Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2015 Jan 1.
Published in final edited form as: J Alzheimers Dis. 2015 Jan 1;43(2):649–655. doi: 10.3233/JAD-140729

Genetic determinants of disease progression in Alzheimer’s disease

Xingbin Wang 1,2, Oscar L Lopez 3,4, Robert A Sweet 3,4,5,7, James T Becker 3,4, Steven T DeKosky 6, Mahmud M Barmada 1, F Yesim Demirci 1, M Ilyas Kamboh 1,4,5
PMCID: PMC4245313  NIHMSID: NIHMS609908  PMID: 25114068

Abstract

There is a strong genetic basis for late-onset of Alzheimer’s disease (LOAD); thus far 22 genes/loci have been identified that affect the risk of LOAD. However, the relationships among the genetic variations at these loci and clinical progression of the disease have not been fully explored. In the present study, we examined the relationships of 22 known LOAD genes to the progression of AD in 680 AD patients recruited from the University of Pittsburgh Alzheimer’s Disease Research Center. Patients were classified as “rapid progressors” if the MMSE changed ≥3 points in 12 months and “slow progressors” if the MMSE changed ≤2 points. We also performed a genome-wide association study in this cohort in an effort to identify new loci for AD progression. Association analysis between SNPs and the progression status of the AD cases was performed using logistic regression model controlled for age, gender, dementia medication use, psychosis, and hypertension. While no significant association was observed with either APOE*4 (p=0.94) or APOE*2 (p=0.33) with AD progression, we found multiple nominally significant associations (p<0.05) either within or adjacent to seven known LOAD genes (INPP5D, MEF2C, TREM2, EPHA1, PTK2B, FERMT2 and CASS4) that harbor both risk and protective SNPs. Genome-wide association analyses identified four suggestive loci (PAX3, CCRN4L, PIGQ and ADAM19) at p<1E-05. Our data suggest that short-term clinical disease progression in AD has genetic basis. Better understanding of these genetic factors could help to improve clinical trial design and potentially affect the development of disease modifying therapies.

Keywords: LOAD, GWAS, MMSE, AD progression

Introduction

Late-onset Alzheimer’s disease (LOAD), is a complex multifactorial neurodegenerative disease and the leading cause of dementia among the elderly [1]. Currently, there are approximately 5 million AD cases in the United States, and about 81.1 million cases worldwide [2]. Due to its long clinical course, AD is a major public health problem. Genetic susceptibility due to multiple genes and interactions among them influence the risk of AD, which has a strong genetic basis with heritability estimates up to 80% [3].

APOE is the major susceptibility gene for LOAD. Genome-wide association studies (GWAS) have identified 21 additional susceptibility loci including BIN1, INPP5D, MEF2C, CD2AP,HLA-DRB1/HLA-DRB5, TREM2, EPHA1,NME8, ZCWPW1, CLU, PTK2B, CELF1, MS4A6A, PICALM, SORL1, FERMT2, SLC2A4, DSG2, ABCA7, CD33, and CASS4[4-9]. Recently rare variants in TREM2 have also been reported to be associated with LOAD risk [10]. In addition to AD risk, genetic variation at these loci may also affect components of the natural history of the clinical dementia. However, the relationship between these known loci and dementia progression has not been explored extensively, highlighting the need to use other approaches in order to identify additional genes involved in the clinical and pathological manifestations of AD.

Large populations of well-characterized and longitudinally followed cases are necessary for such analyses. AD is characterized by gradual cognitive and functional decline, relating to the progressive degeneration of structure and chemistry of the brain over time. The patients’ ability to remember, understand, communicate and reason gradually declines, with largely non-uniform rates of progression[11]. Many factors can affect the rate of clinical progression, including brain atrophy rates[12-14], patterns of regional brain atrophy[15], ventricular enlargement[16], neuropsychological and cerebral profiles[17], vascular factors[18], and immunological factors[19].

Genetic factors may also affect the rate of AD progression [20, 21]. The known AD risk genes are good candidates for assessing whether their genetic variation affects the natural history of AD. In this study we used the rate of AD clinical progression, as indexed by change in MMSE score after 12 months follow-up as a phenotype and hypothesized that like disease risk, disease progression also has a genetic basis. We used our previously described GWAS data set [23, 24] to 1) examine the role of 22 known LOAD genes with AD progression in 680 well-characterized and longitudinally followed-up AD patients, and 2) to perform a GWAS analysis in an effort to identify additional loci for AD progression, irrespective if they are genome-wide significant or not, for hypothesis generation.

Materials and Methods

Subjects

The AD patients were recruited from Alzheimer’s Research Program (ARP; 1983-1988) and the Alzheimer’s Disease Research Center (ADRC) at the University of Pittsburgh (1988 to present). A total of 1,886 Probable AD patients were examined between April 1983 and December 2005; details of the cohort are described elsewhere [22]. All subjects received an extensive neuropsychiatric evaluation including medical history and physical examination, neurological history and examination, semi-structured psychiatric interview, neuroimaging, and neuropsychological assessment.

Follow-up measurements, definition of Rapid Progression

For the purpose of this study, the rate of progression was defined by the change in the Mini Mental State Examination (MMSE) score from baseline evaluation to the clinic visit approximately 1 year later. Subjects whose MMSE scores changed ≥3 points/year were classified as “rapid progressors” and those whose scores change ≤2 points/year were classified as “slow progressors” [22].

Genotyping and quality control (QC) of genotype data

Samples were genotyped using the Illumina Omni1-Quad chip as described previously [23, 24] SNPs with call rate <98% and minor allele frequency (MAF)<1%, and failing to adhere to the Hardy-Weinberg equilibrium (HWE) test (P<1E-06) were removed. Genotypes for two APOE SNPs, rs429358 (E*4) and rs7412 (E*2) were determined either as previously described [25] or using TaqMan SNP genotyping assays. For GWAS, a total of 803,323 QC-passed SNPs were selected for analysis.

Statistical analysis

We used t-tests and χ2-tests to analyze demographic and clinical differences between rapid progressors and slow progressors. The association between AD progression status and SNPs was tested using an additive logistic regression model that included age, dementia medication use (taking any cholinesterase inhibitor (AChEI) or memantine), psychosis (at any time during follow-up), hypertension and the top four principal components derived from our GWAS data as covariates. The Versatile Gene-based Associations (VEGA) analyses [26] were performed for the known 22 LOAD genes and 4 suggestive genes identified in this study. In these genes, LD -Select Tag SNP selection algorithm was implemented in Haploview [27] with an r2 cutoff of 0.8 to select independent SNPs within each gene plus 10kb on either side of the gene. All statistical tests were two-sided. All analyses were done in R and/or PLINK[28].

Results

Characteristics of rapid and normal progressors

There were 373 slow progressors and 307 rapid progressors among the 680 patients included in this analysis. Table 1 shows the demographic and clinical characteristics of the patients by progression type. The rapid progressors were younger (p=0.05), had more hypertension (p=0.04) and less psychotic symptoms (p=0.01) and used less dementia medications (p=6.5E-05) than patients who were classified as slow progressors. Since the effect of genetic factors on AD progression may have been confounded by those variables, they were included in the additive logistic regression model.

Table 1.

Demographic and clinical characteristics of rapidly progressive AD patients and normally progressive AD patients

slower
(n=373)
Rapid
(n=307)
t-test/χ2 p-value
Age 77. 6 + 6.0 76.6 + 6.3 2.0 0.05
Gender (male/female) 136/237 119/188 0.28 0.59
Education ( years) 12.78 + 3.1 12.96 + 3.0 −0.75 0.45
Baseline MMSE 19.00+ 4.75 18.82 + 5.50 0.45 0.65
Medication (Yes/No) 291/82 196/111 15.95 6.50E-05
Psychosis (Yes/No) 125/248 133/174 6.47 0.01
Heart Disease (Yes/No) 74/299 68/239 0.41 0.52
Diabetes Mellitus (Yes/No) 29/344 27/280 0.11 0.73
Hypertension (Yes/No) 196/177 136/171 4.26 0.04
Depression Yes/No) 59/314 52/252 0.08 0.77

*Age: patients’ age at entry; MMSE: the mean Mini-Mental state examination scores; Education: the years of getting education; Medication: taking any cholinesterase inhibitor (AChEI) treatment or not; psychosis: the presence or absence of psychotic symptom

Association of known LOAD genes with AD Progression

The associations of AD progression with genetic variations in known 22 LOAD genes are presented in Table 2. SNPs in 7 genes (INPP5D, MEF2C, TREM2, EPHA1, PTK2B, FERMT2, and CASS4) were associated with AD progression at the nominal cutoff of p<0.05. While the top SNPs in 4 genes were associated with slow AD progression (PERMT2/rs7160582, OR=1.62; p=1.08E-02., INPP5D/rs1057258, OR=1.48; p=0.01, PTK2B/rs4732720, OR=1.34; p=0.01, and TREM2/rs7748777, OR=1.34; p=0.011), SNPs in 3 genes were associated with rapid progression (MEF2C/rs9293505, OR=0.275; p=0.03, EPHA1/rs11768549, OR=0.246; p=0.037, and CASS4/rs16979934, OR=0.596; p=0.033). In the gene-based analysis, 2 of these 7 genes remained significant (PERMT2, p=0.04) or had borderline significance (INPP5D, p=0.07).

Table 2.

Results of Association Analysis between LOAD Genes and the Progression of AD

Single locus analysis Gene-based analysis

CHR Gene Total SNPs Tagger SNPs Lead SNP BP A MAF OR P1 SNPs
(p<0.05)
Test P2



2 BIN1 34 22 rs6750960 127551475 A 0.14 0.7467 0.07 0 26.24557 0.56
2 INPP5D 60 50 rs1057258 233780368 A 0.18 1.476 0.01 8 95.60424 0.07
5 MEF2C 33 25 rs9293505 88222225 A 0.02 0.2751 0.03 1 21.10667 0.58
6 CD2AP 23 11 rs2894740 47689800 G 0.40 1.202 0.12 0 21.57318 0.41
6 HLA-DRB1
/HLA-DRB5
45 34 rs6597017 3902270 A 0.28 1.222 0.13 0 27.95059 0.19
6 TREM2 5 5 rs7748777 41241784 A 0.46 1.34 0.01 1 9.091647 0.15
7 EPHA1 21 18 rs11768549 142805275 A 0.01 0.246 0.04 1 16.09665 0.57
7 NME8 59 29 rs12671838 37906849 A 0.03 1.857 0.06 0 47.42309 0.62
7 ZCWPW1 12 6 rs5015756 99851393 A 0.43 0.8638 0.19 0 10.45819 0.44
8 CLU 15 10 rs9331947 27510794 G 0.04 1.621 0.12 0 8.238591 0.74
8 PTK2B 99 36 rs4732720 27293625 G 0.49 1.336 0.01 14 147.9678 0.17
11 CELF1 10 5 rs2242081 47456843 G 0.46 1.174 0.16 0 9.485295 0.39
11 MS4A6A 6 4 rs12453 59702321 G 0.37 0.8959 0.34 0 3.104919 0.61
11 PICALM 28 16 rs17148741 85443439 A 0.02 0.5551 0.23 0 8.709003 0.93
11 SORL1 61 40 rs2276412 120966056 A 0.02 1.98 0.15 0 28.9085 0.92
14 FERMT2 27 12 rs7160582 52411195 A 0.10 1.629 0.01 9 68.89701 0.04
17 SLC2A4 12 6 rs3744404 7133916 A 0.02 1.376 0.51 0 1.059282 0.99
18 DSG2 30 14 rs12604517 27378422 A 0.24 1.25 0.09 0 18.37447 0.69
19 ABCA7 32 19 rs4147914 1000269 A 0.16 1.244 0.15 0 25.98096 0.56
19 APOE 20 14 lab_rs7412 50103919 A 0.03 0.7496 0.34 0 4.800756 0.98
19 CD33 10 6 rs1803254 56434956 C 0.07 0.6993 0.09 0 6.215891 0.62
20 CASS4 28 18 rs16979934 54460186 G 0.06 0.5956 0.03 1 18.45753 0.67

*CHR: chromosome; Gene:candidate gene associated with AD; Total SNPs: the total number of SNPs located in the region 10kb before and after the gene; Tagger SNPs: the number of tagger SNPs in the region; Lead SNP: most significant SNP in the gene region; BP: base-pair position of the lead SNP under building version 36; A: the minor allele of the lead SNP; MAF: minor allele frequency; OR: odds ratio; P1: the p-value in the single locus analysis; SNPs (p<0.05): total number of SNPs with p < 5E-02 in the gene region; Test: the test statistics in gene-based analysis; P2: the p-value in the gene-based analysis.

New loci associated with AD Progression in GWAS

Next we examined our genome-wide association data in order to identify new loci for disease progression. Quantile-quantile (QQ) plot of the observed and expected p-values is shown in Supplementary Figure 1, and the Manhattan plot showing association signals is presented in Supplementary Figure 2. We identified four suggestive novel loci with p<1E-05. The top SNP, rs348987 (p=3.32E-06), was located near PAX3 on chromosome 2 at position 119kb. There were 19 additional SNPs with p<0.05 in this region (Table 3).The other three top SNPs were, CCRN4L /rs13116075, p=7.94E-06 on chromosome 4, PIGQ /rs2071979, p=8.17E-06 on chromosome 16 and ADAM19 /rs2277027, p=9.55E-06 on chromosome 5. The regional association plots containing SNPs within 500kb on either side of the top SNP in the 4 suggestive loci are shown in Supplementary Figures 3-6. We also performed gene-based analyses on the four genes and three of them (CCRN4L, PIGQ, ADAM19) demonstrated significant associations with AD progression (p<0.05).

Table 3.

Novel loci Associated with AD progression (P<1E-05)

Single locus analysis Gene-based analysis

CHR Gene Total SNPs Tagger SNPs Lead SNP BP A MAF OR P1 SNPs
(p<0.05)
Test P2



2 PAX3 50 26 rs348987 222653295 A 0.46 0.574 3.32E-06 20 64.71445 0.18
4 CCRN4L 4 4 rs13116075 140149482 G 0.15 0.496 7.94E-06 4 32.4932 0.0001
5 ADAM19 55 36 rs2277027 156864954 C 0.35 1.737 9.55E-06 18 192.15 0.002
16 PIGQ 10 4 rs2071979 564115 G 0.40 0.5918 8.17E-06 9 148.9967 0.00002

Discussion

Among the known LOAD genes, INPP5D, MEF2C, TREM2, EPHA1, PTK2B, FERMT2 and CASS4 revealed nominal associations (p<0.05) with dementia progression and two of them (PERMT2 and INPP5D) survived in the gene-based analysis. Although none of the observed associations survived after adjusting for multiple comparisons, we believe they may provide insight for future studies as they are present in confirmed genes for LOAD, which in addition to affecting risk may also affect components of natural history of AD. Our findings, together with a recently published study showing association of PICALM/rs3851179 with dementia progression [29], supports this hypothesis ; Although we did not replicate this result in our samples for the same SNP (p =0.12), the direction of allelic effect was the same, suggesting that this may be a weak, but genuine association.

Our GWAS analysis identified four suggestive loci (PAX3, CCRN4L, PIGQ and ADAM19) with significance of p<1E-05. The most significant association was identified 119kb from the 3′ region of the PAX3 gene on chromosome 2q35 (rs348987; p=3.32E-06). Although the associated SNPs were not present in an annotated gene in this region, the nearby PAX3 is a reasonable candidate gene that codes for a transcription factor. Down-regulation of PAX3 has been attributed to altered signaling pathways involving cell cycle, apoptosis, cell adhesion, cytoskeletal remodeling, and development [30]. Mutations in PAX3 are associated with Waardenberg syndrome [31-33]. Furthermore, an intronic SNP in CASS4, a recently implicated gene for LOAD [9], has been suggested to affect the PAX3 binding motif [34]. The next most significant SNP (rs13116075; p=7.94E-06) was located in the CCRN4L gene on chromosome 4q31, which is expressed in the brain [35], and genetic variation in this gene has been shown previously to affect body mass index [36]. The third top SNP resides on chromosome 16p13 near PIGQ/RAB40C (rs2071979; p=8.17E-06). RAB40C is a member of the Rab family of small GTPases that play important roles in neuronal and glial metabolism [37]. Another nearby gene in this region, RAB11FIP3, interacts with and regulates Rab GTPases, suggesting a potential combined significance of these functionally related genes in AD progression.

Limitations of our study include the relatively small sample sizes in both the rapid and slow AD progression groups, and variability of duration of time of follow-up of the cases for cognitive decline. Dementia medications affect individuals’ rates of decline [22], although we adjusted for this in the logistic regression models. Further, clinical disease progression is very complex, and many unknown demographic and clinical variables (e.g. other medical illnesses and sources of disability) not assessed in this study may have confounded our results. Because of the relatively small sample size, our GWAS findings are meant for only hypothesis generation for future larger studies.

In conclusion, our data suggest that short-term clinical disease progression in AD has genetic basis as we observed nominal associations with some known LOAD genes. Our secondary GWAS analysis identified 4 suggestive loci that, although not meeting the genome-wide significant threshold of p<5E-08, are potential candidate genes for AD clinical progression that warrant follow-up studies in larger data sets.

Supplementary Material

Supplementary Figures

Acknowledgments

This study was supported by the National Institutes of Health grants AG030653, AG041718, AG005133 and AG027224. The content is solely the responsibility of the authors and does not necessarily represent the official views of the Department of Veterans Affairs, the National Institutes of Health, or the United States Government.

Footnotes

Conflict of Interest

The authors declare no conflict of interest.

References

  • [1].Evans DA, Funkenstein HH, Albert MS, Scherr PA, Cook NR, Chown MJ, Hebert LE, Hennekens CH, Taylor JO. Prevalence of Alzheimer’s disease in a community population of older persons. Higher than previously reported. JAMA. 1989;262:2551–2556. [PubMed] [Google Scholar]
  • [2].Ballard C, Gauthier S, Corbett A, Brayne C, Aarsland D, Jones E. Alzheimer’s disease. Lancet. 2011;377:1019–1031. doi: 10.1016/S0140-6736(10)61349-9. [DOI] [PubMed] [Google Scholar]
  • [3].Gatz M, Reynolds CA, Fratiglioni L, Johansson B, Mortimer JA, Berg S, Fiske A, Pedersen NL. Role of genes and environments for explaining Alzheimer disease. Arch Gen Psychiatry. 2006;63:168–174. doi: 10.1001/archpsyc.63.2.168. [DOI] [PubMed] [Google Scholar]
  • [4].Lambert JC, Heath S, Even G, Campion D, Sleegers K, Hiltunen M, Combarros O, Zelenika D, Bullido MJ, Tavernier B, Letenneur L, Bettens K, Berr C, Pasquier F, Fievet N, Barberger-Gateau P, Engelborghs S, De Deyn P, Mateo I, Franck A, Helisalmi S, Porcellini E, Hanon O, de Pancorbo MM, Lendon C, Dufouil C, Jaillard C, Leveillard T, Alvarez V, Bosco P, Mancuso M, Panza F, Nacmias B, Bossu P, Piccardi P, Annoni G, Seripa D, Galimberti D, Hannequin D, Licastro F, Soininen H, Ritchie K, Blanche H, Dartigues JF, Tzourio C, Gut I, Van Broeckhoven C, Alperovitch A, Lathrop M, Amouyel P. Genome-wide association study identifies variants at CLU and CR1 associated with Alzheimer’s disease. Nat Genet. 2009;41:1094–1099. doi: 10.1038/ng.439. [DOI] [PubMed] [Google Scholar]
  • [5].Harold D, Abraham R, Hollingworth P, Sims R, Gerrish A, Hamshere ML, Pahwa JS, Moskvina V, Dowzell K, Williams A, Jones N, Thomas C, Stretton A, Morgan AR, Lovestone S, Powell J, Proitsi P, Lupton MK, Brayne C, Rubinsztein DC, Gill M, Lawlor B, Lynch A, Morgan K, Brown KS, Passmore PA, Craig D, McGuinness B, Todd S, Holmes C, Mann D, Smith AD, Love S, Kehoe PG, Hardy J, Mead S, Fox N, Rossor M, Collinge J, Maier W, Jessen F, Schurmann B, van den Bussche H, Heuser I, Kornhuber J, Wiltfang J, Dichgans M, Frolich L, Hampel H, Hull M, Rujescu D, Goate AM, Kauwe JS, Cruchaga C, Nowotny P, Morris JC, Mayo K, Sleegers K, Bettens K, Engelborghs S, De Deyn PP, Van Broeckhoven C, Livingston G, Bass NJ, Gurling H, McQuillin A, Gwilliam R, Deloukas P, Al-Chalabi A, Shaw CE, Tsolaki M, Singleton AB, Guerreiro R, Muhleisen TW, Nothen MM, Moebus S, Jockel KH, Klopp N, Wichmann HE, Carrasquillo MM, Pankratz VS, Younkin SG, Holmans PA, O’Donovan M, Owen MJ, Williams J. Genome-wide association study identifies variants at CLU and PICALM associated with Alzheimer’s disease. Nat Genet. 2009;41:1088–1093. doi: 10.1038/ng.440. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [6].Seshadri S, Fitzpatrick AL, Ikram MA, DeStefano AL, Gudnason V, Boada M, Bis JC, Smith AV, Carassquillo MM, Lambert JC, Harold D, Schrijvers EM, Ramirez-Lorca R, Debette S, Longstreth WT, Jr., Janssens AC, Pankratz VS, Dartigues JF, Hollingworth P, Aspelund T, Hernandez I, Beiser A, Kuller LH, Koudstaal PJ, Dickson DW, Tzourio C, Abraham R, Antunez C, Du Y, Rotter JI, Aulchenko YS, Harris TB, Petersen RC, Berr C, Owen MJ, Lopez-Arrieta J, Varadarajan BN, Becker JT, Rivadeneira F, Nalls MA, Graff-Radford NR, Campion D, Auerbach S, Rice K, Hofman A, Jonsson PV, Schmidt H, Lathrop M, Mosley TH, Au R, Psaty BM, Uitterlinden AG, Farrer LA, Lumley T, Ruiz A, Williams J, Amouyel P, Younkin SG, Wolf PA, Launer LJ, Lopez OL, van Duijn CM, Breteler MM. Genome-wide analysis of genetic loci associated with Alzheimer disease. JAMA. 2010;303:1832–1840. doi: 10.1001/jama.2010.574. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [7].Hollingworth P, Harold D, Sims R, Gerrish A, Lambert JC, Carrasquillo MM, Abraham R, Hamshere ML, Pahwa JS, Moskvina V, Dowzell K, Jones N, Stretton A, Thomas C, Richards A, Ivanov D, Widdowson C, Chapman J, Lovestone S, Powell J, Proitsi P, Lupton MK, Brayne C, Rubinsztein DC, Gill M, Lawlor B, Lynch A, Brown KS, Passmore PA, Craig D, McGuinness B, Todd S, Holmes C, Mann D, Smith AD, Beaumont H, Warden D, Wilcock G, Love S, Kehoe PG, Hooper NM, Vardy ER, Hardy J, Mead S, Fox NC, Rossor M, Collinge J, Maier W, Jessen F, Ruther E, Schurmann B, Heun R, Kolsch H, van den Bussche H, Heuser I, Kornhuber J, Wiltfang J, Dichgans M, Frolich L, Hampel H, Gallacher J, Hull M, Rujescu D, Giegling I, Goate AM, Kauwe JS, Cruchaga C, Nowotny P, Morris JC, Mayo K, Sleegers K, Bettens K, Engelborghs S, De Deyn PP, Van Broeckhoven C, Livingston G, Bass NJ, Gurling H, McQuillin A, Gwilliam R, Deloukas P, Al-Chalabi A, Shaw CE, Tsolaki M, Singleton AB, Guerreiro R, Muhleisen TW, Nothen MM, Moebus S, Jockel KH, Klopp N, Wichmann HE, Pankratz VS, Sando SB, Aasly JO, Barcikowska M, Wszolek ZK, Dickson DW, Graff-Radford NR, Petersen RC, van Duijn CM, Breteler MM, Ikram MA, DeStefano AL, Fitzpatrick AL, Lopez O, Launer LJ, Seshadri S, Berr C, Campion D, Epelbaum J, Dartigues JF, Tzourio C, Alperovitch A, Lathrop M, Feulner TM, Friedrich P, Riehle C, Krawczak M, Schreiber S, Mayhaus M, Nicolhaus S, Wagenpfeil S, Steinberg S, Stefansson H, Stefansson K, Snaedal J, Bjornsson S, Jonsson PV, Chouraki V, Genier-Boley B, Hiltunen M, Soininen H, Combarros O, Zelenika D, Delepine M, Bullido MJ, Pasquier F, Mateo I, Frank-Garcia A, Porcellini E, Hanon O, Coto E, Alvarez V, Bosco P, Siciliano G, Mancuso M, Panza F, Solfrizzi V, Nacmias B, Sorbi S, Bossu P, Piccardi P, Arosio B, Annoni G, Seripa D, Pilotto A, Scarpini E, Galimberti D, Brice A, Hannequin D, Licastro F, Jones L, Holmans PA, Jonsson T, Riemenschneider M, Morgan K, Younkin SG, Owen MJ, O’Donovan M, Amouyel P, Williams J. Common variants at ABCA7, MS4A6A/MS4A4E, EPHA1, CD33 and CD2AP are associated with Alzheimer’s disease. Nat Genet. 2011;43:429–435. doi: 10.1038/ng.803. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [8].Naj AC, Jun G, Beecham GW, Wang LS, Vardarajan BN, Buros J, Gallins PJ, Buxbaum JD, Jarvik GP, Crane PK, Larson EB, Bird TD, Boeve BF, Graff-Radford NR, De Jager PL, Evans D, Schneider JA, Carrasquillo MM, Ertekin-Taner N, Younkin SG, Cruchaga C, Kauwe JS, Nowotny P, Kramer P, Hardy J, Huentelman MJ, Myers AJ, Barmada MM, Demirci FY, Baldwin CT, Green RC, Rogaeva E, St George-Hyslop P, Arnold SE, Barber R, Beach T, Bigio EH, Bowen JD, Boxer A, Burke JR, Cairns NJ, Carlson CS, Carney RM, Carroll SL, Chui HC, Clark DG, Corneveaux J, Cotman CW, Cummings JL, DeCarli C, DeKosky ST, Diaz-Arrastia R, Dick M, Dickson DW, Ellis WG, Faber KM, Fallon KB, Farlow MR, Ferris S, Frosch MP, Galasko DR, Ganguli M, Gearing M, Geschwind DH, Ghetti B, Gilbert JR, Gilman S, Giordani B, Glass JD, Growdon JH, Hamilton RL, Harrell LE, Head E, Honig LS, Hulette CM, Hyman BT, Jicha GA, Jin LW, Johnson N, Karlawish J, Karydas A, Kaye JA, Kim R, Koo EH, Kowall NW, Lah JJ, Levey AI, Lieberman AP, Lopez OL, Mack WJ, Marson DC, Martiniuk F, Mash DC, Masliah E, McCormick WC, McCurry SM, McDavid AN, McKee AC, Mesulam M, Miller BL, Miller CA, Miller JW, Parisi JE, Perl DP, Peskind E, Petersen RC, Poon WW, Quinn JF, Rajbhandary RA, Raskind M, Reisberg B, Ringman JM, Roberson ED, Rosenberg RN, Sano M, Schneider LS, Seeley W, Shelanski ML, Slifer MA, Smith CD, Sonnen JA, Spina S, Stern RA, Tanzi RE, Trojanowski JQ, Troncoso JC, Van Deerlin VM, Vinters HV, Vonsattel JP, Weintraub S, Welsh-Bohmer KA, Williamson J, Woltjer RL, Cantwell LB, Dombroski BA, Beekly D, Lunetta KL, Martin ER, Kamboh MI, Saykin AJ, Reiman EM, Bennett DA, Morris JC, Montine TJ, Goate AM, Blacker D, Tsuang DW, Hakonarson H, Kukull WA, Foroud TM, Haines JL, Mayeux R, Pericak-Vance MA, Farrer LA, Schellenberg GD. Common variants at MS4A4/MS4A6E, CD2AP, CD33 and EPHA1 are associated with late-onset Alzheimer’s disease. Nat Genet. 2011;43:436–441. doi: 10.1038/ng.801. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [9].Lambert JC, Ibrahim-Verbaas CA, Harold D, Naj AC, Sims R, Bellenguez C, Jun G, Destefano AL, Bis JC, Beecham GW, Grenier-Boley B, Russo G, Thornton-Wells TA, Jones N, Smith AV, Chouraki V, Thomas C, Ikram MA, Zelenika D, Vardarajan BN, Kamatani Y, Lin CF, Gerrish A, Schmidt H, Kunkle B, Dunstan ML, Ruiz A, Bihoreau MT, Choi SH, Reitz C, Pasquier F, Hollingworth P, Ramirez A, Hanon O, Fitzpatrick AL, Buxbaum JD, Campion D, Crane PK, Baldwin C, Becker T, Gudnason V, Cruchaga C, Craig D, Amin N, Berr C, Lopez OL, De Jager PL, Deramecourt V, Johnston JA, Evans D, Lovestone S, Letenneur L, Moron FJ, Rubinsztein DC, Eiriksdottir G, Sleegers K, Goate AM, Fievet N, Huentelman MJ, Gill M, Brown K, Kamboh MI, Keller L, Barberger-Gateau P, McGuinness B, Larson EB, Green R, Myers AJ, Dufouil C, Todd S, Wallon D, Love S, Rogaeva E, Gallacher J, St George-Hyslop P, Clarimon J, Lleo A, Bayer A, Tsuang DW, Yu L, Tsolaki M, Bossu P, Spalletta G, Proitsi P, Collinge J, Sorbi S, Sanchez-Garcia F, Fox NC, Hardy J, Naranjo MC, Bosco P, Clarke R, Brayne C, Galimberti D, Mancuso M, Matthews F, Moebus S, Mecocci P, Del Zompo M, Maier W, Hampel H, Pilotto A, Bullido M, Panza F, Caffarra P, Nacmias B, Gilbert JR, Mayhaus M, Lannfelt L, Hakonarson H, Pichler S, Carrasquillo MM, Ingelsson M, Beekly D, Alvarez V, Zou F, Valladares O, Younkin SG, Coto E, Hamilton-Nelson KL, Gu W, Razquin C, Pastor P, Mateo I, Owen MJ, Faber KM, Jonsson PV, Combarros O, O’Donovan MC, Cantwell LB, Soininen H, Blacker D, Mead S, Mosley TH, Jr., Bennett DA, Harris TB, Fratiglioni L, Holmes C, de Bruijn RF, Passmore P, Montine TJ, Bettens K, Rotter JI, Brice A, Morgan K, Foroud TM, Kukull WA, Hannequin D, Powell JF, Nalls MA, Ritchie K, Lunetta KL, Kauwe JS, Boerwinkle E, Riemenschneider M, Boada M, Hiltunen M, Martin ER, Schmidt R, Rujescu D, Wang LS, Dartigues JF, Mayeux R, Tzourio C, Hofman A, Nothen MM, Graff C, Psaty BM, Jones L, Haines JL, Holmans PA, Lathrop M, Pericak-Vance MA, Launer LJ, Farrer LA, van Duijn CM, Van Broeckhoven C, Moskvina V, Seshadri S, Williams J, Schellenberg GD, Amouyel P. Meta-analysis of 74,046 individuals identifies 11 new susceptibility loci for Alzheimer’s disease. Nat Genet. 2013;45:1452–1458. doi: 10.1038/ng.2802. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [10].Guerreiro R, Wojtas A, Bras J, Carrasquillo M, Rogaeva E, Majounie E, Cruchaga C, Sassi C, Kauwe JS, Younkin S, Hazrati L, Collinge J, Pocock J, Lashley T, Williams J, Lambert JC, Amouyel P, Goate A, Rademakers R, Morgan K, Powell J, St George-Hyslop P, Singleton A, Hardy J. TREM2 variants in Alzheimer’s disease. N Engl J Med. 2013;368:117–127. doi: 10.1056/NEJMoa1211851. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [11].Thalhauser CJ, Komarova NL. Alzheimer’s disease: rapid and slow progression. J R Soc Interface. 2012;9:119–126. doi: 10.1098/rsif.2011.0134. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [12].Jack CR, Jr., Shiung MM, Gunter JL, O’Brien PC, Weigand SD, Knopman DS, Boeve BF, Ivnik RJ, Smith GE, Cha RH, Tangalos EG, Petersen RC. Comparison of different MRI brain atrophy rate measures with clinical disease progression in AD. Neurology. 2004;62:591–600. doi: 10.1212/01.wnl.0000110315.26026.ef. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [13].Ridha BH, Barnes J, Bartlett JW, Godbolt A, Pepple T, Rossor MN, Fox NC. Tracking atrophy progression in familial Alzheimer’s disease: a serial MRI study. Lancet Neurol. 2006;5:828–834. doi: 10.1016/S1474-4422(06)70550-6. [DOI] [PubMed] [Google Scholar]
  • [14].Sluimer JD, Vrenken H, Blankenstein MA, Fox NC, Scheltens P, Barkhof F, van der Flier WM. Whole-brain atrophy rate in Alzheimer disease: identifying fast progressors. Neurology. 2008;70:1836–1841. doi: 10.1212/01.wnl.0000311446.61861.e3. [DOI] [PubMed] [Google Scholar]
  • [15].McEvoy LK, Fennema-Notestine C, Roddey JC, Hagler DJ, Jr., Holland D, Karow DS, Pung CJ, Brewer JB, Dale AM. Alzheimer disease: quantitative structural neuroimaging for detection and prediction of clinical and structural changes in mild cognitive impairment. Radiology. 2009;251:195–205. doi: 10.1148/radiol.2511080924. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [16].Nestor SM, Rupsingh R, Borrie M, Smith M, Accomazzi V, Wells JL, Fogarty J, Bartha R. Ventricular enlargement as a possible measure of Alzheimer’s disease progression validated using the Alzheimer’s disease neuroimaging initiative database. Brain. 2008;131:2443–2454. doi: 10.1093/brain/awn146. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [17].Mann UM, Mohr E, Gearing M, Chase TN. Heterogeneity in Alzheimer’s disease: progression rate segregated by distinct neuropsychological and cerebral metabolic profiles. J Neurol Neurosurg Psychiatry. 1992;55:956–959. doi: 10.1136/jnnp.55.10.956. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [18].Mielke MM, Rosenberg PB, Tschanz J, Cook L, Corcoran C, Hayden KM, Norton M, Rabins PV, Green RC, Welsh-Bohmer KA, Breitner JC, Munger R, Lyketsos CG. Vascular factors predict rate of progression in Alzheimer disease. Neurology. 2007;69:1850–1858. doi: 10.1212/01.wnl.0000279520.59792.fe. [DOI] [PubMed] [Google Scholar]
  • [19].Prolo P, Chiappelli F, Angeli A, Dovio A, Perotti P, Pautasso M, Sartori ML, Saba L, Mussino S, Fraccalini T, Fanto F, Mocellini C, Rosso MG, Grasso E. Physiologic modulation of natural killer cell activity as an index of Alzheimer’s disease progression. Bioinformation. 2007;1:363–366. doi: 10.6026/97320630001363. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [20].Farrer LA, Cupples LA, van Duijn CM, Connor-Lacke L, Kiely DK, Growdon JH. Rate of progression of Alzheimer’s disease is associated with genetic risk. Arch Neurol. 1995;52:918–923. doi: 10.1001/archneur.1995.00540330100021. [DOI] [PubMed] [Google Scholar]
  • [21].Murphy GM, Jr., Claassen JD, DeVoss JJ, Pascoe N, Taylor J, Tinklenberg JR, Yesavage JA. Rate of cognitive decline in AD is accelerated by the interleukin-1 alpha -889 *1 allele. Neurology. 2001;56:1595–1597. doi: 10.1212/wnl.56.11.1595. [DOI] [PubMed] [Google Scholar]
  • [22].Lopez OL, Becker JT, Saxton J, Sweet RA, Klunk W, DeKosky ST. Alteration of a clinically meaningful outcome in the natural history of Alzheimer’s disease by cholinesterase inhibition. J Am Geriatr Soc. 2005;53:83–87. doi: 10.1111/j.1532-5415.2005.53015.x. [DOI] [PubMed] [Google Scholar]
  • [23].Kamboh MI, Barmada MM, Demirci FY, Minster RL, Carrasquillo MM, Pankratz VS, Younkin SG, Saykin AJ, Sweet RA, Feingold E, DeKosky ST, Lopez OL. Genome-wide association analysis of age-at-onset in Alzheimer’s disease. Mol Psychiatry. 2012;17:1340–1346. doi: 10.1038/mp.2011.135. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [24].Kamboh MI, Demirci FY, Wang X, Minster RL, Carrasquillo MM, Pankratz VS, Younkin SG, Saykin AJ, Jun G, Baldwin C, Logue MW, Buros J, Farrer L, Pericak-Vance MA, Haines JL, Sweet RA, Ganguli M, Feingold E, Dekosky ST, Lopez OL, Barmada MM. Genome-wide association study of Alzheimer’s disease. Transl Psychiatry. 2012;2:e117. doi: 10.1038/tp.2012.45. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [25].Kamboh MI, Aston CE, Hamman RF. The relationship of APOE polymorphism and cholesterol levels in normoglycemic and diabetic subjects in a biethnic population from the San Luis Valley, Colorado. Atherosclerosis. 1995;112:145–159. doi: 10.1016/0021-9150(94)05409-c. [DOI] [PubMed] [Google Scholar]
  • [26].Liu JZ, McRae AF, Nyholt DR, Medland SE, Wray NR, Brown KM, Hayward NK, Montgomery GW, Visscher PM, Martin NG, Macgregor S. A versatile gene-based test for genome-wide association studies. Am J Hum Genet. 2010;87:139–145. doi: 10.1016/j.ajhg.2010.06.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [27].Barrett JC, Fry B, Maller J, Daly MJ. Haploview: analysis and visualization of LD and haplotype maps. Bioinformatics. 2005;21:263–265. doi: 10.1093/bioinformatics/bth457. [DOI] [PubMed] [Google Scholar]
  • [28].Purcell S, Neale B, Todd-Brown K, Thomas L, Ferreira MA, Bender D, Maller J, Sklar P, de Bakker PI, Daly MJ, Sham PC. PLINK: a tool set for whole-genome association and population-based linkage analyses. Am J Hum Genet. 2007;81:559–575. doi: 10.1086/519795. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [29].Ruiz A, Hernandez I, Ronsende-Roca M, Gonzalez-Perez A, Rodriguez-Noriega E, Ramirez-Lorca R, Mauleon A, Moreno-Rey C, Boswell L, Tune L, Valero S, Alegret M, Gayan J, Becker JT, Real LM, Tarraga L, Ballard C, Terrin M, Sherman S, Payami H, Lopez OL, Mintzer JE, Boada M. Exploratory analysis of seven Alzheimer’s disease genes: disease progression. Neurobiol Aging. 2013;34:1310, e1311–1317. doi: 10.1016/j.neurobiolaging.2012.08.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [30].Fang WH, Wang Q, Li HM, Ahmed M, Kumar P, Kumar S. PAX3 in neuroblastoma: oncogenic potential, chemosensitivity and signalling pathways. J Cell Mol Med. 2014;18:38–48. doi: 10.1111/jcmm.12155. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [31].Baldwin CT, Hoth CF, Macina RA, Milunsky A. Mutations in PAX3 that cause Waardenburg syndrome type I: ten new mutations and review of the literature. Am J Med Genet. 1995;58:115–122. doi: 10.1002/ajmg.1320580205. [DOI] [PubMed] [Google Scholar]
  • [32].Baldwin CT, Lipsky NR, Hoth CF, Cohen T, Mamuya W, Milunsky A. Mutations in PAX3 associated with Waardenburg syndrome type I. Hum Mutat. 1994;3:205–211. doi: 10.1002/humu.1380030306. [DOI] [PubMed] [Google Scholar]
  • [33].Hoth CF, Milunsky A, Lipsky N, Sheffer R, Clarren SK, Baldwin CT. Mutations in the paired domain of the human PAX3 gene cause Klein-Waardenburg syndrome (WS-III) as well as Waardenburg syndrome type I (WS-I) Am J Hum Genet. 1993;52:455–462. [PMC free article] [PubMed] [Google Scholar]
  • [34].Rosenthal SL, Barmada MM, Wang X, Demirci FY, Kamboh MI. Connecting the dots: potential of data integration to identify regulatory SNPs in late-onset Alzheimer’s disease GWAS findings. PLoS One. 2014;9:e95152. doi: 10.1371/journal.pone.0095152. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [35].Dupressoir A, Barbot W, Loireau MP, Heidmann T. Characterization of a mammalian gene related to the yeast CCR4 general transcription factor and revealed by transposon insertion. J Biol Chem. 1999;274:31068–31075. doi: 10.1074/jbc.274.43.31068. [DOI] [PubMed] [Google Scholar]
  • [36].Chang YC, Chiu YF, Liu PH, Hee SW, Chang TJ, Jiang YD, Lee WJ, Lee PC, Kao HY, Hwang JJ, Chuang LM. Genetic variation in the NOC gene is associated with body mass index in Chinese subjects. PLoS One. 2013;8:e69622. doi: 10.1371/journal.pone.0069622. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [37].Ng EL, Tang BL. Rab GTPases and their roles in brain neurons and glia. Brain Res Rev. 2008;58:236–246. doi: 10.1016/j.brainresrev.2008.04.006. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary Figures

RESOURCES