• Aucun résultat trouvé

Genetic risk factors for the posterior cortical atrophy variant of Alzheimer's disease

N/A
N/A
Protected

Academic year: 2021

Partager "Genetic risk factors for the posterior cortical atrophy variant of Alzheimer's disease"

Copied!
11
0
0

Texte intégral

(1)

HAL Id: hal-01289466

https://hal.sorbonne-universite.fr/hal-01289466

Submitted on 16 Mar 2016

HAL is a multi-disciplinary open access

archive for the deposit and dissemination of

sci-entific research documents, whether they are

pub-lished or not. The documents may come from

teaching and research institutions in France or

abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est

destinée au dépôt et à la diffusion de documents

scientifiques de niveau recherche, publiés ou non,

émanant des établissements d’enseignement et de

recherche français ou étrangers, des laboratoires

publics ou privés.

Distributed under a Creative Commons Attribution| 4.0 International License

variant of Alzheimer’s disease

Jonathan M. Schott, Sebastian J. Crutch, Minerva M. Carrasquillo, James

Uphill, Tim J. Shakespeare, Natalie S. Ryan, Keir X.X. Yong, Manja

Lehmann, Nilufer Ertekin-Taner, Neil R. Graff-Radford, et al.

To cite this version:

Jonathan M. Schott, Sebastian J. Crutch, Minerva M. Carrasquillo, James Uphill, Tim J. Shakespeare,

et al.. Genetic risk factors for the posterior cortical atrophy variant of Alzheimer’s disease. Alzheimer’s

and Dementia, Elsevier, 2016, �10.1016/j.jalz.2016.01.010�. �hal-01289466�

(2)

Featured Article

Genetic risk factors for the posterior cortical atrophy variant of

Alzheimer’s disease

Q10

Jonathan M. Schott

a,

*

, Sebastian J. Crutch

a

, Minerva M. Carrasquillo

b

, James Uphill

c

,

Tim J. Shakespeare

a

, Natalie S. Ryan

a

, Keir X. X. Yong

a

, Manja Lehmann

a

,

Nilufer Ertekin-Taner

b,d

, Neil R. Graff-Radford

d

, Bradley F. Boeve

e

, Melissa E. Murray

b

,

Qurat ul Ain Khan

b

, Ronald C. Petersen

e

, Dennis W. Dickson

b

, David S. Knopman

e

,

Gil D. Rabinovici

f

, Bruce L. Miller

f

, Aida Suarez Gonzalez

a,g

, Eulogio Gil-N

eciga

g

,

Julie S. Snowden

h

, Jenny Harris

h

, Stuart M. Pickering-Brown

h

, Eva Louwersheimer

i

,

Wiesje M. van der Flier

i

, Philip Scheltens

i

, Yolande A. Pijnenburg

i

, Douglas Galasko

j,k

,

Marie Sarazin

l

, Bruno Dubois

m

, Eloi Magnin

n

, Daniela Galimberti

o

, Elio Scarpini

o

,

Stefano F. Cappa

p

, John R. Hodges

q

, Glenda M. Halliday

q

, Lauren Bartley

q

, Maria C. Carrillo

r

,

Jose T. Bras

s

, John Hardy

s

, Martin N. Rossor

a

, John Collinge

c

, Nick C. Fox

a

, Simon Mead

c

Q2

aDepartment of Neurodegenerative Disease, Dementia Research Centre, UCL Institute of Neurology, UK

Q3

bDepartment of Neuroscience, Mayo Clinic, Jacksonville, FL, USA

cDepartment of Neurodegenerative Disease, MRC Prion Unit, UCL Institute of Neurology, UK

dDepartment of Neurology, Mayo Clinic, Jacksonville, FL, USA

eDepartment of Neurology, Mayo Clinic, Rochester, MN, USA

fUCSF, San Francisco, CA, USA

gUniversity Hospital Virgen del Rocıo, Seville, Spain

h

Institute of Brain, Behaviour and Mental Health, University of Manchester, UK

i

Alzheimer center, Department of Neurology, VU University Medical Center, Neuroscience Campus, Amsterdam, Netherlands

j

Department of Epidemiology & biostatistics, VU University Medical Center, Amsterdam, The Netherlands

k

UC San Diego/VA San Diego Healthcare System, San Diego, CA, USA

l

INSERM U610, H^opital de la Salp^etriere, Paris, France

m

Centre des Maladies Cognitives et Comportementales, IM2A, ICM, Paris 6 University, France

n

Regional Memory Centre (CMRR), CHU Besanc¸on, Besanc¸on, France

oUniversity of Milan, Fondazione Ca Granda, IRCCS Ospedale Policlinico, Italy

p

Vita-Salute San Raffaele University, Milan, Italy

q

University of New South Wales, Sydney, Australia

rAlzheimer’s Association, Chicago, IL, USA

sDepartment of Molecular Neurosciences, UCL Institute of Neurology, UK

Abstract Introduction: The genetics underlying posterior cortical atrophy (PCA), typically a rare variant of Alzheimer’s disease (AD), remain uncertain.

Methods: We genotyped 302 PCA patients from 11 centers, calculated risk at 24 loci for AD/DLB and performed an exploratory genome-wide association study.

Results: We confirm that variation in/near APOE/TOMM40 (P5 6 ! 10214) alters PCA risk, but with smaller effect than for typical AD (PCA: odds ratio [OR]5 2.03, typical AD: OR 5 2.83, P5 .0007). We found evidence for risk in/near CR1 (P 5 7 ! 1024), ABCA7 (P5 .02) and BIN1 (P5 .04). ORs at variants near INPP5D and NME8 did not overlap between PCA and typical AD.

*Corresponding author. Tel.:11 020 3448 3011; Fax: ---.

Q4

E-mail address:j.schott@ucl.ac.uk

http://dx.doi.org/10.1016/j.jalz.2016.01.010

1552-5260/Ó 2016 The Alzheimer’s Association. Published by Elsevier Inc. All rights reserved.

Alzheimer’s & Dementia- (2016) 1-10

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 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 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 109

(3)

Exploratory genome-wide association studies confirmed APOE and identified three novel loci:

rs76854344 near CNTNAP5 (P5 8 ! 10210OR 5 1.9 [1.5–2.3]); rs72907046 near FAM46A

(P5 1 ! 1029OR5 3.2 [2.1–4.9]); and rs2525776 near SEMA3C (P 5 1 ! 1028, OR5 3.3

[2.1–5.1]).

Discussion:We provide evidence for genetic risk factors specifically related to PCA. We identify three candidate loci that, if replicated, may provide insights into selective vulnerability and pheno-typic diversity in AD.

Ó 2016 The Alzheimer’s Association. Published by Elsevier Inc. All rights reserved.

Keywords: Posterior cortical atrophy; Alzheimer’s disease; Genetics; GWAS; Selective vulnerability; ApoE

1. Introduction Posterior

Q5 cortical atrophy (PCA) is a rare

neurodegenera-tive syndrome, typically a variant of Alzheimer’s disease (AD), although occasionally due to other pathologies including dementia with Lewy bodies, corticobasal degener-ation, and prion disease[1]. Patients with PCA present with combinations of cognitive problems attributable to posterior cortical dysfunction and in particular difficulties with higher level visual processing including simultanagnosia, optic apraxia, optic ataxia, and visual disorientation; other fea-tures may include dyslexia, dyscalculia, dysgraphia, and limb dyspraxia. In contrast with typical, amnestic AD, mem-ory is relatively spared until the disease becomes advanced. MR brain imaging in PCA typically shows parieto-occipital lobe atrophy with relative preservation of medial temporal lobe structures[2]; fluoro-deoxyglucose positron emission tomography (PET) shows prominent posterior cortical hypo-metabolism[3]; and in a single case study using the AV1451 tau PET tracer, posterior cortical tau deposition [4]. By contrast, PET imaging using amyloid-binding ligands typi-cally shows global amyloid deposition [3]. Aside from the imaging and cognitive differences, patients with PCA are typically younger than those with typical amnestic late-onset AD, usually with disease late-onset in the sixth or seventh decade[1]. PCA is almost invariably a sporadic disorder, and the risk factors for developing the syndrome are unknown. Understanding the genetic architecture of the PCA variant of AD may provide insights both into factors predisposing to young onset AD, as well as mechanisms underlying regional vulnerability in AD.

To date, only a few, single-center studies have addressed genetic risk for PCA [5–10], and due to the rarity of the syndrome all have been relatively small, the largest being a maximum of 81 cases [9]. Some, but not all, of these studies have suggested that despite their early-disease onset, patients with PCA may be less likely than expected to have an APOE E4 allele, the commonest risk factor for late-onset AD. Other studies have suggested that there may be differ-ences in some of the more recently identified genetic risks for AD in patients with PCA[9]. Recognizing the rarity of this AD variant, we formed an international consortium comprising eleven centers, using clinical diagnostic criteria

to define cases of PCA, with the principal aim of determining whether APOE E4 and genetic risks from recent genome-wide association studies (GWAS) of AD and dementia with Lewy bodies (DLB, see below) are the risk factors for the PCA variant of AD. In a second, exploratory analysis, we performed a pilot GWAS analysis to identify novel puta-tive genetic risk factors for PCA.

2. Methods

2.1. PCA patients and controls Q6

After an inaugural multidisciplinary meeting of PCA re-searchers[11], latterly formalized as the Alzheimer’s Asso-ciation’s International Society to Advance Alzheimer’s Research and Treatment (ISTAART) Professional Interest Area in Atypical AD and Associated Syndromes, an interna-tional collaborative group was established to assess genetic risk factors for PCA. Researchers identified individuals with PCA, in whom a deoxyribo nucleic acid (DNA) sample was available. Patients who the referring physician had diag-nosed with AD, had multidomain cognitive impairment ful-filling criteria for AD dementia, and had one or both of two published criteria for PCA, as proposed by Tang-Wai[5]and Mendez [12] (Table 1) were included. Additional data collected included gender, age at disease onset, age at death (where applicable), and whether there was molecular (cere-brospinal fluid or amyloid PET using locally defined ranges) evidence or pathologic confirmation of underlying AD pa-thology. Each site had appropriate local ethical approvals in place, and all participants gave informed written consent. Controls were from UK, USA, and Germany (see below).

2.2. Genetic and statistical analyses

DNA samples were analyzed at the MRC Prion Unit, Department of Neurodegenerative Disease, Institute of Neurology, UCL. PCA samples were genotyped on Illumina 660 arrays (n5 54, UCL cohort only) and OmniExpress ar-rays (n5 239, all cohorts); in total, 293 passed sample qual-ity control, implemented using PLINK. Controls were genotyped on Illumina 550 (n5 809, KORA F4 German, PMID: 16032514), OmniExpress (n5 1185, Geisinger US

110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 159 160 161 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 184 185 186 187 188 189 190 191 192 193 194 195 196 197 198 199 200 201 202 203 204 205 206 207 208 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231

(4)

http://www.geisinger.org), Illumina 2.5 M (n 5 1882, KORA F3 German), Illumina 5M (n5 1651, Framingham US, see acknowledgments), and Illumina 1.2 M (n 5 5020, WTCCC2 UK, see acknowledgments) arrays. Physical locations refer to the Feb 2009 (GRCh37/hg19) as-sembly. We excluded SNPs with a minor allele frequency (MAF), 1% (n 5 70,042 from OmniExpress case arrays); genotyping rate,99% (n 5 85,086 from OmniExpress case arrays); or Hardy-Weinberg Equilibrium (HWE) exact test P , 1023in controls. Cases with call rate,98% (n 5 6) as well as ethnic outliers (n5 2) were excluded after visual-ization of multidimensional scaling plots. Related and dupli-cate cases were removed by IBS/IBD calculation (n5 1) and re-examination of patient data, as they became available post genotyping (n 5 2). A Pi-Hat (proportion identity by descent) threshold of .0.1875 was used, which should exclude first and second degree relatives. Two duplicates removed due to later availability of patient data were also included within the six cases removed for low call rate thus bringing the total number of cases removed to nine. Owing to the multitude of different genotyping platforms, control comparisons were carried out sequentially and ex-clusions removed at each stage. A total of 840 KORAF4 German controls were originally genotyped; 17 with call rate,98%, six with Pi-Hat . 0.1875 and eight MDS out-liers were removed. A total of 1950 KORAF3 German con-trols were originally downloaded; three with call rate,98%,

57 with Pi-Hat . 0.1875, and eight MDS outliers were removed. A total of 1264 Geisinger US controls were origi-nally downloaded; two with call rate,98%, 69 with Pi-Hat .0.1875, and eight MDS outliers were removed. 2467 FHS US controls were originally downloaded; 16 with call rate ,98%, 793 with Pi-Hat .0.1875 and seven MDS outliers were removed. 5050 WTCCC2 UK controls were generated from available raw IDAT files; 26 with call rate,98%, four with Pi-Hat. 0.1875, and zero MDS outliers were removed (see Supplementary Table 1). All remaining cases and controls were finally visualized on an MDS plot (see

Supplementary Fig. 1), outlier detection was performed us-ing PLINK v1.07, and no further outliers were detected. IBS/ IBD estimation of the final cases and controls also lead to no further exclusions based on relatedness. Shapeitv2 was used, in conjunction with the 1000 Genomes Phase 1 Integrated variant set (b37 March 2012 release), to align all data rela-tive to the posirela-tive strand [13]. To avoid potential down-stream cross platform confusion, however, we removed any A/T or G/C transversions to phase each chromosome from each platform separately before imputation using Impute2 (v2.3.0). GTOOL (v0.6.6) was used to extract and collate samples into their respective cohorts for association testing [14]. Association testing was performed using SNPtest_v2.5-beta4 employing the frequentist (additive model) score method which involves weighting by the likeli-hood of each imputed genotype [15]. Four population

Table 1

Clinical criteria for PCA

Tang-Wai et al, 2004[5] Mendez et al, 2002[11]

Core features

Insidious onset and gradual progression

Presentation of visual complaints in the absence of significant primary ocular disease

Relative preservation of anterograde memory and insight early in the disorder

Disabling visual impairment throughout the disorder Absence of stroke or tumor

Absence of early parkinsonism and hallucinations Any of the following findings

Simultanagnosia with or without optic ataxia or ocular apraxia Constructional dyspraxia

Visual field defect

Environmental disorientation

Any of the elements of Gerstmann syndrome Supportive features

Alexia Presenile onset

Ideomotor or dressing apraxia Prosopagnosia

Investigations

Neuropsychological deficits referable to parietal and/or occipital regions

Focal or asymmetric atrophy in parietal and/or occipital regions on structural imaging

Focal or asymmetric hypoperfusion/hypometabolism in parietal and/ or occipital regions on functional

imaging.

Core diagnostic features (all must be present) Insidious onset and gradual progression

Presentation with visual complaints with intact primary visual functions

Evidence of predominant complex visual disorder on examination Elements of Balint’s syndrome

Visual agnosia Dressing apraxia

Environmental disorientation

Proportionally less impaired deficits in memory and verbal fluency Relatively preserved insight with or without depression

Supportive diagnostic features Presenile onset

Alexia

Elements of Gerstmann’s syndrome Ideomotor apraxia

Physical examination within normal limits Investigations

Neuropsychology: predominantly impaired perceptual deficits Brain imaging: predominantly occipitoparietal

abnormality (especially on functional neuroimaging) with relative sparing of frontal and mesiotemporal

regions.

J.M. Schott et al. / Alzheimer’s & Dementia- (2016) 1-10 3

232 233 234 235 236 237 238 239 240 241 242 243 244 245 246 247 248 249 250 251 252 253 254 255 256 257 258 259 260 261 262 263 264 265 266 267 268 269 270 271 272 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 293 294 295 296 297 298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 353

(5)

covariates derived from IBS/IBD analysis in PLINK were used in the association analysis[16,17]. The case-control as-sociation test statistic inflation factor was 1.06 (Supplementary Fig. 2). Any association statistics mentioned in the results section are shown with standard genomic control (PGC) corrected and uncorrected P values.

Post association-testing QC excluded markers with a MAF ,1% and departure from HWE, in both combined and any single control cohort P, 1024. We also excluded makers with SNPtestv2 derived “info metric” and “add info metric” below 0.9. The final autosomal analysis thus included 5.9 M markers.

We assessed whether the genetic risks for typical AD and PCA were different by comparing odds ratios (OR) between our study and those published in typical AD

[18]. We employed a Wald-type test, first calculating the standard error of the difference in log (OR) from the re-ported CIs. We then divided the empirical difference in log (OR) by its standard error, and thus derived a z-score and P value, relying on the approximate normality of log (OR) estimates from large samples. In the main analysis of candidate SNPs, we used a Bonferroni corrected asso-ciation threshold of P, .002 based on the testing of a lead SNP from 24 independent loci derived from studies of AD

[18] and DLB[19].

Although the typical samples sizes that are required to discover novel genome-wide significant risk factors in complex disorders are in the thousands, there are some precedents of strong genetic effects detected with small but phenotypically homogenous samples, including vari-ants at APOE in AD [20], PRNP in prion disease [21], and complement factor H in age related macular degener-ation [22]. We therefore performed an exploratory genome-wide association study using established method-ologies that account for population structure and with imputation of SNPs not present on the genotyping arrays. As PCA is a clinical syndrome that may be due to pathol-ogies other than AD, we also assessed the odds ratios at SNPs of interest in a subset of patients with biomarker/ pathological evidence for AD.

3. Results

3.1. Patients and demographics

A total of 302 samples fulfilling entry criteria to the study were available from eleven centers: University College London (n 5 94); Mayo Clinic, USA (n 5 77); University of California San Francisco, USA (n5 25); University of Sev-ille, Spain (n5 25); University of Manchester, UK (n 5 20); VU University, Netherlands (n5 17); University of California San Diego, USA (n5 16); INSERM, France (n 5 10); Univer-sity of Milan, Italy (n5 6); CHU Besanc¸on, France (n 5 6); and University of New South Wales, Australia (n5 6). Demo-graphics are shown inTable 2. All samples fulfilled Tang-Wai clinical criteria for PCA[5], and in the 225 samples for which data were available, 97.3% also fulfilled Mendez criteria[12];

41% of the cohort was male. Mean (6SD) age at symptom onset was 58.96 6.9 years, and 82.5% had young onset de-mentia, as defined by age at onset of,65 years. Thirty-four patients had died, with a mean age at death of 68.0 (67.7) years. Molecular or pathologic evidence for underlying Alz-heimer pathology was available for 82 (27%), of whom 52 had a CSF profile compatible with AD; 32 had a positive am-yloid PET scan; and 15 had autopsy proven AD. None had evidence for pathology or biomarkers for non-AD pathology. DNA from nine individuals failed array quality control, and statistical analyses were done on the remaining 293/ 302 samples (seeTable 2). We also considered the associa-tions in the sub-sample of 82 with biomarker/autopsy evidence for underlying AD pathology.

3.2. Comparisons at known genetic loci for AD and DLB Results of the genetic analysis of candidate risk factors for the whole PCA cohort are shown inTable 2. First, we consid-ered 24 SNPs known to be genetic risk factors in AD and/or DLB. The best proxy genotyped for the APOE E4 AD-risk allele, rs2075650, located on chromosome 19 in the TOMM40 gene and 13kb upstream of APOE, was identified as a strong risk factor for PCA (OR 2.03, [95% CI5 1.68– 2.46], P 5 6 ! 10214, PGC 5 3 ! 10213; Fig. 1 and

Fig. 2A). rs3818361 located on chromosome 1 in CR1 was also significantly associated (ORs 5 1.38 [1.14–1.67], P5 7 ! 1024, PGC5 1 ! 1023; Fig. 2B). rs3764650 in

ABCA7 (OR5 1.39 [1.07–1.8], P 5 .02, PGC5 .02) and Table 2

Clinical features and demographics

Total number of DNA samples received 302

Number (%) male 124 (41%)

Mean6 SD age at onset (y) 58.96 6.9

Number (%) with young onset dementia (onset,65 y) 249 (83%)

Number with biomarker/path evidence for AD* 82 (27%)

Number (%) with known age of death 34

Mean6 SD age at death (y) 67.96 7.7

9 samples failed genetic QC

Total number of DNA samples passing QC and entering analysis

293

Number (%) male 120 (41%)

Mean6 SD age at onset (y) 58.86 6.9

Number (%) with young onset dementia (onset,65 y) 243 (83%)

Number with biomarker/path evidence for AD* 77 (26%)

Number (%) with known age of death 33

Mean6 SD age at death (y) 67.86 7.8

*No individual with biomarker/path evidence for a non-AD diagnosis were included. 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475

(6)

rs744373 upstream of BIN1 (OR5 1.2 [1.01–1.43], P 5 .04, PGC5 .05) reached nominal significance but did not surpass

our Bonferroni corrected threshold of P, .002. Other candi-date SNPs showed no evidence of association.

In the subset of individuals with either biomarker (CSF or amyloid PET) or pathologic evidence for underlying AD (n 5 82), rs2075650 (at the APOE/TOMM40 locus, subse-quently referred to as APOE) was again identified as a risk

factor with a similar OR to the whole group (OR5 2.00 [1.39–2.89] P5 9 ! 1025, PGC5 1 ! 1024). rs3818361

(CR1) and rs3764650 (ABCA7) both showed nominally sig-nificant differences compared with controls (CR1 OR5 1.7 [1.20–2.41], P 5 .003, PGC 5 .004; ABCA7 OR 5 1.83

[1.17–2.86], P5 .009, PGC5 .01). There was no evidence

for an effect of BIN1 (OR5 1.08 [0.76–1.52], P 5 .65) in the biomarker cohort.

web 4C =FPO web 4C =FPO

Fig. 1. Manhattan plot of autosomes with threshold for genome-wide significance (P, 5 ! 1028) indicated by the red line. Four loci achieved statistical

sig-nificance at APOE (chromosome 19), SEMA3C (chromosome 7), FAM46A (chromosome 6), and CNTNAP5 (chromosome 2).

web 4C =FPO web 4C =FPO

Fig. 2. SNPannotation and proxy (SNAP) plots for five regions of interest. These plots illustrate the statistical evidence of association at a locus together withQ1

information about nearby genes and linkage disequilibrium between the most strongly associated SNP and its neighbors on the chromosome.

J.M. Schott et al. / Alzheimer’s & Dementia- (2016) 1-10 5

476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597

(7)

3.3. Comparing risk for PCA to typical AD

Comparing odds ratios and the nominal risk conferred in the whole PCA cohort against the most recently published meta-meta studies of typical AD [16], the effect size seen at the APOE locus was significantly less strong in PCA than for typical AD (PCA: OR 2.03, typical AD: OR 2.83, P5 .0007, PGC5 .001, see methods). Although there was

no evidence for risk in PCA vs controls, the risk effects at rs35349669 in INPP5D (P 5 .02, PGC 5 .02) and

rs2718058 upstream of NME8 (P 5 .03, PGC5 .04) both

were nominally different in PCA than typical AD. 3.4. Exploratory GWAS

As several different array platforms were used in the study, only 210,670 SNPs were genotyped across all samples. In this set, there was little evidence of inflation in the association test statistic (l 5 1.05). Only the proxy for the APOE E4 AD-risk allele, rs2075650, achieved genome-wide significance. We went on to analyze 5.9-M SNPs in the imputed data set. Aside from chromosome 19 (APOE locus), three loci on chromosomes 7, 2, and 6 respec-tively were of interest (Fig. 1). rs2525776 on chromosome 7, upstream of SEMA3C (OR 3.3 [2.1–5.1], P5 1.4 ! 1028, PGC5 4 ! 1028; Fig. 2C); rs76854344 on chromosome

2, upstream of CNTNAP5, (OR 1.9 [1.5–2.3], P 5 8.0 !

10210, PGC5 2 ! 1029;Fig. 2D), and rs72907046 on

chro-mosome 6, downstream of FAM46A (OR 3.2 [2.1–4.9], P5 1.1 ! 1029, PGC5 3 ! 1029;Fig. 2E) were all

asso-ciated with PCA. Restricting the analysis to the 82 individ-uals with biomarker/pathology evidence for underlying AD pathology, the corresponding odds ratios were similar: 3.8 [1.8–8.2], P 5 2.8 ! 1024 for rs2525776; 1.8 [1.2– 2.7], P5 2.1 ! 1023for rs76854344; and 2.5 [1.0–6.1], P5 2.7 ! 1022)for rs72907046. None of these three lociQ7

showed any evidence of association with typical AD on the IGAP AD-risk GWAS meta-analysis. A full list of sug-gestive associations P , 1024 based on the SNPs repre-sented on case and control arrays (the intersection SNPs) is available in a Supplementary Table 2, and the entire data set is available at the NHGRI-EBI GWAS Catalog.

4. Discussion

We report findings from a consortium to study genetic risk factors in PCA, a rare predominantly early-onset cogni-tive disorder characterized by progressive and dispropor-tionately posterior cortical dysfunction and atrophy, and usually associated with AD-type pathology. Our primary aim was to explore the relationship between PCA and a pre-determined list of candidate SNPs derived from studies of

Table 3

Results of the main analysis for candidate SNPs which were discovered in typical AD (ref[16], except *Seshadri et al. JAMA 2010:303; 1832–40,yHollingworth

et al. Nat Genet 2011:43:429–35) or DLB[17]

Candidate

SNP Chr Nearest gene

Typical AD/DLB OR

PCA vs control

P value PCA or PCA or CI

P Value for comparison of OR (PCA vs typical AD)

Cases MAF Control MAF rs3818361 1 CR1 1.17 6.71E204 1.38 (1.14–1.67) .09 0.24 0.19 rs744373 2 BIN1 1.17 0.04 1.20 (1.01–1.43) .77 0.32 0.28 rs35349669 2 INPP5D 1.08 0.21 0.89 (0.75–1.04) .02 0.46 0.49 rs6825004 4 SCARB2 0.78 0.42 0.92 (0.77–1.1) .38 0.29 0.31 rs7687945 4 SNCA 0.75 0.92 1.00 (0.85–1.18) .08 0.49 0.49 rs190982 5 MEF2C 0.93 0.66 0.99 (0.83–1.18) .50 0.40 0.40 rs10948363 6 CD2AP 1.10 0.67 0.97 (0.8–1.17) .19 0.27 0.27 rs11767557 7 EPHA1 0.90 0.39 0.92 (0.74–1.13) .73 0.19 0.20 rs2718058 7 NME8 0.93 0.17 1.12 (0.95–1.33) .03 0.39 0.36 rs1476679 7 ZCWPW1 0.91 0.73 1.05 (0.88–1.25) .12 0.31 0.30 rs11136000 8 CLU 0.87 0.27 0.91 (0.77–1.09) .58 0.37 0.40 rs28834970 8 PTK2B 1.10 0.27 1.10 (0.93–1.3) .98 0.37 0.35 rs10838725 11 CELF1 1.08 0.98 1.01 (0.84–1.2) .45 0.32 0.31 rs670139y 11 MS4A4E 1.08 0.77 0.97 (0.82–1.14) .19 0.40 0.41 rs983392 11 MS4A6A 0.90 0.71 1.04 (0.88–1.22) .09 0.42 0.41 rs3851179 11 PICALM 0.87 0.39 0.94 (0.79–1.11) .40 0.35 0.37 rs11218343 11 SORL1 0.77 0.57 0.81 (0.4–1.65) .88 0.01 0.02 rs17125944 14 FERMT2 1.14 0.73 0.94 (0.7–1.27) .21 0.08 0.09 rs10498633 14 SLC24A4 RIN3 0.91 0.14 0.86 (0.7–1.06) .62 0.20 0.23 rs3764650 19 ABCA7 1.20 0.02 1.39 (1.07–1.8) .28 0.12 0.09 rs2075650 19 APOE 2.83 6.24E214 2.03 (1.68–2.46) .0007 0.25 0.14 rs3865444y 19 CD33 0.91 0.69 0.95 (0.8–1.14) .61 0.30 0.31 rs597668* 19 EXOC3L2 BLOC1S3 MARK4 1.18 0.59 1.04 (0.84–1.3) .30 0.17 0.16 rs7274581 20 CASS4 0.88 0.52 1.12 (0.85–1.48) .09 0.10 0.09

OR, odds ratio; MAF, minor allele frequency; CI, confidence interval.

598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719

(8)

typical AD and DLB. Our main findings are the identifica-tion of genetic risk for PCA at some of the known AD-risk loci, but not at the two DLB-risk loci that were tested (Table 3). We demonstrate PCA-risk association with vari-ants in or near APOE, CR1, ABCA7 and BIN1 in the whole cohort, but only those at APOE, CR1, and ABCA7 remain nominally significant in the small molecularly defined sub-group. We also show evidence for nonoverlapping CIs of ge-netic risk at APOE, INPP5D, and NME8. Although this is by far the largest study of PCA to date, our relatively small sam-ple size remains underpowered to detect genome-wide significant associations at risk loci with small effect sizes, such as those shown to associate in AD-risk GWAS. Howev-er, our exploratory GWAS nominates the three novel loci, SEMA3C, CNTNAP5, and FAM46A, which achieved genome-wide significance, as potential genes of interest in PCA.

Although still small for genetic studies, this collection is the result of a global collaborative effort and is considerably larger than any previous studies of PCA reporting varying evidence that APOE is a risk factor [5–10]. Although the vast majority of samples in previous studies overlap with the present study, individual studies comprise,w25% of the current total. We found robust genetic association at APOE but with an odds ratio significantly smaller than those seen in typical AD. APOE is the best established and strongest risk factor for sporadic AD and has been associated not only with increased risk per se, but also with earlier disease onset[23], the rate of hippocampal atro-phy[24]and more memory led disease. However, the situa-tion is more complex in early-onset AD which is associated both with a greater proportion of nonamnestic presentations, and perhaps a relatively reduced proportion of APOE E4 car-riers[25]. Studies investigating APOE risk for PCA (or as it was sometimes previously defined, biparietal AD[6]), have shown mixed results, perhaps due to differences in case defi-nition, and the limited sample size of each study. In this study of PCA, by far the largest yet published, we confirm that APOE is a risk factor for PCA, but that it is a weaker risk factor than for typical AD.

Addressing risk factors other than APOE, the largest sin-gle previous study—which included samples also used in this analysis—found nominally significant association with SNPs in or near CLU, BIN1, and ABCA7[9]. We found ev-idence of association in the same direction for the SNP at ABCA7 both in the whole cohort and those with biomarker evidence for AD; for BIN1 in the whole cohort alone; but could not confirm that SNPs near to CLU are risk factors. Aside from APOE, only variants in CR1 surpassed our statis-tical threshold for multiple testing in the whole cohort; these variants also showed nominal significance in the biomarker positive subgroup. Estimates of effect size were greater in PCA than those in typical AD at each associated locus aside from APOE. Although these differences were only nomi-nally statistically significant, the four risk loci we report are among the strongest known common genetic risk factors

for typical AD (Table 3). CR1 has multiple functions including the regulation of complement and phagocytosis of immune complexes and pathogens, which are increas-ingly though to be relevant to AD pathogenesis [26]. ABCA7 may play a role in AD through regulation of phago-cytosis or lipid metabolism. BIN1 mechanisms in AD are un-clear but may be involved in endocytosis and the recycling of endocytic vesicles[27].

Although they did not confer significant alteration of PCA risk, we found that odds ratio confidence intervals for SNPs at or near to INPP5D and NME8 in PCA did not over-lap those of typical AD and showed directionally opposite effects, INPP5D was identified as a risk factor for AD in a recent large meta-analysis and plays an important role in a number of inflammatory processes. There is little evidence for the function of NME8 in the central nervous system, although a role in modification of oxidative stress has been proposed[28]. Although these findings were only nominally significant and need independent replication, they do raise the possibility that syndromic variants of AD may be differ-entially associated with alterations in certain risk genes, perhaps through altered responses to inflammation or stress. The results of our exploratory genome-wide study impli-cate three potential strong risk loci, near to CNTNAP5, FAM46A, and upstream of SEMA3C. The regions of strong LD with these associations did not include directly geno-typed SNPs across all platforms, and therefore false-positive associations related to differential accuracy of imputation between case and control arm of the study remain possible. With the caveat that these findings must therefore be considered preliminary and require follow-up replication in an independent sample and by direct genotyping, it is notable that all three genes have roles in processes poten-tially relevant to PCA. Contactin-associated protein-like 5 gene (CNTNAP5) belongs to a subgroup of the neurexin family of multidomain transmembrane proteins involved in cell adhesion and intercellular communication in the central nervous system and has been implicated as a risk factor for bipolar disorder and autism spectrum disorders[29]. Family with sequence similarity 46, member A1 (FAM46A), origi-nally C6orf37, is preferentially expressed within the neural retina[30] and has been implicated in cell signaling path-ways related to retinal neurodegeneration [31]. Class III semaphorins including Semaphorin 3C (SEMA3C) have been examined as potential modifying factors in neurode-generation through interactions with plexins and neuropi-lins. SEMA3C has been identified as a chemotrophic molecule influencing attractive guidance for cortical axon development[32]; the expression of SEMA3C and its recep-tors have been shown to influence the maturation of the vi-sual system [33]; and SEMA3C is also expressed in the hippocampus, where it has a role in influencing the afferent connections of the developing hippocampus and in partic-ular the ingrowth of septo-hippocampal connections [34], the major cholinergic connections implicated in learning and memory [35]. Finally, SEMA3C expression has been

J.M. Schott et al. / Alzheimer’s & Dementia- (2016) 1-10 7

720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768 769 770 771 772 773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792 793 794 795 796 797 798 799 800 801 802 803 804 805 806 807 808 809 810 811 812 813 814 815 816 817 818 819 820 821 822 823 824 825 826 827 828 829 830 831 832 833 834 835 836 837 838 839 840 841

(9)

shown to correlate with functional network connectivity within the brain[36]. Although at this stage speculative, it is possible therefore that perhaps subtle differences in cortical development might influence where pathology starts and/or how it spreads through the brain if a neurodegenera-tion process is initiated later in life. The fact that all three of these novel genetic risks showed nominal associations in the relatively small subset of individuals with biomarker evi-dence for AD and the absence of similar evievi-dence of associ-ation in any of these genes with IGAP studies of typical AD suggest that if confirmed, these loci may be specific risks for PCA due to Alzheimer’s disease.

The main limitation of our study is the necessarily modest sample size of this very rare disorder, noting that the case numbers presented here were only achievable through the establishment of an international consortium. We plan to continue to collect further samples to allow for replication in due course. Based on standard power cal-culations in case-control studies, even in the favorable sit-uation of completely accurate imputation of the functional SNP we are only adequately powered to detect effect sizes of OR .1.5, for common candidate SNPs, and the only common genetic risk factor of this strength in typical AD is APOE. We made comparisons with studies of patients diagnosed with typical AD; however, these patients and/or studies are different in multiple ways including the genotyping platforms used, later age at clinical onset, potential pathologic heterogeneity, and most probably dif-ferences in geographical location, all of which could confound the comparison. Although PCA is underpinned by AD pathology in most cases, we only had evidence for underlying AD in a proportion (w1/4), and we cannot confirm that the genetic risks we have determined are spe-cific for the AD variant of PCA rather than the syndrome of PCA or for young onset AD. However, allowing for the fact that the confidence intervals are inevitably large, it is notable that the estimates for the odds ratios for APOE, CR1, and ABCA7, and the putative genes identified in our exploratory GWAS were similar or larger in the proportion with molecular evidence for AD, suggesting that the risk we identify are likely to be for the AD variant of PCA, rather than for the syndrome per se.

One of the major outstanding issues in neurodegenerative disease research is an explanation for the often very striking phenotypic heterogeneity underpinned by the same broad core pathology. Possibilities for phenotype modification include demographic and environmental factors, including age at onset, or perhaps more likely complex gene and/or environment interactions and/or factors related to the mis-folded proteins and their propagation, tissue or network selectivity and toxicity. The results of this study suggest that subtle differences in established risk factors may be associated with some of this heterogeneity and provide test-able suggestions for novel genes that may influence the development of the hippocampal and visual system, which may influence the development of the PCA phenotype

rela-tive to other syndromes. If confirmed in future studies, next generation sequencing may be useful in determining whether these findings might be underpinned by rare variants with large effect sizes. More broadly, genetic investigation of well-phenotyped AD variants may provide important in-sights into disease biology in typical AD.

Acknowledgments

We acknowledge the support of the Alzheimer’s Association’s International Society to Advance Alzheimer’s Research and Treatment (ISTAART), Alzheimer’s Research UK, the UK Alzheimer’s Society and the Raena Frankel-Pollen Fund. This work was supported by the NIHR UCL/UCLH Biomed-ical Research Centre, the NIHR Queen Square Dementia Biomedical Research Unit, the Wolfson Foundation, and the UK Medical Research Council. The Dementia Research Centre is an Alzheimer’s Research UK Coordinating Center. N.C.F. and M.N.R. are NIHR senior investigators. J.T.B. re-ceives funding from the UK Alzheimer’s Society. This work was also supported by the Medical Research Council Demen-tias Platform UK (MR/L023784/1 and MR/009076/1); Alz-heimer’s Research UK (Senior Research Fellowship to S.C., Research Fellowship to T.S.); the Dunhill Medical Trust (grant number R337/0214 to S.C.); ESRC/NIHR (ES/K006711/1 to S.C.); EPSRC (EP/M006093/1 to S.C.); Alzheimer’s Associa-tion (MNIRGD 2013 award to M.M.C.); Mayo Alzheimer’s Disease Research Center (P50 AG0016574 to D.W.D., N.E.-T., N.R.G.-R., B.F.B., D.S.K., R.C.P.); National Institute on Aging (R01 AG032990 and U01 AG046139 to N.E.-T.; R01-AG045611 to G.D.R.; and P50-AG023501 to B.L.M. and G.D.R.); National Institute of Neurological Disorders and Stroke (R01 NS080820 to N.E.-T.); John Douglas French Alzheimer’s Foundation (to G.D.R., B.L.M.); andState of Cal-Q8

ifornia Department of Health Services Alzheimer’s Disease Research Centre of California (04-33516, B.L.M.). D.G. and E.S. are supported by Italian Ministry of Health (Ricerca Cor-rente). W.M.v.d.F., P.S. receive funding from PERADES-JPND. E.L. receives funding from Stichting Dioraphte. Case ascertainment in Australia was funded by grants from the Na-tional Health and Medical Research Council of Australia (NHMRC program grant #1037746) and the Australian Research Council Centre of Excellence in Cognition and its Disorders Memory Node (#CE110001021); G.H. is a NHMRC senior principal research fellow (#1079679). We would like to thank the participants in the Framingham, Geisinger, WTCCC, and KORA studies used as controls in this article. The Fra-mingham Heart Study is conducted and supported by the Na-tional Heart, Lung, and Blood Institute (NHLBI) in collaboration with Boston University (Contract No. N01-HC-25195). This article was not prepared in collaboration with investigators of the Framingham Heart Study and does not necessarily reflect the opinions or views of the Framing-ham Heart Study, Boston University, or NHLBI. SHARe Illu-mina genotyping was provided under an agreement between Illumina and Boston University. Samples and data in this study

842 843 844 845 846 847 848 849 850 851 852 853 854 855 856 857 858 859 860 861 862 863 864 865 866 867 868 869 870 871 872 873 874 875 876 877 878 879 880 881 882 883 884 885 886 887 888 889 890 891 892 893 894 895 896 897 898 899 900 901 902 903 904 905 906 907 908 909 910 911 912 913 914 915 916 917 918 919 920 921 922 923 924 925 926 927 928 929 930 931 932 933 934 935 936 937 938 939 940 941 942 943 944 945 946 947 948 949 950 951 952 953 954 955 956 957 958 959 960 961 962 963

(10)

were also provided by the Geisinger MyCode Project. Funding for the MyCode project was provided by a grant from Commonwealth of Pennsylvania and the Clinic Research Fund of Geisinger Clinic. Funding support for the genotyping of the MyCode cohort was provided by a Geisinger Clinic operating funds and an award from the Clinic Research Fund. The data sets used for the analyses described in this article were obtained from dbGaP at http://www.ncbi. nlm.nih.gov/gap through dbGaP accession number phs000381.v1.p1. This study also makes use of the data gener-ated by the Wellcome Trust Case Control Consortium. A full list of the investigators who contributed to the generation of the data is available from www.wtccc.org.uk. Funding for the project was provided by the Wellcome Trust under award 076113.

Q9

Supplementary data

Supplementary data related to this article can be found online athttp://dx.doi.org/10.1016/j.jalz.2016.01.010.

RESEARCH IN CONTEXT

1. Systematic review: Reviewing the literature for pub-lications investigating the genetics of posterior cortical atrophy (PCA), there is conflicting evidence for the role of APOE in the PCA variant of Alz-heimer’s disease AD), and limited evidence for the more recently identified genetic risks for AD. 2. Interpretation: Through the establishment of an

inter-national consortium to create the largest study exploring the genetics of PCA to date, we demon-strate that (1) APOE is a risk factor for PCA but con-fers a smaller risk than for typical AD; (2) some of the genetic risks for typical AD are also associated with PCA risk; and (3) nominate three novel risk loci for PCA.

3. Future directions: These data provide clear directions and testable hypotheses for future studies, including (1) the establishment of a replication cohort and (2) investigation of the identified genes as factors influ-encing selective vulnerability in AD.

References

[1] Crutch SJ, Lehmann M, Schott JM, Rabinovici GD, Rossor MN, Fox NC. Posterior cortical atrophy. Lancet Neurol 2012;11:170–8. [2] Lehmann M, Koedam EL, Barnes J, Bartlett JW, Ryan NS,

Pijnenburg YAL, et al. Posterior cerebral atrophy in the absence of medial temporal lobe atrophy in pathologically-confirmed Alz-heimer’s disease. Neurobiol Aging 2012;33:627.

[3] Lehmann M, Ghosh PM, Madison C, Laforce R. Diverging patterns of amyloid deposition and hypometabolism in clinical variants of prob-able Alzheimer’s disease. Brain 2013;136:844–58.

[4] Ossenkoppele R, Schonhaut DR, Baker SL, O’Neil JP, Janabi M, Ghosh PM, et al. Tau, amyloid, and hypometabolism in a patient with posterior cortical atrophy. Ann Neurol 2015;77:338–42. [5] Tang-Wai DF, Graff-Radford NR, Boeve BF, Dickson DW, Parisi JE,

Crook R, et al. Clinical, genetic, and neuropathologic characteristics of posterior cortical atrophy. Neurology 2004;63:1168–74.

[6] Schott JM, Ridha BH, Crutch SJ, Healy DG, Uphill JB, Warrington EK, et al. Apolipoprotein e genotype modifies the pheno-type of Alzheimer disease. Arch Neurol 2006;63:155–6.

[7] van der Flier WM, Schoonenboom SN, Pijnenburg YA, Fox NC, Scheltens P. The effect of APOE genotype on clinical phenotype in Alzheimer disease. Neurology 2006;67:526–7.

[8] Snowden JS, Stopford CL, Julien CL, Thompson JC, Davidson Y, Gibbons L, et al. Cognitive Phenotypes in Alzheimer’s Disease and Genetic Risk. Cortex 2007;43:835–45.

[9] Carrasquillo MM, Khan QU, Murray ME, Krishnan S, Aakre J, Pankratz VS, et al. Late-onset Alzheimer disease genetic variants in posterior cortical atrophy and posterior AD. Neurology 2014; 82:1455–62.

[10] Carrasquillo MM, Barber I, Lincoln SJ, Murray ME, Camsari GB, Khan QU, et al. Evaluating pathogenic dementia variants in posterior cortical atrophy. Neurobiol Aging 2016;37:38–44.

[11] Crutch SJ, Schott JM, Rabinovici GD, Boeve BF, Cappa SF, Dickerson BC, et al. Shining a light on posterior cortical atrophy. Alz-heimers Dement 2013;9:463–5.

[12] Mendez MF, Ghajarania M, Perryman KM. Posterior cortical atrophy: clinical characteristics and differences compared to Alzheimer’s dis-ease. Dement Geriatr Cogn Disord 2002;14:33–40.

[13] Delaneau O, Zagury JF, Marchini J. Improved whole-chromosome phasing for disease and population genetic studies. Nat Methods 2013;10:5–6.

[14] Howie BN, Donnelly P, Marchini J. A flexible and accurate genotype imputation method for the next generation of genome-wide association studies. PLoS Genet 2009;5:e1000529.

[15] Marchini J, Howie B. Genotype imputation for genome-wide associa-tion studies. Nat Rev Genet 2010;11:499–511.

[16] Marchini J, Howie B, Myers S, McVean G, Donnelly P. A new multi-point method for genome-wide association studies by imputation of genotypes. Nat Genet 2007;39:906–13.

[17] Purcell S, Neale B, Todd-Brown K, Thomas L, Ferreira MA, Bender D, et al. PLINK: a tool set for whole-genome association and population-based linkage analyses. Am J Hum Genet 2007;81:559–75.

[18] Lambert JC, Ibrahim-Verbaas CA, Harold D, Naj AC, Sims R, Bellenguez C, et al. Meta-analysis of 74,046 individuals identifies 11 new susceptibility loci for Alzheimer’s disease. Nat Genet 2013; 45:1452–8.

[19] Bras J, Guerreiro R, Darwent L. Genetic analysis implicates APOE, SNCA and suggests lysosomal dysfunction in the etiology of dementia with Lewy bodies. Hum Mol Genet 2014;23:6139–46.

[20] Corder EH, Saunders AM, Strittmatter WJ. Gene dose of apolipopro-tein E type 4 allele and the risk of Alzheimer’s disease in late onset families. Science 1993;261:921–3.

[21] Mead S, Poulter M, Uphill J, Beck J, Whitfield J, Webb TE, et al. Ge-netic risk factors for variant Creutzfeldt–Jakob disease: a genome-wide association study. Lancet Neurol 2009;8:57–66.

[22] Klein RJ, Zeiss C, Chew EY, Tsai JY, Sackler RS, Haynes C, et al. Complement factor H polymorphism in age-related macular degener-ation. Science 2005;308:385–9.

[23] Thambisetty M, An Y, Tanaka T. Alzheimer’s disease risk genes

and the age-at-onset phenotype. Neurobiol Aging 2013;34:

2696.e1–5.

[24] Mori E, Lee K, Yasuda M, Hashimoto M, Kazui H, Hirono N, et al. Accelerated hippocampal atrophy in Alzheimer’s disease with apoli-poprotein E epsilon4 allele. Ann Neurol 2002;51:209–14.

J.M. Schott et al. / Alzheimer’s & Dementia- (2016) 1-10 9

964 965 966 967 968 969 970 971 972 973 974 975 976 977 978 979 980 981 982 983 984 985 986 987 988 989 990 991 992 993 994 995 996 997 998 999 1000 1001 1002 1003 1004 1005 1006 1007 1008 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 1074 1075 1076 1077 1078 1079 1080 1081 1082 1083 1084 1085

(11)

[25] van der Flier WM, Pijnenburg YA, Fox NC, Scheltens P. Early-onset versus late-onset Alzheimer’s disease: the case of the missing APOE

3

4 allele. Lancet Neurol 2011;10:280–8.

[26] Crehan H, Holton P, Wray S, Pocock J, Guerreiro R, Hardy J. Comple-ment receptor 1 (CR1) and Alzheimer’s disease. Immunobiology 2012;217:244–50.

[27] Prokic I, Cowling BS, Laporte J. Amphiphysin 2 (BIN1) in physiology and diseases. J Mol Med 2014;92:453–63.

[28] Rosenthal SL, Kamboh MI. Late-Onset Alzheimer’s Disease Genes and the Potentially Implicated Pathways. Curr Genet Med Rep 2014; 2:85–101.

[29] Pagnamenta AT, Bacchelli E, de Jonge MV, Mirza G, Scerri TS, Minopoli F, et al. Characterization of a family with rare deletions in CNTNAP5 and DOCK4 suggests novel risk loci for autism and dyslexia. Biol Psychiatry 2010;68:320–8.

[30] Lagali PS, Kakuk LE, Griesinger IB, Wong PW, Ayyagari R. Identifi-cation and characterization of C6orf37, a novel candidate human retinal disease gene on chromosome 6q14. Biochem Biophys Res Commun 2002;293:356–65.

[31] Barragan I, Borrego S, Abd El-Aziz MM, El-Ashry MF, Abu-Safieh L, Bhattacharya SS, et al. Genetic Analysis of FAM46A

in Spanish Families with Autosomal Recessive Retinitis Pigmen-tosa: Characterisation of Novel VNTRs. Ann Hum Genet 2008; 72:26–34.

[32] Bagnard D, Thomasset N, Lohrum M, P€uschel AW, Bolz J. Spatial dis-tributions of guidance molecules regulate chemorepulsion and chemo-attraction of growth cones. J Neurosci 2000;20:1030–5.

[33] Sharma A, LeVaillant CJ, Plant GW, Harvey AR. Changes in expression of Class 3 Semaphorins and their receptors during devel-opment of the rat retina and superior colliculus. BMC Dev Biol 2014;14:34.

[34] Steup A, Lohrum M, Hamscho N, Savaskan NE, Ninnemann O, Nitsch R, et al. Sema3C and Netrin-1 Differentially Affect Axon Growth in the Hippocampal Formation. Mol Cell Neurosci 2000; 15:141–55.

[35] Dutar P, Bassant MH, Senut MC. The septohippocampal pathway: structure and function of a central cholinergic system. Physiol Rev 1995;75:393–427.

[36] Richiardi J, Altmann A, Milazzo AC, Chang C, Chakravarty MM, Banaschewski T, et al. Brain Networks. Correlated gene expression supports synchronous activity in brain networks. Science 2015; 348:1241–4. 1086 1087 1088 1089 1090 1091 1092 1093 1094 1095 1096 1097 1098 1099 1100 1101 1102 1103 1104 1105 1106 1107 1108 1109 1110 1111 1112 1113 1114 1115 1116 1117 1118 1119 1120 1121 1122 1123 1124 1125 1126 1127 1128 1129 1130 1131 1132 1133 1134 1135 1136 1137 1138 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 1204 1205 1206 1207

Figure

Fig. 1. Manhattan plot of autosomes with threshold for genome-wide significance (P , 5 ! 10 28 ) indicated by the red line

Références

Documents relatifs

We therefore investi- gated whether individuals with ASD carrying a rare CNTN6 variant had other rare (minor allele frequencyo1% in 1000 genomes or ExAC) and deleterious (considered

(*) Subjects with no missing value for the risk factors in the multivariate model; Number of subjects excluded from analyses because of missing values: 129 men and 112 women; OR,

Moreover, many studies have focused on selected occupational populations recruited from companies characterized by high levels of exposure to risk factors for

This section provides a brief overview of current knowledge on the relationships between key NCD behavioural risk factors (e.g., physical inactivity, poor nutrition, smoking,

However, it is important to anticipate weight gain after tobacco cessation [42], given that weight gain can be associated with metabolic disorders and overweightness and obesity,

El- evated plasma total homocysteine (tHcy) concentration is now widely accepted as a major independent risk factor for cerebrovascular, peripheral vascular disease and may explain,

Einfahrung und Etablierung eines Qualit~itsmanagements muss man sich zun~ichst mit dem Gedanken vertraut machen, dass dies ein un- aufhOrlicher Prozess sein wird,

Multiplication/Division Avec Les Chiffres Romains (A) Réponses.. Répondez à chaque question en