180 1181 1182 1183 1184 1185 1186 1187 1188 1189 1190 1191 1192 1193 1194 1195 1196 1197 1198 1199 1200 1201 1202 1203

M.M.M. Wilhelmus et al. / Tissue Transglutaminase in Alzheimer’s Disease

1211 1212 1213 1214

[111]

1215 1216 1217 1218 1219 1220

[112]

1221 1222 1223

[113]

1224 1225 1226

[114]

1227 1228 1229 1230

[115]

1231 1232 1233 1234

[116]

1235 1236 1237 1238

[117]

1239 1240 1241 1242

[118]

1243 1244 1245 1246 1247

[119]

1248 1249 1250 1251 1252

[120]

1253 1254 1255 1256

[121]

[123]

oo f

[110]

Pr

1210

[122]

[124]

[125]

or

1209

uth

1208

fibrinogen, or fibronectin into high molecular weight complexes on the extracellular surface of isolated hepatocytes. Use of 2-[(2-oxopropyl)thio] imidazolium derivatives as cellular transglutaminase inactivators. J Biol Chem 266, 22501-22509. Matlung HL, VanBavel E, van den Akker J, de Vries CJ, Bakker EN (2010) Role of transglutaminases in cuffinduced atherosclerotic lesion formation in femoral arteries of ApoE3 Leiden mice. Atherosclerosis 213, 77-84. Matlung HL, Neele AE, Groen HC, van GK, Tuna BG, van WA, de VJ, Wentzel JJ, Hoogenboezem M, van Buul JD, VanBavel E, Bakker EN (2012) Transglutaminase activity regulates atherosclerotic plaque composition at locations exposed to oscillatory shear stress. Atherosclerosis 224, 355362. van den Akker J, VanBavel E, van GR, Matlung HL, Guvenc TB, Janssen GM, van Veelen PA, Boelens WC, De Mey JG, Bakker EN (2011) The redox state of transglutaminase 2 controls arterial remodeling. PLoS One 6, e23067 Yuan L, Siegel M, Choi K, Khosla C, Miller CR, Jackson EN, Piwnica-Worms D, Rich KM (2007) Transglutaminase 2 inhibitor, KCC009, disrupts fibronectin assembly in the extracellular matrix and sensitizes orthotopic glioblastomas to chemotherapy. Oncogene 26, 2563-2573. Matic I, Sacchi A, Rinaldi A, Melino G, Khosla C, Falasca L, Piacentini M (2010) Characterization of transglutaminase type II role in dendritic cell differentiation and function. J Leukoc Biol 88, 181-188. DiRaimondo TR, Klock C, Warburton R, Herrera Z, Penumatsa K, Toksoz D, Hill N, Khosla C, Fanburg B (2014) Elevated transglutaminase 2 activity is associated with hypoxia-induced experimental pulmonary hypertension in mice. ACS Chem Biol 9, 266-275. Sollid LM, Khosla C (2005) Future therapeutic options for celiac disease. Nat Clin Pract Gastroenterol Hepatol 2, 140147. Borlikova GG, Trejo M, Mably AJ, Mc Donald JM, Sala FC, Regan CM, Murphy KJ, Masliah E, Walsh DM (2013) Alzheimer brain-derived amyloid beta-protein impairs synaptic remodeling and memory consolidation. Neurobiol Aging 34, 1315-1327. Duyckaerts C (2004) Looking for the link between plaques and tangles. Neurobiol Aging 25, 735-739. Korponay-Szabo IR, Laurila K, Szondy Z, Halttunen T, Szalai Z, Dahlbom I, Rantala I, Kovacs JB, Fesus L, Maki M (2003) Missing endomysial and reticulin binding of coeliac antibodies in transglutaminase 2 knockout tissues. Gut 52, 199-204. Nardacci R, Lo IO, Ciccosanti F, Falasca L, Addesso M, Amendola A, Antonucci G, Craxi A, Fimia GM, Iadevaia V, Melino G, Ruco L, Tocci G, Ippolito G, Piacentini M (2003) Transglutaminase type II plays a protective role in hepatic injury. Am J Pathol 162, 1293-1303.

[126]

dA

1207

[127]

cte

[109]

rre

1206

5-isoxazoleacetic acid (AT-125). J Biol Chem 255, 6734-6738. Choi K, Siegel M, Piper JL, Yuan L, Cho E, Strnad P, Omary B, Rich KM, Khosla C (2005) Chemistry and biology of dihydroisoxazole derivatives: Selective inhibitors of human transglutaminase 2. Chem Biol 12, 469-475. Watts RE, Siegel M, Khosla C (2006) Structure-activity relationship analysis of the selective inhibition of transglutaminase 2 by dihydroisoxazoles. J Med Chem 49, 7493-7501. Killackey JJ, Bonaventura BJ, Castelhano AL, Billedeau RJ, Farmer W, DeYoung L, Krantz A, Pliura DH (1989) A new class of mechanism-based inhibitors of transglutaminase enzymes inhibits the formation of cross-linked envelopes by human malignant keratinocytes. Mol Pharmacol 35, 701706. Pardin C, Gillet SM, Keillor JW (2006) Synthesis and evaluation of peptidic irreversible inhibitors of tissue transglutaminase. Bioorg Med Chem 14, 8379-8385. Halim D, Caron K, Keillor JW (2007) Synthesis and evaluation of peptidic maleimides as transglutaminase inhibitors. Bioorg Med Chem Lett 17, 305-308. Marrano C, de MP, Gagnon P, Lapierre D, Gravel C, Keillor JW (2001) Synthesis and evaluation of novel dipeptidebound 1,2,4-thiadiazoles as irreversible inhibitors of guinea pig liver transglutaminase. Bioorg Med Chem 9, 3231-3241. Marrano C, de MP, Keillor JW (2001) Evaluation of novel dipeptide-bound alpha,beta-unsaturated amides and epoxides as irreversible inhibitors of guinea pig liver transglutaminase. Bioorg Med Chem 9, 1923-1928. Bruce SE, Bjarnason I, Peters TJ (1985) Human jejunal transglutaminase: Demonstration of activity, enzyme kinetics and substrate specificity with special relation to gliadin and coeliac disease. Clin Sci (Lond) 68, 573-579. Hausch F, Halttunen T, Maki M, Khosla C (2003) Design, synthesis, and evaluation of gluten peptide analogs as selective inhibitors of human tissue transglutaminase. Chem Biol 10, 225-231. Verhaar R, Jongenelen CA, Gerard M, Baekelandt V, van Dam AM, Wilhelmus M, Drukarch B (2011) Blockade of enzyme activity inhibits tissue transglutaminase-mediated transamidation of alpha-synuclein in a cellular model of Parkinson’s disease. Neurochem Int 58, 785-793. Freund KF, Doshi KP, Gaul SL, Claremon DA, Remy DC, Baldwin JJ, Pitzenberger SM, Stern AM (1994) Transglutaminase inhibition by 2-[(2-oxopropyl)thio]imidazolium derivatives: Mechanism of factor XIIIa inactivation. Biochemistry 33, 10109-10119. Maiuri L, Ciacci C, Ricciardelli I, Vacca L, Raia V, Rispo A, Griffin M, Issekutz T, Quaratino S, Londei M (2005) Unexpected role of surface transglutaminase type II in celiac disease. Gastroenterology 129, 1400-1413. Barsigian C, Stern AM, Martinez J (1991) Tissue (type II) transglutaminase covalently incorporates itself,

co

1205

Un

1204

[128]

[129]

[130] [131]

[132]

15

1257 1258 1259 1260 1261 1262 1263 1264 1265 1266 1267 1268 1269 1270 1271 1272 1273 1274 1275 1276 1277 1278 1279 1280 1281 1282 1283 1284 1285 1286 1287 1288 1289 1290 1291 1292 1293 1294 1295 1296 1297 1298 1299 1300 1301 1302 1303 1304 1305 1306 1307 1308 1309

Lihat lebih banyak...

Comentarios

Copyright © 2017 DATOSPDF Inc.