• Microbial-generated amyloids and Alzheimer's disease (AD) (1/2)

    From =?UTF-8?B?4oqZ77y/4oqZ?=@21:1/5 to All on Wed Sep 2 13:32:24 2015
    NCBI
    PMC

    http://www.ncbi.nlm.nih.gov/pmc/articles/PMC4322713/


    US National Library of Medicine
    National Institutes of Health
    Search databaseSearch termSearch
    Limits Advanced Journal list Help
    Journal ListFront Aging Neurosciv.7; 2015PMC4322713
    Logo of frontagineuro
    Front Aging Neurosci. 2015; 7: 9.
    Published online 2015 Feb 10. doi: 10.3389/fnagi.2015.00009
    PMCID: PMC4322713


    Microbial-generated amyloids and Alzheimer's disease (AD)



    James M. Hill1,2,3,4 and Walter J. Lukiw1,2,5,*
    Author information ► Article notes ► Copyright and License information ► This article has been cited by other articles in PMC.
    Go to:
    Introduction
    Atypical amyloid generation, folding, aggregation and impaired clearance are characteristic pathological features of human neurodegenerative disorders including Alzheimer's disease (AD). What is generally not appreciated is that a major secretory product
    of microbes is amyloid, and that the contribution of microbial amyloid to the pathophysiology of the human central nervous system (CNS) is potentially substantial. While earlier findings suggested that these amyloids may serve some immune-evasive
    strategy, it has recently become evident that humans have a tremendously heavy systemic burden of amyloid which may contribute to the pathology of progressive neurological diseases with an amyloidogenic component. This perspective will highlight some
    recent inroads made into our understanding of the enigmatic role that microbial amyloids may play in the homeostasis and pathology of the CNS with particular reference to AD wherever possible.

    Go to:
    Amyloid: microbial and CNS sources
    “Amyloid” is a generic term for any aggregated, insoluble, lipoprotein-rich deposit exhibiting β-pleated sheet structures oriented perpendicular to the fibrillar axis (Steensma, 2001; Badtke et al., 2009; Blanco et al., 2012; Buxbaum and Linke, 2012)
    . Amyloids are characterized by an apple-green birefringence (λmax ~555 nm) when stained with the secondary diazo-dye Congo-red (λmax ~498 nm) when viewed under polarized light (upon binding to amyloid, Congo Red displays bright fluorescence emission
    at λmax ~614 nm after excitation at λmax ~497 nm; Alexandrov et al., 2011; O'Brien and Wong, 2011). Amyloid fibrillation is initiated by self-aggregation of protein monomers-into-dimers, oligomers and fibrils, which accumulate over time, and this
    process is thought to result from the hydrophobic nature of the aromatic amino-acid peptides comprising the primary sequence of the amyloid (O'Brien and Wong, 2011; Lukiw, 2012). The Congo red dye-based intercalation of β-pleated sheets, induction of a
    positive anisotropy that is polarized and directionally dependent, and generation of a measureable wavelength shift and apple-green birefringence is the hallmark of all amyloids and is the “gold standard” in the diagnosis of amyloidogenic disease (
    Linke, 2006; Buxbaum and Linke, 2012). The polymerization of amyloidogenic proteins is cooperative, and can be accelerated by amyloid aggregates derived from the same protein in a selective “seeding” process. The identification of the “amylome,”
    a classification of amino acid sequences within proteins with internal, self-complementary interfaces and high fiber-forming propensity has improved our understanding of the capability of different proteins to form amyloids that contribute to “dense-
    deposit” disease (Goldschmidt et al., 2010; O'Brien and Wong, 2011; Lukiw, 2012). The pathogenesis of diseases that accumulate amyloid, including AD, all involve a marked inflammatory response at sites of amyloid deposition, and this is mediated by
    microglial cells, the “roving macrophages” of the CNS. Microglia appear to utilize molecular sensors on their external surface, such as the Toll-like receptor 2, TLR2, to recognize abnormal forms of amyloid and initiate a phagocytic or “clearance”
    response (Zhao et al., 2013; Ferrera et al., 2014; Jones et al., 2014). Here we describe a relatively recent collection of stimulating research at the crossroads of microbial and AD amyloids highlighting 5 recent, specific and illustrative insights into
    the potential contribution of microbial-derived amyloids to CNS amyloidogenesis and AD.

    Microbiome-derived amyloids

    The microbiome is the aggregate of all microorganisms that reside on and within the human body, forming a complex ecosystem that includes the skin, oral and nasal mucosa, the urogenital and gastrointestinal (GI) tracts. The microbiome of the GI tract is
    by far the largest reservoir of microbes in the human body, containing about 1014 microbes; over 99% of microbiota in the GI tract are anaerobic bacteria, with fungi, protozoa, archaebacteria and other microorganisms making up the remainder (
    Bhattacharjee and Lukiw, 2013; Hill et al., 2014; Lin et al., 2014). Prokaryotic cells of the human microbiome outnumber human eukaryotic host cells by about 100 to 1, and collectively microbial genes outnumber host genes by about 150 to 1 (Hill et al.,
    2014; Lin et al., 2014). Recent microbiome analysis has revealed that the majority (98%) of all GI tract species belongs to only 4 major bacterial divisions: the gram-positive Firmicutes (64%) and Actinobacteria (3%) and the gram-negative Bacteroidetes (
    23%) and Proteobacteria (8%). The remaining 2% consist of minor taxonomic divisions which are quite diverse (Hattori and Taylor, 2009; Bhattacharjee and Lukiw, 2013; Schwartz and Boles, 2013). Many different microbiome species including fungi and
    bacteria secrete amyloid (Hill et al., 2014; Syed and Boles, 2014). For example, amyloids are associated with fungal surface-structures and the recent observation of amyloidogenic fungal proteins and diffuse mycoses in the blood of AD patients suggest
    that chronic fungal infection increases AD risk (Alonso et al., 2014a,b; Hill et al., 2014). To cite one other recent example, in Escherichia coli extracellular amyloids known as curli fibers, composed of the major structural subunit CsgA, are a common
    secretory component that facilitate surface attachment and adhesion, biofilm development and protection against host defenses (Schwartz and Boles, 2013; Asti and Gioglio, 2014). Biofilms represent a matrix of extracellular polymeric amyloids and other
    lipoproteins in various structural forms. Interestingly, the extracellular 17.7 kDa CsgA amyloid precursor contains a pathogen-associated molecular pattern (PAMP) that, like the Aβ42 peptide, is recognized by the human immune system TLR2 (Zhou et al.,
    2012). An expanding list of bacterial amyloid systems include those associated with gram-negative species of Streptomyces, Bacillus, Pseudomonas, Staphylococcus and others, suggesting that functional amyloids are a widespread phenomenon utilized by a
    wide diversity of microbiome-bacteria (Schwartz and Boles, 2013; Asti and Gioglio, 2014; Hill et al., 2014). Indeed the extremely large number and variety of microbiome bacteria and their capability to produce vast quantities of amyloids indicates that
    human physiology may be potentially exposed to a tremendous systemic amyloid burden, especially during aging when the GI tract epithelial and blood-brain barriers become significantly more restructured and permeable (Bhattacharjee and Lukiw, 2013;
    Marques et al., 2013; Oakley and Tharakan, 2014).

    The amyloid peptides of AD and endotoxin-mediated inflammation

    A ~770 amino acid type 3 transmembrane β-amyloid precursor protein (βAPP), through interaction with membrane proteins and tandem secretase cleavage yields a series of ragged Aβpeptide monomers, two of the most abundant being the 40 and 42 amino acid
    peptides Aβ40 and Aβ42 (Van Broeck et al., 2007; O'Brien and Wong, 2011; Zhang et al., 2012). Aβ40 peptides associate with the endothelium of the cerebral vasculature, and the more neurotoxic, albeit less abundant, hydrophobic Aβ42 species form the
    nuclei of the senile plaque (SP) lesions of AD (Alexandrov et al., 2011; O'Brien and Wong, 2011). Interestingly, the extra two hydrophobic amino acids in Aβ42 (vs. Aβ40) appear to convey many of the toxic biophysical attributes and self-aggregation of
    this slightly larger molecule (Zhou et al., 2011; Teng et al., 2012; Zhang et al., 2012). The recognition of Aβ42 peptides and their misfolded aggregates by microglial-surveillance systems, and the inability of microglial cells to deal with these toxic,
    pro-inflammatory inclusions are thought to form the molecular basis for the elevated oxidative stress, aberrant immune-activation and chronic inflammation characteristic of AD brain (Armstrong, 2006; Cui et al., 2007; Van Broeck et al., 2007;
    Boutajangout and Wisniewski, 2013; Furukawa and Nukina, 2013; Ferrera et al., 2014; Serpente et al., 2014; Takeda et al., 2014). Interestingly (i) βAPP-derived Aβ42 peptides induce a pattern of expression of inflammatory genes typical of the classical
    immune- and inflammatory-response induced by infectious agents such as bacterial lipopolysaccharide (LPS), a common endotoxin of the outer membrane of gram-negative bacteria (Colangelo et al., 2002; Ferrera et al., 2014), and (ii) the presence of
    bacteria, LPS or endotoxin-mediated inflammation strongly contributes to amyloid neurotoxicity (Hammer et al., 2008; Dasari et al., 2011; Blanco et al., 2012; Zhou et al., 2012; Serpente et al., 2014).

    Congo red staining to identify amyloid deposits

    Congo red [3,3′-([1,1′-biphenyl]-4,4′-diyl)bis(4-aminonaphthalene-1-sulfonic-acid-disodium salt)], a toxic, water soluble secondary diazo dye, was first synthesized in 1883 and used as a textile dye (Steensma, 2001; Linke, 2006). Due to its
    toxicity, Congo red's use in textile dying was discontinued, but because of its important spectroscopic properties gained wide use in investigative microbiology (Linke, 2006). The proposed Congo red-mediated staining mechanism suggests a substrate-
    mediated hydrophobic pi-pi orbital stacking interaction between the aromatic rings of the dye molecules and β-pleated sheet structures of both amyloids and textiles (Buxbaum and Linke, 2012). Congo red was classically used in microbiological
    epidemiology as a bacterial-stain, for example, to rapidly identify the presence of virulent forms of gram-negative Shigella where the dye binds the bacterium's unique LPS surface structures (Linke, 2006; Buxbaum and Linke, 2012). Congo red's apple-green
    birefringent fluorescence under polarized light when bound to amyloid fibrils is currently used as a sensitive diagnostic tool for amyloidosis including the Aβ42-enriched SPs of AD (Steensma, 2001; Linke, 2006; Buxbaum and Linke, 2012). Interestingly it
    has very recently been found that (i) LPS is capable of inducing a pathogenic Congo red-sensitive β-pleated sheet conformation in prion amyloids (Saleem et al., 2014); and (ii) the infectious microbial burden is significantly associated with both AD
    development and the propensity of AD amyloids to be stained by Congo red (Bu et al., 2014).

    Molecular mimicry between mitochondria and bacteria

    Because mitochondria appear to have originated from bacteria via endosymbiotic relationships that formed very early in the evolutionary history of eukaryotes, cross-reactivity of mitochondria and immunological responses to bacterial amyloid or LPS may
    have deleterious effects on mitochondrial function. This “molecular mimicry” is partially evidenced by extragastric diseases such as the basal ganglia disorder Sydenham's chorea, rheumatic fever, low-grade systemic-inflammatory states and the link to
    the Firmicute Streptococcus and/or the gram-negative microaerophilic Proteobacteria Helicobacter pylori (Douglas-Escobar et al., 2013; Hayashi, 2013; Hornig, 2013; Roubaud Baudron et al., 2013; Hill et al., 2014). Previous bacterial infection resulting
    in antibody formation to amyloids or bacterial endotoxins may predispose CNS mitochondria or amyloids to subsequent attack by antibodies resulting in the up-regulation of inflammatory signaling (Bu et al., 2014).

    Activation of TLR2 by amyloids

    TLRs are type I membrane-spanning protein receptors expressed in microglial cells. TLRs play key roles in host protection from microbial-invasion via the activation of the innate-immune system by sensing structurally conserved PAMPs from microbes that
    are distinguishable from, and not innate to, the host organism (Tükel et al., 2009; Harry, 2013; Yu and Ye, 2014). Of the 13 currently identified TLRs (TLR1 to TLR13) the microglial TLR2s are activated by amyloid, lipoproteins and other microbial
    triggers that subsequently induce cytokine production, inflammation, phagocytosis and innate immune defense responses that directly impact CNS homoeostasis and drive neuropathology. More specifically the TLR2/TLR1 complex can recognize biofilm-associated
    amyloids produced by Firmicutes, Bacteroidetes, and Proteobacteria (Nishimori et al., 2012; Bhattacharjee and Lukiw, 2013; Asti and Gioglio, 2014). Remarkably, the Aβ42 peptides of AD that are associated with robust microglia-mediated inflammatory
    responses also activate TLR2 (Gustot et al., 2013; Yu and Ye, 2014). Of further interest is (i) that microbial amyloids induce pro-inflammatory interleukin IL-17A, a driver of NF-kB signaling and cyclooxygenase-2 activation, and other potent mediators of
    inflammatory-responses such as IL-22 via direct TLR2 activation (Nishimori et al., 2012); and (ii) that increased levels of IL-17A and IL-22 are associated with chronic inflammatory diseases including AD (Zhang et al., 2013).

    Go to:
    Concluding remarks
    Diverse microbes of the human microbiome generate functional amyloids. Their ability to bind Congo red has provided useful tools for characterizing both microbial- and CNS-derived amyloids (Zhou et al., 2012; Serpente et al., 2014). The large amount of
    microbial-generated GI amyloid implicates high potential systemic exposure to bacterial amyloid, and the bioavailability of amyloid to the CNS increases as humans age (Bhattacharjee and Lukiw, 2013; Marques et al., 2013; Tran and Greenwood-Van Meerveld,
    2013; Oakley and Tharakan, 2014). Microbial and CNS amyloids are biologically similar in their structure and immunogenicity and complex mechanistic interrelationships between these amyloids are beginning to emerge. It is remarkable (i) that human
    microbes that produce amyloids such as CsgA and curli, and the Aβ42 peptides that accumulate in AD, are recognized by the same TLR2/TLR1 immune sensor-receptor system of the 13 different TLR-type receptors available; and (ii) that they direct the same
    up-regulation of IL-17A- and IL-22-mediated pro-inflammatory-signaling (Rapsinski et al., 2013; Zhang et al., 2013; Yu and Ye, 2014). Interestingly, CsgA and Aβ42 peptides do not share common amino-acid sequences, only structural similarity in their
    PAMPs. Microbes or their secretory or degradation products including their amyloids and LPSs are powerful inflammatory activators and inducers of cytokines and complement proteins, affecting vascular permeability and generating free-radicals that further
    support amyloidogenesis (Hill et al., 2014; Lin et al., 2014). These pathogenic signaling features are also highly characteristic of AD neuropathology. A more detailed understanding of human microbial ecosystems and their amyloids should give insight
    into amyloid-misfolding and their contribution to inflammatory-signaling in health, aging and disease. It will certainly be interesting to see: (i) if any microbial-generated amyloids co-localize with the amyloid-dense SP deposits of AD; (ii) if GI tract
    microbiome-derived amyloids become more available systemically as humans age; and (iii) what the evolution and nature of amyloid-related communication between the GI-tract and the CNS has on the development or propagation of amyloidogenesis in pro-
    inflammatory degenerative disease.

    Conflict of interest statement

    The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

    Go to:
    Acknowledgments
    This work was presented in part at the Alzheimer Association International Conference 2013 (AAIC 2013) Annual Meeting held in Boston MA, USA and at the AAIC 2014 held in Copenhagen, Denmark. Sincere thanks are extended to Drs. PN Alexandrov, F Culicchia,
    C Eicken and C Hebel for short post-mortem interval (PMI) human brain tissues or extracts, miRNA array work and initial data interpretation, and to AI Pogue, D Guillot and J Lockwood for expert technical assistance. Additional thanks are extended to the
    many physicians and neuropathologists of Canada and the USA who have provided high quality, short post-mortem interval human brain and retinal tissues for study. Additional human control and AD brain tissues were provided by the Memory Impairments and
    Neurological Disorders (MIND) Institute and the University of California, Irvine Alzheimer's Disease Research Center (UCI-ADRC; NIA P50 AG16573). Research on miRNA in the Lukiw laboratory involving the innate-immune response in AD, amyloidogenesis and
    neuro-inflammation was supported through a COBRE III Pilot Project NIH/NIGMS Grant P30-GM103340, an unrestricted grant to the LSU Eye Center from Research to Prevent Blindness (RPB); the Louisiana Biotechnology Research Network (LBRN) and NIH grants NEI
    EY006311, NIA AG18031 and NIA AG038834.

    Go to:
    References
    Alexandrov P. N., Pogue A., Bhattacharjee S., Lukiw W. J. (2011). Retinal amyloid peptides and complement factor H in transgenic models of Alzheimer's disease. Neuroreport 22, 623–627. 10.1097/WNR.0b013e3283497334 [PMC free article] [PubMed] [Cross Ref]
    Alonso R., Pisa D., Marina A. I., Morato E., Rábano A., Carrasco L. (2014b). Fungal infection in patients with Alzheimer's disease. J. Alzheimers Dis. 41, 301–311. 10.3233/JAD-132681 [PubMed] [Cross Ref]
    Alonso R., Pisa D., Rábano A., Carrasco L. (2014a). Alzheimer's disease and disseminated mycoses. Eur. J. Clin. Microbiol. Infect. Dis. 33, 1125–1132. 10.1007/s10096-013-2045-z [PubMed] [Cross Ref]
    Armstrong R. A. (2006). Plaques and tangles and the pathogenesis of Alzheimer's disease. Folia Neuropathol. 44, 1–11. [PubMed]
    Asti A., Gioglio L. (2014). Can a bacterial endotoxin be a key factor in the kinetics of amyloid fibril formation? J. Alzheimers. Dis. 39, 169–179. 10.3233/JAD-131394 [PubMed] [Cross Ref]
    Badtke M. P., Hammer N. D., Chapman M. R. (2009). Functional amyloids signal their arrival. Sci. Signal. 2:pe43. 10.1126/scisignal.280pe43 [PMC free article] [PubMed] [Cross Ref]
    Bhattacharjee S., Lukiw W. J. (2013). Alzheimer's disease and the microbiome. Front. Cell. Neurosci. 7:153. 10.3389/fncel.2013.00153 [PMC free article] [PubMed] [Cross Ref]
    Blanco L. P., Evans M. L., Smith D. R., Badtke M. P., Chapman M. R. (2012). Diversity, biogenesis and function of microbial amyloids. Trends Microbiol. 20, 66–73. 10.1016/j.tim.2011.11.005 [PMC free article] [PubMed] [Cross Ref]
    Boutajangout A., Wisniewski T. (2013). The innate immune system in Alzheimer's disease. Int. J. Cell Biol. 2013, 576383. 10.1155/2013/576383 [PMC free article] [PubMed] [Cross Ref]
    Bu X. L., Yao X. Q., Jiao S. S., Zeng F., Liu Y. H., Xiang Y., et al. . (2014). A study on the association between infectious burden and Alzheimer's disease. Eur. J. Neurol. [Epub ahead of print]. 10.1111/ene.12477 [PubMed] [Cross Ref]
    Buxbaum J. N., Linke R. P. (2012). A molecular history of the amyloidoses. J. Mol. Biol. 421, 142–159. 10.1016/j.jmb.2012.01.024 [PubMed] [Cross Ref]
    Colangelo V., Schurr J., Ball M. J., Pelaez R. P., Bazan N. G., Lukiw W. J. (2002). Gene expression profiling of 12633 genes in Alzheimer hippocampal CA1: transcription and neurotrophic factor down-regulation and up-regulation of apoptotic and pro-
    inflammatory signaling. J. Neurosci. Res. 70, 462–473. 10.1002/jnr.10351 [PubMed] [Cross Ref]
    Cui J. G., Hill J. M., Zhao Y., Lukiw W. J. (2007). Expression of inflammatory genes in the primary visual cortex of late-stage Alzheimer's disease. Neuroreport 18, 115–119. 10.1097/WNR.0b013e32801198bc [PubMed] [Cross Ref]
    Dasari M., Espargaro A., Sabate R., Lopez del Amo J. M., Fink U., Grelle G., et al. . (2011). Bacterial inclusion bodies of Alzheimer's disease β-amyloid peptides can be employed to study native-like aggregation intermediate states. Chembiochem 12, 407
    423. 10.1002/cbic.201000602 [PubMed] [Cross Ref]
    Douglas-Escobar M., Elliott E., Neu J. (2013). Effect of intestinal microbial ecology on the developing brain. JAMA Pediatr. 167, 374–379. 10.1001/jamapediatrics.2013.497 [PubMed] [Cross Ref]
    Ferrera D., Mazzaro N., Canale C., Gasparini L. (2014). Resting microglia react to Aβ42 fibrils but do not detect oligomers or oligomer-induced neuronal damage. Neurobiol. Aging. 35, 2444–2457. 10.1016/j.neurobiolaging.2014.05.023 [PubMed] [Cross Ref]
    Furukawa Y., Nukina N. (2013). Functional diversity of protein fibrillar aggregates from physiology to RNA granules to neurodegenerative diseases. Biochim. Biophys. Acta. 1832, 1271–1278. 10.1016/j.bbadis.2013.04.011 [PubMed] [Cross Ref]
    Goldschmidt L., Teng P. K., Riek R., Eisenberg D. (2010). Identifying the amylome, proteins capable of forming amyloid-like fibrils. Proc. Natl. Acad. Sci. U.S.A. 107, 3487–3492. 10.1073/pnas.0915166107 [PMC free article] [PubMed] [Cross Ref]
    Gustot A., Raussens V., Dehousse M., Dumoulin M., Bryant C. E., Ruysschaert J. M., et al. . (2013). Activation of innate immunity by lysozyme fibrils is critically dependent on cross-β sheet structure. Cell. Mol. Life Sci. 70, 2999–3012. 10.1007/
    s00018-012-1245-5 [PubMed] [Cross Ref]
    Hammer N. D., Wang X., McGuffie B. A., Chapman M. R. (2008). Amyloids: friend or foe? J. Alzheimers Dis. 13, 407–419. [PMC free article] [PubMed]
    Harry G. J. (2013). Microglia during development and aging. Pharmacol. Ther. 139, 313–326 10.1016/j.pharmthera.2013.04.013 [PMC free article] [PubMed] [Cross Ref]
    Hattori M., Taylor T. D. (2009). The human intestinal microbiome: a new frontier of human biology. DNA Res. 16, 1–12. 10.1093/dnares/dsn033 [PMC free article] [PubMed] [Cross Ref]
    Hayashi M. (2013). [Anti-basal ganglia antibody]. Brain Nerve 65, 377–384. [PubMed]
    Hill J. M., Clement C., Pogue A. I., Bhattacharjee S., Zhao Y., Lukiw W. J. (2014). Pathogenic microbes, the microbiome, and Alzheimer's disease (AD). Front. Aging Neurosci. 6:127. 10.3389/fnagi.2014.00127 [PMC free article] [PubMed] [Cross Ref]
    Hornig M. (2013). The role of microbes and autoimmunity in the pathogenesis of neuropsychiatric illness. Curr. Opin. Rheumatol. 25, 488–795. 10.1097/BOR.0b013e32836208de [PubMed] [Cross Ref]
    Jones B. M., Bhattacharjee S., Dua P., Hill J. M., Zhao Y., Lukiw W. J. (2014). Regulating amyloidogenesis through the natural triggering receptor expressed in myeloid/microglialcells 2 (TREM2). Front. Cell. Neurosci. 8:94. 10.3389/fncel.2014.00094 [PMC
    free article] [PubMed] [Cross Ref]
    Lin C. S., Chang C. J., Lu C. C., Martel J., Ojcius D. M., Ko Y. F., et al. . (2014). Impact of the gut microbiota, prebiotics, and probiotics on human health and disease. Biomed. J. 37, 259–268. 10.4103/2319-4170.138314 [PubMed] [Cross Ref]
    Linke R. P. (2006). Congo red staining of amyloid; improvements and practical guide for a more precise diagnosis of amyloid and the different amyloidoses. Chapter 11.1, in Protein Misfolding, Aggregation, and Conformational Diseases Part A: Protein
    Aggregation and Conformational Diseases, eds Uversky V. N., Fink A. L., editors. (New York, NY: Springer; ), 239–276 10.1007/0-387-25919-8_12 [Cross Ref]
    Lukiw W. J. (2012). Amyloid beta (Aβ) peptide modulators and other current treatment strategies for Alzheimer's disease (AD). Expert Opin. Emerg. Drugs 17, 43–60. 10.1517/14728214.2012.672559 [PMC free article] [PubMed] [Cross Ref]
    Marques F., Sousa J. C., Sousa N., Palha J. A. (2013). Blood-brain-barriers in aging and in Alzheimer's disease. Mol. Neurodegener. 8, 38. 10.1186/1750-1326-8-38 [PMC free article] [PubMed] [Cross Ref]
    Nishimori J. H., Newman T. N., Oppong G. O., Rapsinski G. J., Yen J. H., Biesecker S. G., et al. . (2012). Microbial amyloids induce interleukin 17A (IL-17A) and IL-22 responses via Toll-like receptor 2 (TLR2) activation in the intestinal mucosa. Infect.
    Immun. 80, 4398–4408. 10.1128/IAI.00911-12 [PMC free article] [PubMed] [Cross Ref]
    Oakley R., Tharakan B. (2014). Vascular hyperpermeability and aging. Aging Dis. 5, 114–125. 10.14336/AD.2014.0500114 [PMC free article] [PubMed] [Cross Ref]
    O'Brien R. J., Wong P. C. (2011). Amyloid precursor protein processing and Alzheimer's disease. Annu. Rev. Neurosci. 34, 185–204. 10.1146/annurev-neuro-061010-113613 [PMC free article] [PubMed] [Cross Ref]
    Rapsinski G. J., Newman T. N., Oppong G. O., van Putten J. P., Tükel Ç. (2013). CD14 protein acts as an adaptor molecule for the immune recognition of Salmonella curli fibers. J. Biol. Chem. 288, 14178–14188. 10.1074/jbc.M112.447060 [PMC free article]
    [PubMed] [Cross Ref]
    Roubaud Baudron C., Franceschi F., Salles N., Gasbarrini A. (2013). Extragastric diseases and Helicobacter pylori. Helicobacter. 18(Suppl. 1), 44–51. 10.1111/hel.12077 [PubMed] [Cross Ref]
    Saleem F., Bjorndahl T. C., Ladner C. L., Perez-Pineiro R., Ametaj B. N., Wishart D. S. (2014). Lipopolysaccharide induced conversion of recombinant prion protein. Prion 8, 221–232. 10.4161/pri.28939 [PMC free article] [PubMed] [Cross Ref]
    Schwartz K., Boles B. R. (2013). Microbial amyloids - functions and interactions within the host. Curr. Opin. Microbiol. 16, 93–99. 10.1016/j.mib.2012.12.001 [PMC free article] [PubMed] [Cross Ref]
    Serpente M., Bonsi R., Scarpini E., Galimberti D. (2014). Innate immune system and inflammation in Alzheimer's disease: from pathogenesis to treatment. Neuroimmunomodulation 21, 79–87. 10.1159/000356529 [PubMed] [Cross Ref]
    Steensma D. P. (2001). “Congo” red: out of Africa? Arch. Pathol. Lab. Med. 125, 250–252. [PubMed]
    Syed A. K., Boles B. R. (2014). Fold modulating function: bacterial toxins to functional amyloids. Front. Microbiol. 5:401. 10.3389/fmicb.2014.00401 [PMC free article] [PubMed] [Cross Ref]
    Takeda S., Sato N., Morishita R. (2014). Systemic inflammation, blood-brain barrier vulnerability and cognitive/non-cognitive symptoms in Alzheimer disease: relevance to pathogenesis and therapy. Front. Aging Neurosci. 6:171. 10.3389/fnagi.2014.00171 [
    PMC free article] [PubMed] [Cross Ref]
    Teng P. K., Anderson N. J., Goldschmidt L., Sawaya M. R., Sambashivan S., Eisenberg D. (2012). Ribonuclease A suggests how proteins self-chaperone against amyloid fiber formation. Protein Sci. 21, 26–37. 10.1002/pro.754 [PMC free article] [PubMed] [
    Cross Ref]
    Tran L., Greenwood-Van Meerveld B. (2013). Age-associated remodeling of the intestinal epithelial barrier. J. Gerontol. ABiol. Sci. Med. Sci. 68, 1045–1056. 10.1093/gerona/glt106 [PMC free article] [PubMed] [Cross Ref]
    Tükel C., Wilson R. P., Nishimori J. H., Pezeshki M., Chromy B. A., Bäumler A. J. (2009). Responses to amyloids of microbial and host origin are mediated through toll-like receptor 2. Cell Host Microbe. 6, 45–53. 10.1016/j.chom.2009.05.020 [PMC free
    article] [PubMed] [Cross Ref]
    Van Broeck B., Van Broeckhoven C., Kumar-Singh S. (2007). Current insights into molecular mechanisms of Alzheimer disease and their implications for therapeutic approaches. Neurodegener. Dis. 4, 349–365. 10.1159/000105156 [PubMed] [Cross Ref]
    Yu Y., Ye R. D. (2014). Microglial Aβreceptors in Alzheimer's disease. Cell. Mol. Neurobiol. 35, 71–83. 10.1007/s10571-014-0101-6 [PubMed] [Cross Ref]
    Zhang H., Ma Q., Zhang Y. W., Xu H. (2012). Proteolytic processing of Alzheimer's β-amyloid precursor protein. J. Neurochem. 120, 9–21. 10.1111/j.1471-4159.2011.07519.x [PMC free article] [PubMed] [Cross Ref]
    Zhang J., Ke K. F., Liu Z., Qiu Y. H., Peng Y. P. (2013). Th17 cell-mediated neuroinflammation is involved in neurodegeneration of Aβ42-induced Alzheimer's disease model rats. PLoS ONE 8:e75786. 10.1371/journal.pone.0075786 [PMC free article] [PubMed] [
    Cross Ref]
    Zhao Y., Bhattacharjee S., Jones B. M., Dua P., Alexandrov P. N., Hill J. M., et al. . (2013). Regulation of TREM2 expression by an NF-κB-sensitive miRNA-34a. Neuroreport 24, 318–323. 10.1097/WNR.0b013e32835fb6b0 [PMC free article] [PubMed] [Cross Ref]
    Zhou Y., Blanco L. P., Smith D. R., Chapman M. R. (2012). Bacterial amyloids. Methods Mol. Biol. 849, 303–320. 10.1007/978-1-61779-551-0_21 [PubMed] [Cross Ref]
    Zhou Z. D., Chan C. H., Ma Q. H., Xu X. H., Xiao Z. C., Tan E. K. (2011). The roles of amyloid precursor protein in neurogenesis: implications to pathogenesis and therapy of Alzheimer disease. Cell Adh. Migr. 5, 280–292. 10.4161/cam.5.4.16986 [PMC free
    article] [PubMed] [Cross Ref]
    Articles from Frontiers in Aging Neuroscience are provided here courtesy of Frontiers Media SA
    PubReader format: click here to try
    Formats:

    Article | PubReader | ePub (beta) | PDF (377K) | Citation
    Share

    Share on Facebook Facebook Share on Twitter Twitter Share on Google Plus Google+
    Save items
    Add to Favorites
    View more options
    Similar articles in PubMed
    Microbial Sources of Amyloid and Relevance to Amyloidogenesis and Alzheimer's Disease (AD).
    [J Alzheimers Dis Parkinsonism....]
    Microbiome-generated amyloid and potential impact on amyloidogenesis in Alzheimer's disease (AD).
    [J Nat Sci. 2015]
    Toll-like receptor 2 acts as a natural innate immune receptor to clear amyloid beta 1-42 and delay the cognitive decline in a mouse model of Alzheimer's disease.
    [J Neurosci. 2008]
    The innate immune system in Alzheimer's disease.
    [Int J Cell Biol. 2013]
    Inflammation in Alzheimer's disease: amyloid-beta oligomers trigger innate immunity defence via pattern recognition receptors.
    [Prog Neurobiol. 2009]
    See reviews...
    See all...
    Cited by other articles in PMC
    Microbiome-generated amyloid and potential impact on amyloidogenesis in Alzheimer’s disease (AD)
    [Journal of nature and science....]
    See all...
    Links
    PubMed
    Taxonomy
    Recent Activity
    ClearTurn Off
    Microbial-generated amyloids and Alzheimer's disease (AD)
    Microbial-generated amyloids and Alzheimer's disease (AD)
    Frontiers in Aging Neuroscience. 2015; 7()
    Alzheimer's disease and the microbiome
    Alzheimer's disease and the microbiome
    Frontiers in Cellular Neuroscience. 2013; 7()
    Recycling Metchnikoff: Probiotics, the Intestinal Microbiome and the Quest for L...
    Recycling Metchnikoff: Probiotics, the Intestinal Microbiome and the Quest for Long Life
    Frontiers in Public Health. 2013; 1()
    Microorganisms with Claimed Probiotic Properties: An Overview of Recent Literatu...
    Microorganisms with Claimed Probiotic Properties: An Overview of Recent Literature
    International Journal of Environmental Research and Public Health. 2014 May; 11(5)4745
    The microbiota and microbiome in aging: potential implications in health and age...
    The microbiota and microbiome in aging: potential implications in health and age-related diseases.
    J Am Geriatr Soc. 2015 Apr ;63(4):776-81. doi: 10.1111/jgs.13310. Epub 2015 Apr 8 .
    PubMed
    See more...
    "Congo" red: out of Africa?
    [Arch Pathol Lab Med. 2001]
    Review Functional amyloids signal their arrival.
    [Sci Signal. 2009]
    Review Diversity, biogenesis and function of microbial amyloids.
    [Trends Microbiol. 2012]
    Review A molecular history of the amyloidoses.
    [J Mol Biol. 2012]
    Retinal amyloid peptides and complement factor H in transgenic models of Alzheimer's disease.
    [Neuroreport. 2011]
    Review Amyloid precursor protein processing and Alzheimer's disease.
    [Annu Rev Neurosci. 2011]
    Amyloid beta (Aβ) peptide modulators and other current treatment strategies for Alzheimer's disease (AD).
    [Expert Opin Emerg Drugs. 2012]
    Identifying the amylome, proteins capable of forming amyloid-like fibrils. [Proc Natl Acad Sci U S A. 2010]

    [continued in next message]

    --- SoupGate-Win32 v1.05
    * Origin: fsxNet Usenet Gateway (21:1/5)