Back to Journals » Journal of Inflammation Research » Volume 9

Targeting neuroinflammation in Alzheimer’s disease

Authors Bronzuoli MR, Iacomino A, Steardo L, Scuderi C

Received 28 July 2016

Accepted for publication 1 October 2016

Published 3 November 2016 Volume 2016:9 Pages 199—208

DOI https://doi.org/10.2147/JIR.S86958

Checked for plagiarism Yes

Review by Single anonymous peer review

Peer reviewer comments 3

Editor who approved publication: Professor Ning Quan



Maria Rosanna Bronzuoli,1 Aniello Iacomino,2 Luca Steardo,1 Caterina Scuderi1

1Department of Physiology and Pharmacology “Vittorio Erspamer”, Sapienza University of Rome, Rome, Italy; 2Faculty of Psychology, University of Rome “G. Marconi”, Rome, Italy

Abstract: Almost 47 million people suffer from dementia worldwide, with an estimated new case diagnosed every 3.2 seconds. Alzheimer’s disease (AD) accounts for approximately 60%–80% of all dementia cases. Given this evidence, it is clear dementia represents one of the greatest global public health challenges. Currently used drugs alleviate the symptoms of AD but do not treat the underlying causes of dementia. Hence, a worldwide quest is under way to find new treatments to stop, slow, or even prevent AD. Besides the classic targets of the oldest therapies, represented by cholinergic and glutamatergic systems, β-amyloid (Aβ) plaques, and tau tangles, new therapeutic approaches have other targets. One of the newest and most promising strategies is the control of reactive gliosis, a multicellular response to brain injury. This phenomenon occurs as a consequence of a persistent glial activation, which leads to cellular dysfunctions and neuroinflammation. Reactive gliosis is now considered a key abnormality in the AD brain. It has been demonstrated that reactive astrocytes surround both Aβ plaques and tau tangles. In this condition, glial cells lose some of their homeostatic functions and acquire a proinflammatory phenotype amplifying neuronal damage. So, molecules that are able to restore their physiological functions and control the neuroinflammatory process offer new therapeutic opportunities for this devastating disease. In this review, we describe the role of neuroinflammation in the AD pathogenesis and progression and then provide an overview of the recent research with the aim of developing new therapies to treat this disorder.

Keywords: reactive gliosis, astrocyte, microglia, Alzheimer’s disease

Introduction

Dementia is a chronic condition characterized by a progressive cognitive impairment that leads to functional disability.1 In 2015, it was estimated that approximately 47 million people worldwide were affected by dementia, and this number is expected to increase, reaching 131.5 million by 2050.2 As such, it represents a veritable public health challenge. Alzheimer’s disease (AD), a pathology first described by Alois Alzheimer in 1907,3 is the most frequent cause of dementia in elderly. Knowledge about the etiology and pathogenesis of the disease is continuously updated,4 but there are still limitations in diagnostic capability5 and in the discovery of pharmacological treatments that would be able to stop or better prevent the disease. At present, AD is incurable. Despite the huge amount of preclinical and clinical investigation, medications currently used provide only a modest symptomatic relief to a subset of patients and do not treat the underlying causes of this disease. The reasons for this failure are probably due to the scant knowledge of the cellular and molecular mechanisms implicated in AD pathogenesis and of the approved therapies that coarsely affect both cholinergic and glutamatergic neurotransmission. Conversely, many of the new drugs in development aim to modify the disease process itself by impacting one or more of the many wide-ranging brain changes caused by AD. These changes offer potential targets for new drugs to stop or slow down the disease progression. It is now well recognized that AD is a multifactorial disorder. It is pathologically characterized by widespread oxidative stress, mitochondrial damage, glutamate excitotoxicity, neuroinflammation, neurofibrillary tangle (NFT) formation, and β-amyloid (Aβ) deposition creating senile plaques (SPs).6 These latter are constituted by Aβ peptide, and their genesis is followed by intracellular deposition of NFTs,7 as a consequence of tau protein hyperphosphorylation. The results are synaptic and neuronal dysfunction and loss.8 Over the years, it has been demonstrated that other factors play an important role in the pathogenesis and progression of AD. Among them, the key role of neuroinflammation has been affirmed.9

Physiologically, the inflammatory process is aimed at controlling injuries through several mechanisms to repair tissues.10 However, an increasing amount of literature confirms its role in the pathogenesis and exacerbation of AD.1114 Inflammation acts to remove both the initial cause of the infliction and to eliminate the destroyed tissues and dead cells resulting from the original injury.

In fact, inflammation is emerging as the real cause of the associated disease, more than a mere contribution to the exacerbation of tissue damage. Indeed, some studies have revealed that the injection of lipopolysaccharide in transgenic mice induces neuroinflammation, triggering intracellular Aβ deposit and tau phosphorylation.15,16

The molecular processes are not necessarily the primary events. The inflammatory machine could also be triggered by traumatic or surgical causes. The microglial priming model suggests that the presymptomatic AD pathology, characterized by low levels of proinflammatory mediators, can act on microglia for long periods of time.17 Furthermore, stress, inflammation, and infection can operate as secondary triggers, causing changes in these primed cells: they reach an activated state establishing an inflammatory response contributing to AD pathogenesis.18

From an immunological point of view, the central nervous system was always seen as a highly protected tissue, exposed to inflammatory phenomena solely in cases of infection or disruption of the blood–brain barrier (BBB). Nowadays, we know that there are several cells expressing pattern recognition receptors able to induce inflammatory signaling pathways.13 These pattern recognition receptors can recognize molecular signals of microbial molecules, called pathogen-associated molecular patterns, as well as endogenous damage-associated molecular patterns (DAMPs), that typically accumulate in infected tissues. DAMPS are present in diseased brains as misfolded proteins (eg, SPs and NFTs), aggregated peptides, or nucleic acids.19 It is clear that DAMPs can trigger neuroinflammation by deflecting proinflammatory reactions from their helpful purpose, and this is the reason why our way of viewing neurodegenerative diseases has changed over the years.

The role of the neuroinflammatory process is not exclusively attributable to innate immunity (which in the brain is constituted by microglia), but it is also caused by other brain resident cells that constitute, in one word, macroglia (ie, astrocytes, NG2-positive cells, and oligodendrocytes), as well as endothelial cells and neurons.2023

Hence, it is clear that there are many characters involved in this inflammatory process. Thus, a better knowledge of the mechanisms underlying the role of neuroinflammation in AD can be an excellent starting point for the development of molecules able to counteract it.

The pathophysiology of neuroinflammation and its role in Alzheimer’s disease

Even if Aβ deposits can alone induce an inflammatory response that subsequently leads to AD development, it is well established that the neuroinflammatory pathophysiology is more complex and driven by the activation of different brain cells. In particular, growing evidence suggests that this phenomenon is mainly supported by glial cells, which respond quickly to brain injuries, activating a series of repair mechanisms to restore brain physiology. Glial cells are nonexcitable cells of the central nervous system. These cells are a highly heterogeneous population, responsible for many important brain functions.24 While microglia acts as the first form of immune defense in the brain, astrocytes are an essential neurosupportive cell type. Indeed, astrocytes finely control the environment by regulating pH, ion homeostasis, oxidative stress, and blood flow.25,26 These cells together with microglia, oligodendrocytes, neurons, pericytes, and endothelial cells constitute the neurovascular unit, responsible for the proper functioning of the BBB.27 In addition, astrocytes contribute importantly to synaptogenesis and dynamically modulate information processing and signal transmission, regulate neural and synaptic plasticity, and provide trophic and metabolic support to neurons.28,29 Interestingly, data from animal models and human autopsy revealed that both SPs and NFTs cause an immune response in the brain and colocalize close to activated glial cells. Astrocyte and microglia acquire a reactive phenotype20 and rapidly act in response to pathology undergoing important changes in their morphology and functioning.30,31 Such an activation is fundamentally a protective response aimed at removing injurious stimuli. The neuroprotective action of reactive astrocytes takes place by modulating Aβ-mediated neurotoxicity, degrading, internalizing, and removing Aβ, thus creating a protective barrier that surrounds plaques.3234 However, uncontrolled and prolonged activation goes beyond physiological control, and detrimental effects override the beneficial ones. In this condition, glial cells foster neuroinflammatory response, accounting for the synthesis of different cytokines and proinflammatory mediators.35,36 This condition is called reactive gliosis and is a characteristic event of AD brains (Figure 1). For example, activated microglia reduces Aβ accumulation by increasing its phagocytosis, clearance, and degradation,37 as well as by secreting factors such as the glia-derived neurotrophic factor, helpful for neuronal survival.38 Recently, microglia functions aimed at Aβ clearance were attributed to the presence of triggering receptor expressed on myeloid cells 2 (TREM2), a transmembrane receptor;39,40 indeed not long ago, TREM2 was identified as a risk gene for AD.41,42 It is reasonable to think that this association between TREM2 and AD is due to the many functions carried out through the activation of different pathways ranging from phagocytosis to encouraging survival and proliferation, and finally promoting secretion of cytokines and chemokines.4345

Figure 1 Schematic representation of glial activation.


Notes: As a result of brain damage (eg, brain trauma, ischemia, Aβ accumulation, NFTs, etc) microglia and astrocytes acquire a so-called reactive phenotype losing their physiological functions. Morphofunctional changes, loss of three-dimensional network, and neurovascular unit alterations contribute to cause a homeostatic imbalance. Moreover, after activation, these cells produce a wide range of cytokines and proinflammatory mediators, leading to chronic inflammation. Even if the initial intent of these modifications is reparative, such long-lasting and uncontrolled activation causes further neurodegeneration.


Abbreviations: ROS, reactive oxygen species, NO, nitric oxide; Aβ, β-.amyloid; NFTs, neurofibrillary tangles; BBB, blood–brain barrier.

Even astrocytes play an important role in the maintenance of the cerebral homeostasis. These cells are responsible for the proper functioning of the BBB, provide nutrients to neurons, preserve the extracellular ion balance, and remove and degrade Aβ.46 However, glial functions are deeply altered whenever tissue physiology is not restored. In these circumstances, the inability to counteract Aβ and NFTs accumulation constantly stimulates the machinery needed to remove debris; in this way, astrocytes actively support inflammation.47,48

Several studies demonstrated that their action becomes relevant from early stages of the pathogenic process, turning to a cycle independent from Aβ presence, neural dysfunction, cell death, and disease progression.4951 The resulting chronic inflammation is due to the release of proinflammatory molecules that act not only in an autocrine manner, allowing the perpetuation of the reactive gliosis, but also in a paracrine one, the main cause of the neuronal death that increases the pathological damage.52,53 Neuronal death is determined by the release of not only inflammatory mediators, but also of reactive oxygen species, nitric oxide (NO), proteolytic enzymes, complement factors, and/or excitatory aminoacids.54 At the molecular level, the release of these mediators affects neuron–glia crosstalk, influencing redox enzyme sensors, receptors, and transcription factors.55

In physiological conditions, microglia protects the brain from pathogens, and, together with macroglia, helps maintain homeostasis of the tissue. In AD, all these cells became more reactive and change their morphology surrounding SPs.56 This is possible because of the presence of proinflammatory receptors on their surface. Microglia is able to identify and bind Aβ oligomers and fibrils and the amyloid precursor protein (APP)57 through a large number of receptors, including scavenger receptor class A type 1, MARCO, scavenger receptor class B member 1, CD36, and the receptor for advanced glycation end product,58,59 G protein-coupled receptors formyl peptide receptor 2 60 and chemokine-like receptor 1,61 toll-like receptors (TLRs) TLR2,62 TLR4, and the CD14 coreceptor, and α6β1 integrin.63 The outcome of the bond between Aβ and these receptors is the production of inflammatory mediators such as cytokines (interleukin [IL]-1α, IL-1β, IL-6, IL-8, IL-12, IL-18, and IL-23, interferon (IFN)-γ, tumor necrosis factor [TNF]-α, and granulocyte-macrophage colony-stimulating factor [GM-CSF]),64,65 chemokines (monocyte chemotactic protein 1 (MCP1), MCP-113, fractalkine),66,67 chemoattractant proteins, prostaglandins, complement factors, thromboxanes, pentraxins, NO, reactive oxygen species, leukotrienes, proteases, protease inhibitors, adhesion molecules (interaction between CD40-CD40 ligand CD40L),68 coagulation factors, and C-reactive protein, most of which are detectable in AD animal and/or in the brain or cerebrospinal fluid of AD patients.25,69,70 However, glial cells are also capable of producing some regulatory cytokines, such as IL-10 and transforming growth factor-β (TGF-β), but in AD their release is modified, exacerbating the disease.7173 Among anti-inflammatory factors, we also recall the cluster of differentiation-200 (CD200) regulated by the anti-inflammatory IL-4 and expressed by neurons, T- and B-cells, whose receptor is expressed by glia. Both AD patients and mouse models show an age-related or Aβ-induced CD200 reduction.7476

Upstream of cytokines production is the activation of the nuclear factor-kappa B (NF-κB) pathway,77 and the subsequent activation of mitogen-activated protein kinase (MAPK) pathways, whose proinflammatory gene expression is Aβ dependent.78 Extracellular signal-regulated protein kinases (ERKs), stress-activated protein kinases c-Jun NH2-terminal kinase (JNK), and p38 constitute the set of MAPKs whose action is exerted both in the cytoplasm and in the nucleus, thereby phosphorylating transcription factors. For example, p38 can contribute to neuroinflammation by inducing TNF-α gene transcription, which increases the activator protein-1 (AP-1) activity,79 besides being directly responsible for tau phosphorylation.80

In turn, proinflammatory mediators increase the activity and the products of amyloidogenic pathway, especially Aβ(1–42). For instance, the γ-secretase cell-based assays showed that TNF-α, IL-1β, and IFN-γ cause the initiation of APP cleavage through the MAPK pathway,81 and a more recent study demonstrated that NF-κB signaling, activated by TNF-α, results in an increased Aβ synthesis driven by the β-secretase (BACE-1) transcription.82

To the vicious circle driven by cytokines and MAPKs,83,84 the resulting activation of the complement cascade has to be added,85 as well as the induction of proinflammatory enzymes, such as cyclooxygenase-2 (COX-2)86 and the inducible nitric oxide synthase (iNOS).87 Induction of these enzymes may also be linked to the excessive release of S100B (ββ form of the S100), a neurotrophin expressed by activated astrocytes,88 which is able to induce NF-κB activation,89 as well as encourage tauopathy.23 Two more proinflammatory proteins, implicated in the pathophysiology of AD, belong to the S100 family: S100A9 and S100A12. These proteins, produced by activated microglia and macrophages, are increased in AD brain and are responsible for protein complex formation.90,91 S100A9 is present within SPs and Aβ deposits surrounding blood vessels, and it is also abundant in tissues neighboring Aβ deposits, confirming that increased S100A9 levels can stimulate peptides aggregation and deposition.92

Studies report that tau hyperphosphorylation is directly affected by inflammatory mediators, including the cyclin-dependent kinase 5 (CDK5):93 IL-6 stimulates neuronal protein p35, which in turn is responsible for the kinase activation that can act on tau.94 CDK5 is not the only kinase related to neuroinflammation. Recently, the role of protein kinase 2 (CK2, former casein kinase II) has been described. In fact, CK2 immunopositive astrocytes have been found to be associated with amyloid deposits in AD brains, suggesting its involvement in the neuroinflammatory response.95

Inflammatory mediators, in particular cytokines, are also responsible for increased BBB permeabilization driven by chemokines, allowing leukocyte penetration in the brain.96,97 This is possible because of altering the resistance of tight junctions, upregulation of cytokines expression, and COX-2 transcription in endothelial cells.98 For example, IL-6, IL-10, IL-13, and prostaglandins stimulated by lipopolysaccharide may increase the influx of Aβ across the BBB, besides upregulating APP processing in the brain.99,100

Another mechanism underlying the pathogenic process led by neuroinflammation is the blockage of neurogenesis, which is inhibited by some proinflammatory cytokines such as IL-6, TNF-α, and IL-18, responsible for neural progenitor cells death, and inhibition of their differentiation.101 Interestingly, these cells are located in the subgranular layer of the dentate gyrus of the hippocampus, in the subventricular zone of the lateral ventricles, and amygdala – areas mainly affected by AD and cognitive impairment.102

One of the still poorly explored mechanisms that might govern the relationship between AD and neuroinflammation, but definitely is in charge of the neurodegenerative processes, involves the glycogen synthase kinase-3 (constitutively active serine/threonine protein kinase) pathway.103 This idea comes from the observation of the results obtained by blocking this kinase, which causes an increase of the anti-inflammatory IL-10 and a decrease of proinflammatory cytokines as a consequence of TLRs stimulation and NO production.104106

The salient events reported in this paragraph are summarized in Figure 2.

Figure 2 Integrated pathways between glial activation, neuroinflammation, and neuronal death after brain injury.


Notes: Whenever a brain injury occurs, glial activation takes place with the aim of removing injurious stimuli. To this aim, activated cells undergo a series of morphofunctional changes and acquire a reactive phenotype. Activation causes, among other things, glial hypertrophy, astrocyte endfeet retraction, and gain of amoeboid microglial structure. These changes, if not stopped, can induce synaptic dysfunction, homeostatic imbalance, neurovascular unit dysfunction, loss of three-dimensional network, and BBB dysfunction. In addition, reactive microglia and astrocytes release a wide range of proinflammatory mediators aimed at removing the primary injury. The occurrence of a reactive state is very probably a protective response. However, uncontrolled and prolonged activation goes beyond physiological control, and detrimental effects override the beneficial ones. Solid arrows indicate complete pathways. Dashed arrows indicate pathways that that occur partially or do not.


Abbreviations: BBB, blood–brain barrier; Aβ, β-amyloid; NFTs, neurofibrillary tangles; SCARA-1, scavenger receptor class A Type 1; SCARB-1, scavenger receptor class B member 1; RAGE, receptor for advanced glycation end product; GPCRs, G protein-coupled receptors; FPR2, formyl peptide receptor 2; CMKLR1, chemokine-like receptor 1; TLRs, toll-like receptors; IL, interleukin; IFN-γ, interferon-γ; GM-CSF, granulocyte-macrophage colony-stimulating factor; NO, nitric oxide; ROS, reactive oxygen species; SPs, senile plaques; IBA1, ionized calcium-binding adapter molecule 1; CD, cluster of differentiation.

Neuroinflammatory targets in Alzheimer’s disease

Because of the knowledge acquired so far and the failure of so many antiamyloid trials, scientific interest has shifted to other features of neurodegeneration including neuroinflammation.107

Evidence mentioned in “The Pathophysiology of Neuroinflammation and Its Role in Alzheimer’s Disease” section shows how neuroinflammation is driven by a large number of events apparently different but strongly dependent one on the other.108 For this reason, it is difficult to identify the best target upon which to act. Recently, much work has been done, but much more research still needs to be done.

It is now clear that the AD neurodegenerative process is also orchestrated by proinflammatory cytokines and their receptors, which therefore become promising targets on which to focus by means of different approaches. Blocking gene expression of cytokines, releasing or binding their receptors, or better regulating the functioning of cells implicated in the neuroinflammation are definitely strategies still in exploration.109 The possibility of of reducing tau kinase activity and oligomeric and fibrillary Aβ accumulation by neutralizing IL-1β or TNF-α/TNF-α receptor through antibodies has been demonstrated in murine models of AD.64,110113 In this context, the role of molecules with anti-inflammatory properties (such as minocycline) that are able to decrease astrocyte release of proinflammatory cytokines and reduce both tau and amyloid pathogenesis,114 as well as improve AD behavioral symptoms, is not less important.115

Interestingly, both in vitro and in vivo studies have shown that pharmacological inhibition of COX-2 and inducible NO synthase has positive outcomes.38,116120

Lastly, in AD models it was observed that it is possible to obtain satisfactory results by modulating kinases that are not only directly related to tau hyperphosphorylation but also to neuroinflammation. One example is the modulation of glycogen synthase kinase-3β. Experimental studies have shown that it is possible to exert anti-inflammatory effect by inhibiting this enzyme, giving us another potential therapeutic target to consider.87,88,121

Currently available products and products in research and development focusing on neuroinflammatory targets

In the past decades, several epidemiological and clinical studies were carried out to demonstrate the neuroprotective potential of several nonsteroidal anti-inflammatory drugs.122,123 After the pioneering work with indometacin demonstrated the ability to restore cognitive functions in the enrolled subjects, many other clinical trials have shown only unsatisfactory results.124127 Since the failure of trials with classical nonsteroidal anti-inflammatory drugs, scientists tested COX-2-selective compounds effects. Once again, results were disappointing.128,129 Evidence from a clinical trial with naproxen suggests its ability to reduce tau and Aβ levels in cerebrospinal fluid and plasma.130 AD is a multifactorial disease and the inflammatory outcome, driven by glial activation, depends on the context and on the stage of the pathology. For these reasons, an ideal anti-inflammatory compound should be able to control the detrimental effects and, at the same time, preserve the physiological glial activation.

An alternative and recent therapeutic approach is represented by nutraceuticals (eg, curcumin, apigenin, docosahexaenoic acid, resveratrol, and n-3 fatty acids).131134 Despite encouraging preclinical results, the success rate in humans has been very low.135

Complex results were obtained after vaccinating AD patients against Aβ and NFTs. A large number of studies have been done in this field, and promising data were obtained in preclinical models. Unfortunately, these encouraging findings were not replicated in clinical trials, and promising vaccines were stopped because of adverse effects such as meningoencephalitis.136 Some of these studies revealed that immunization halts glial activation.137 By the physiological importance of this phenomenon, this is probably why the immunization has caused severe adverse reactions.

Presently, several competing hypotheses (especially related to time of intervention) may help explain the failure of translating preclinical studies into the clinical ones, but so far there is no way to confirm which of these explanations is correct.

Future research direction

The pathogenic role of neuroinflammation in AD is now well recognized and accepted. Nevertheless, the underlying mechanisms have not been sufficiently elucidated. Several factors contribute to this failure. First of all, there is a lack of adequate preclinical models that best mimic the disease and, in particular, the processes of glial activation and neuroinflammation. Then, another important factor is the comprehension of the role of each cellular component in the inflammatory process, for example, the identification of cell-specific biomarkers. Indeed, specifically clarifying changes in both immune system and inflammatory machinery would make available different pathways for pharmacological manipulations aimed at delaying the onset and/or the progression of the disease. Finally, it is important to define the inflammatory stages to correlate each phase to AD progression and to clarify which processes are protective and which ones are detrimental.

The achievement of these goals will allow scientists to practice many other experimental approaches. The hope is to get satisfactory results from clinical studies with compounds that have been successful in vitro, ex vivo, and/or in vivo experiments, such as the administration of molecules like acetylpuerarin,138 edaravone,139 palmitoylethanolamide,38 N-[2-(4-hydroxyphenyl)ethyl]-2-(2,5-dimethoxyphenyl)-3-(3-methoxy-4-hydroxyphenyl) acrylamide (compound FLZ),140 oleuropeinaglycone,141 oridonin,142 protocatechuic acid,143 resveratrol,110 rutin,144 or immunotherapies145,146 and vaccinations.147

Conclusion

Growing evidence confirms that neuroinflammation, finely orchestrated by neuronal, glial, and immune components, is a contributing cause of Aβ aggregation, tau hyperphosphorylation, and neuronal damage and death. The resulting production of cytokines and proinflammatory molecules has initially a neuroprotective role, but subsequently becomes the cause of further neurodegeneration.

Unfortunately, because of the lack of appropriate animal models, we still lack a complete understanding of the relationship between inflammatory process stages and AD progression. This could explain, at least in part, the unsuccessful results of clinical trials performed with anti-inflammatory molecules whose efficacy was significantly proven in preclinical investigations.

Therefore, future experimental studies must intensively investigate the intricate paths of the neuroinflammatory process and define the best time to control it. In this way, it will be possible to achieve more focused and functional therapeutic strategies in the hope of not only alleviating but also modifying AD progression.

Acknowledgments

This work was supported by the SAPIENZA University Grant to CS (prot. C26A15X58E). The authors thank Miss Federica Bellifemine for her helpful contribution in preparing the figure 1.

Disclosure

The authors report no conflicts of interest in this work.

References

1.

Stefaniak J, O’Brien J. Imaging of neuroinflammation in dementia: a review. J Neurol Neurosurg Psychiatry. 2016;87(1):21–28.

2.

Prince M, Wimo A, Guerchet M, Ali GC, Wu YT, Prina M. World Alzheimer Report 2015. The Global Impact of Dementia: An Analysis of Prevalence, Incidence, Cost and Trends. Alzheimer’s Disease International; 2015.

3.

Alzheimer A. Uber eine eigenartige Erkangkung der Hirnrinde (An unusual illness of the cerebral cortex). Allgemeine Zeitschr Psychisch-Gerichtliche Medizin. 1907;64:146–148.

4.

Sperling RA, Aisen PS, Beckett LA, et al. Toward defining the preclinical stages of Alzheimer’s disease: recommendations from the National Institute on Aging-Alzheimer’s Association workgroups on diagnostic guidelines for Alzheimer’s disease. Alzheimers Dement. 2011;7:280–292.

5.

Lopez OL, Schwam E, Cummings J, et al. Predicting cognitive decline in Alzheimer’s disease: an integrated analysis. Alzheimers Dement. 2010;6:431–439.

6.

Querfurth HW, LaFerla FM. Alzheimer’s disease. N Engl J Med. 2010;362(4):329–344.

7.

Braak H, Braak E. Morphological criteria for the recognition of Alzheimer’s disease and the distribution pattern of cortical changes related to this disorder. Neurobiol Aging. 1994;355–356; (discussion 379–380).

8.

Blennow K, de Leon MJ, Zetterberg H. Alzheimer’s disease. Lancet. 2006;368:387–403.

9.

Ferreira ST, Clarke JR, Bomfim TR, De Felice FG. Inflammation, defective insulin signaling, and neuronal dysfunction in Alzheimer’s disease. Alzheimers Dement. 2014;10(1 Suppl):S76–S83.

10.

Brown KL, Cosseau C, Gardy JL, Hancock RE. Complexities of targeting innate immunity to treat infection. Trends Immunol. 2007;28:260–266.

11.

Prokop S, Miller KR, Heppner FL. Microglia actions in Alzheimer’s disease. Acta Neuropathol. 2013;126:461–477.

12.

Perry VH, Holmes C. Microglial priming in neurodegenerative disease. Nature Rev. Neurol. 2014;10:217–224.

13.

Heneka MT, Kummer MP, Latz E. Innate immune activation in neurodegenerative disease. Nature Rev Immunol. 2014;14:463–477.

14.

Zhang B, Gaiteri C, Bodea LG, et al. Integrated systems approach identifies genetic nodes and networks in late-onset Alzheimer’s disease. Cell. 2013;153:707–720.

15.

Sheng JG, Bora SH, Xu G, Borchelt DR, Price DL, Koliatsos VE. Lipopolysaccharide induced-neuroinflammation increases intracellular accumulation of amyloid precursor protein and amyloid beta peptide in APPswe transgenic mice. Neurobiol Dis. 2003;14:133–145.

16.

Kitazawa M, Oddo S, Yamasaki TR, et al. Lipopolysaccharide-induced inflammation exacerbates tau pathology by a cyclin-dependent kinase 5-mediated pathway in a transgenic model of Alzheimer’s disease. J Neurosci. 2005;25:8843–8853.

17.

Vukic V, Callaghan D, Walker D, et al. Expression of inflammatory genes induced by beta-amyloid peptides in human brain endothelial cells and in Alzheimer’s brain is mediated by the JNK-AP1 signaling pathway. Neurobiol Dis. 2009;34:95–106.

18.

Cunningham C, Campion S, Lunnon K, et al. Systemic inflammation induces acute behavioral and cognitive changes and accelerates neurodegenerative disease. Biol Psychiatry. 2009;65:304–312.

19.

Vezzani M, Maroso M, Balosso S, Sanchez MA, Bartfai T. IL-1 receptor/Toll-like receptor signaling in infection, inflammation, stress and neurodegeneration couples hyperexcitability and seizures. Brain Behav Immun. 2011;25(7):1281–1289.

20.

Verkhratsky A, Rodríguez JJ, Steardo L. Astrogliopathology: a central element of neuropsychiatric diseases? Neuroscientist. 2014;20(6):576–588.

21.

Steardo L Jr, Bronzuoli MR, Iacomino A, Esposito G, Steardo L, Scuderi C. Does neuroinflammation turn on the flame in Alzheimer’s disease? Focus on astrocytes. Front Neurosci. 2015;9:259.

22.

Lécuyer MA, Kebir H, Prat A. Glial influences on BBB functions and molecular players in immune cell trafficking. Biochim Biophys Acta. 2016;1862(3):472–482.

23.

Esposito G, Scuderi C, Lu J, et al. S100B induces tau protein hyperphosphorylation via Dickopff-1 up-regulation and disrupts the Wnt pathway in human neural stem cells. J Cell Mol Med. 2008;12(3):914–927.

24.

Verkhratsky A, Nedergaard M, Hertz L. Why are astrocytes important? Neurochem Res. 2015;40(2):389–401.

25.

Deitmer JW, Rose CR. pH regulation and proton signalling by glial cells. Prog Neurobiol. 1996;48(2):73–103.

26.

Iadecola C, Nedergaard M. Glial regulation of the cerebral microvasculature. Nat Neurosci. 2007;10(11):1369–1376.

27.

Zlokovic BV. The blood-brain barrier in health and chronic neurodegenerative disorders. Neuron. 2008;57(2):178–201.

28.

Perea G, Navarrete M, Araque A. Tripartite synapses: astrocytes process and control synaptic information. Trends Neurosci. 2009;32(8):421–431.

29.

Sofroniew MV. Astrogliosis: Diversity of astrocyte functions and phenotypes in neural circuits. Cold Spring Harb Perspect Biol. 2014;7(2):a020420.

30.

Sofroniew MV, Vinters HV. Astrocytes: biology and pathology. Acta Neuropathol. 2010;119(1):7–35.

31.

Scuderi C, Stecca C, Iacomino A, Steardo L. Role of astrocytes in major neurological disorders: the evidence and implications. IUBMB Life. 2013;65(12):957–961.

32.

Paradisi S, Sacchetti B, Balduzzi M, Gaudi S, Malchiodi-Albedi F. Astrocyte modulation of in vitro beta-amyloid neurotoxicity. Glia. 2004;46(3):252–260.

33.

Thal DR. The role of astrocytes in amyloid beta-protein toxicity and clearance. Exp Neurol. 2012;236:1–5.

34.

Mathur R, Ince PG, Minett T, et al. A reduced astrocyte response to β-amyloid plaques in the ageing brain associates with cognitive impairment. PLoS One. 2015;10(2):e0118463.

35.

Mrak RE, Griffin WS. The role of activated astrocytes and of the neurotrophic cytokine S100B in the pathogenesis of Alzheimer’s disease. Neurobiol Aging. 2001;22(6):915–922.

36.

Tuppo EE, Arias HR. The role of inflammation in Alzheimer’s disease. Int J Biochem Cell Biol. 2005;37(2):289–305.

37.

Frautschy SA, Yang F, Irrizarry M, et al. Microglial response to amyloid plaques in APPsw transgenic mice. Am J Pathol. 1998;152:307–317.

38.

Liu B, Hong JS. Role of Microglia in inflammation-mediated neurodegenerative diseases: mechanisms and strategies for therapeutic intervention. J Pharm Exp Ther. 2003;304:1–7.

39.

Melchior B, Garcia AE, Hsiung BK, et al. Dual induction of TREM2 and tolerance-related transcript, Tmem176b, in amyloid transgenic mice: implications for vaccine-based therapies for Alzheimer’s disease. ASN Neuro. 2010;2(3):e00037.

40.

Jiang T, Tan L, Zhu XC, et al. Upregulation of TREM2 ameliorates neuropathology and rescues spatial cognitive impairment in a transgenic mouse model of Alzheimer’s disease. Neuropsychopharmacol. 2014;39(13):2949–2962.

41.

Guerreiro R, Wojtas A, Bras J, et al. TREM2 variants in Alzheimer’s disease. N Engl J Med. 2013;368:117–127.

42.

Jonsson T, Stefansson H, Steinberg S, et al. Variant of TREM2 associated with the risk of Alzheimer’s disease. N Engl J Med. 2013;368:107–116.

43.

Takahashi K, Rochford CD, Neumann H. Clearance of apoptotic neurons without inflammation by microglial triggering receptor expressed on myeloid cells-2. J Exp Med. 2005;201:647–657.

44.

Otero K, Shinohara M, Zhao H, et al. TREM2 and β-catenin regulate bone homeostasis by controlling the rate of osteoclastogenesis. J Immunol. 2012;188(6):2612–2621.

45.

Bouchon A, Hernandez-Munain C, Cella M, Colonna MA. DAP12-mediated pathway regulates expression of CC chemokine receptor 7 and maturation of human dendritic cells. J Exp Med. 2001;194:1111–1122.

46.

Wyss-Coray T, Loike JD, Brionne TC, et al. Adult mouse astrocytes degrade amyloid-beta in vitro and in situ. Nat Med. 2003;9:453–457.

47.

Rubio-Perez JM, Morillas-Ruiz JM. A review: inflammatory process in Alzheimer’s disease, role of cytokines. Scientific World J. 2012;2012:756357.

48.

Town T, Nikolic V, Tan J. The microglial “activation” continuum: from innate to adaptive responses. J Neuroinflammation. 2005;2:24.

49.

Gandy S, Heppner FL. Microglia as dynamic and essential components of the amyloid hypothesis. Neuron. 2013;78:575–577.

50.

Sudduth TL, Schmitt FA, Nelson PT, Wilcock DM. Neuroinflammatory phenotype in early Alzheimer’s disease. Neurobiol Aging. 2013;34:1051–1059.

51.

Holmes C, Cunningham C, Zotova E, et al. Systemic inflammation and disease progression in Alzheimer disease. Neurology. 2009;73:768–774.

52.

Scuderi C, Stecca C, Valenza M, et al. Palmitoylethanolamide controls reactive gliosis and exerts neuroprotective functions in a rat model of Alzheimer’s disease. Cell Death Dis. 2014;5:e1419.

53.

Block ML, Hong JS. Microglia and inflammation-mediated neurodegeneration: multiple triggers with a common mechanism. Prog Neurobiol. 2005;76:77–98.

54.

Christov A, Ottman JT, Grammas P. Vascular inflammatory, oxidative and protease-based processes: implications for neuronal cell death in Alzheimer’s disease. Neurol Res. 2004;26(5):540–546.

55.

Liu X, Wu Z, Hayashi Y, Nakanishi H. Age-dependent neuroinflammatory responses and deficits in long-term potentiation in the hippocampus during systemic inflammation. Neuroscience. 2012;216:133–142.

56.

Olabarria M, Noristani HN, Verkhratsky A, Rodríguez JJ. Concomitant astroglial atrophy and astrogliosis in a triple transgenic animal model of Alzheimer’s disease. Glia. 2010;58(7):831–838.

57.

Barger SW, Harmon AD. Microglial activation by Alzheimer amyloid precursor protein and modulation by apolipoprotein E. Nature. 1997;388(6645):878–881.

58.

Yu Y, Ye RD. Microglial Aβ receptors in Alzheimer’s disease. Cell Mol Neurobiol. 2015;35(1):71–83.

59.

Cai Z, Liu N, Wang C, et al. Role of RAGE in Alzheimer’s Disease. Cell Mol Neurobiol. 2016;36(4):483–495.

60.

Le Y, Gong W, Tiffany HL, et al. Amyloid (beta)42 activates a G-protein-coupled chemoattractant receptor, FPR-like-1. J Neurosci. 2001;21(2):RC123.

61.

Peng L, Yu Y, Liu J, et al. The chemerin receptor CMKLR1 is a functional receptor for amyloid-beta peptide. J Alzheimers Dis. 2014;43(1):227–242.

62.

Stewart CR, Stuart LM, Wilkinson K, et al. CD36 ligands promote sterile inflammation through assembly of a Toll-like receptor 4 and 6 heterodimer. Nature Immunol. 2010;11:155–161.

63.

Koenigsknecht J, Landreth G. Microglial phagocytosis of fibrillar β-amyloid through a β1 integrin-dependent mechanism. J Neurosci. 2004;4:9838–9846.

64.

von Bernhardi R, Tichauer JE, Eugenín J. Aging dependent changes of microglial cells and their relevance for neurodegenerative disorders. J Neurochem. 2010;112:1099–1114.

65.

Li K, Liu S, Yao S, Wang B, Dai D, Yao L. Interaction between interleukin-8 and methylenetetrahydrofolate reductase genes modulates Alzheimer’s disease risk. Dement Geriatr Cogn Disord. 2009;27:286–291.

66.

Westin K, Buchhave P, Nielsen H, Minthon L, Janciauskiene S, Hansson O. CCL2 is associated with a faster rate of cognitive decline during early stages of Alzheimer’s disease. PLoS One. 2012;7:e30525.

67.

Wu J, Bie B, Yang H, Xu JJ, Brown DL, Naguib M. Suppression of central chemokine fractalkine receptor signaling alleviates amyloid-induced memory deficiency. Neurobiol Aging. 2013;34:2843–2852.

68.

Yu S, Liu YP, Liu YH, et al. Diagnostic utility of VEGF and soluble CD40L levels in serum of Alzheimer’s patients. Yu Clin Chim Acta. 2016;453:154–159.

69.

Alam Q, Alam MZ, Mushtaq G, et al. Inflammatory process in Alzheimer and Parkinson’s diseases: central role of cytokines. Curr Pharm Des. 2016;22(5):541–548.

70.

Shen Y, Meri S. Yin and Yang: complement activation and regulation in Alzheimer’s disease. Prog Neurobiol. 2003;70(6):463–472.

71.

Wang WY, Tan MS, Yu JT, Tan L. Role of pro-inflammatory cytokines released from microglia in Alzheimer’s disease. Ann Transl Med. 2015;3(10):136.

72.

Sierra A, Gottfried-Blackmore AC, Mcewen BS, Bulloch K. Microglia derived from aging mice exhibit an altered inflammatory profile. Glia. 2007;55:412–424.

73.

Welser-Alves JV, Milner R. Microglia are the major source of TNF-alpha and TGF-beta1 in post natal glial cultures; regulation by cytokines, lipopolysaccharide and vitronectin. Neurochem Int. 2013;63:47–53.

74.

Varnum MM, Kiyota T, Ingraham KL, Ikezu S, Ikezu T. The anti-inflammatory glycoprotein, CD200, restores neurogenesis and enhances amyloid phagocytosis in a mouse model of Alzheimer’s disease. Neurobiol Aging. 2015;36(11):2995–3007.

75.

Lyons A, McQuillan K, Deighan B, et al. Decreased neuronal CD200 expression in IL-4-deficient mice results in increased neuroinflammation in response to lipopolysaccharide. Brain Behav Immunol. 2009;23(7):1020–1027.

76.

Walker DG, Dalsing-Hernandez JE, Campbell NA, Lue LF. Decreased expression of CD200 and CD200 receptor in Alzheimer’s disease: a potential mechanism leading to chronic inflammation. Exp Neurol. 2009;215(1):5–19.

77.

Wu D, Zhang X, Zhao M, Zhou AL. The role of the TLR4/NF-κB signaling pathway in Aβ accumulation in primary hippocampal neurons. Acta Physiologica Sinica. 2015;67(3):319–328.

78.

Feld M, Krawczyk MC, Sol Fustiñana M, et al. Decrease of ERK/MAPK overactivation in prefrontal cortex reverses early memory deficit in a mouse model of Alzheimer’s disease. J Alzheimers Dis. 2014;40(1):69–82.

79.

Spriggs DR, Deutsch S, Kufe DW. Genomic structure, induction and production of TNF-alpha. Immunol Ser. 1992;56:3–34.

80.

Feijoo C, Campbell DG, Jakes R, Goedert M, Cuenda A. Evidence that phosphorylation of the microtubule-associated protein Tau by SAPK4/p38delta at Thr50 promotes microtubule assembly. J Cell Sci. 2005;118:397–408.

81.

Liao YF, Wang BJ, Cheng HT, Kuo LH, Wolfe MS. Tumor necrosis factor-alpha, interleukin-1beta, and interferon-gamma stimulate gamma-secretase-mediated cleavage of amyloid precursor protein through a JNK-dependent MAPK pathway. J Biol Chem. 2004;279(47):49523–4932.

82.

Chen CH, Zhou W, Liu S, et al. Increased NF-kappaB signalling up-regulates BACE1 expression and its therapeutic potential in Alzheimer’s disease. Int J Neuropsychopharmacol. 2012;15(1):77–90.

83.

Munoz L, Ammit AJ. Targeting p38 MAPK pathway for the treatment of Alzheimer’s disease. Neuropharmacology. 2010;58(3):561–568.

84.

Glass CK, Saijo K, Winner B, Marchetto MC, Gage FH. Mechanisms underlying inflammation in neurodegeneration. Cell. 2010;140(6):918–934.

85.

Lian H, Litvinchuk A, Chiang AC, Aithmitti N, Jankowsky JL, Zheng H. Astrocyte-microglia cross talk through complement activation modulates amyloid pathology in mouse models of Alzheimer’s disease. J Neurosci. 2016;36(2):577–589.

86.

Fattahi MJ, Mirshafiey A. Positive and negative effects of prostaglandins in Alzheimer’s disease. Psychiatry Clin Neurosci. 2014;68(1):50–60.

87.

Malinski T. Nitric oxide and nitroxidative stress in Alzheimer’s disease. J Alzheimers Dis. 2007;11:207–218.

88.

Hu J, Ferreira A, Van Eldik LJ. S100β induces neuronal cell death through nitric oxide release from astrocytes. J Neurochem. 1997;69:2294–2301.

89.

Mori T1, Koyama N, Arendash GW, Horikoshi-Sakuraba Y, Tan J, Town T. Overexpression of human S100B exacerbates cerebral amyloidosis and gliosis in the Tg2576 mouse model of Alzheimer’s disease. Glia. 2010;58(3):300–314.

90.

Shepherd CE, Goyette J, Utter V, et al. Inflammatory S100A9 and S100A12 proteins in Alzheimer’s disease. Neurobiol Aging. 2006;27(11):1554–1563.

91.

Vogl T, Gharibyan AL, Morozova-Roche LA. Proinflammatory S100A8 and S100A9 proteins: self-assembly into multifunctional native and amyloid complexes. Int J Mol Sci. 2012;13(3):2893–2917.

92.

Wang C, Klechikov AG, Gharibyan AL, et al. The role of pro-inflammatory S100A9 in Alzheimer’s disease amyloid-neuroinflammatory cascade. Acta Neuropathol. 2014;127(4):507–522.

93.

Czapski GA, Gąssowska M, Wilkaniec A, Chalimoniuk M, Strosznajder JB, Adamczyk A. The mechanisms regulating cyclin-dependent kinase 5 in hippocampus during systemic inflammatory response: the effect on inflammatory gene expression. Neurochem Int. 2016;93:103–112.

94.

Quintanilla RA, Orellana D, Gonzalez-Billault C, Maccioni RB. Interleukin-6 induces Alzheimer-type phosphorylation of tau protein by deregulating the cdk5/p35 pathway. Exp Cell Res. 2004;295(1):245–257.

95.

Rosenberger AF, Morrema TH, Gerritsen WH, et al. Increased occurrence of protein kinase CK2 in astrocytes in Alzheimer’s disease pathology. J Neuroinflammation. 2016;13(1):4.

96.

Terrando N, Eriksson LI, Ryu JK, et al. Resolving postoperative neuroinflammation and cognitive decline. Ann Neurol. 2011;70(6):986–995.

97.

Engelhardt B. T cell migration into the central nervous system during health and disease: different molecular keys allow access to different central nervous system compartments. Clin Exp Neuroimmunol. 2010;1:79–93.

98.

Won D, Dorovini-Zis K, Vincent SR. Cytokines, nitric oxide, and cGMP modulate the permeability of an in vitro model of the human blood–brain barrier. Exp Neurol. 2004;190(2):446–455.

99.

Deane R, Du Yan S, Submamaryan RK, et al. RAGE mediates amyloid-beta peptide transport across the blood–brain barrier and accumulation in brain. Nat Med. 2003;9(7):907–913.

100.

Jaeger LB, Dohgu S, Sultana R, et al. Lipopolysaccharide alters the blood–brain barrier transport of amyloid beta protein: a mechanism for inflammation in the progression of Alzheimer’s disease. Brain Behav Immun. 2009;23(4):507–517.

101.

Liu YP, Lin HI, Tzeng SF. Tumor necrosis factor-alpha and interleukin-18 modulate neuronal cell fate in embryonic neural progenitor culture. Brain Res. 2005;1054(2):152–158.

102.

Bernier PJ, Bedard A, Vinet J, Levesque M, Parent A. Newly generated neurons in the amygdala and adjoining cortex of adult primates. Proc Natl Acad Sci USA. 2002;99(17):11464–11469.

103.

Phiel CJ, Wilson CA, Lee VM, Klein PS. GSK-3alpha regulates production of Alzheimer’s disease amyloid-beta peptides. Nature. 2003;423(6938):435–439.

104.

Martin M, Rehani K, Jope RS, Michalek SM. Toll-like receptor-mediated cytokine production is differentially regulated by glycogen synthase kinase 3. Nat Immunol. 2005;6(8):777–784.

105.

Wang MJ, Huang HY, Chen WF, Chang HF, Kuo JS. Glycogen synthase kinase-3beta inactivation inhibits tumor necrosis factor-alpha production in microglia by modulating nuclear factor κB and MLK3/JNK signaling cascades. J Neuroinflammation. 2010;7:99.

106.

Huang WC, Lin YS, Wang CY, et al. Glycogen synthase kinase-3negatively regulates anti-inflammatory interleukin-10 for lipopolysaccharide-induced iNOS/NO biosynthesis and RANTES production in microglial cells. Immunology. 2009;128(1 Suppl):S275–S286.

107.

Castello MA, Jeppson JD, Soriano S. Moving beyond anti-amyloid therapy for the prevention and treatment of Alzheimer’s disease. BMC Neurol. 2014;14:169.

108.

Frautschy SA, Cole GM. Why pleiotropic interventions are needed for Alzheimer’s disease. Mol Neurobiol. 2010;41:392–409.

109.

von Bernhardi R, Cornejo F, Parada GE, Eugenín J. Role of TGFβ signaling in the pathogenesis of Alzheimer’s disease. Front Cell Neurosci. 2015;9:426.

110.

Kitazawa M, Cheng D, Tsukamoto MR, et al. Blocking IL-1 signaling rescues cognition, attenuates tau pathology, and restores neuronal β-catenin pathway function in an Alzheimer’s disease model. J Immunol. 2011;187:6539–6549.

111.

Shi JQ, Shen W, Chen J, et al. Anti-TNF-α reduces amyloid plaques and tau phosphorylation and induces CD11c-positive dendritic-like cell in the APP/PS1 transgenic mouse brains. Brain Res. 2011;1368:239–247.

112.

He P, Cheng X, Staufenbiel M, et al. Long-term treatment of thalidomide ameliorates amyloid-like pathology through inhibition of β-secretase in a mouse model of Alzheimer’s disease. PLoS One. 2013;8(2):e55091.

113.

Tweedie D, Ferguson RA, Fishman K, et al. Tumor necrosis factor-α synthesis inhibitor 3,6’-dithiothalidomide attenuates markers of inflammation, Alzheimer pathology and behavioral deficits in animal models of neuroinflammation and Alzheimer’s disease. J Neuroinflammation. 2012;9:106.

114.

Garwood CJ, Cooper JD, Hanger DP, Noble W. Antiinflammatory impact of minocycline in a mouse model of tauopathy. Front Psychiatry. 2010;1:136.

115.

Parachikova A, Vasilevko V, Cribbs DH, LaFerla FM, Green KN. Reductions in amyloid-beta-derived neuroinflammation, with minocycline, restore cognition but do not significantly affect tau hyperphosphorylation. J Alzheimers Dis. 2010;21(2):527–542.

116.

Scuderi C, Stecca C, Bronzuoli MR, et al. Sirtuin modulators control reactive gliosis in an in vitro model of Alzheimer’s disease. Front Pharmacol. 2014;5:89.

117.

Bicca MA, Costa R, Loch-Neckel G, Figueiredo CP, Medeiros R, Calixto JB. B2 receptor blockage prevents Aβ-induced cognitive impairment by neuroinflammation inhibition. Behav Brain Res. 2015;278:482–491.

118.

Gan P, Zhang L, Chen Y, et al. Anti-inflammatory effects of glaucocalyxin B in microglia cells. J Pharmacol Sci. 2015;128(1):35–46.

119.

Scuderi C, Steardo L. Neuroglial roots of neurodegenerative diseases: therapeutic potential of palmitoylethanolamide in models of Alzheimer’s disease. CNS NeurolDisord Drug Targets. 2013;12(1):62–69.

120.

Cirillo C, Capoccia E, Iuvone T, et al. S100B inhibitor pentamidine attenuates reactive gliosis and reduces neuronal loss in a mouse model of Alzheimer’s disease. Biomed Res Int. 2015;2015:508342.

121.

Green HF, Nolan YM. GSK-3 mediates the release of IL-1β, TNF-α and IL-10 from cortical glia. Neurochem Int. 2012;61(5):666–671.

122.

Andersen K, Launer LJ, Ott A, Hoes AW, Breteler MM, Hofman A. Do nonsteroidal anti-inflammatory drugs decrease the risk for Alzheimer’s disease? The Rotterdam study. Neurology. 1995;45(8):1441–1445.

123.

Gasparini L, Ongini E, Wenk G. Non-steroidal antiinflammatory drugs (NSAIDs) in Alzheimer’s disease: old and new mechanisms of action. J Neurochem. 2004;91(3):521–536.

124.

Feldman HH, Doody RS, Kivipelto M, et al. Randomized controlled trial of atorvastatin in mild to moderate Alzheimer disease: LEADe. Neurology. 2010;74(12):956–964.

125.

Lyketsos CG, Breitner JC, Green RC, et al. Naproxen and celecoxib do not prevent AD in early results from a randomized controlled trial. Neurology. 2007;68:1800–1808.

126.

Martin BK, Szekely C, Brandt J, et al. Cognitive function over time in the Alzheimer’s Disease Anti-inflammatory Prevention Trial (ADAPT): results of a randomized, controlled trial of naproxen and celecoxib. Arch Neurol. 2008;65:896–905.

127.

de Jong D, Jansen R, Hoefnagels W, et al. No effect of one-year treatment with indomethacin on Alzheimer’s disease progression: a randomized controlled trial. PLoS One. 2008;3(1):e1475.

128.

Reines SA, Block GA, Morris JC, et al. Rofecoxib: no effect on Alzheimer’s disease in a 1-year, randomized, blinded, controlled study. Neurology. 2004;62(1):66–71.

129.

Thal LJ, Ferris SH, Kirby L, et al. Rofecoxib Protocol 078 study group. A randomized, double-blind, study of rofecoxib in patients with mild cognitive impairment. Neuropsychopharmacol. 2005;30(6):1204–1215.

130.

Breitner JC, Baker LD, Montine TJ, et al. Extended results of the Alzheimer’s disease anti-inflammatory prevention trial. Alzheimers Dement. 2011;7(4):402–411.

131.

Baum, L, Lam, CW, Cheung SK, et al. Six-month randomized, placebo controlled, double-blind, pilot clinical trial of curcumin in patients with Alzheimer disease. J Clin Psychopharmacol. 2008;28(1):110–113.

132.

Shinto L, Quinn J, Montine T, et al. A randomized placebo-controlled pilot trial of omega-3 fatty acids and alpha lipoic acid in Alzheimer’s disease. J Alzheimers Dis. 2014;38(1):111–120.

133.

Fava A, Pirritano D, Plastino M, et al. The effect of lipoic acid therapy on cognitive functioning in patients with Alzheimer’s disease. J Neurodegener Dis. 2013;2013:454253.

134.

Witte AV, Kerti L, Margulies DS, Floel A. Effects of resveratrol on memory performance, hippocampal functional connectivity, and glucose metabolism in healthy older adults. J Neurosci. 2014;34(23):7862–7870.

135.

Venigalla M, Sonego S, Gyengesi E, Sharman MJ, Münch G. Novel promising therapeutics against chronic neuroinflammation and neurodegeneration in Alzheimer’s disease. Neurochem Int. 2015;95:63–74.

136.

Pohanka M. Vaccination to Alzheimer Disease. Is it a promising tool or a blind way? Curr Med Chem. 2016;23(14):1432–1441.

137.

Zotova E, Bharambe V, Cheaveau M, et al. Inflammatory components in human Alzheimer’s disease and after active amyloid-β42 immunization. Brain. 2013;136(9):2677–2696.

138.

Meng QH, Lou FL, Hou WX, Liu M, Guo H, Zhang XM. Acetylpuerarin reduces inflammation and improves memory function in a rat model of Alzheimer’s disease induced by Abeta1-42. Pharmazie. 2013;68(11):904–908.

139.

Jiao SS, Yao XQ, Liu YH, et al. Edaravone alleviates Alzheimer’s disease-type pathologies and cognitive deficits. Proc Natl Acad Sci USA. 2015;112(16):5225–5230.

140.

Wu LY, Bao XQ, Pang HY, Sun H, Zhang D. FLZ attenuates learning and memory deficits via suppressing neuroinflammation induced by LPS in mice. J Asian Nat Prod Res. 2015;17(3):306–317.

141.

Casamenti F, Grossi C, Rigacci S, Pantano D, Luccarini I, Stefani M. Oleuropeinaglycone: a possible drug against degenerative conditions. In vivo evidence of its effectiveness against Alzheimer’s disease. J Alzheimers Dis. 2015;45(3):679–688.

142.

Wang S, Yang H, Yu L, et al. Oridonin attenuates Aβ1-42-induced neuroinflammation and inhibits NF-κB pathway. PLoS One. 2014;9(8):e104745.

143.

Song Y, Cui T, Xie N, Zhang X, Qian Z, Liu J. Protocatechuic acid improves cognitive deficits and attenuates amyloid deposits, inflammatory response in aged AβPP/PS1 double transgenic mice. Int Immunopharmacol. 2014;20(1):276–281.

144.

Xu PX, Wang SW, Yu XL, et al. Rutin improves spatial memory in Alzheimer’s disease transgenic mice by reducing Aβ oligomer level and attenuating oxidative stress and neuroinflammation. Behav Brain Res. 2014;264:173–180.

145.

Cantarella G, Di Benedetto G, Puzzo D, et al. Neutralization of TNFSF10 ameliorates functional outcome in a murine model of Alzheimer’s disease. Brain. 2015;138(1):203–216.

146.

Counts SE, Ray B, Mufson EJ, Perez SE, He B, Lahiri DK. Intravenous immunoglobulin (IVIG) treatment exerts antioxidant and neuropreservatory effects in preclinical models of Alzheimer’s disease. J Clin Immunol. 2014;34(1 Suppl):S80–S85.

147.

Xing XN, Zhang WG, Sha S, et al. Amyloid β 3-10 DNA vaccination suggests a potential new treatment for Alzheimer’s disease in BALB/c mice. Chin Med J (Engl). 2011;124(17):2636–2641.

Creative Commons License © 2016 The Author(s). This work is published and licensed by Dove Medical Press Limited. The full terms of this license are available at https://www.dovepress.com/terms.php and incorporate the Creative Commons Attribution - Non Commercial (unported, v3.0) License. By accessing the work you hereby accept the Terms. Non-commercial uses of the work are permitted without any further permission from Dove Medical Press Limited, provided the work is properly attributed. For permission for commercial use of this work, please see paragraphs 4.2 and 5 of our Terms.