Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2010 Jun 1.
Published in final edited form as: Curr Alzheimer Res. 2009 Jun;6(3):196–204. doi: 10.2174/156720509788486608

A common biological mechanism in cancer and Alzheimer’s disease?

Maria I Behrens 1,*, Corinne Lendon 2, Catherine M Roe 3
PMCID: PMC2810550  NIHMSID: NIHMS165753  PMID: 19519301

Abstract

Cancer and Alzheimer’s disease (AD) are two common disorders for which the final pathophysiological mechanism is not yet clearly defined. In a prospective longitudinal study we have previously shown an inverse association between AD and cancer, such that the rate of developing cancer in general with time was significantly slower in participants with AD, while participants with a history of cancer had a slower rate of developing AD. In cancer, cell regulation mechanisms are disrupted with augmentation of cell survival and/or proliferation, whereas conversely, AD is associated with increased neuronal death, either caused by, or concomitant with, beta amyloid (Aβ) and tau deposition. The possibility that perturbations of mechanisms involved in cell survival/death regulation could be involved in both disorders is discussed. Genetic polymorphisms, DNA methylation or other mechanisms that induce changes in activity of molecules with key roles in determining the decision to “repair and live”- or “die” could be involved in the pathogenesis of the two disorders. As examples, the role of p53, Pin1 and the Wnt signaling pathway are discussed as potential candidates that, speculatively, may explain inverse associations between AD and cancer.

Keywords: Alzheimer, cancer, tumor suppressors, Pin1, Wnt signaling pathway

An inverse association between cancer and AD

While attending patients in nursing homes where most residents have some type of dementia, we were puzzled by the observation that a history of cancer was not a common finding among residents who were demented, whereas many residents that were cognitively normal had had cancer in the past. This anecdotal observation was followed by a longitudinal prospective study in which we found an inverse association between cancer and AD [1]. The study was done using archival data of longitudinal studies of memory and aging study at the Alzheimer’s disease Research Center at Washington University School of Medicine in St Louis. In these studies, participants are cognitively evaluated annually with the Clinical Dementia Rating CDR [2], and a thorough medical history is obtained including a history of cancer, its type, treatment and date of diagnosis. Results showed an inverse association between cancer and AD. Specifically, we found that of the 594 participants, who at their first visit had no history of cancer, 395 had dementia of the Alzheimer type [DAT] and 199 were cognitively normal. Data from subsequent visits revealed that the rate of receiving a cancer diagnosis was significantly slower in the DAT group (p < 0.001). Conversely, of the 249 participants who were cognitively normal at study entry we found that 50 had a history of cancer and 199 did not. Those with a history of cancer had a slower rate of receiving a DAT diagnosis with time, although this result did not reach statistical significance at an alpha level of .05 (p = 0.0600). Cox proportional hazard models indicated that the inverse relationship between cancer and DAT was not accounted for by demographic factors such as age, sex and education [1]. These results are in accordance with previous cross-sectional and case-control studies which report a reduced prevalence of cancer among individuals with AD [38].

The incident cancers in the demented and non-demented groups were similar to those in the general US population [9]; 57% of all neoplasias in our study were skin cancers, most of them benign, ~80% basal or squamous cell with the analysis of carcinomas. The survival analyses were repeated examining the association between DAT at study entry and the development of skin cancer specifically. When skin cancer alone was analyzed as the dependent variable we found that, as with all cancers, the rate of skin cancer diagnosis was slower for participants with DAT compared to non-demented individuals in both the log-rank test (p <0.001) and the Cox proportional hazard models (p < 0.005; hazard ratios associated with DAT versus no dementia ranged between 0.22–0.26). Other cancer types were too infrequent in our sample to be analyzed separately. When combining all non-skin cancers into a single group, the rate of cancer diagnosis was slower with time for the DAT group, although the difference did not reach statistical significance (all p>0.05; hazard ratios associated with the DAT versus no dementia ranged between 0.65–0.83) [10].

The relationship of dementia to the different types of cancers individually (other than skin cancers) remains to be analyzed in detail, as well as the role of environmental factors. However, there is an interesting possibility that one or more biological mechanisms may link AD and cancer. If such a mechanism can be identified, it might lead to better understanding of these two disorders, as well as strategies to protect us from them. The purpose of this speculative review is to describe some potential biological links between the development of cancer and AD.

A common biological mechanism with opposing effects?

The neuropathological hallmarks of Alzheimer’s disease include senile plaques, and neurofibrillary tangles. Senile plaques consist of extracellular deposits mainly composed of the beta amyloid (Aβ) peptide. Neurofibrillary tangles are intracellular deposits of an abnormally hyperphosphorylated tau, a microtubule-associated protein which is involved in axonal transport and other functions. In addition to amyloid plaques and neurofibrillary tangles, AD is characterized by extensive neuritic and synaptic degeneration and neuronal cell death. The type of cell death in AD is still controversial, but it is clear that in AD there is progressive atrophy of the brain due to cell and synaptic loss. The leading explanation for the pathologic changes associated with AD is the “amyloid hypothesis” which states that neuronal dysfunction and death, neurofibrillary degeneration, microglial activation and the full manifestation of Alzheimer pathology are initiated by Aβ deposition [11]. The abnormal phosphorylation of tau in AD and other neurodegenerative disorders induces a decreased capability of tau to promote tubulin polymerization and bind to microtubules [12], leading to a generalized loss of microtubule stability [13] and eventually retrograde neurodegeneration. Although the amyloid and tau hypotheses are the favored ones, apoptosis, synaptic loss, or neuronal dysfunction prior to cell death might also play a role in the physiopathology of AD [14, 15].

Normal tissue function requires that the rate of cell loss is matched by the rate of renewal. The maintenance of tissue homeostasis is achieved by efficient mechanisms that control genomic stability to prevent aberrant proliferation [16,17,18]. After DNA damage, intracellular stress pathways which are able to recognize the damage are activated, and either recruit DNA repair factors to mend the damage, or induce apoptosis or senescence [19]. In cancer this reparative process is defective, leading to excessive cell growth. To simplify, cancer is a disorder typified by uncontrolled, excessive cell growth, whereas conversely, AD and other neurodegenerative disorders are characterized by progressive dysfunction and eventual loss of neuronal loss.

One possible explanation for why AD and cancer appear to be inversely associated is that both diseases arise via malfunction of an underlying common mechanism that regulates cell survival. This hypothetical mechanism could regulate the capability of the cells of switching the cell machinery from a prone-to-death state (AD phenotype) to a prone-to-survive/grow state (cancer phenotype). If cells are in the prone-to die state, then neurons will be more susceptible to cell death under stressors such as Aβ, tau hyperphosphorylation, oxidation, inflammation, or other unknown risk factors (the AD phenotype). At the same time, cells will respond to initiating cancer stimuli (UV radiation, for example) by death, reducing cancer susceptibility. Conversely, if cells are shifted to a survival/growth state, neurons would have a greater likelihood of surviving if subjected to stressors, while concomitantly becoming more susceptible to cancer development. Genetic polymorphisms in several key molecules that determine changes in their activity or variance in their DNA methylation could explain such opposing effects [20]. Here we discuss the possible involvement of the tumor suppressor p53, of Pin1 and the Wnt signaling pathway in the pathophysiology of both cancer and AD.

Tumor suppressors in cancer and aging

In tissues with the capability of renewal the repair or regeneration of cells has evolved as an advantage in terms of increased longevity compared to postmitotic tissues which lack this capability. However, this versatility to regenerate has the inherent risk of hyperproliferation, among which the most dangerous is cancer. Thus, together with the renewal capacity that poses increased longevity there evolved mechanisms to suppress tumor formation [21]. Tumor suppressors are the main actors of the surveillance mechanism to avoid aberrant proliferation under normal conditions. Tumor suppressors were given this name because, as regulators of diverse cellular activities, their loss enhances tumor formation. They regulate cell cycle checkpoint responses, detection and repair of DNA damage, protein ubiquitination and degradation, mitogenic signaling, cell specification, differentiation and migration, and angiogenesis [16]. Tumor suppressors can either eliminate potential cancer cells by inducing programmed cell death (apoptosis), or alternatively, they can induce permanent withdrawal from the cell cycle (cellular senescence). Apoptosis could give rise to a depletion of irreplaceable postmitotic cells in nonrenewable tissues; and in depletion of proliferating or stem cell pools in renewable tissues. In the same way, senescence could deplete tissues of proliferating or stem cell pools, resulting in the accumulation of senescent cells. Thus, tumor suppressor mechanisms may be an example of evolutionary antagonistic pleiotropy [21, 22, 23], that is, they promote early-life survival by preventing the development of cancer, but eventually limiting longevity.

The p53 gene is the prototypical tumor suppressor and its pathway is inactivated in most human cancers [18]. p53 is at the hub of numerous signaling pathways that are triggered in response to particular stresses and in this way is a major regulator of cellular stress. p53 can be described as a stress response gene; its product (the p53 protein) acts to induce apoptosis or cell-cycle arrest in response to DNA damage, thereby maintaining genetic stability in the organism by transcriptional and nontranscriptional mechanisms [18]. Mice engineered to be deficient in p53 are developmentally normal, but susceptible to spontaneous tumors [24]. In addition, knocked out p53 mice (p53−/−) have a significantly higher number of proliferating cells, as assessed by the incorporation of the nucleotide analogue bromodeoxyuridine in the lateral ventricle wall, compared to wild-type littermates [25].

As mentioned above, the anticancer roles of p53 may come at the cost of proliferative reserve and thereby compromise tissue repair and promote the aging phenotype [23, 2628]. Tyner et al [26] created transgenic mice with a p53-mutation that confers phenotypes consistent with an activated form of p53. The animals had very few cancers, but developed several features of aging and showed reduced longevity. Further support for the role of p53 in accelerating aging comes from a report showing that transgenic mice that overexpress p44 (a protein that enhances p53 activity) have a shortened lifespan and accelerated aging, and an extraordinary low incidence of cancer compared to wild-type mice [29]. Additional support for the concept that hyperactivation of the tumor suppressor p53 may be related to accelerated aging comes from a study showing that mice deficient in Zmpste24, a metalloproteinase involved in the maturation of lamin A, show a senescence phenotype at the cellular level and accelerated aging at the organism level, together with a marked upregulation of p53 target genes [28]. If p53 enhancement leads to aging and reduced longevity, the opposite should be expected in p53 knockout mice. Until recently, no effect of p53 on longevity had been detected, because the predisposition to tumors precluded an analysis of the role of p53 in longevity. However, genetically manipulated mice showing increased activity of the Arf/p53 pathway but conserving the normal regulation of p53, show both cancer resistance and decreased aging [30], which is what would be expected if cancer, a major cause of death in mice, is prevented. Also, transgenic mouse models with elevated p53 activity, but under normal regulatory control, show reduced tumor formation without accelerated aging [31, 32]. Therefore, the regulation of p53 activity is crucial to determine the role that p53 will assume. For example, depending on the tissue, the same acetylation of the Lys amino acid at position 320 of p53 can promote neurite outgrowth in neuronal cells [33] or cell cycle arrest in other tissues [34].

Taken together these results show that the regulation of p53’s actions in cells is extremely important to determine the fate of the cells or tissues. p53 is activated and integrates the different incoming signals that sense different forms of cellular stress. Therefore, it is conceivable that small deregulations towards one or the other side, could favor survival/regeneration or death/senescence of the cells. The inactivation of p53 is implicated in the development of cancer, and p53 activation might play a role in promoting aging. Elevated p53 levels have been detected in the central nervous system of patients diagnosed with neurodegenerative diseases, such as Huntington’s disease and Amyotrophic Lateral Sclerosis [35, 36] and in mouse models [37] and in the brains of Parkinson’s disease patients [38]. The intracellular expression of the Aβ protein under a neuron-specific promoter led progressively to degeneration and death of neurons in the brains of transgenic mice and Aβ accumulation was correlated with activation of p53 [39]. Also, elevated levels of p73, a member of the p53 family, have been described in mice injected with fibrils of Aβ and in mice models of AD [40]. In neuroblastoma cell lines, intracellular Aβ42 directly activated the p53 promoter, resulting in p53-dependent apoptosis [41]. Intracellular Aβ40 had a similar but smaller effect in the same study [41]. Several reports have described upregulation of p53 in the brains of patients with AD [41, 42, 43]. The presenilins (PS1 and PS2) form part of the gamma secretase complex that cleaves the amyloid precursor protein (APP) to generate Aβ. Mutations in the presenilins cause familial forms of AD and have also been shown to trigger p53-dependent cell death [44, 45]. The intracellular C-terminal fragments of the gamma secretase cleavage of APP trigger the activation of caspase-3 and an increase in p53activity and mRNA [45].

Another important tumor suppressor pathway is the pRB/p16 pathway. The p16 gene functions as a negative regulator of cell cycle and is therefore considered to represent a tumor suppressor. As with p53, deletion of p16 is frequently observed in cancer cell lines and some malignant tumors including acute lymphoblastic leukemia of childhood, melanomas, gliomas, as well as carcinoma of the pancreas, esophagus, lung, bladder, head, and neck [46, 47]. On the other hand, increased levels of p16 expression and of the incidence of p16-positive cells are associated with age in many mouse and rat tissues [48, 49]. Until recently however, there was no evidence that enhanced pRB function accelerated aging. Three reports show that increased levels of p16 contribute to aging by limiting self-renewal of regenerative cells in different tissues such as brain, endocrine pancreas and bone marrow [5053].

Alterations in p53, or other tumor suppressors, could play a role in explaining the opposing results we found with regard to the development of cancer and DAT (Figure 1). Patients with slight increases in the activity of tumor suppressor proteins would have lower risks of developing cancers, but instead would be at higher risk of developing AD because of an increased susceptibility to cell death or senescence against stressors such as Aβ, tau hyperphosphorylation, and/or oxidative stress. These slight alterations could be due to polymorphisms or differences in DNA methylation in tumor suppressors that confer an increased risk across the lifespan. Augmented levels of tumor suppressors could also limit the renewal capacity of stem cells, and in this way induce AD by preventing, in the long run, the replenishment of apoptotic neurons, or the reparation of dendrites and spines in damaged neurons, processes which are compatible with the slow course of the disease. Similarly, the low incidence of AD in the cancer group could be explained by the presence of an inactivated form of a tumor suppressor, which in addition to favoring previous cancer development, could also favor a decreased susceptibility to neuronal death, conferring protection against AD. Additional support for a role of tumor suppressors in aging and AD comes from a microarray study, in which an up-regulation of a disproportionately high number of tumor suppressors or tumor suppressor co-factors, including several of the retinoblastoma (Rb) family, were found in the CA1 region of the hippocampus of AD patients [54].

Figure 1.

Figure 1

Role of p53 in cancer and AD. In response to toxic or stress signals, p53 is activated through a number of post-translational modifications and induces cell cycle arrest among other functions. The decision is made whether to induce DNA repair or apoptosis of damaged cells to maintain genomic stability. If the cell machinery in the whole organism were shifted to high p53 in response to stressors, the cells would be more prone to cell death and AD could develop. If, on the contrary, the cell machinery were shifted to low or no p53, the cells would be more prone to develop a cancer. ROS, reactive oxygen species.

Propyl isomerase (Pin1)

Variations in factors upstream of p53 that ultimately result in its altered activity could also contribute to the observed opposing presentation of cancer and AD phenotypes. One such mechanism could involve the action of Pin1. Pin1 (protein interacting with NIMA 1) is a ubiquitous enzyme that catalyses cis/trans isomerization of phosphorylated serine or threonine residues that immediately precede a proline [5558]. Pin1 is conserved from yeast to humans and has been shown to regulate a diverse array of cellular processes of cell proliferation and differentiation, such as cell-cycle control, transcription and splicing regulation, DNA replication checkpoint control, DNA damage response, neuronal survival, and germ cell development [55]. Aberrant Pin1 function has been implicated in both cancer and AD [57, 58]. Investigations show that Pin1 binds to and isomerizes hyperphosphorylated tau, specifically at the Thr231-Pro site, to restore the ability of tau to bind microtubules and promote their assembly and facilitate tau dephosphorylation [5961]. Furthermore, in addition to this role on tau, Pin1 is also involved in regulating APP processing and Aβ production. Pastorino et al [62] have shown that Pin1 binds to APP and accelerates its cis/trans isomerization. Overexpression of Pin1 reduces Aβ secretion from cell cultures, whereas Pin1 KO increases its secretion. In addition, Pin1 KO alone, or in combination with overexpression of mutant APP in mice selectively elevates insoluble brain Aβ 42 in an age-dependent manner. Therefore, deletion of the Pin1 gene alone in mice causes progressive deposition of tau and Aβ, and neuronal degeneration [56, 60]. Pin1 expression is induced during neuronal differentiation and is highly expressed in most neurons in the brain [6163]. In accordance with these results down-regulation of Pin1 has been reported in the hippocampus of AD patients [64]. However, a compensatory activation or up-regulation of Pin1 may also be induced in AD brains [65]. The complex regulation of Pin1 is strengthened by recent studies showing opposite effects of Pin1 on tau protein stability and tauopathy phenotype depending on whether the tau is wild-type (WT) or has the P301L mutation [65]. Pin1 knockdown or KO increased stability of WT tau protein stability and Pin1 overexpression suppressed the tauopathy phenotype in WT tau transgenic mice. In contrast, the opposite was found with mutant P301L tau; Pin1 knockdown or KO decreased P301L tau protein stability and abolished its robust tauopathy phenotype in the mutant mice, whereas its overexpression exacerbated the tauopathy phenotype in P301L tau mice [65]. Pin1 promoter polymorphisms appear to associate with reduced Pin1 levels and increased risk for late-onset Alzheimer’s disease [6669], but not all case-control studies agree (70, 71]. Evidence for a participation of Pin1 in ALS has also been reported [72].

A prevalent overexpression of Pin1 has been shown in most human cancers including prostate, breast, lung, colon and liver [7376], but not in others such as renal cancer [77]. Pin1 is important for the activation of multiple oncogenic pathways involved in tumorigenesis, such as cyclin D1, Wnt/β-catenin, NF-κB, p53, and p73 [73, 76, 7883]. Accumulating evidence suggests that Pin1 regulates the timing of p53 activation, modulating its interaction with DNA and cofactors [84]. In response to toxic stimuli, the interaction between p53 and Pin1 markedly increases with phosphorylation of a subset of Ser/Thr-Pro motifs of p53 and its subsequent isomerization. Cells lacking Pin1 fail to efficiently stabilize p53 and are then able to escape cell cycle arrest and apoptotic responses [83, 85], thus promoting the ‘prone to cancer’ direction in figure 1. Also, a role of Pin1 in promoting the mitochondrial apoptotic machinery has been described in neurons [86, 87]. Taken together, these data suggest that alterations in Pin1 activity could explain an inverse association between cancer and AD. Patients with less active Pin1 would be at a greater risk of developing AD and not cancer and, conversely, those with an active Pin1 would be more prone to develop cancer and not AD.

Wnt signaling pathway

The Wnt (wingless-type murine-mammary tumour virus integration site) signaling pathway is important for many developmental and adult processes, such as gastrulation, axis formation, cell polarity, organ development and maintenance of stem cell pluripotency and is remarkably conserved in a wide range of organisms, from Caenorhabditis elegans to humans [88]. In the canonical pathway, wnt proteins bind to cell-surface receptors composed of members of the Frizzled family and a low density lipoprotein receptor 5/6 (LRP 5/6). The receptor complex in turn is associated with a large cytoplasmic protein complex comprised of axin, (axis inhibition protein), APC (adenomatosis polyposis coli), CK1α (casein kinase 1 alpha), GSK-3β (glycogen synthase kinase 3 beta) and GβP/frat [88]. The activation of the pathway by the binding of Wnt proteins ultimately stabilizes cytoplasmic β-catenin that translocates to the nucleus and is involved in gene expression regulation that promotes several physiological functions, among them cell survival and proliferation, through the binding to TCF/LEF transcription factors and the expression of wnt-target genes. In the absence of Wnt binding GSK-3β phosphorylates β-catenin molecules which are then directed to the ubiquitin-mediated degradation pathway, thus preventing their survival-promoting action.

The Wnt signaling pathway has been related to cancer and neurodegeneration [8891]. Several components of the Wnt pathway have been implicated in carcinogenesis. Perturbations of the Wnt signaling pathway are best known to be involved in colorectal cancer [92, 93] and are associated with several other cancers including lung, prostate, breast [9296]. Recent evidence also shows that an upregulation of the Wnt signalling pathway is a key step in skin cancers, both for melanomas and for basal and squamous cell carcinomas [97102].

A role of the Wnt signaling pathway has also been implicated in AD [91, 103108]. The initial work of Inestrosa and collaborators found a relationship between Aβ-induced neurotoxicity and a loss of the wnt signaling pathway activity, with decreased cytoplasmic levels of β-catenin. They demonstrated that inhibition by lithium of GSK-3β, a central modulator of the Wnt pathway, protected rat hippocampal neurons from Abeta-induced damage [109]. Also, pre-treatment of neurons with wnt-3a conditioned media preserved neurons from the neurotoxic effect of Aβ [110]. In primary cultures of cortical and hippocampal neurons Aβ neurotoxicity increases the activation of GSK-3β, the hyperphospholylation of tau proteins, and loss of microtubule network [111, 112]. Wnt ligands are able to prevent the Aβ-induced decrease in the number of neurites on hippocampal primary cultures [113]. Therefore, defects in wnt signaling have been proposed in the pathogenesis of AD [90, 114, 115]. The role of β-catenin as a survival element in AD is reinforced by results showing that phosphorylation of tau stabilizes beta catenin, antagonizing apoptosis, and the knock down of β-catenin produces an increase in the number of apoptotic cells [116].

In accordance with a role of Wnt signaling in AD, a recent study among bipolar patients Nunes et al [117] reported a lower incidence of AD in those patients who had been taking lithium than in those without lithium therapy, suggesting that the inhibition of GSK3 might also have effects in clinical grounds. In a recent report they show a dose dependent reduction of GSK3β expression in hippocampal cells in culture and in the brain and leucocytes of rats treated with lithium [118]. β-catenin levels are markedly reduced in AD patients carrying presenilin-1 (PS-1) inherited mutations [119]. Furthermore, recent studies have shown that Apolipoprotein E4, known to be a risk factor for AD, inhibits the Wnt signaling pathway in PC12 cells [120], and an association between a highly conserved LRP6 polymorphism and the risk of developing late-onset Alzheimer’s disease in ApoE 4 allele carriers was found in a case-control and a large family-based study of AD patients [121]. LRP5/6 is a component of the receptor complex on Wnt, and interestingly, functional analyses revealed that the associated polymorphism of LRP6 has decreased β-catenin signaling in HEK293T cells [121].

In all, these results suggest that a deregulation of the Wnt signaling pathway could possibly explain our inverse association between cancer and AD. A subtle deregulation favoring Wnt activation could explain a greater tendency to develop tumors, and at the same time protect against degeneration, favoring neuronal survival. On the other hand, a small change towards suppression of Wnt signaling could explain a greater susceptibility to neuronal death or loss of dendritic spines, while at the same time protect against the development of cancer (Fig 2). Polymorphisms or perturbations of the epigenome [20] in key molecules in the pathway that might favor or unfavor the activity of the pathway could determine the chances of developing a cancer, thus avoiding neurodegeneration, or development of AD, and avoiding hyperproliferation.

Figure 2.

Figure 2

The wnt signaling pathway involvement in cancer and neurodegeneration. When wnt binds to the LRP-frizzled receptor in the surface of the cell,β-catenin is stabilized promoting expression of wnt target genes and proliferation. Subtle disequilibrium in any step of the pathway in a manner that determines activation of the pathway, such as increased expression or polimorphisms that induce activation of wnt or β-catenin would favor cancer development, preventing neurodegeneration. On the contrary, conditions that induce inactivation of the pathway would favor the development of Alzheimer’s disease or other degenerative disorder, and as a consequence protect from cancer development.

Since the inverse association between cancer and AD that was found in our study was present in cancers of different organs, a speculative biological mechanism should be applicable to all cells in the organism. That is, the alteration in the survival mechanisms that theoretically could protect from AD should be present in neurons as well as in other cells in the body, that would be then be more predisposed to develop cancer. And vice versa, defects in survival mechanisms that would favor AD development would protect against cancer development in all the cells in the organism. In favor of a systemic deregulation of cell survival mechanisms, it has been reported that lymphocytes from AD patients are more susceptible to cell death caused by apoptosis-inducing factors, compared to a similarly-aged control group [122]. Also, lymphocytes and fibroblasts from AD patients show increased levels of p53 compared to healthy controls of comparable age [123]. Furthermore, fibroblasts derived from the p53-mutated mice with enhanced p53 activity were more resistant to transformation by activated ras plus myc oncogene [26].

A hypothetical common biological mechanism explaining an inverse association between the development of cancer and AD could be extended to other neurodegenerative diseases characterized by increased cell death. A “prone to die” status of cells could favor all those diseases characterized by progressive neurodegeneration, such as Parkinson’s disease (PD), frontotemporal dementia and other tauopathies. In support of this, two longitudinal studies and a case control study suggest that overall cancer mortality risk and tumor frequency are reduced in individuals with PD, for both smoking-related and non-smoking related cancers [124126]. Interestingly, several of the genes that are now known to be associated with PD were studied in cancer research before their involvement in PD was recognized [127, 128]. In frontotemporal lobar degeneration loss-of-function mutations have been recently identified in progranulin in chromosome 17 [129/132]. Progranulin is a multifunctional protein expressed in peripheral tissues and in the central nervous system, both in neurons and glia, involved in wound healing and inflammation. Interestingly, it contributes to tumorigenesis in diverse cancers when overexpressed, including breast cancer, clear cell renal carcinoma, invasive ovarian carcinoma and glioblastoma, [133]. Not just neuronal cells would be in the prone-to-die state, but also cells in other tissues could be more susceptible to degeneration. Following this idea, it is conceivable to speculate that other degenerative systemic disorders, such as osteoarthritis or osteoporosis, could also be associated with a reduced risk of cancer, and vice versa, patients with a history of cancer could have a reduced risk of systemic degenerative disorders.

Concluding remarks

The finding of an inverse association between cancer and AD opens up several avenues of investigation that may lead to clues about the nature of both AD and cancer. A putative common biological mechanism that inversely operates in the two disorders, one leading to increased cell growth or survival, and the other to a higher risk of cell death, could explain these results. Understanding the basis of the association between cancer and AD is made more imperative considering that treatments currently under investigation to prevent and treat Alzheimer’s disease might lead to a greater risk of cancer development, and inversely, treatments to prevent cancer could predispose to the development of AD. Although much work remains to be done to determine whether cancer and AD are in fact linked via a common biological mechanism, the eventual identification of such a mechanism may provide insight into therapeutic strategies that could aid in preventing both disorders.

Acknowledgments

Supported by grants Fondecyt # 1080569 (MIB) and grants P50-AG05681 and P01-AG03991 from the National Institute on Aging, and the Charles and Joanne Knight Alzheimer Research Initiative of Washington University’s Alzheimer Disease Research Center (CR).

I thank Alison Goate, Rafi Kopan and John Morris for discussions, Osvaldo Alvarez and Sergio Lavandero for comments in revising the manuscript and Nibaldo Inestrosa for encouraging support and providing the Wnt figure.

References

  • 1.Roe CM, Behrens MI, Xiong C, Miller JP, Morris JC. Alzheimer’s disease and cancer. Neurology. 2005;64:895–898. doi: 10.1212/01.WNL.0000152889.94785.51. [DOI] [PubMed] [Google Scholar]
  • 2.Morris JC. The Clinical Dementia Rating (CDR): current version and scoring rules. Neurology. 1993;43:2412–2414. doi: 10.1212/wnl.43.11.2412-a. [DOI] [PubMed] [Google Scholar]
  • 3.Corsellis JAN. Mental Illness and the Ageing Brain. London: Oxford University Press; 1962. [Google Scholar]
  • 4.Tirumalasetti F, Han L, Birkett DP. The relationship between cancer and Alzheimer’s disease. J Am Geriatr Soc. 1991;39:840. doi: 10.1111/j.1532-5415.1991.tb02713.x. [DOI] [PubMed] [Google Scholar]
  • 5.DeSouky AL. The relationship between cancer and Alzheimer’s disease. J Am Geriatr Soc. 1992;40:1075. doi: 10.1111/j.1532-5415.1992.tb04490.x. [DOI] [PubMed] [Google Scholar]
  • 6.Yamada M, Sasaki H, Mimori Y Kasagi F, Sudoh S, Ikeda J, et al. Prevalence and risks of dementia in the Japanese population: RERF’s adult health study Hiroshima subjects. Radiation Effects Research Foundation. J Am Geriatr Soc. 1999;47:189–195. doi: 10.1111/j.1532-5415.1999.tb04577.x. [DOI] [PubMed] [Google Scholar]
  • 7.Beard CM, Kokmen E, Sigler C, Smith GE, Petterson T, O’Brien PC. Cause of death in Alzheimer’s disease. Ann Epidemiol. 1996;6:195–200. doi: 10.1016/1047-2797(95)00068-2. [DOI] [PubMed] [Google Scholar]
  • 8.Ganguli M, Dodge HH, Shen C, Pandav RS, DeKosky ST. Alzheimer Disease and Mortality: A 15-Year Epidemiological Study. Arch Neurol. 2005;62:779–784. doi: 10.1001/archneur.62.5.779. [DOI] [PubMed] [Google Scholar]
  • 9.Parker SL, Tong T, Bolden S, Wingo PA. Cancer statistics. CA Cancer J Clin. 1996;46:5–27. doi: 10.3322/canjclin.46.1.5. [DOI] [PubMed] [Google Scholar]
  • 10.Behrens MI, Roe CM, Morris JC. Inverse association between cancer and dementia of the Alzheimers type. In: von Bernhardy R, Inestrosa NC, editors. Neurodegenerative diseases: From Molecular Concepts to Therapeutic Targets. Nova Science Publishers, Inc; New York: 2008. pp. 111–120. [Google Scholar]
  • 11.Hardy J. Alzheimer’s disease: the amyloid cascade hypothesis: an update and reappraisal. J Alzheimers Dis. 2006;9:151–153. doi: 10.3233/jad-2006-9s317. [DOI] [PubMed] [Google Scholar]
  • 12.Billingsley ML, Kincaid RL. Regulated phosphorylation and dephosphorylation of tau protein: effects on microtubule interaction, intracellular trafficking and neurodegeneration. Biochem J. 1997;323:577–591. doi: 10.1042/bj3230577. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Alonso AD, Grundke-Iqbal I, Barra HS, Iqbal K. Abnormal phosphorylation of tau and the mechanism of Alzheimer neurofibrillary degeneration: sequestration of microtubule-associated proteins 1 and 2 and the disassembly of microtubules by the abnormal tau. Proc Natl Acad Sci USA. 1997;94:298–303. doi: 10.1073/pnas.94.1.298. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Green KN, LaFerla FM. Linking calcium to Abeta and Alzheimer’s disease. Neuron. 2008;59:190–194. doi: 10.1016/j.neuron.2008.07.013. [DOI] [PubMed] [Google Scholar]
  • 15.Shankar GM, Li S, Mehta TH, Garcia-Munoz A, Shepardson NE, Smith I, Brett FM, et al. Amyloid-beta protein dimers isolated directly from Alzheimer’s brains impair synaptic plasticity and memory. Nat Med. 2008;14:837–842. doi: 10.1038/nm1782. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Sherr CJ. Principles of tumor suppression. Cell. 2004;116:235–246. doi: 10.1016/s0092-8674(03)01075-4. [DOI] [PubMed] [Google Scholar]
  • 17.Riley T, Sontag E, Chen P, Levine A. Transcriptional control of human p53-regulated genes. Nat Rev Mol Cell Biol. 2008;9:402–412. doi: 10.1038/nrm2395. [DOI] [PubMed] [Google Scholar]
  • 18.Murray-Zmijewski F, Slee EA, Lu X. A complex barcode underlies the heterogeneous response of p53 to stress. Nat Rev Mol Cell Biol. 2008;9:702–712. doi: 10.1038/nrm2451. [DOI] [PubMed] [Google Scholar]
  • 19.Sharpless NE, DePinho RA. How stem cells age and why this makes us grow old. Nat Rev Mol Cell Biol. 2007;8:703–713. doi: 10.1038/nrm2241. [DOI] [PubMed] [Google Scholar]
  • 20.Tremolizzo L, Rodriguez-Menendez V, Brighina L, Ferrarese C. Is the inverse association between Alzheimer’s disease and cancer the result of a different propensity to methylate DNA? Medical Hypothesis. 2006;66:1251–1252. doi: 10.1016/j.mehy.2005.12.022. [DOI] [PubMed] [Google Scholar]
  • 21.Campisi J. Aging and cancer cell biology, 2008. Aging Cell. 2008;7:281–284. doi: 10.1111/j.1474-9726.2008.00383.x. [DOI] [PubMed] [Google Scholar]
  • 22.Papazoglu C, Mills AA. p53: at the crossroad between cancer and ageing. J Pathol. 2007;211:124–133. doi: 10.1002/path.2086. [DOI] [PubMed] [Google Scholar]
  • 23.Campisi J. Aging, tumor suppression and cancer: high wire-act! Mech Ageing Dev. 2005;126:51–58. doi: 10.1016/j.mad.2004.09.024. [DOI] [PubMed] [Google Scholar]
  • 24.Donehower LA, Harvey M, Slagle BL, McArthur MJ, Montgomery CA, Jr, Butel JS, Bradley A. Mice deficient for p53 are developmentally normal but susceptible to spontaneous tumors. Nature. 1992;356:215–221. doi: 10.1038/356215a0. [DOI] [PubMed] [Google Scholar]
  • 25.Meletis K, Wirta V, Hede SM, Nister M, Lundeberg J, Frisen J. p53 suppresses the self-renewal of adult neural stem cells. Development. 2006;133:363–369. doi: 10.1242/dev.02208. [DOI] [PubMed] [Google Scholar]
  • 26.Tyner SD, Venkatachalam S, Choi J, Jones S, Ghebranious N, Igelmann H, et al. p53 mutant mice that display early ageing-associated phenotypes. Nature. 2002;415:45–53. doi: 10.1038/415045a. [DOI] [PubMed] [Google Scholar]
  • 27.Sharpless NE, DePinho RA. Telomeres, stem cells, senescence, and cancer. J Clin Invest. 2004;113:160–168. doi: 10.1172/JCI20761. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Varela I, Cadinanos J, Pendas AM, Gutierrez-Fernandez A, Folgueras AR, Sanchez LM, et al. Accelerated ageing in mice deficient in Zmpste24 protease is linked to p53 signalling activation. Nature. 2005;437:564–568. doi: 10.1038/nature04019. [DOI] [PubMed] [Google Scholar]
  • 29.Maier B, Gluba W, Bernier B, Turner T, Mohammad K, Guise T, et al. Modulation of mammalian life span by the short isoform of p53. Genes Dev. 2004;18:306–319. doi: 10.1101/gad.1162404. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Matheu A, Maraver A, Klatt P, Flores I, Garcia-Cao I, Borras C, et al. Delayed ageing through damage protection by the Arf/p53 pathway. Nature. 2007;448:375–379. doi: 10.1038/nature05949. [DOI] [PubMed] [Google Scholar]
  • 31.Garcia-Cao I, Garcia-Cao M, Martin-Caballero J, Criado LM, Klatt P, Flores JM, et al. “Super p53” mice exhibit enhanced DNA damage response, are tumor resistant and age normally. EMBO J. 2002;21:6225–6235. doi: 10.1093/emboj/cdf595. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Mendrysa SM, O’Leary KA, McElwee MK, Michalowski J, Eisenman RN, Powell DA, Perry ME. Tumor suppression and normal aging in mice with constitutively high p53 activity. Genes Dev. 2006;20:16–21. doi: 10.1101/gad.1378506. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Di Giovanni S, Knights CD, Rao M, Yakovlev A, Beers J, Catania J, et al. The tumor suppressor protein p53 is required for neurite outgrowth and axon regeneration. EMBO J. 2006;25:4084–4096. doi: 10.1038/sj.emboj.7601292. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Knights CD, Catania J, Di Giovanni S, Muratoglu S, Perez R, Swartzbeck A, et al. Distinct p53 acetylation cassettes differentially influence gene-expression patterns and cell fate. J Cell Biol. 2006;173:533–544. doi: 10.1083/jcb.200512059. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Bae BI, Xu H, Igarashi S, Fujimuro M, Agrawal N, Taya Y, et al. p53 mediates cellular dysfunction and behavioral abnormalities in Huntington’s disease. Neuron. 2005;47:29–41. doi: 10.1016/j.neuron.2005.06.005. [DOI] [PubMed] [Google Scholar]
  • 36.Martin LJ. p53 is abnormally elevated and active in the CNS of patients with amyotrophic lateral sclerosis. Neurobiol Dis. 2000;7:613–622. doi: 10.1006/nbdi.2000.0314. [DOI] [PubMed] [Google Scholar]
  • 37.Martin LJ. Transgenic mice with human mutant genes causing Parkinson’s disease and amyotrophic lateral sclerosis provide common insight into mechanisms of motor neuron selective vulnerability to degeneration. Rev Neurosci. 2007;18:115–136. doi: 10.1515/revneuro.2007.18.2.115. [DOI] [PubMed] [Google Scholar]
  • 38.Mogi M, Kondo T, Mizuno Y, Nagatsu T. p53 protein, interferon-gamma, and NF-kappaB levels are elevated in the parkinsonian brain. Neurosci Lett. 2007;414:94–97. doi: 10.1016/j.neulet.2006.12.003. [DOI] [PubMed] [Google Scholar]
  • 39.LaFerla FM, Hall CK, Ngo L, Jay G. Extracellular deposition of beta-amyloid upon p53-dependent neuronal cell death in transgenic mice. J Clin Invest. 1996;98:1626–1632. doi: 10.1172/JCI118957. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Cancino GI, Toledo EM, Leal NR, Hernandez DE, Yevenes LF, Inestrosa NC, Alvarez AR. STI571 prevents apoptosis, tau phosphorylation and behavioural impairments induced by Alzheimer’s beta-amyloid deposits. Brain. 2008;131:2425–2442. doi: 10.1093/brain/awn125. [DOI] [PubMed] [Google Scholar]
  • 41.Ohyagi Y, Asahara H, Chui DH, Tsuruta Y, Sakae N, Miyoshi K, et al. Intracellular Abeta42 activates p53 promoter: a pathway to neurodegeneration in Alzheimer’s disease. FASEB J. 2005;19:255–257. doi: 10.1096/fj.04-2637fje. [DOI] [PubMed] [Google Scholar]
  • 42.Cenini G, Sultana R, Memo M, Butterfield DA. Elevated levels of pro-apoptotic p53 and its oxidative modification by the lipid peroxidation product, HNE, in brain from subjects with amnestic mild cognitive impairment and Alzheimer’s disease. J Cell Mol Med. 2008;12:987–994. doi: 10.1111/j.1582-4934.2008.00163.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Hooper C, Meimaridou E, Tavassoli M, Melino G, Lovestone S, Killick R. p53 is upregulated in Alzheimer’s disease and induces tau phosphorylation in HEK293a cells. Neurosci Lett. 2007;418:34–37. doi: 10.1016/j.neulet.2007.03.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Nguyen HN, Lee MS, Hwang DY, Kim YK, Yoon do Y, Lee JW, Yun YP, et al. Mutant presenilin 2 increased oxidative stress and p53 expression in neuronal cells. Biochem Biophys Res Commun. 2007;357:174–180. doi: 10.1016/j.bbrc.2007.03.119. [DOI] [PubMed] [Google Scholar]
  • 45.Alves da Costa C, Sunyach C, Pardossi-Piquard R, Sevalle J, Vincent B, Boyer N, et al. Presenilin-dependent gamma-secretase-mediated control of p53-associated cell death in Alzheimer’s disease. J Neurosci. 2006;26:6377–6385. doi: 10.1523/JNEUROSCI.0651-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Polager S, Ginsberg D. E2F - at the crossroads of life and death. Trends Cell Biol. 2008;18:528–535. doi: 10.1016/j.tcb.2008.08.003. [DOI] [PubMed] [Google Scholar]
  • 47.Fridman AL, Tainsky MA. Critical pathways in cellular senescence and immortalization revealed by gene expression profiling. Oncogene. 2008;27:5975–5987. doi: 10.1038/onc.2008.213. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Zindy F, Quelle DE, Roussel MF, Sherr CJ. Expression of the p16INK4a tumor suppressor versus other INK4 family members during mouse development and aging. Oncogene. 1997;15:203–211. doi: 10.1038/sj.onc.1201178. [DOI] [PubMed] [Google Scholar]
  • 49.Krishnamurthy J, Torrice C, Ramsey MR, Kovalev GI, Al-Regaiey K, Su L, Sharpless NE. Ink4a/Arf expression is a biomarker of aging. J Clin Invest. 2004;114:1299–1307. doi: 10.1172/JCI22475. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Molofsky AV, Slutsky SG, Joseph NM, He S, Pardal R, Krishnamurthy J, Sharpless NE, Morrison SJ. Increasing p16INK4a expression decreases forebrain progenitors and neurogenesis during ageing. Nature. 2006;443:448–452. doi: 10.1038/nature05091. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Krishnamurthy J, Ramsey MR, Ligon KL, Torrice C, Koh A, Bonner-Weir S, Sharpless NE. p16INK4a induces an age-dependent decline in islet regenerative potential. Nature. 2006;443:453–457. doi: 10.1038/nature05092. [DOI] [PubMed] [Google Scholar]
  • 52.Janzen V, Forkert R, Fleming HE, Saito Y, Waring MT, Dombkowski DM, et al. Stem-cell ageing modified by the cyclin-dependent kinase inhibitor p16INK4a. Nature. 2006;443:421–426. doi: 10.1038/nature05159. [DOI] [PubMed] [Google Scholar]
  • 53.Sharpless NE, DePinho RA. How stem cells age and why this makes us grow old. Nat Rev Mol Cell Biol. 2007;8:703–713. doi: 10.1038/nrm2241. [DOI] [PubMed] [Google Scholar]
  • 54.Blalock EM, Geddes JW, Chen KC, Porter NM, Markesbery WR, Landfield PW. Incipient Alzheimer’s disease: microarray correlation analyses reveal major transcriptional and tumor suppressor responses. Proc Natl Acad Sci USA. 2004;101:2173–2178. doi: 10.1073/pnas.0308512100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Yaffe MB, Schutkowski M, Shen M, Zhou XZ, Stukenberg PT, Rahfeld JU, et al. Sequence-specific and phosphorylation-dependent proline isomerization: a potential mitotic regulatory mechanism. Science. 1997;278:1957–1960. doi: 10.1126/science.278.5345.1957. [DOI] [PubMed] [Google Scholar]
  • 56.Lu KP, Zhou XZ. The prolyl isomerase PIN1: a pivotal new twist in phosphorylation signalling and disease. Nat Rev Mol Cell Biol. 2007;8:904–916. doi: 10.1038/nrm2261. [DOI] [PubMed] [Google Scholar]
  • 57.Lu KP. Pinning down cell signaling, cancer and Alzheimer’s disease. Trends Biochem Sci. 2004;29:200–209. doi: 10.1016/j.tibs.2004.02.002. [DOI] [PubMed] [Google Scholar]
  • 58.Takahashi K, Uchida C, Shin RW, Shimazaki K, Uchida T. Prolyl isomerase, Pin1: new findings of post-translational modifications and physiological substrates in cancer, asthma and Alzheimer’s disease. Cell Mol Life Sci. 2008;65:359–375. doi: 10.1007/s00018-007-7270-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Lu PJ, Wulf G, Zhou XZ, Davies P, Lu KP. The prolyl isomerase Pin1 restores the function of Alzheimer-associated phosphorylated tau protein. Nature. 1999;399:784–788. doi: 10.1038/21650. [DOI] [PubMed] [Google Scholar]
  • 60.Zhou XZ, Kops O, Werner A, Lu PJ, Shen M, Stoller G, et al. Pin1-dependent prolyl isomerization regulates dephosphorylation of Cdc25C and tau proteins. Mol Cell. 2000;6:873–883. doi: 10.1016/s1097-2765(05)00083-3. [DOI] [PubMed] [Google Scholar]
  • 61.Liou YC, Sun A, Ryo A, Zhou XZ, Yu ZX, Huang HK, Uchida T, Bronson R, Bing G, Li X, Hunter T, Lu KP. Role of the prolyl isomerase Pin1 in protecting against age-dependent neurodegeneration. Nature. 2003;424:556–561. doi: 10.1038/nature01832. [DOI] [PubMed] [Google Scholar]
  • 62.Pastorino L, Sun A, Lu PJ, Zhou XZ, Balastik M, Finn G, et al. The prolyl isomerase Pin1 regulates amyloid precursor protein processing and amyloid-beta production. Nature. 2006;440:528–534. doi: 10.1038/nature04543. [DOI] [PubMed] [Google Scholar]
  • 63.Hamdane M, Dourlen P, Bretteville A, Sambo AV, Ferreira S, Ando K, et al. Pin1 allows for differential Tau dephosphorylation in neuronal cells. Mol Cell Neurosci. 2006;32:155–160. doi: 10.1016/j.mcn.2006.03.006. [DOI] [PubMed] [Google Scholar]
  • 64.Sultana R, Boyd-Kimball D, Poon HF, Cai J, Pierce WM, Klein JB, et al. Oxidative modification and down-regulation of Pin1 in Alzheimer’s disease hippocampus: A redox proteomics analysis. Neurobiol Aging. 2006;27:918–925. doi: 10.1016/j.neurobiolaging.2005.05.005. [DOI] [PubMed] [Google Scholar]
  • 65.Wang S, Simon BP, Bennett DA, Schneider JA, Malter JS, Wang DS. The significance of Pin1 in the development of Alzheimer’s disease. J Alzheimers Dis. 2007;11:13–23. doi: 10.3233/jad-2007-11105. [DOI] [PubMed] [Google Scholar]
  • 66.Lim J, Balastik M, Lee TH, Nakamura K, Liou YC, Sun A, Finn G, Pastorino L, Lee VM, Lu KP. Pin1 has opposite effects on wild-type and P301L tau stability and tauopathy. J Clin Invest. 2008;118:1877–1889. doi: 10.1172/JCI34308. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Segat L, Pontillo A, Annoni G, Trabattoni D, Vergani C, Clerici M, et al. PIN1 promoter polymorphisms are associated with Alzheimer’s disease. Neurobiol Aging. 2007;28:69–74. doi: 10.1016/j.neurobiolaging.2005.11.009. [DOI] [PubMed] [Google Scholar]
  • 68.Wijsman EM, Daw EW, Yu CE, Payami H, Steinbart EJ, Nochlin D, et al. Evidence for a novel late-onset Alzheimer disease locus on chromosome 19p13.2. Am J Hum Genet. 2004;75:398–409. doi: 10.1086/423393. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Poli M, Gatta LB, Dominici R, Lovati C, Mariani C, Albertini A, Finazzi D. DNA sequence variations in the prolyl isomerase Pin1 gene and Alzheimer’s disease. Neurosci Lett. 2005;389:66–70. doi: 10.1016/j.neulet.2005.07.027. [DOI] [PubMed] [Google Scholar]
  • 70.Nowotny P, Bertelsen S, Hinrichs AL, Kauwe JS, Mayo K, Jacquart S, et al. Association studies between common variants in prolyl isomerase Pin1 and the risk for late-onset Alzheimer’s disease. Neurosci Lett. 2007;419:15–17. doi: 10.1016/j.neulet.2007.03.071. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Lambert JC, Bensemain F, Chapuis J, Cottel D, Amouyel P. Association study of the PIN1 gene with Alzheimer’s disease. Neurosci Lett. 2006;402:259–261. doi: 10.1016/j.neulet.2006.04.010. [DOI] [PubMed] [Google Scholar]
  • 72.Kesavapany S, Patel V, Zheng YL, Pareek TK, Bjelogrlic M, Albers W, et al. Inhibition of Pin1 reduces glutamate-induced perikaryal accumulation of phosphorylated neurofilament-H in neurons. Mol Biol Cell. 2007;18:3645–3655. doi: 10.1091/mbc.E07-03-0237. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Wulf GM, Ryo A, Wulf GG, Lee SW, Niu T, Petkova V, et al. Pin1 is overexpressed in breast cancer and cooperates with Ras signaling in increasing the transcriptional activity of c-Jun towards cyclin D1. EMBO J. 2001;20:3459–3472. doi: 10.1093/emboj/20.13.3459. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Ryo A, Nakamura M, Wulf G, Liou YC, Lu KP. Pin1 regulates turnover and subcellular localization of beta-catenin by inhibiting its interaction with APC. Nat Cell Biol. 2001;3:793–801. doi: 10.1038/ncb0901-793. [DOI] [PubMed] [Google Scholar]
  • 75.He J, Zhou F, Shao K, Hang J, Wang H, Rayburn E, et al. Overexpression of Pin1 in non-small cell lung cancer (NSCLC) and its correlation with lymph node metastases. Lung Cancer. 2007;56:51–58. doi: 10.1016/j.lungcan.2006.11.024. [DOI] [PubMed] [Google Scholar]
  • 76.Pang R, Lee TK, Poon RT, Fan ST, Wong KB, Kwong YL, Tse E. Pin1 interacts with a specific serine-proline motif of hepatitis B virus X-protein to enhance hepatocarcinogenesis. Gastroenterology. 2007;132:1088–1103. doi: 10.1053/j.gastro.2006.12.030. [DOI] [PubMed] [Google Scholar]
  • 77.Bao L, Kimzey A, Sauter G, Sowadski JM, Lu KP, Wang DG. Prevalent overexpression of prolyl isomerase Pin1 in human cancers. Am J Pathol. 2004;164:1727–1737. doi: 10.1016/S0002-9440(10)63731-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Dougherty MK, Muller J, Ritt DA, Zhou M, Zhou XZ, Copeland TD, et al. Regulation of Raf-1 by direct feedback phosphorylation. Mol Cell. 2005;17:215–224. doi: 10.1016/j.molcel.2004.11.055. [DOI] [PubMed] [Google Scholar]
  • 79.Kim MR, Choi HS, Heo TH, Hwang SW, Kang KW. Induction of vascular endothelial growth factor by peptidyl-prolyl isomerase Pin1 in breast cancer cells. Biochem Biophys Res Commun. 2008;369:547–553. doi: 10.1016/j.bbrc.2008.02.045. [DOI] [PubMed] [Google Scholar]
  • 80.Ryo A, Suizu F, Yoshida Y, Perrem K, Liou YC, Wulf G, et al. Regulation of NF-kappaB signaling by Pin1-dependent prolyl isomerization and ubiquitin-mediated proteolysis of p65/RelA. Mol Cell. 2003;12:1413–1426. doi: 10.1016/s1097-2765(03)00490-8. [DOI] [PubMed] [Google Scholar]
  • 81.van Drogen F, Sangfelt O, Malyukova A, Matskova L, Yeh E, Means AR, Reed SI. Ubiquitylation of cyclin E requires the sequential function of SCF complexes containing distinct hCdc4 isoforms. Mol Cell. 2006;23:37–48. doi: 10.1016/j.molcel.2006.05.020. [DOI] [PubMed] [Google Scholar]
  • 82.Suizu F, Ryo A, Wulf G, Lim J, Lu KP. Pin1 regulates centrosome duplication, and its overexpression induces centrosome amplification, chromosome instability, and oncogenesis. Mol Cell Biol. 2006;26:1463–1479. doi: 10.1128/MCB.26.4.1463-1479.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Mantovani F, Tocco F, Girardini J, Smith P, Gasco M, Lu X, et al. The prolyl isomerase Pin1 orchestrates p53 acetylation and dissociation from the apoptosis inhibitor iASPP. Nat Struct Mol Biol. 2007;14:912–920. doi: 10.1038/nsmb1306. [DOI] [PubMed] [Google Scholar]
  • 84.Braithwaite AW, Del Sal G, Lu X. Some p53-binding proteins that can function as arbiters of life and death. Cell Death Differ. 2006;13:984–993. doi: 10.1038/sj.cdd.4401924. [DOI] [PubMed] [Google Scholar]
  • 85.Zacchi P, Gostissa M, Uchida T, Salvagno C, Avolio F, Volinia S, Ronai Z, Blandino G, Schneider C, Del Sal G. The prolyl isomerase Pin1 reveals a mechanism to control p53 functions after genotoxic insults. Nature. 2002;419:853–857. doi: 10.1038/nature01120. [DOI] [PubMed] [Google Scholar]
  • 86.Becker EB, Bonni A. Pin1 in neuronal apoptosis. Cell Cycle. 2007;6:1332–1335. doi: 10.4161/cc.6.11.4316. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Barone MC, Desouza LA, Freeman RS. Pin1 promotes cell death in NGF-dependent neurons through a mechanism requiring c-Jun activity. J Neurochem. 2008;106:734–745. doi: 10.1111/j.1471-4159.2008.05427.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Coombs GS, Covey TM, Virshup DM. Wnt signaling in development, disease and translational medicine. Curr Drug Targets. 2008;9:513–531. doi: 10.2174/138945008784911796. [DOI] [PubMed] [Google Scholar]
  • 89.Maiese K, Li F, Chong ZZ, Shang YC. The Wnt signaling pathway: aging gracefully as a protectionist? Pharmacol Ther. 2008;118:58–81. doi: 10.1016/j.pharmthera.2008.01.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Caricasole A, Bakker A, Copani A, Nicoletti F, Gaviraghi G, Terstappen GC. Two sides of the same coin: Wnt signaling in neurodegeneration and neuro-oncology. Biosci Rep. 2005;25:309–327. doi: 10.1007/s10540-005-2893-6. [DOI] [PubMed] [Google Scholar]
  • 91.Inestrosa NC, Toledo EM. The role of Wnt signaling in neuronal dysfunction in Alzheimer’s Disease. Mol Neurodegener. 2008;3:9. doi: 10.1186/1750-1326-3-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Buongiorno P, Pethe VV, Charames GS, Esufali S, Bapat B. Rac1 GTPase and the Rac1 exchange factor Tiam1 associate with Wnt-responsive promoters to enhance beta-catenin/TCF-dependent transcription in colorectal cancer cells. Mol Cancer. 2008;7:73. doi: 10.1186/1476-4598-7-73. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Firestein R, Bass AJ, Kim SY, Dunn IF, Silver SJ, Guney I, et al. CDK8 is a colorectal cancer oncogene that regulates beta-catenin activity. Nature. 2008;455:547–551. doi: 10.1038/nature07179. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Karim R, Tse G, Putti T, Scolyer R, Lee S. The significance of the Wnt pathway in the pathology of human cancers. Pathology. 2004;36:120–128. doi: 10.1080/00313020410001671957. [DOI] [PubMed] [Google Scholar]
  • 95.Fiorentino M, Zadra G, Palescandolo E, Fedele G, Bailey D, Fiore C, et al. Overexpression of fatty acid synthase is associated with palmitoylation of Wnt1 and cytoplasmic stabilization of beta-catenin in prostate cancer. Lab Invest. 2008 Oct 6; doi: 10.1038/labinvest.2008.97. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Veeck J, Bektas N, Hartmann A, Kristiansen G, Heindrichs U, Knuchel R, Dahl E. Wnt signaling in human breast cancer: expression of the putative Wnt inhibitor Dickkopf-3 (DKK3) is frequently suppressed by promoter hypermethylation in mammary tumors. Breast Cancer Res. 2008;10:R82. doi: 10.1186/bcr2151. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Coe BP, Lockwood WW, Girard L, Chari R, Macaulay C, Lam S, et al. Differential disruption of cell cycle pathways in small cell and non-small cell lung cancer. Br J Cancer. 2006;94:1927–1935. doi: 10.1038/sj.bjc.6603167. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Da Forno PD, Pringle JH, Hutchinson P, Osborn J, Huang Q, Potter L, et al. WNT5A expression increases during melanoma progression and correlates with outcome. Clin Cancer Res. 2008;14:5825–5832. doi: 10.1158/1078-0432.CCR-07-5104. [DOI] [PubMed] [Google Scholar]
  • 99.Liu ZL, Li Y, Kong QY, Zhan C, Wang Q, Chen XY, et al. Immunohistochemical profiling of Wnt, NF-kappaB, Stat3 and Notch signaling in human epidermal tumors. J Dermatol Sci. 2008;52:133–136. doi: 10.1016/j.jdermsci.2008.06.011. [DOI] [PubMed] [Google Scholar]
  • 100.Hoseong Yang S, Andl T, Grachtchouk V, Wang A, Liu J, Syu LJ, et al. Pathological responses to oncogenic Hedgehog signaling in skin are dependent on canonical Wnt/beta-catenin signaling. Nat Genet. 2008 Aug 1; doi: 10.1038/ng.192. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Malanchi I, Peinado H, Kassen D, Hussenet T, Metzger D, Chambon P, et al. Cutaneous cancer stem cell maintenance is dependent on beta-catenin signalling. Nature. 2008;452:650–653. doi: 10.1038/nature06835. [DOI] [PubMed] [Google Scholar]
  • 102.Asplund A, Gry Bjorklund M, Sundquist C, Stromberg S, Edlund K, Ostman A, et al. Expression profiling of microdissected cell populations selected from basal cells in normal epidermis and basal cell carcinoma. Br J Dermatol. 2008;158:527–538. doi: 10.1111/j.1365-2133.2007.08418.x. [DOI] [PubMed] [Google Scholar]
  • 103.Adams PD, Enders GH. Wnt signaling and senescence: A tug of war in early neoplasia? Cancer Biol Ther. 2008 Nov 7;7(11) doi: 10.4161/cbt.7.11.6943. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Toledo EM, Colombres M, Inestrosa NC. Wnt signaling in neuroprotection and stem cell differentiation. Prog Neurobiol. 2008 Aug 19; doi: 10.1016/j.pneurobio.2008.08.001. [DOI] [PubMed] [Google Scholar]
  • 105.De Ferrari GV, Moon RT. The ups and downs of Wnt signaling in prevalent neurological disorders. Oncogene. 2006;25:7545–7553. doi: 10.1038/sj.onc.1210064. [DOI] [PubMed] [Google Scholar]
  • 106.Magdesian MH, Carvalho MM, Mendes FA, Saraiva LM, Juliano MA, Juliano L, Garcia-Abreu J, Ferreira ST. Amyloid-beta binds to the extracellular cysteine-rich domain of Frizzled and inhibits Wnt/beta-catenin signaling. J Biol Chem. 2008;283:9359–9368. doi: 10.1074/jbc.M707108200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Caricasole A, Copani A, Caraci F, Aronica E, Rozemuller AJ, Caruso A, et al. Induction of Dickkopf-1, a negative modulator of the Wnt pathway, is associated with neuronal degeneration in Alzheimer’s brain. J Neurosci. 2004;24:6021–6027. doi: 10.1523/JNEUROSCI.1381-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Ghanevati M, Miller CA. Phospho-□-Catenin Accumulation in Alzheimer’s Disease and in Aggresomes Attributable to Proteasome Dysfunction. J Mol Neurosci. 2005;25:79–94. doi: 10.1385/JMN:25:1:079. [DOI] [PubMed] [Google Scholar]
  • 109.De Ferrari GV, Chacon MA, Barria MI, Garrido JL, Godoy JA, Olivares G, et al. Activation of Wnt signaling rescues neurodegeneration and behavioral impairments induced by β-amyloid fibrils. Mol Psychiatry. 2003;8:195–208. doi: 10.1038/sj.mp.4001208. [DOI] [PubMed] [Google Scholar]
  • 110.Alvarez AR, Godoy JA, Mullendorff K, Olivares GH, Bronfman M, Inestrosa NC. Wnt-3a overcomes β-amyloid toxicity in rat hippocampal neurons. Exp Cell Res. 2004;297:186–196. doi: 10.1016/j.yexcr.2004.02.028. [DOI] [PubMed] [Google Scholar]
  • 111.Busciglio J, Lorenzo A, Yeh J, Yankner BA. b-Amyloid fibrils induce tau phosphorylation and loss of microtubule binding. Neuron. 1995;14:879–888. doi: 10.1016/0896-6273(95)90232-5. [DOI] [PubMed] [Google Scholar]
  • 112.Takashima A, Honda T, Yasutake K, Michel G, Murayama O, Murayama M, et al. Activation of tau protein kinase I/glycogen synthase kinase-3b by amyloid b peptide (25–35) enhances phosphorylation of tau in hippocampal neurons. Neurosci Res. 1998;31:317–323. doi: 10.1016/s0168-0102(98)00061-3. [DOI] [PubMed] [Google Scholar]
  • 113.Inestrosa NC, Varela-Nallar L, Grabowski CP, Colombres M. Synaptotoxicity in Alzheimer’s disease: the Wnt signaling pathway as a molecular target. IUBMB Life. 2007;59:316–321. doi: 10.1080/15216540701242490. [DOI] [PubMed] [Google Scholar]
  • 114.De Ferrari GV, Inestrosa NC. Wnt signaling function in Alzheimer’s disease. Brain Res Brain Res Rev. 2000;33:1–12. doi: 10.1016/s0165-0173(00)00021-7. [DOI] [PubMed] [Google Scholar]
  • 115.Mudher A, Lovestone S. Alzheimer’s disease-do tauists and baptists finally shake hands? Trends Neurosci. 2002;25:22–26. doi: 10.1016/s0166-2236(00)02031-2. [DOI] [PubMed] [Google Scholar]
  • 116.Li HL, Wang HH, Liu SJ, Deng YQ, Zhang YJ, Tian Q, et al. Phosphorylation of tau antagonizes apoptosis by stabilizing beta-catenin, a mechanism involved in Alzheimer’s neurodegeneration. Proc Natl Acad Sci U S A. 2007;104:3591–3596. doi: 10.1073/pnas.0609303104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Nunes PV, Forlenza OV, Gattaz WF. Lithium and risk for Alzheimer’s disease in elderly patients with bipolar disorder. Br J Psychiatry. 2007;190:359–360. doi: 10.1192/bjp.bp.106.029868. [DOI] [PubMed] [Google Scholar]
  • 118.Mendes CT, Mury FB, de Sa Moreira E, Alberto FL, Forlenza OV, Dias-Neto E, Gattaz WF. Lithium reduces Gsk3b mRNA levels: implications for Alzheimer Disease. Eur Arch Psychiatry Clin Neurosci. 2008 Oct 17; doi: 10.1007/s00406-008-0828-5. [DOI] [PubMed] [Google Scholar]
  • 119.Zhang Z, Hartmann H, Do VM, Abramowski D, Sturchler-Pierrat C, Staufenbiel M, Sommer B, van de Wetering M, Clevers H, Saftig P, et al. Destabilization of β-catenin by mutations in presenilin-1 potentiates neuronal apoptosis. Nature. 1998;395:698–702. doi: 10.1038/27208. [DOI] [PubMed] [Google Scholar]
  • 120.Caruso A, Motolese M, Iacovelli L, Caraci F, Copani A, Nicoletti F, Terstappen GC, Gaviraghi G, Caricasole A. Inhibition of the canonical Wnt signaling pathway by apolipoprotein E4 in PC12 cells. J Neurochem. 2006;98:364–371. doi: 10.1111/j.1471-4159.2006.03867.x. [DOI] [PubMed] [Google Scholar]
  • 121.De Ferrari GV, Papassotiropoulos A, Biechele T, Wavrant De-Vrieze F, Avila ME, Major MB, Myers A, Saez K, Henriquez JP, Zhao A, et al. Common genetic variation within the low-density lipoprotein receptor-related protein 6 and late-onset Alzheimer’s disease. Proc Natl Acad Sci USA. 2007;104:9434–9439. doi: 10.1073/pnas.0603523104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Lombardi VR, Garcia M, Rey L, Cacabelos R. Characterization of cytokine production, screening of lymphocyte subset patterns and in vitro apoptosis in healthy and Alzheimer’s Disease (AD) individuals. J Neuroimmunol. 1999;97:163–171. doi: 10.1016/s0165-5728(99)00046-6. [DOI] [PubMed] [Google Scholar]
  • 123.Uberti D, Lanni C, Racchi M, Govoni S, Memo M. Conformationally altered p53: a putative peripheral marker for Alzheimer’s disease. Neurodegener Dis. 2008;5:209–211. doi: 10.1159/000113704. [DOI] [PubMed] [Google Scholar]
  • 124.Vanacore N, Spila-Alegiani S, Raschetti R, Meco G. Mortality cancer risk in parkinsonian patients: a population-based study. Neurology. 1999;52:395–398. doi: 10.1212/wnl.52.2.395. [DOI] [PubMed] [Google Scholar]
  • 125.D’Amelio M, Ragonese P, Morgante L, Epifanio A, Callari G, Salemi G, et al. Tumor diagnosis preceding Parkinson’s disease: a case-control study. Mov Disord. 2004;19:807–811. doi: 10.1002/mds.20123. [DOI] [PubMed] [Google Scholar]
  • 126.Olsen JH, Friis S, Frederiksen K, McLaughlin JK, Mellemkjaer L, Moller H. Atypical cancer pattern in patients with Parkinson’s disease. Br J Cancer. 2005;92:201–205. doi: 10.1038/sj.bjc.6602279. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.West AB, Dawson VL, Dawson TM. To die or grow: Parkinson’s disease and cancer. Trends Neurosci. 2005;28:348–352. doi: 10.1016/j.tins.2005.05.002. [DOI] [PubMed] [Google Scholar]
  • 128.Staropoli JF. Tumorigenesis and neurodegeneration: two sides of the same coin? Bioessays. 2008;30:719–727. doi: 10.1002/bies.20784. [DOI] [PubMed] [Google Scholar]
  • 129.Baker M, Mackenzie IR, Pickering-Brown SM, Gass J, Rademakers R, Lindholm C, et al. Mutations in progranulin cause tau-negative frontotemporal dementia linked to chromosome 17. Nature. 2006;442:916–919. doi: 10.1038/nature05016. [DOI] [PubMed] [Google Scholar]
  • 130.Mukherjee O, Pastor P, Cairns NJ, Chakraverty S, Kauwe JS, Shears S, et al. HDDD2 is a familial frontotemporal lobar degeneration with ubiquitin-positive, tau-negative inclusions caused by a missense mutation in the signal peptide of progranulin. Ann Neurol. 2006;60:314–322. doi: 10.1002/ana.20963. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Behrens MI, Mukherjee O, Tu PH, Liscic RM, Grinberg LT, Carter D, et al. Neuropathologic heterogeneity in HDDD1: a familial frontotemporal lobar degeneration with ubiquitin-positive inclusions and progranulin mutation. Alzheimer Dis Assoc Disord. 2007;21:1–7. doi: 10.1097/WAD.0b013e31803083f2. [DOI] [PubMed] [Google Scholar]
  • 132.Skoglund L, Brundin R, Olofsson T, Kalimo H, Ingvast S, Blom ES, et al. Frontotemporal dementia in a large Swedish family is caused by a progranulin null mutation. Neurogenetics. 2008 Oct 15; doi: 10.1007/s10048-008-0155-z. [DOI] [PubMed] [Google Scholar]
  • 133.Ong CH, Bateman A. Progranulin (granulin-epithelin precursor, PC-cell derived growth factor, acrogranin) in proliferation and tumorigenesis. Histol Histopathol. 2003;18:1275–1288. doi: 10.14670/HH-18.1275. [DOI] [PubMed] [Google Scholar]

RESOURCES