Next Article in Journal
Systemic Inflammation Indices, Chemokines, and Metabolic Markers in Perimenopausal Women
Previous Article in Journal
Clinical Association of Haptoglobin with Oxidized LDL in Obese Patients with Type 2 Diabetes Mellitus
Previous Article in Special Issue
The Neuroprotective Potential of Ocimum Plant Species: Seasoning the Mind with Sweet and Holy Basil
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Neuroprotective Role of Curcumin: From Molecular Pathways to Clinical Translation—A Narrative Review

1
Institute of Preventive Medicine and Public Health, Semmelweis University, 1085 Budapest, Hungary
2
Fodor Center for Prevention and Healthy Aging, Semmelweis University, 1085 Budapest, Hungary
3
Health Sciences Division, Doctoral College, Semmelweis University, 1085 Budapest, Hungary
4
Department of Neurology and Stroke, Saint John’s Central Hospital of North Buda, 1125 Budapest, Hungary
5
Heart and Vascular Center, Semmelweis University, 1122 Budapest, Hungary
6
Department of Rheumatology and Clinical Immunology, Semmelweis University, 1085 Budapest, Hungary
7
Department of Pulmonology, Semmelweis University, 1083 Budapest, Hungary
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Nutrients 2025, 17(17), 2884; https://doi.org/10.3390/nu17172884
Submission received: 12 August 2025 / Revised: 31 August 2025 / Accepted: 5 September 2025 / Published: 6 September 2025
(This article belongs to the Special Issue Therapeutic Potential of Phytochemicals in Neurodegenerative Diseases)

Abstract

Neurodegenerative disorders, including Alzheimer’s disease (AD), Parkinson’s disease (PD), and post-stroke cognitive impairment (PSCI), represent an escalating global health and economic challenge. In the quest for disease-modifying interventions, natural polyphenols—most notably curcumin, the principal bioactive compound of Curcuma longa—have attracted considerable interest due to their pleiotropic neuroprotective effects. This narrative review critically synthesizes findings from a selection of peer-reviewed articles published between 2000 and 2025, chosen for their relevance to curcumin’s molecular mechanisms and translational potential. Curcumin’s complex chemical structure confers antioxidant, anti-inflammatory, and epigenetic modulatory properties; however, its clinical application is limited by poor oral bioavailability. Mechanistically, curcumin attenuates oxidative stress and suppresses key inflammatory mediators, including nuclear factor kappa B (NF-κB), cyclooxygenase-2 (COX-2), and inducible nitric oxide synthase (iNOS). Additionally, it modulates apoptosis, inhibits amyloid-beta aggregation, and enhances cellular quality control processes such as autophagy and mitophagy, while upregulating neurotrophic factors such as brain-derived neurotrophic factor (BDNF). Preclinical studies employing rodent models of AD, PD, and ischemic stroke have demonstrated curcumin’s dose-dependent neuroprotective efficacy, with improved outcomes observed using nanoparticle-based delivery systems. Early-phase clinical trials further support curcumin’s favorable safety profile and potential cognitive benefits, although challenges remain regarding pharmacokinetics, formulation standardization, and therapeutic reproducibility. Future directions include the development of advanced drug delivery platforms, combinatory therapeutic regimens, and personalized medicine approaches integrating curcumin within multifaceted neurotherapeutic strategies. Collectively, this narrative review highlights curcumin as a promising multi-targeted candidate for combating neurodegenerative diseases, while emphasizing the need for further translational and clinical validation.

1. Introduction

Neurodegenerative diseases such as Alzheimer’s disease (AD), Parkinson’s disease (PD), and post-stroke neurodegeneration pose a significant and growing global health challenge, with profound social and economic consequences [1]. These disorders are characterized by progressive neuronal loss and functional decline, and current therapeutic options largely focus on symptom management rather than disease modification [2]. Central to neurodegeneration are interrelated molecular mechanisms, including oxidative stress, chronic neuroinflammation, mitochondrial dysfunction, impaired proteostasis, and apoptosis, all contributing to irreversible neuronal damage [3].
Recently, naturally derived polyphenols with pleiotropic bioactivity have garnered considerable attention for their capacity to modulate these interconnected pathological pathways [4,5]. Among them, curcumin—the principal bioactive compound in Curcuma longa (turmeric)—has emerged as a promising neuroprotective agent due to its antioxidant, anti-inflammatory, anti-apoptotic, and metal-chelating properties [6]. Long utilized in traditional Asian medicine, curcumin has recently attracted biomedical interest for its ability to target key molecular drivers of neurodegeneration [7].
Curcumin exerts diverse neuroprotective effects, including modulation of intracellular signaling cascades, preservation of mitochondrial integrity, suppression of microglial activation, and enhancement of synaptic plasticity [8,9]. Notably, curcumin interacts with nicotinic acetylcholine receptors, particularly the α7 subtype (α7-nAChRs), which regulate dopaminergic neurotransmission, neuroinflammation, and cellular resilience [10]. Experimental models of PD suggest that α7-nAChR modulation by curcumin may contribute to improved motor function and dopaminergic neuron survival [11,12].
Despite promising in vitro, in vivo, and in silico evidence, clinical translation is hindered by curcumin’s low oral bioavailability, rapid metabolism, and limited central nervous system penetration [13]. Nevertheless, preclinical and early clinical studies report beneficial effects on cognitive function, neuroinflammation, and neuronal survival across multiple neurodegenerative models [14,15].
This article presents a narrative review that synthesizes and critically evaluates current evidence on the neuroprotective properties of curcumin, with particular emphasis on its molecular mechanisms in neurodegenerative disorders such as Alzheimer’s disease, Parkinson’s disease, and post-stroke cognitive impairment. The literature was identified through a broad search of peer-reviewed publications in major biomedical databases (2000–2025), with selection based on thematic relevance and scientific quality rather than strict systematic criteria. By clarifying the mechanistic basis of curcumin’s neuroprotective effects and assessing its translational potential, this review aims to provide a balanced overview of both the opportunities and limitations of curcumin-based interventions in neurodegenerative diseases.

2. Methods

2.1. Literature Search Strategy

A comprehensive literature search was conducted between March and May 2025 across four major databases—PubMed/MEDLINE, Web of Science, Scopus, and Embase—supplemented by Google Scholar to capture grey literature and additional sources not indexed in traditional databases.
The search strategy combined Medical Subject Headings (MeSH) with free-text keywords to ensure broad coverage. The primary search strings included:
  • “curcumin” AND (“neurodegeneration” OR “neurodegenerative disease” OR “Parkinson’s disease” OR “Alzheimer’s disease” OR “stroke” OR “cognitive impairment”)
  • “curcumin” AND (“oxidative stress” OR “neuroinflammation” OR “BDNF” OR “NF-κB” OR “mitophagy” OR “apoptosis”)
  • “curcumin” AND (“bioavailability” OR “nanoparticles” OR “clinical trial”)
Searches were limited to English-language publications between January 2000 and May 2025. The temporal restriction was chosen because research on the molecular mechanisms of curcumin in neurodegenerative disorders has significantly expanded since 2000, coinciding with advances in molecular biology techniques, biomarker discovery, and the initiation of early-phase clinical trials. Earlier literature prior to 2000 primarily addressed curcumin’s general pharmacological and anti-inflammatory effects, without a strong focus on neurodegeneration. Thus, this timeframe was selected to ensure both scientific relevance and methodological rigor. No restrictions were placed on study design, enabling the inclusion of in vitro, in vivo, clinical studies, and systematic or narrative reviews. Reference lists of key articles were also manually screened to identify additional eligible studies.

2.2. Study Selection and Inclusion Criteria

Studies were included if they:
(i)
investigated molecular or cellular mechanisms of curcumin relevant to neurodegenerative disorders;
(ii)
reported preclinical data from in vitro or in vivo models, or clinical evidence of neuroprotective effects;
(iii)
addressed pharmacokinetics, bioavailability, or delivery strategies within a neurological context.
Exclusion criteria were:
(i)
non-peer-reviewed publications (unless of significant foundational value);
(ii)
studies focused solely on non-neurological conditions;
(iii)
duplicate or non-English publications.

2.3. Guiding Research Question

To provide focus and clarity, the following guiding research question was formulated: “What are the key molecular mechanisms by which curcumin exerts neuroprotective effects, and how can these mechanisms be translated into clinical applications in neurodegenerative diseases?”

2.4. Data Extraction and Synthesis

For each included study, the following data were extracted:
(i)
molecular pathways modulated by curcumin (oxidative stress, neuroinflammation, autophagy, apoptosis, mitochondrial function);
(ii)
pharmacokinetic characteristics and bioavailability enhancement strategies;
(iii)
experimental models used, particularly animal models of Alzheimer’s disease, Parkinson’s disease, and ischemic stroke;
(iv)
clinical efficacy and safety outcomes.
The evidence was synthesized narratively, integrating mechanistic insights with preclinical and clinical findings. Special attention was given to identifying translational gaps, methodological limitations, and promising future directions for curcumin-based therapeutic interventions in neurodegenerative disorders.

3. Biological Properties of Curcumin

3.1. Chemical Structure and Physical Characteristics

Curcumin (diferuloylmethane), the principal polyphenolic constituent of Curcuma longa, was first isolated in the early 19th century, and its chemical structure was elucidated in 1910 [16]. It is a symmetric molecule composed of two aromatic rings substituted with ortho-methoxy and phenolic groups, linked by a seven-carbon chain containing a conjugated β-diketone moiety [17,18]. The molecule exists in tautomeric keto and enol forms, with the keto form predominating under acidic or neutral pH conditions, whereas the enol form is favored in alkaline environments [19].
The curcuminoid fraction of turmeric primarily comprises three naturally occurring analogs: curcumin (I), demethoxycurcumin (II), and bisdemethoxycurcumin (III), each contributing differently to its biological activity [20]. Curcumin is a highly lipophilic compound that is practically insoluble in water but soluble in organic solvents such as ethanol, dimethyl sulfoxide (DMSO), and acetone [21]. Its hydrophobic nature and chemical instability under physiological conditions pose significant challenges for its pharmacological application [22].

3.2. Pharmacokinetics and Bioavailability

Following oral administration, curcumin demonstrates poor systemic bioavailability due to limited intestinal absorption, rapid hepatic metabolism, and swift systemic elimination [23]. Clinical pharmacokinetic studies have shown that even high oral doses (4–8 g/day) yield low plasma concentrations, typically ranging from 0.4 to 1.7 micromoles per liter (μmol/L) [24]. These low levels are largely attributed to extensive first-pass hepatic metabolism, including reductive metabolism and subsequent conjugation with glucuronic acid and sulfate, producing curcumin glucuronides and sulfates (Figure 1) [25].
To overcome these limitations, a wide range of formulation strategies has been developed to enhance curcumin’s oral bioavailability [26]. These include nanoemulsions, liposomal encapsulation, and polymer-based nanoparticle systems, such as poly(lactic-co-glycolic acid) (PLGA), alginate, and lactoferrin nanoparticles [27]. Cyclodextrin complexes have also been shown to improve curcumin’s aqueous solubility and chemical stability. In addition, co-administration with piperine (an alkaloid derived from Piper nigrum) significantly inhibits curcumin glucuronidation, thereby increasing its plasma concentration by up to 2000% and extending its elimination half-life [28]. These advanced delivery systems not only increase systemic exposure but also enhance blood–brain barrier (BBB) penetration, which is crucial for achieving therapeutic concentrations in the central nervous system, particularly in the context of neurodegenerative diseases such as Alzheimer’s and Parkinson’s disease [29,30].

3.3. Antioxidant, Anti-Inflammatory, and Epigenetic Effects

Curcumin exhibits a broad spectrum of biological activities, with its antioxidant and anti-inflammatory properties underpinning its neuroprotective potential [31]. As an antioxidant, curcumin directly scavenges reactive oxygen and nitrogen species (ROS/RNS) and upregulates endogenous antioxidant enzymes, including superoxide dismutase, catalase, and glutathione peroxidase [32]. It further inhibits lipid peroxidation and preserves mitochondrial integrity under oxidative stress conditions [33].
The anti-inflammatory effects of curcumin are mediated via downregulation of pivotal inflammatory signaling pathways, notably nuclear factor-kappa B (NF-κB), alongside suppression of pro-inflammatory enzymes such as cyclooxygenase-2 (COX-2) and inducible nitric oxide synthase (iNOS) [34]. Additionally, curcumin attenuates microglial activation, thereby mitigating neuroinflammation in pathological contexts [8].
At the epigenetic level [35], curcumin modulates DNA methylation, histone acetylation, and microRNA expression, regulating gene networks associated with neuronal survival, apoptosis inhibition, and synaptic plasticity [36]. These epigenetic modifications contribute to the restoration of cellular homeostasis and enhanced resilience in neurodegenerative milieus [37]. Collectively, curcumin’s capacity to attenuate oxidative damage, suppress neuroinflammation, and promote neuronal survival underscores its therapeutic promise in central and peripheral neurodegenerative disorders [38,39]. Its multi-targeted mechanisms of action render curcumin a compelling natural candidate in the development of effective neuroprotective interventions [11].

4. Molecular Mechanisms Underlying the Neuroprotective Effects of Curcumin

The neuroprotective potential of curcumin arises from its capacity to modulate multiple interconnected pathophysiological processes within the central nervous system (CNS) [14,40]. Curcumin is widely recognized for its potent, multifaceted anti-aging [41,42,43], pro-survival, and pro-longevity properties [41,42] exerted via synergistic modulation of diverse cellular and molecular mechanisms associated with aging and neurodegeneration [44]. Its key molecular targets encompass oxidative stress, inflammatory signaling pathways, apoptosis, amyloid protein aggregation, autophagy/mitophagy, and neurotrophic factor regulation.
Emerging evidence suggests that the gut–brain axis plays a critical role in the neuroprotective effects of curcumin, a pathway that has received limited attention in both preclinical and clinical analyses. Given curcumin’s minimal oral bioavailability, a substantial portion of the ingested compound remains in the gastrointestinal tract, where it can directly modulate gut microbiota composition and function. This alternative mechanism provides a plausible explanation for curcumin’s neuroprotective effects independent of systemic plasma levels.
Preclinical studies have demonstrated that curcumin favorably alters gut microbial communities, promoting the proliferation of beneficial taxa such as Lactobacillus and Bifidobacterium, while inhibiting pro-inflammatory bacteria [45,46,47,48]. Curcumin also enhances the production of microbial metabolites, particularly short-chain fatty acids (SCFAs) such as butyrate, which exert systemic anti-inflammatory and neuroprotective effects by modulating microglial activation and tight junction protein expression in the gut epithelium [46,49]. Furthermore, curcumin has been shown to improve intestinal barrier integrity, reducing the so-called “leaky gut” phenomenon that is increasingly implicated in systemic inflammation and neuroinflammation [50,51].
Mechanistically, SCFAs produced under curcumin-induced microbial modulation can influence central nervous system function through several pathways, including histone deacetylase inhibition, enhancement of regulatory T cell differentiation, and modulation of inflammatory signaling via NF-κB and inflammasome pathways. This gut-mediated signaling may complement curcumin’s direct CNS actions, providing a comprehensive explanation for its observed neuroprotective efficacy despite low systemic bioavailability. Integrating these findings offers a more complete mechanistic framework, suggesting that curcumin’s neuroprotective potential is mediated not only by direct antioxidant and anti-inflammatory effects in the CNS but also through modulation of peripheral microbiota–gut–brain interactions [49,52].

4.1. Modulation of Oxidative Stress by Curcumin: Molecular Mechanisms and Neuroprotective Potential

Oxidative stress, characterized by the excessive intracellular accumulation of reactive oxygen species (ROS) and reactive nitrogen species (RNS), plays a central role in the etiology of numerous neurodegenerative disorders, including Alzheimer’s disease, Parkinson’s disease, and amyotrophic lateral sclerosis (ALS) [53,54]. The resulting oxidative damage contributes to aberrant protein aggregation, mitochondrial dysfunction, DNA strand breaks, and the activation of neuroinflammatory and apoptotic pathways [55]. Curcumin, a natural polyphenolic antioxidant, mitigates these deleterious effects through multiple mechanisms, including direct radical scavenging, upregulation of endogenous antioxidant defenses, and preservation of mitochondrial function.

4.1.1. Direct Free Radical-Scavenging Activity

Curcumin’s chemical structure, featuring phenolic hydroxyl groups, enables it to function as an effective hydrogen or electron donor, thereby directly neutralizing reactive free radicals such as superoxide anion (O2), hydroxyl radical (•OH), peroxynitrite (ONOO), and hydrogen peroxide (H2O2) [56]. This potent radical-scavenging capacity has been extensively validated in vitro, in animal models, and in clinical contexts [57,58]. For example, in benzo[a]pyrene-exposed murine models, oral administration of curcumin at 60 mg/kg significantly attenuated ROS production and lipid peroxidation in pulmonary tissue, concomitant with restoration of intracellular glutathione (GSH) levels [59]. Similarly, a human intervention study involving subjects with chronic exposure to arsenic-contaminated water demonstrated that supplementation with 500 mg curcumin twice daily markedly reduced systemic ROS, lipid peroxidation, and DNA damage in peripheral blood mononuclear cells [60].

4.1.2. Activation of Endogenous Antioxidant Enzymes

In addition to its direct radical-scavenging activity, curcumin enhances endogenous cellular antioxidant defenses by upregulating the activity and expression of critical antioxidant enzymes. Preclinical studies in rodent models have demonstrated that curcumin significantly increases the activity of superoxide dismutase (SOD), glutathione peroxidase (GPx), glutathione reductase (GR), catalase (CAT), and glutathione-S-transferase (GST) across multiple tissues, including liver, kidney, and nervous system [61,62]. These enzymes play vital roles in maintaining intracellular redox balance by detoxifying reactive oxygen species and preventing lipid peroxidation, thereby contributing to cellular protection against oxidative injury.

4.1.3. Activation of the Nrf2–ARE Pathway

Curcumin also exerts its antioxidant effects through activation of the nuclear factor erythroid 2–2-related factor 2 (Nrf2) signaling pathway, a master regulator of cellular defense against oxidative stress [63,64]. Upon activation, Nrf2 dissociates from its cytoplasmic inhibitor Keap1, translocates into the nucleus, and binds to antioxidant response elements (ARE) in the promoter regions of target genes. This induces the transcription of cytoprotective enzymes such as heme oxygenase-1 (HO-1), NAD(P)H quinone oxidoreductase-1 (NQO1), and glutathione synthetase, which collectively enhance cellular resilience to oxidative damage. Curcumin has been shown to promote Nrf2 nuclear translocation and increase HO-1 expression in neuronal, endothelial, and renal cell models [65,66]. This pathway is especially relevant in the central nervous system, where basal antioxidant capacity is often compromised in neurodegenerative conditions.

4.1.4. Mitochondrial Protection and Inhibition of Lipid Peroxidation

Mitochondria serve both as a primary source and a principal target of oxidative stress. Curcumin preserves mitochondrial membrane integrity, maintains mitochondrial membrane potential, and attenuates mitochondrial ROS generation [67]. Additionally, it stabilizes the electron transport chain complexes, particularly Complexes I and III, and inhibits cardiolipin oxidation, an early event triggering apoptotic cascades [68]. The anti–lipid peroxidation properties of curcumin have been corroborated in various in vivo models, as evidenced by significant reductions in malondialdehyde (MDA) and 4-hydroxynonenal (4-HNE) levels, both established biomarkers of oxidative membrane damage [59,69].

4.2. Anti-Inflammatory Mechanisms of Curcumin: Inhibition of NF-κB, COX-2, and iNOS

Curcumin exerts broad-spectrum anti-inflammatory effects predominantly via modulation of the nuclear factor kappa B (NF-κB) signaling pathway [70]. By inhibiting NF-κB activation, curcumin suppresses the transcription of pivotal pro-inflammatory mediators, including cytokines such as interleukin-1β (IL-1β), tumor necrosis factor-alpha (TNF-α), interleukin-6 (IL-6), COX-2, and iNOS. This downregulation reduces the biosynthesis of nitric oxide (NO) and prostaglandins, culminating in attenuation of both systemic and central nervous system inflammation. Consequently, curcumin is regarded as a promising adjunct therapeutic agent in inflammatory and neurodegenerative disorders (Figure 2).
Chronic NF-κB activation is a pathological hallmark of neurodegenerative diseases, driven mainly by sustained microglial activation, oxidative stress, apoptosis, and persistent release of inflammatory mediators. Curcumin’s inhibition of NF-κB signaling mitigates these deleterious processes, thereby protecting neuronal integrity. Thus, NF-κB inhibition represents a critical mechanism underlying curcumin’s neuroprotective efficacy.

4.2.1. Inhibition of the NF-κB Signaling Pathway

NF-κB is a master transcription factor that orchestrates the expression of numerous pro-inflammatory genes, including IL-1β, TNF-α, IL-6, iNOS, and COX-2. Under basal conditions, NF-κB is sequestered in the cytoplasm as an inactive complex with the inhibitor protein IκBα. Upon stimulation by agents such as lipopolysaccharide (LPS), IκB kinase (IKK) phosphorylates IκBα, triggering its ubiquitination and proteasomal degradation. This degradation liberates NF-κB, allowing its nuclear translocation and subsequent initiation of inflammatory gene transcription [71].
Curcumin modulates this signaling cascade at multiple checkpoints. It inhibits IKKβ activity, thereby preventing IκBα phosphorylation and degradation. Additionally, curcumin stabilizes IκBα, preserving the NF-κB/IκBα complex in the cytoplasm and inhibiting NF-κB nuclear translocation. Furthermore, curcumin attenuates the DNA-binding activity of NF-κB, cumulatively resulting in suppressed transcription of pro-inflammatory target genes. These multi-level inhibitory effects effectively reduce inflammation in both peripheral tissues and the central nervous system.

4.2.2. Inhibition of COX-2 Expression

COX-2 is an inducible enzyme predominantly upregulated in response to inflammatory stimuli and is critical for the biosynthesis of pro-inflammatory prostaglandins, notably prostaglandin E2 (PGE2). Curcumin exerts dose-dependent inhibition of COX-2 mRNA and protein expression across various experimental inflammation models, mainly through the suppression of transcription factors NF-κB and activator protein 1 (AP-1). This suppression is particularly relevant in neuroinflammation, where elevated PGE2 levels contribute to microglial activation and exacerbate neuronal damage [72,73]. For instance, in LPS-stimulated murine models, curcumin administration at 50–100 mg/kg significantly reduced COX-2 expression in brain tissue, correlating with attenuated neuroinflammatory responses.

4.2.3. Downregulation of iNOS Expression and Nitric Oxide Production

Inducible nitric oxide synthase (iNOS) is a key inflammatory enzyme responsible for high-output production of nitric oxide under pathological conditions. Excessive NO, particularly when reacting with superoxide anion (O2) to generate peroxynitrite (ONOO), exerts cytotoxic effects through mechanisms such as DNA damage, lipid peroxidation, and protein nitration. Curcumin effectively suppresses iNOS gene transcription and protein expression, leading to a significant reduction in NO production. These inhibitory effects have been extensively validated in multiple in vitro models, including lipopolysaccharide (LPS)-stimulated BV2 microglial cultures, as well as in in vivo models of inflammation [74,75,76,77,78,79]. For example, curcumin at concentrations of 10–20 µM markedly decreased iNOS expression and NO release in LPS-treated BV2 cells. Similarly, oral administration of curcumin at 100 mg/kg in murine models of colitis resulted in reduced intestinal iNOS expression, which correlated with improved histopathological outcomes [80].

4.3. Role of Curcumin in Apoptosis Regulation: Molecular Mechanisms and Neuroprotective Relevance

Programmed cell death (apoptosis) plays a central role in the pathophysiology of neurodegenerative diseases. Oxidative stress, inflammatory mediators, mitochondrial dysfunction, and DNA damage contribute to pathological neuronal loss, particularly in Alzheimer’s disease, Parkinson’s disease, and ischemic brain injury. Curcumin has been demonstrated to modulate major apoptotic signaling pathways, thereby promoting neuronal survival and preservation of neural tissue.

4.3.1. Inhibition of Intrinsic (Mitochondrial) Apoptosis

The intrinsic apoptotic pathway is predominantly activated by oxidative stress or mitochondrial dysfunction, with the B-cell lymphoma 2 (Bcl-2) protein family serving as key regulators. Curcumin has been reported to downregulate the expression of pro-apoptotic Bax, while upregulating anti-apoptotic proteins such as Bcl-2 and myeloid cell leukemia 1 (Mcl-1). This shift in the Bcl-2/Bax ratio prevents mitochondrial outer membrane permeabilization (MOMP), thereby inhibiting the release of cytochrome c into the cytosol and subsequent activation of the apoptosome and caspase-9—caspase-3 cascade. This mechanism is especially relevant in neurons, where curcumin selectively targets pathologically altered or dysfunctional cells for apoptosis while exerting minimal cytotoxicity on healthy neuronal populations. Such selective modulation underscores curcumin’s therapeutic potential in neurodegenerative disorders [81,82,83,84,85,86].

4.3.2. Modulation of Extrinsic, Death Receptor-Mediated Apoptosis

Curcumin also modulates the extrinsic apoptotic pathway, which is triggered by activation of death receptors on the cell surface, such as Fas (CD95) and tumor necrosis factor-related apoptosis-inducing ligand receptor (TRAIL-R). Activation of these receptors leads to recruitment of adaptor proteins and activation of initiator caspase-8, which directly cleaves and activates executioner caspase-3. Alternatively, caspase-8 cleaves BH3-interacting domain death agonist (Bid), linking to the intrinsic mitochondrial apoptotic pathway.
Preclinical studies have shown that curcumin enhances the cleavage of caspase-8 and caspase-3, suggesting a dual mechanism promoting apoptosis. This may facilitate selective elimination of inflamed, damaged, or dysfunctional cells within the CNS, contributing to the restoration of tissue homeostasis in neurodegenerative conditions [87,88,89,90].

4.3.3. Modulation of Apoptosis-Related Signaling Pathways: NF-κB, p53, and MAPK

Curcumin exerts complex regulatory effects on key apoptosis-related signaling pathways, including nuclear factor kappa B (NF-κB), tumor protein p53, and mitogen-activated protein kinases (MAPKs) [91,92,93].
  • NF-κB pathway: Curcumin inhibits NF-κB activation by stabilizing its inhibitor IκBα, preventing NF-κB nuclear translocation. This results in decreased transcription of anti-apoptotic genes such as Bcl-2, Mcl-1, and X-linked inhibitor of apoptosis protein (XIAP). In the context of chronic neuroinflammation, this may restore apoptosis susceptibility in damaged neurons.
  • p53 pathway: Curcumin upregulates p53, a key mediator of the cellular stress response. Activated p53 enhances transcription of pro-apoptotic genes such as Bax, promoting mitochondrial-mediated apoptosis in response to oxidative stress or DNA damage, thereby supporting neuronal quality control.
  • MAPK pathway: Curcumin modulates multiple MAPK signaling branches. It inhibits p38 MAPK, implicated in inflammation and apoptosis, and regulates c-Jun N-terminal kinase (JNK) and extracellular signal-regulated kinase (ERK) pathways, exerting context-dependent effects on neuronal survival and stress response.

4.3.4. Bax/Bcl-2 Ratio as a Molecular Marker

Curcumin’s modulation of the Bax/Bcl-2 ratio is well documented and serves as a reliable indicator of intrinsic apoptosis activation. An increased Bax/Bcl-2 ratio correlates with mitochondrial membrane permeabilization and progression toward apoptosis. Numerous in vitro and in vivo studies have confirmed the relevance of this parameter in assessing curcumin’s neuroprotective efficacy [94].

4.4. Modulation of Pathological Protein Aggregation

Curcumin exerts neuroprotective effects, in part, by modulating the aggregation of key pathological proteins implicated in neurodegenerative disorders, including amyloid-β (Aβ) in Alzheimer’s disease, α-synuclein in Parkinson’s disease, and Tau in tauopathies.

4.4.1. Amyloid-β Aggregation

Among curcumin’s diverse neuroprotective properties, its capacity to inhibit Aβ aggregation remains particularly significant in AD pathogenesis. The progression of AD is closely linked to misfolding and self-assembly of Aβ peptides into soluble oligomers and insoluble fibrils, which contribute to synaptic dysfunction, neuronal toxicity, and extracellular plaque deposition.
Curcumin has been shown to bind directly to Aβ oligomers and fibrils, stabilizing them in more soluble and less neurotoxic conformations, thereby inhibiting further fibrillogenesis [95,96]. This anti-aggregation activity is primarily attributed to curcumin’s molecular structure—especially its aromatic rings and β-diketone moiety—which facilitate π–π stacking interactions, hydrophobic contacts, and hydrogen bonding with Aβ residues. Structure–activity relationship studies revealed that curcumin derivatives bearing hydroxyl or trifluoromethoxy substituents enhance inhibition of Aβ fibril formation [97,98]. Furthermore, curcumin chelates redox-active metal ions such as Cu2+, Zn2+, and Fe3+, which are known to accelerate Aβ aggregation and promote ROS generation [99,100].

4.4.2. α-Synuclein Aggregation

α-Synuclein aggregation is a central pathological mechanism in PD, leading to the formation of Lewy bodies and neurodegeneration [101]. Curcumin has been reported to inhibit α-synuclein fibrillation and stabilize non-toxic oligomeric species, reducing cytotoxicity in cellular and animal PD models [102,103]. This effect appears to involve direct binding to α-synuclein monomers and oligomers, modulation of hydrophobic interactions, and interference with β-sheet formation, positioning curcumin as a potential disease-modifying agent in synucleinopathies.

4.4.3. Tau Aggregation and Hyperphosphorylation

Hyperphosphorylation of the Tau protein contributes to neurofibrillary tangle formation in AD and other tauopathies [104]. Preclinical and clinical evidence suggest that curcumin can interfere with Tau hyperphosphorylation and aggregation [105,106,107]. Notably, a Longvida® study demonstrated reductions in Tau dimer levels following curcumin supplementation [106,108,109]. Mechanistically, curcumin may modulate Tau kinases and phosphatases, reduce oxidative stress, and inhibit inflammatory pathways that promote Tau pathology [107]. Through these complementary mechanisms—including direct interactions with Aβ, α-synuclein, and Tau, inhibition of fibrillogenesis, and metal ion chelation—curcumin exhibits broad anti-protein aggregation activity. This multi-targeted mode of action supports its potential as an adjunctive or preventive therapeutic strategy across neurodegenerative disorders characterized by pathological protein aggregation.

4.5. Regulation of Autophagy and Mitophagy

Curcumin exerts neuroprotective effects partly through the modulation of autophagy and mitophagy—two fundamental cellular quality control mechanisms essential for maintaining neuronal homeostasis. Autophagy is a highly conserved lysosome-dependent catabolic process responsible for the degradation of damaged organelles, misfolded proteins, and cytoplasmic debris. In post-mitotic neurons, which lack proliferative capacity, efficient autophagic activity is crucial for proteostasis and the prevention of neurodegenerative pathology.
A substantial body of evidence indicates that curcumin enhances autophagic flux, as demonstrated by increased expression of key autophagy markers such as LC3-II (microtubule-associated protein 1A/1B-light chain 3, form II) and Beclin-1, which are involved in autophagosome formation and initiation, respectively. Curcumin promotes the conversion (lipidation) of LC3-I to LC3-II and upregulates Beclin-1 expression, reflecting the activation of the autophagic machinery [110,111].
Moreover, curcumin facilitates mitophagy, the selective degradation of damaged or depolarized mitochondria via autophagy, thus preserving mitochondrial integrity and cellular bioenergetics. It activates key mitophagy regulators, including PTEN-induced kinase 1 (PINK1) and the E3 ubiquitin ligase Parkin, which coordinate the ubiquitination and lysosomal targeting of dysfunctional mitochondria [112]. This enhancement of mitophagy contributes to increased cellular resilience against stress, mitigates apoptotic signaling pathways, and supports neuronal survival. Importantly, curcumin’s ability to induce autophagy and mitophagy extends beyond neuroprotection, bearing implications in oncology. By promoting the clearance of toxic protein aggregates and damaged mitochondria, curcumin sensitizes cancer cells to chemotherapeutic agents and disrupts tumor cell survival mechanisms [113].

4.6. Induction of Neurotrophic Factors

Brain-derived neurotrophic factor (BDNF) is a crucial neurotrophin in the central nervous system, playing a vital role in neuronal survival, differentiation, synaptogenesis, long-term potentiation (LTP), and overall neuroplasticity [114,115,116,117,118,119]. BDNF is abundantly expressed in brain regions integral to cognitive and emotional processing, including the hippocampus, prefrontal cortex, and amygdala [120].
Numerous studies have demonstrated that systemic administration of curcumin significantly upregulates both BDNF mRNA and protein expression, particularly within the hippocampus. This modulation likely underpins the cognitive enhancement and neuroprotective effects observed in preclinical models [121]. The upregulation of BDNF is mediated through activation of intracellular signaling pathways, notably the extracellular signal-regulated kinase/mitogen-activated protein kinase (ERK/MAPK) cascade and the cAMP response element-binding protein (CREB), both well-established regulators of BDNF gene transcription [11].
Elevated BDNF levels enhance synaptic plasticity and promote dendritic spine formation, processes essential for memory consolidation and learning. Correspondingly, curcumin treatment is associated with improved cognitive performance and memory restoration in models of neurodegeneration and chronic stress [122].
Beyond Alzheimer’s disease, increased BDNF expression may also confer neuroprotective effects in Parkinson’s disease by supporting the survival and function of dopaminergic neurons in the substantia nigra pars compacta. Curcumin-induced elevation of BDNF in this context may contribute to stabilization of dopaminergic neurotransmission and amelioration of motor deficits [11]. In summary, induction of BDNF constitutes a pivotal mechanism underlying curcumin’s neuroprotective actions, reinforcing its therapeutic potential as an adjunctive or preventive strategy in Alzheimer’s disease, Parkinson’s disease, and other CNS disorders.

4.7. Putative Cerebrovascular Effects of Curcumin and Their Contribution to Neuroprotection

Cerebrovascular health is a critical determinant of brain function, particularly in aging, where structural and functional alterations in the microvasculature contribute substantially to cognitive decline and neurodegeneration [123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140]. Age-related changes in cerebral microcirculation include impaired neurovascular coupling (NVC) [124,127,129,132,134,136,137,138,139,141,142,143,144,145,146,147,148,149,150], blood–brain barrier disruption [123,124,127,128,129,131,135,136,141,146,151,152,153,154,155,156,157,158,159,160,161,162,163,164,165,166], dysregulation of cerebral blood flow (CBF), endothelial dysfunction [141,167], endothelial senescence [168,169,170,171,172,173,174,175,176], microvascular rarefaction [128,177], and chronic low-grade microvascular inflammation [164,178,179]. These pathological processes compromise neuronal homeostasis by impairing oxygen and nutrient delivery, increasing exposure to neurotoxic blood-derived molecules, and disrupting the tightly regulated cellular milieu necessary for synaptic function and neuronal survival. Importantly, these microvascular disturbances are now recognized as major contributors not only to vascular cognitive impairment and dementia [180] but also to AD, PD [181,182], and post-stroke cognitive impairment [183,184,185,186,187,188,189,190,191,192], where mixed vascular–neurodegenerative pathology is common [127,129,193,194].
Curcumin’s neuroprotective effects may be partly mediated through its capacity to preserve and restore cerebrovascular health. A growing body of preclinical and in vitro evidence indicates that curcumin exerts a broad range of vasoprotective actions that are particularly relevant in the context of age-related microvascular alterations. NVC impairment—a hallmark of cerebrovascular aging—reduces the brain’s ability to match local blood flow to neuronal activity, thereby limiting oxygen and nutrient delivery to metabolically active regions [124,141,195]. By virtue of its potent anti-inflammatory and antioxidant effects, particularly through inhibition of NF-κB [196] and activation of Nrf2 signaling [197,198,199,200], curcumin can restore endothelial nitric oxide synthase (eNOS) activity, normalize vascular tone, and improve stimulus-induced cerebral blood flow responses. Another critical target is the BBB, which becomes increasingly vulnerable to disruption with age and in neurodegenerative diseases, leading to the extravasation of plasma proteins, infiltration of immune cells, and amplification of neuroinflammation [127,136,157,160]. Curcumin has been shown to upregulate tight junction proteins such as claudin-5, occludin, and ZO-1, while suppressing matrix metalloproteinase-9 (MMP-9), a key mediator of BBB breakdown [201]. Through these actions, curcumin reduces vascular permeability and protects the delicate microenvironment of the brain parenchyma.
Age-related endothelial dysfunction and reduced nitric oxide bioavailability also contribute to cerebral hypoperfusion, a well-recognized risk factor for neurodegeneration. Curcumin’s ability to scavenge reactive oxygen species and suppress vascular NADPH oxidase activity helps preserve nitric oxide signaling, thereby supporting endothelium-dependent vasodilation and stabilizing cerebral blood flow regulation. In addition, curcumin mitigates endothelial senescence—a process that drives a pro-inflammatory, pro-thrombotic, and vasoconstrictive endothelial phenotype—by reducing oxidative DNA damage, suppressing the senescence-associated secretory phenotype (SASP), and promoting endothelial cell survival and proliferation, ultimately maintaining microvascular integrity in the aging brain.
Furthermore, curcumin may counteract microvascular rarefaction, the age-related reduction in capillary density that limits metabolic support to neurons and increases their vulnerability to injury. By modulating angiogenic pathways, particularly through vascular endothelial growth factor (VEGF) and phosphoinositide 3-kinase/protein kinase B (PI3K/Akt) signaling, curcumin supports the maintenance of a functional capillary network. Furthermore, its suppression of chronic, low-grade microvascular inflammation—achieved through inhibition of pro-inflammatory cytokines such as interleukin-1 beta (IL-1β), tumor necrosis factor-alpha (TNF-α), and interleukin-6 (IL-6), as well as downregulation of adhesion molecules including intercellular adhesion molecule-1 (ICAM-1) and vascular cell adhesion molecule-1 (VCAM-1)—reduces leukocyte–endothelial interactions and prevents secondary vascular injury.
Taken together, these vasoprotective effects complement curcumin’s direct neuroprotective actions, such as inhibition of amyloid aggregation, enhancement of autophagy, and upregulation of neurotrophic factors. In age-related central nervous system disorders, the preservation of microvascular function may therefore represent a crucial mechanistic link between curcumin’s systemic vascular benefits and its ability to promote neuronal resilience and cognitive health.

5. Preclinical Evidence

Curcumin has garnered significant scientific interest as a potential adjunct therapy for various neurological and psychiatric disorders [202]. Extensive preclinical research has demonstrated its robust anti-inflammatory, antioxidant, anti-apoptotic, and neuroprotective properties across diverse models of nervous system diseases, including Alzheimer’s disease, Parkinson’s disease, depression, anxiety, stroke, and autism spectrum disorders.
These neuroprotective effects are mediated by multiple mechanisms, such as the downregulation of pro-inflammatory cytokines (e.g., TNF-α, IL-1β), reduction of reactive oxygen species (ROS), and inhibition of apoptosis through modulation of the Bax/Bcl-2 ratio. Additionally, curcumin suppresses pathological protein aggregation, notably amyloid-β and α-synuclein, which are pivotal in the pathogenesis of neurodegenerative diseases [15,203,204,205]. The following section offers a comprehensive overview of preclinical findings, with a focus on Alzheimer’s disease, Parkinson’s disease, and ischemic stroke models.

5.1. Animal Model Outcomes

5.1.1. Alzheimer’s Disease

Numerous studies have demonstrated curcumin’s neuroprotective effects in various animal models of Alzheimer’s disease. Sun et al. [45] reported that oral administration of curcumin in amyloid precursor protein/presenilin-1 (APP/PS1) transgenic mice improved spatial memory and reduced hippocampal amyloid-beta (Aβ) plaque deposition, effects linked to modulation of gut microbiota composition. Lin et al. [206] found that combined treatment with curcumin and berberine decreased Aβ1–42 levels, neuroinflammation, and oxidative stress, thereby enhancing cognition through modulation of amyloid precursor protein processing and activation of the AMPK–autophagy pathway. Similarly, Su et al. [207] showed that curcumin improved learning performance and reduced Aβ burden while attenuating microglial activation; these effects were attributed to enhanced bioavailability and regulation of NF-κB, mTOR, and Nrf2 signaling pathways.
In aluminum chloride-induced AD rat models, curcumin ameliorated cognitive deficits, reduced anxiety-like behavior, restored cholinergic function, and mitigated oxidative stress and inflammation [208]. In vitro, curcumin reversed Aβ-induced mitochondrial dysfunction and promoted mitochondrial biogenesis and synaptic protein preservation [209]. Wang et al. [210] demonstrated that curcumin suppressed astrocyte activation and neuroinflammation, resulting in improved memory in Aβ40-induced rats.
Nanocurcumin formulations have shown improved brain bioavailability, reduced amyloid plaque load, and enhanced cognitive performance in transgenic mice [211]. Likewise, NanoCurc™ decreased oxidative stress and apoptosis while augmenting antioxidant defenses both in vitro and in vivo [212].
Begum et al. [213] and Xiong et al. [214] reported that curcumin and its metabolite tetrahydrocurcumin reduced neuroinflammation and oxidative damage; notably, only curcumin prevented amyloid plaque deposition and protein oxidation, underscoring the essential role of its dienone bridge in anti-amyloidogenic activity.
Additional preclinical studies have further confirmed curcumin’s neuroprotective effects across diverse AD models. In intracerebroventricular streptozotocin (ICV-STZ)-induced rats, curcumin improved memory performance, attenuated oxidative stress, and enhanced cholinergic activity [215]. In D-galactose-induced aged mice, curcumin promoted hippocampal neurogenesis and increased brain-derived neurotrophic factor levels, leading to improved cognitive function [216]. Moreover, in behavioral assays including the object recognition test (ORT), object location test (OLT), and Y-maze, administration of curcumin at doses ranging from 50 to 100 mg/kg enhanced recognition memory, accompanied by modest anti-inflammatory effects, although no significant changes in neurogenesis were observed [217].
In human Tau (hTau) transgenic mice, curcumin formulated as Longvida significantly reduced Tau dimer levels, enhanced synaptic function, and upregulated the expression of heat shock proteins 70 and 90 (HSP70/90) [108]. Aerosolized curcumin administration decreased amyloid-beta plaque burden and improved spatial memory performance in 5XFAD transgenic mice [218]. In vitro studies under Aβ-induced oxidative stress conditions demonstrated that curcumin reduced lipid peroxidation, increased activities of antioxidant enzymes, and elevated levels of nerve growth factor (NGF) [219]. In amyloid precursor protein Swedish mutation (APPSw) transgenic mice, dietary curcumin treatment dose-dependently decreased interleukin-1 beta (IL-1β), oxidative stress markers, amyloid plaque accumulation, and microglial activation [220]. Notably, nanoformulated curcumin showed superior efficacy in attenuating neuroinflammation and enhancing cognitive function compared to conventional curcumin in streptozotocin (STZ)-induced Alzheimer’s disease models [221]. A summary of these neuroprotective effects is provided in Table 1.

5.1.2. Parkinson’s Disease

Numerous preclinical studies highlight the neuroprotective potential of curcumin across various animal and cellular models of Parkinson’s disease. Curcumin exerts its effects through multiple mechanisms, including attenuation of oxidative stress and inflammation, mitochondrial protection, and modulation of autophagy and protein aggregation.
In a copper-induced parkinsonism model, Abbaoui et al. [222] demonstrated that curcumin restored tyrosine hydroxylase (TH) expression in dopaminergic brain regions and improved motor function, likely through its antioxidant and anti-inflammatory properties. Similarly, Rajeswari et al. [223] reported that curcumin and its metabolite tetrahydrocurcumin reversed reductions in dopamine and 3,4-dihydroxyphenylacetic acid (DOPAC) levels, and inhibited monoamine oxidase-B (MAO-B) activity in a 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) mouse model. Pan et al. [224] found that curcumin suppressed c-Jun N-terminal kinase (JNK) phosphorylation, thereby preventing mitochondrial apoptosis events such as Bcl-2-associated X protein (Bax) translocation and cytochrome c release, ultimately preserving dopaminergic neurons.
In a human neuroblastoma SH-SY5Y cell line overexpressing the A53T mutant α-synuclein, Jiang et al. [225] showed that curcumin restored autophagic activity by inhibiting the mechanistic target of rapamycin (mTOR)/ribosomal protein S6 kinase beta-1 (p70S6K) signaling pathway, reducing α-synuclein accumulation. Similarly, Wang et al. [226] demonstrated that curcumin decreased reactive oxygen species production and apoptosis in SH-SY5Y cells treated with α-synuclein oligomers, while stabilizing these aggregates.
In a rotenone-induced mouse model, Ramires Júnior et al. [227] reported that nanoemulsion-formulated curcumin improved motor function, reduced lipid peroxidation, preserved mitochondrial complex I activity, and enhanced endogenous antioxidant defenses. Yang et al. [228] observed that curcumin enhanced learning and memory performance and increased hippocampal dopamine and norepinephrine levels, likely via activation of the brain-derived neurotrophic factor (BDNF)/tropomyosin receptor kinase B (TrkB)/phosphoinositide 3-kinase (PI3K) signaling pathway. Fikry et al. [229] found that curcumin protected cerebellar Purkinje neurons, attenuated astrogliosis, and restored motor function following prolonged rotenone exposure.
Jagatha et al. [230] demonstrated that curcumin increased glutathione synthesis by activating γ-glutamylcysteine ligase (γ-GCL), reduced protein oxidation, and stabilized mitochondrial function both in vitro and in a glutathione-deficient mouse model. Sharma et al. [231] reported that in a lipopolysaccharide (LPS)-induced rat model, curcumin suppressed inflammatory markers (GFAP, NF-κB, TNF-α, IL-1β), improved redox balance, reduced iron accumulation, and prevented α-synuclein aggregation.
A systematic review encompassing 13 preclinical studies with 298 animals confirmed curcumin’s efficacy in reducing dopaminergic neuron loss, elevating striatal dopamine levels, and decreasing oxidative stress, inflammation, and apoptosis in the substantia nigra and striatum (p < 0.05) [232]. Liu et al. [233] demonstrated that high-dose curcumin (160 mg/kg/day) significantly improved motor function and preserved TH-positive dopaminergic neurons in a 6-hydroxydopamine (6-OHDA) rat model, indicating dose-dependent neuroprotection via reduction in oxidative stress and neuroinflammation. Similarly, curcumin oil solution (Curoil) enhanced motor performance and dopaminergic neuron survival in an MPTP-induced PD mouse model, likely due to improved bioavailability [234]. Collectively, these findings support curcumin’s multifaceted neuroprotective actions across key pathogenic mechanisms in PD, underscoring its potential as a candidate for future disease-modifying therapies. A summary of these effects is presented in Table 2.

5.1.3. The Effects of Curcumin in Preclinical Models of Ischemic and Hemorrhagic Stroke

Extensive preclinical studies have demonstrated curcumin’s neuroprotective effects in various cerebrovascular disorder models, including ischemic and hemorrhagic stroke [39]. Despite variability in animal models, dosing, and treatment regimens, these studies consistently highlight curcumin’s potent anti-inflammatory, antioxidant, and anti-apoptotic properties, along with its ability to modulate microglial polarization.
In a middle cerebral artery occlusion (MCAO) model, Li et al. [235], showed that co-administration of 100 mg/kg curcumin with human umbilical cord mesenchymal stem cells (hUC-MSCs) significantly reduced brain edema and infarct volume while improving neurological function. This neuroprotection was mediated via activation of the AKT/GSK-3β/β-TrCP/Nrf2 signaling pathway, suppression of oxidative stress, and promotion of M2 microglial polarization. Jia et al. [236] reported that 300 mg/kg curcumin upregulated peroxiredoxin 6 (Prdx6) through SP1 induction, which was critical for its antioxidant and neuroprotective effects, as inhibition of Prdx6 abolished these benefits.
In a global cerebral ischemia model, Altinay et al. [237], demonstrated that combined pre- and post-treatment with curcumin enhanced antioxidant defenses (SOD, CAT, GPx), reduced oxidative and inflammatory markers (IL-6, TNF-α, MDA), and attenuated neuronal apoptosis. Marques et al. [238] showed that nanoemulsified curcumin improved motor function, reduced hematoma volume, and minimized weight loss more effectively than free curcumin in hemorrhagic stroke models, attributed to superior tissue penetration and bioavailability.
Ran et al. [239] found that 150 mg/kg curcumin mitigated white matter injury and improved sensorimotor function via inhibition of the NF-κB/NLRP3 inflammasome pathway and suppression of microglial pyroptosis. Xie et al. [240] in both in vivo and in vitro studies, reported that curcumin protected neurons by inhibiting mitochondrial apoptotic pathways—suppressing Bax activation, upregulating Bcl-2, decreasing caspase-3 activity, and preserving mitochondrial membrane potential.
Li et al. [241] showed that curcumin (300 mg/kg, administered 30 min before reperfusion) significantly reduced infarct size and brain edema and improved neurological scores, effects linked to decreased NF-κB, ICAM-1, MMP-9, and caspase-3 expression, suggesting preservation of blood–brain barrier (BBB) integrity. Liu et al. [242] demonstrated that 7-day curcumin treatment (300 mg/kg) promoted neurogenesis and improved behavioral outcomes through Notch pathway activation, evidenced by increased BrdU+ and BrdU/DCX+ cell counts. Wu et al. [243] reported that curcumin pretreatment upregulated tight junction proteins (ZO-1, occludin, claudin-5) and inhibited NF-κB and MMP-9, reducing BBB permeability and inflammation.
In a distal MCAO model, Liu et al. [244], found that curcumin (150 mg/kg, twice intraperitoneally) decreased infarct volume, enhanced M2 microglial activation (CD206+Iba1+), suppressed M1 phenotype (CD16+Iba1+), and lowered pro-inflammatory cytokines (TNF-α, IL-6, IL-12p70). Wu et al. [245] confirmed that curcumin reduced neuronal injury and oxidative stress in oxygen-glucose deprivation/reperfusion (OGD/R) and MCAO models by activating the AKT/Nrf2 signaling axis. Jiang et al. [246] demonstrated that intravenous curcumin protected against BBB disruption, brain edema, and neuroinflammation by inhibiting the iNOS/NO pathway in MCAO rats. Zhang et al. [247] further validated curcumin’s anti-inflammatory and anti-apoptotic actions by reducing TNF-α and p53, and increasing Sirt1 and Bcl-2 expression.
Collectively, these studies (summarized in Table 3) provide robust evidence that curcumin exerts multifaceted neuroprotective effects in experimental ischemic and hemorrhagic stroke models. Future clinical trials are warranted to evaluate the translational potential of these findings for human stroke therapy.

5.2. Dose–Response Relationships

Preclinical animal studies have generally characterized the dose–response relationship of curcumin, revealing that moderate but detectable neuroprotective effects typically occur at doses between 25 and 50 mg/kg. At these levels, curcumin primarily enhances antioxidant activity, reduces apoptosis, and improves behavioral outcomes in models of stroke and neurodegenerative diseases. More pronounced and consistent neuroprotection is observed at higher doses, particularly ≥100 mg/kg, reflecting stronger activation of mechanisms such as oxidative stress reduction, anti-inflammatory responses, and cytoprotective signaling pathways.
Several investigations report a sigmoidal dose–response curve, with biologically significant effects becoming measurable above a threshold dose of approximately 80 mg/kg. Below this threshold, curcumin’s effects may be subtle or difficult to quantify, whereas efficacy sharply increases beyond it. This dose-dependence corresponds to curcumin’s modulation of multiple cellular pathways, including the attenuation of reactive oxygen species (ROS) production, mitigation of oxidative stress-induced damage, and regulation of key signaling cascades such as AMP-activated protein kinase (AMPK), AKT/mTOR, and NF-κB. These molecular actions contribute to improved cellular function and may delay the progression of age-related disorders, including neurodegenerative and cardiovascular diseases.
Despite curcumin’s inherently low bioavailability and rapid metabolism necessitating higher doses, toxicological studies generally support a favorable safety profile for high-dose administration. To enhance efficacy, various nanoformulations—such as nanoemulsions and nanoparticles—are increasingly employed to improve targeted delivery and bioavailability.
Future research should prioritize the development of curcumin analogs with superior pharmacokinetic properties (e.g., TML-6) and the optimization of dose–response outcomes through nanotechnology and pharmaceutical innovation. Given that most data derive from preclinical models, further well-designed human clinical trials are critical to safely translate these dose-dependent effects into clinical practice, particularly for the prevention and treatment of age-related diseases [248]. In summary, the dose–response relationship underscores that curcumin’s therapeutic efficacy is strongly dose-dependent. Identifying the optimal dosing regimen remains essential to maximize its neuroprotective and antioxidant potential, especially in the context of aging and neurodegeneration.

5.3. Nanoparticle-Based Formulations

Despite curcumin’s promising therapeutic potential, its clinical utility is limited by poor bioavailability, primarily due to low absorption, rapid metabolism, and systemic clearance when administered in conventional forms. To overcome these pharmacokinetic challenges, a variety of nanotechnology-based delivery systems have been developed, substantially enhancing curcumin’s bioavailability and efficacy. Notably, encapsulation within liposomes, poly(lactic-co-glycolic acid) (PLGA) nanoparticles, and saponin-coated nanocarriers has been shown to increase bioavailability up to nine-fold compared to traditional curcumin formulations, even when co-administered with bioenhancers such as piperine [249].
Key advantages of these nanoformulations include:
  • Improved brain targeting: Nanoparticles facilitate curcumin’s transport across the blood–brain barrier, ensuring more effective delivery to brain tissues. This capability is crucial for therapeutic applications in neurodegenerative diseases, stroke, and related disorders.
  • Reduced clearance: Nanoformulations prolong systemic circulation time by slowing curcumin’s elimination, thereby increasing its bioactive presence and therapeutic impact.
  • Enhanced neuroprotective effects: Preclinical studies demonstrate that PLGA- or liposome-encapsulated curcumin can achieve similar or superior neuroprotective outcomes at lower doses than free curcumin, improving treatment consistency and reproducibility.
Moreover, liposomal hydrogels have been engineered for injectable, localized, and controlled release of curcumin, such as in promoting wound healing post-tumor resection. Clinical investigations in humans further confirm that nanocurcumin formulations yield superior therapeutic effects compared to conventional preparations, owing to more rapid absorption and improved tissue accumulation [249,250,251].
Overall, nanoparticle-based curcumin formulations represent a promising strategy to circumvent bioavailability limitations, enabling targeted and efficacious interventions for age-related diseases, neurodegenerative disorders, and cancer. Continued clinical research is essential to identify optimal formulations and administration routes to maximize therapeutic benefits.

6. Neurological Effects of Curcumin: Clinical Outcomes

Clinical trials investigating curcumin’s neurological effects have yielded variable results, though encouraging trends have emerged across several domains. In the management of depression, curcumin has demonstrated efficacy particularly in mild to moderate cases, likely via modulation of monoaminergic neurotransmission and inflammatory pathways [203]. For neuropathic and postoperative pain, diverse formulations—including nanocurcumin—have shown analgesic effects by attenuating inflammation and oxidative stress [204]. Regarding cognitive decline, Alzheimer’s disease, and aging-related impairments, multiple studies report that curcumin supplementation improves memory, attention, and learning, especially with long-term use of highly bioavailable formulations [14,15,205,252]. Although clinical data on stroke remain limited, preclinical models suggest curcumin may reduce brain injury and promote neural regeneration [205]. Additionally, both individuals with neurocognitive disorders and healthy adults have exhibited enhanced information processing, memory, and attention following curcumin intake, despite many studies involving relatively small cohorts [253]. Overall, curcumin represents a multifaceted natural compound with compelling neuroprotective potential established in preclinical research. While clinical evidence is promising, it is not yet conclusive. Future well-powered, rigorously controlled randomized trials are warranted to define optimal formulations, dosing strategies, and therapeutic indications to safely and effectively harness curcumin’s neurological benefits [203,204,254]. A detailed summary of clinical findings is presented in Table 4.

Human Clinical Trials

Numerous human clinical trials have investigated the neurological and systemic effects of curcumin, yielding varied but promising results. Small et al. [255] conducted an 18-month study involving 40 non-demented middle-aged and older adults (51–84 years), administering Theracurmin® at 90 mg twice daily (180 mg/day). Significant improvements were observed in verbal and visual memory as well as attention. PET imaging revealed reduced amyloid and tau accumulation, likely mediated by anti-inflammatory and anti-amyloid mechanisms involving the Akt/Nrf2 signaling pathway.
Cox et al. [256] evaluated the effects of 400 mg/day Longvida® curcumin in healthy adults aged 60–85 over 4 weeks. Acute supplementation (1 h post-dose) improved attention and working memory, while chronic use reduced fatigue and LDL cholesterol and improved mood, attributed to antioxidant, anti-inflammatory, and lipid metabolism-modulating effects.
Baum et al. [257] administered 1 g or 4 g/day of curcumin to 34 patients with probable or possible Alzheimer’s disease over 6 months. While no significant cognitive improvements were detected via MMSE, curcuminoids were measurable in plasma, with a trend toward increased Aβ40 levels. This study underscored the low bioavailability of curcumin and better absorption from capsule formulations.
A double-blind, placebo-controlled trial with 36 patients [258], using Curcumin C3 Complex® at 2 g and 4 g/day for 24 weeks reported no significant adverse events. Although no cognitive benefits were observed (ADAS-Cog), the study reinforced curcumin’s anti-inflammatory and antioxidant properties, as well as its capacity to inhibit amyloid-beta aggregation.
Conversely, a larger trial of 96 older adults [259] showed that daily supplementation with 1500 mg Biocurcumax™ (BCM-95®CG) for 12 months prevented cognitive decline measured by Montreal Cognitive Assessment (MoCA) scores compared to placebo. However, other cognitive metrics showed no significant difference. The authors highlighted curcumin’s antioxidant and anti-inflammatory effects and called for further biomarker research.
In Parkinson’s disease, a 9-month study with 60 patients [260], administering 80 mg nanomicellar curcumin daily found no significant improvement in motor symptoms (MDS-UPDRS) or quality of life (PDQ-39) compared to placebo, despite a trend toward improvement in the MDS-UPDRS Part III subscore. Some participants reported gastrointestinal side effects, including nausea and reflux.
In contrast, a 3-month trial with 50 PD patients [261], administering 160 mg/day nanomicellar curcumin observed significant improvements in sleep quality and overall quality of life (PDQ-39), although fatigue remained unchanged. These effects were likely mediated by antioxidant, anti-inflammatory, and neuroprotective mechanisms.
A 12-month study [262], involving 19 PD patients treated with 2 g/day of a curcumin–phospholipid complex (Meriva®) showed reductions in autonomic dysfunction (COMPASS-31) and non-motor symptoms (NMSS), with slower clinical progression (MDS-UPDRS, Hoehn and Yahr staging). Skin biopsies revealed decreased phosphorylated alpha-synuclein deposition, supporting curcumin’s blood–brain barrier penetration and anti-amyloidogenic, anti-inflammatory effects.
Lastly, a study of 56 post-stroke rehabilitation patients [263], supplementing with 500 mg curcumin plus 5 mg piperine daily for 12 weeks demonstrated significant reductions in hs-CRP, total cholesterol, triglycerides, carotid intima-media thickness (CIMT), body weight, waist circumference, and blood pressure, alongside increased total antioxidant capacity (TAC). Patients receiving curcumin also reported less pain progression relative to placebo. A detailed overview of these clinical trials in Alzheimer’s disease, Parkinson’s disease, and stroke is provided in Table 4.

7. Neurological Effects of Curcumin: A Critical Synthesis of Clinical Evidence

The outcomes of human clinical trials investigating the neuroprotective potential of curcumin are heterogeneous. Nevertheless, several studies suggest that beneficial effects may be observed under specific conditions. The aim of this synthesis is to critically examine the main factors influencing trial outcomes.

7.1. The Role of Bioavailability

Trials reporting negative or non-significant findings predominantly employed conventional, poorly bioavailable formulations, such as standard curcumin powder or Curcumin C3 Complex® [257,258]. In these studies, plasma concentrations remained low, thereby limiting the potential for measurable clinical effects.
In contrast, trials reporting beneficial outcomes generally used advanced formulations designed to enhance absorption, including Theracurmin® [255], Longvida® [256], Biocurcumax™ [259], and Meriva® [262]. These findings underscore bioavailability as a critical determinant of clinical efficacy in curcumin trials.

7.2. Characteristics of the Study Population

The characteristics of the study population also strongly influenced outcomes. Studies involving cognitively healthy participants or those with mild cognitive impairment and early-stage disease [255,256,259] were more likely to report favorable effects, such as improvements in memory, attention, and mood.
By contrast, trials conducted in patients with moderate Alzheimer’s disease [257,258] did not demonstrate consistent cognitive improvements, even at high doses. This suggests that curcumin’s primary value may lie in prevention or early intervention, while its therapeutic potential in advanced neurodegenerative stages appears limited.

7.3. Sensitivity of Endpoints and Biomarkers

Methodological differences across trials further contribute to inconsistent findings. Studies employing more sensitive cognitive tests and biomarker-based endpoints—such as PET imaging to assess amyloid and tau deposition [255] or skin biopsies to detect α-synuclein aggregates [262] were more likely to report beneficial effects.
In contrast, studies relying solely on conventional cognitive scales such as the MMSE or ADAS-Cog [257,258] often failed to detect significant changes. This suggests that curcumin’s effects may be more readily captured using sensitive and targeted outcome measures. Overall, current clinical evidence suggests that curcumin possesses neuroprotective potential, though the supporting data remain preliminary. The most consistent benefits are observed when:
  • formulations with enhanced bioavailability are used [255,256,259,262]
  • interventions are initiated in early or preclinical stages [255,256,259]
  • and sensitive cognitive or biomarker-based endpoints are employed [255,262]
Nonetheless, the overall level of evidence remains low, and methodological heterogeneity across studies significantly limits generalizability. Future research should prioritize larger, longer-duration, randomized, and well-controlled clinical trials to establish the optimal formulation, dosage, and target population for curcumin in the prevention and treatment of neurodegenerative diseases.

8. Efficacy and Safety

Curcumin has emerged as a promising neuroprotective agent against neurodegenerative disorders such as Alzheimer’s disease, Parkinson’s disease, and stroke. It exerts multifaceted effects by modulating key cellular and molecular signaling pathways involved in neuronal damage, including amyloid-β aggregation, oxidative stress, and neuroinflammation. Curcumin primarily enhances antioxidant defenses and suppresses inflammatory responses through activation of the nuclear factor erythroid 2–related factor 2 (Nrf2) pathway and inhibition of NF-κB, which underlie many of its neuroprotective properties. Additionally, curcumin promotes neuronal health and function by stimulating autophagy and neurotrophic factor production. Clinical studies support curcumin’s capacity to cross the blood–brain barrier, which is critical for its central nervous system efficacy.
Despite its potent neuroprotective effects and generally low toxicity, curcumin’s clinical translation is limited by poor bioavailability, characterized by low absorption, rapid metabolism, and systemic elimination. Nanotechnology-based delivery strategies, including liposomal and nanoparticle formulations, as well as curcumin structural analogs, are emerging approaches to enhance solubility, sustain release, and improve targeted brain delivery. These nanoformulations represent a promising avenue to overcome pharmacokinetic limitations and optimize therapeutic outcomes.
Curcumin is generally well tolerated, with low toxicity and minimal adverse effects reported in conventional formulations. However, high-bioavailability formulations have occasionally been associated with liver toxicity. Reported adverse effects include nausea, vomiting, gastrointestinal discomfort, and, rarely, hepatotoxicity. Use of curcumin during pregnancy or breastfeeding may not be safe, and further studies are needed to confirm its safety in these populations.
Nonetheless, further investigations are warranted to confirm long-term safety, optimal dosing, and efficacy, particularly in combination therapies or nanoformulated systems. The precise molecular mechanisms underlying curcumin’s neuroprotective actions also require further elucidation. Collectively, these considerations highlight the need for well-designed clinical trials and continued research into nano-based delivery strategies to fully harness curcumin’s therapeutic potential in neurodegenerative disease contexts [255,256,257].

9. Bioavailability, Standardization

Despite its considerable therapeutic potential, curcumin’s clinical translation is significantly limited by its poor bioavailability, characterized by low oral absorption, rapid metabolism, and systemic elimination. These pharmacokinetic challenges reduce curcumin’s effective concentration in target tissues and limit its clinical efficacy, particularly in neurodegenerative and metabolic disorders. To address these limitations, a variety of advanced delivery strategies and novel formulations have been developed. Nanotechnology-based approaches—including liposomal curcumin, solid lipid nanoparticles (SLNP), polymeric nanoparticles, nanomicelles, dendrimers, nanoemulsions, and nanocrystals—have emerged as particularly promising due to their ability to enhance solubility, protect curcumin from rapid metabolism, facilitate sustained release, and improve targeted tissue and blood–brain barrier delivery [264].
Clinical and preclinical evidence supports the pharmacological benefits of these formulations [265,266,267]. Liposomal curcumin has been shown to maintain sustained plasma concentrations and elicited favorable tumor marker responses in metastatic cancer patients, indicating its potential for targeted delivery. Nanomicelle formulations improved glycemic control, lipid profiles, and inflammatory markers in diabetic patients undergoing hemodialysis, while also demonstrating bone-strengthening effects in postmenopausal women. Solid lipid curcumin particles (SLCP, e.g., Longvida®) increased plasma curcumin levels and were associated with cognitive improvements in healthy older adults. Nanocurcumin formulations have also shown neuroprotective efficacy in amyotrophic lateral sclerosis patients and exhibited immunomodulatory effects in COVID-19, reducing pro-inflammatory cytokine levels and improving clinical outcomes without significant adverse effects [264,265,266,267].
Beyond nanoformulations, other innovative delivery strategies have been investigated to further enhance curcumin’s pharmacokinetic and pharmacodynamic properties. Curcumin-galactomannoside (CGM) formulations utilize natural dietary fibers to improve absorption and tissue distribution [268]. Phytosomal curcumin formulations (e.g., Curserin®) leverage amphipathic phospholipids to increase gastrointestinal stability and bioavailability [269]. Colloidal and amorphous formulations, such as Theracurmin® and CurcuRougeTM, demonstrate superior plasma concentrations and area-under-curve metrics compared to conventional curcumin extracts, translating into improved therapeutic outcomes in musculoskeletal, metabolic, and neurodegenerative disorders [270,271].
Collectively, these data underscore that optimizing curcumin formulations is critical to overcoming its intrinsic pharmacokinetic limitations. The integration of nano- and other advanced delivery systems not only enhances bioavailability but also maximizes therapeutic efficacy across a range of disease contexts, including neurodegeneration, metabolic disorders, cardiovascular disease, cancer, and inflammatory conditions. Future research should continue to elucidate the mechanisms by which these formulations enhance curcumin’s bioavailability, determine optimal dosing regimens, evaluate long-term safety, and validate clinical efficacy in well-controlled, large-scale trials [272,273].

10. Future Perspectives

Curcumin has demonstrated considerable neuroprotective potential; however, its clinical translation is hindered by pharmacokinetic limitations, including poor bioavailability, rapid metabolism, and limited blood–brain barrier penetration. To address these challenges, ongoing research is focused on advanced drug delivery strategies, such as nanoparticle-based systems, liposomal formulations, and structural analogues. These approaches have shown promise in improving curcumin’s solubility, stability, and BBB permeability, thereby enhancing its therapeutic profile in neurodegenerative diseases [23]. Prodrug strategies, which involve chemical modification of curcumin to increase systemic stability and prolong circulation time, may further improve its pharmacokinetics and pharmacodynamics, resulting in more consistent therapeutic exposure at target sites [274].
Combination therapy is another promising avenue, whereby curcumin is administered alongside pharmacological agents [161,172,275,276,277,278,279,280,281,282], nutraceuticals [283], dietary interventions [284,285,286,287,288,289,290,291,292] and lifestyle modifications [293,294,295,296,297,298,299,300,301,302,303,304,305,306,307,308,309,310,311] to exploit synergistic effects. This multitarget approach can modulate several pathogenic processes simultaneously—oxidative stress, neuroinflammation, and protein misfolding—thereby potentially improving overall treatment efficacy [312]. The integration of curcumin into personalized medicine frameworks also holds potential. Biomarker-driven and genotype-based patient stratification could allow treatments to be tailored to individual pathophysiological profiles, maximizing therapeutic benefit while minimizing adverse effects, especially in heterogeneous disorders such as Alzheimer’s and Parkinson’s diseases [313,314].
Realizing these advances will require multidisciplinary research bridging nanotechnology, neuropharmacology, molecular biology, and clinical sciences. Long-term, well-designed randomized controlled trials are necessary to establish optimal formulations, dosing regimens, and safety profiles. Ultimately, embedding curcumin into combination and precision medicine strategies may substantially enhance its clinical utility in neurodegenerative disease management.

11. Limitations

This review is narrative in nature and was not conducted as a systematic review. Consequently, it does not claim to comprehensively cover all available literature, and relevant studies may have been inadvertently omitted. The primary aim was to summarize and interpret current evidence to highlight key mechanistic insights and potential therapeutic implications, rather than to exhaustively analyze all published data. As such, the conclusions should be interpreted with caution, taking into account the potential for selection bias and the qualitative rather than quantitative synthesis of the findings.

12. Conclusions

Curcumin’s neuroprotective potential is increasingly supported by both preclinical and clinical evidence, particularly in Alzheimer’s disease, Parkinson’s disease, and stroke. Preclinical models have shown that curcumin modulates multiple disease-relevant pathways, including reduction in oxidative stress, suppression of neuroinflammation, and inhibition of amyloid-β aggregation. Additional mechanisms—such as stimulation of autophagy and neurotrophic factor production—further contribute to neuronal survival, regeneration, and improved cognitive performance.
Despite this strong preclinical foundation, clinical outcomes remain variable, largely due to curcumin’s low bioavailability, rapid metabolism, and restricted BBB penetration. Advances in nanotechnology, including nanoparticle encapsulation, liposomal delivery, and prodrug development, are promising solutions to these challenges, offering improved pharmacokinetics, stability, and targeted delivery. Combination therapies, leveraging curcumin’s antioxidant and anti-inflammatory properties alongside conventional pharmacological agents, may yield synergistic benefits and improve patient outcomes.
Personalized medicine approaches, guided by biomarker profiling and genetic background, may further optimize curcumin therapy for individual patients. While the evidence base is expanding, successful translation into routine clinical practice will require continued multidisciplinary research to refine formulations, elucidate mechanisms of action, and confirm long-term safety and efficacy. As a natural, multifunctional compound with a favorable safety profile, curcumin remains a highly promising candidate for the prevention and management of neurodegenerative disorders—provided its pharmacological limitations can be effectively addressed.

Author Contributions

A.L., T.C., Á.L., D.M. and V.F.-P. designed the study, and wrote and published the manuscript. A.L., T.J., J.T.V. and V.Z. supplemented and reviewed the manuscript. P.V., J.T.V. and M.F. prepared the manuscript for publication. Á.S. contributed to the interpretation of results and critically revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

Open access funding provided by Semmelweis University. Project: no. TKP2021-NKTA-47 was funded by the National Research, Development and Innovation Fund under TKP2021-NKTA, with support from the Ministry of Innovation and Technology of Hungary. The project was funded by the Ministry of Innovation and Technology under the National Cardiovascular Laboratory Program (RRF-2.3.1-21-2022-00003) from the National Research, Development and Innovation Fund. This work was also supported by the European University for Well-Being (EUniWell) program (grant agreement number: 101004093/EUniWell/EAC-A02-2019/EAC-A02-2019-1) and the National Institute on Aging (NIA R03AG070479, NIA K01AG073614), the American Heart Association AHA CDA941290, and the NIA-supported Geroscience Training Program in Oklahoma (T32AG052363). A.L. was supported by the EKÖP-2024-9 New National Excellence Program of the Ministry for Culture and Innovation from the National Research, Development and Innovation Fund. The funding sources had no role in the study design; in the collection, analysis and interpretation of data; in the writing of the report; or in the decision to submit the article for publication.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

Conflicts of Interest

The authors declare that they have no competing interests.

Abbreviations

6-OHDA—6-Hydroxydopamine, Aβ—Amyloid-beta, AD—Alzheimer’s disease, AKT/Akt—Protein kinase B, APP—Amyloid precursor protein, BDNF—Brain-derived neurotrophic factor, BBB—Blood–brain barrier, CAT—Catalase, Cur—Curcumin, Curoil—Curcumin oil, DOPAC—3,4-Dihydroxyphenylacetic acid, GFAP—Glial fibrillary acidic protein, GPx—Glutathione peroxidase, GSH/GSSG—Reduced/oxidized glutathione, HSP70/HSP90—Heat shock proteins, ICV/i.c.v.—Intracerebroventricular, i.p.—Intraperitoneal, IL-1β—Interleukin-1 beta, IL-6—Interleukin-6, iNOS—Inducible nitric oxide synthase, JNK—c-Jun N-terminal kinase, MDA—Malondialdehyde, MCAO—Middle cerebral artery occlusion, MMP-9—Matrix metalloproteinase-9, MPTP—1-Methyl-4-phenyl-1,2,3,6-tetrahydropyridine, Nrf2—Nuclear factor erythroid 2-related factor 2, NMSS—Non-Motor Symptoms Scale, NGF—Nerve growth factor, NO—Nitric oxide, OGD/R—Oxygen-glucose deprivation/reperfusion, PET—Positron emission tomography, PD—Parkinson’s disease, p-α-syn—Phosphorylated alpha-synuclein, PI3K—Phosphoinositide 3-kinase, Prdx6—Peroxiredoxin 6, ROS—Reactive oxygen species, SOD—Superoxide dismutase, STZ—Streptozotocin, TH—Tyrosine hydroxylase, TH+ neurons—Tyrosine hydroxylase-positive neurons, TNF-α—Tumor necrosis factor-alpha, Y-maze—Behavioral test for spatial working memory, ZO-1—Zonula occludens-1.

References

  1. Zaib, S.; Javed, H.; Khan, I.; Jaber, F.; Sohail, A.; Zaib, Z.; Mehboob, T.; Tabassam, N.; Ogaly, H.A. Neurodegenerative diseases: Their onset, epidemiology, causes and treatment. ChemistrySelect 2023, 8, e202300225. [Google Scholar] [CrossRef]
  2. Lamptey, R.N.; Chaulagain, B.; Trivedi, R.; Gothwal, A.; Layek, B.; Singh, J. A review of the common neurodegenerative disorders: Current therapeutic approaches and the potential role of nanotherapeutics. Int. J. Mol. Sci. 2022, 23, 1851. [Google Scholar] [CrossRef]
  3. Teleanu, D.M.; Niculescu, A.-G.; Lungu, I.I.; Radu, C.I.; Vladâcenco, O.; Roza, E.; Costăchescu, B.; Grumezescu, A.M.; Teleanu, R.I. An overview of oxidative stress, neuroinflammation, and neurodegenerative diseases. Int. J. Mol. Sci. 2022, 23, 5938. [Google Scholar] [CrossRef]
  4. Di Meo, F.; Valentino, A.; Petillo, O.; Peluso, G.; Filosa, S.; Crispi, S. Bioactive polyphenols and neuromodulation: Molecular mechanisms in neurodegeneration. Int. J. Mol. Sci. 2020, 21, 2564. [Google Scholar] [CrossRef]
  5. Bhullar, K.S.; Rupasinghe, H.V. Polyphenols: Multipotent therapeutic agents in neurodegenerative diseases. Oxidative Med. Cell. Longev. 2013, 2013, 891748. [Google Scholar] [CrossRef]
  6. Panda, P.; Mohanty, S.; Gouda, S.R.; Baral, T.C.; Mohanty, A.; Nayak, J.; Mohapatra, R. Advanced strategies for enhancing the neuroprotective potential of curcumin: Delivery systems and mechanistic insights in neurodegenerative disorders. Nutr. Neurosci. 2025, 28, 1151–1176. [Google Scholar] [CrossRef]
  7. Akaberi, M.; Sahebkar, A.; Emami, S.A. Turmeric and curcumin: From traditional to modern medicine. In Studies on Biomarkers and New Targets in Aging Research in Iran: Focus on Turmeric and Curcumin; Springer: Cham, Switzerland, 2021; pp. 15–39. [Google Scholar]
  8. Cianciulli, A.; Calvello, R.; Ruggiero, M.; Panaro, M.A. Inflammaging and brain: Curcumin and its beneficial potential as regulator of microglia activation. Molecules 2022, 27, 341. [Google Scholar] [CrossRef]
  9. Bagheri, H.; Ghasemi, F.; Barreto, G.E.; Rafiee, R.; Sathyapalan, T.; Sahebkar, A. Effects of curcumin on mitochondria in neurodegenerative diseases. Biofactors 2020, 46, 5–20. [Google Scholar] [CrossRef]
  10. Piovesana, R.; Salazar Intriago, M.S.; Dini, L.; Tata, A.M. Cholinergic modulation of neuroinflammation: Focus on α7 nicotinic receptor. Int. J. Mol. Sci. 2021, 22, 4912. [Google Scholar] [CrossRef] [PubMed]
  11. Nebrisi, E.E. Neuroprotective activities of curcumin in Parkinson’s disease: A review of the literature. Int. J. Mol. Sci. 2021, 22, 11248. [Google Scholar] [CrossRef] [PubMed]
  12. Islam, M.R.; Rauf, A.; Akter, S.; Akter, H.; Al-Imran, M.I.K.; Fakir, M.N.H.; Thufa, G.K.; Islam, M.T.; Hemeg, H.A.; Abdulmonem, W.A.; et al. Neuroprotective Potential of Curcumin in Neurodegenerative Diseases: Clinical Insights Into Cellular and Molecular Signaling Pathways. J. Biochem. Mol. Toxicol. 2025, 39, e70369. [Google Scholar] [CrossRef] [PubMed]
  13. Liu, W.; Zhai, Y.; Heng, X.; Che, F.Y.; Chen, W.; Sun, D.; Zhai, G. Oral bioavailability of curcumin: Problems and advancements. J. Drug Target. 2016, 24, 694–702. [Google Scholar] [CrossRef]
  14. Sarker, M.R.; Franks, S.F. Efficacy of curcumin for age-associated cognitive decline: A narrative review of preclinical and clinical studies. Geroscience 2018, 40, 73–95. [Google Scholar] [CrossRef]
  15. Voulgaropoulou, S.D.; Van Amelsvoort, T.; Prickaerts, J.; Vingerhoets, C. The effect of curcumin on cognition in Alzheimer’s disease and healthy aging: A systematic review of pre-clinical and clinical studies. Brain Res. 2019, 1725, 146476. [Google Scholar] [CrossRef]
  16. Malik, P.; Mukherjee, T.K. Structure-function elucidation of antioxidative and prooxidative activities of the polyphenolic compound curcumin. Chin. J. Biol. 2014, 2014, 396708. [Google Scholar] [CrossRef]
  17. Noureddin, S.A.; El-Shishtawy, R.M.; Al-Footy, K.O. Curcumin analogues and their hybrid molecules as multifunctional drugs. Eur. J. Med. Chem. 2019, 182, 111631. [Google Scholar] [CrossRef]
  18. Priyadarsini, K.I. The chemistry of curcumin: From extraction to therapeutic agent. Molecules 2014, 19, 20091–20112. [Google Scholar] [CrossRef]
  19. Rege, S.A.; Arya, M.; Momin, S.A. Structure activity relationship of tautomers of curcumin: A review. Ukr. Food J. 2019, 8, 45–60. [Google Scholar] [CrossRef]
  20. Shahrajabian, M.H.; Sun, W. The golden spice for life: Turmeric with the pharmacological benefits of curcuminoids components, including curcumin, bisdemethoxycurcumin, and demethoxycurcumins. Curr. Org. Synth. 2024, 21, 665–683. [Google Scholar] [CrossRef] [PubMed]
  21. Zielińska, A.; Alves, H.; Marques, V.; Durazzo, A.; Lucarini, M.; Alves, T.F.; Morsink, M.; Willemen, N.; Eder, P.; Chaud, M.V. Properties, extraction methods, and delivery systems for curcumin as a natural source of beneficial health effects. Medicina 2020, 56, 336. [Google Scholar] [CrossRef]
  22. Stanić, Z. Curcumin, a compound from natural sources, a true scientific challenge—A review. Plant Foods Hum. Nutr. 2017, 72, 1–12. [Google Scholar] [CrossRef]
  23. Anand, P.; Kunnumakkara, A.B.; Newman, R.A.; Aggarwal, B.B. Bioavailability of curcumin: Problems and promises. Mol. Pharm. 2007, 4, 807–818. [Google Scholar] [CrossRef] [PubMed]
  24. Hsieh, C. Phase I clinical trial of curcumin, a chemopreventive agent, in patients with high-risk or pre-malignant lesions. Anticancer. Res. 2001, 21, e2900. [Google Scholar]
  25. Jamil, Q.A. The Natural Compound Curcumin: Impact of Cellular Uptake and Metabolism on In Vitro Activity. Ph.D. Thesis, University of Vienna, Vienna, Austria, 2018. [Google Scholar]
  26. Tabanelli, R.; Brogi, S.; Calderone, V. Improving curcumin bioavailability: Current strategies and future perspectives. Pharmaceutics 2021, 13, 1715. [Google Scholar] [CrossRef] [PubMed]
  27. Pan-On, S.; Dilokthornsakul, P.; Tiyaboonchai, W. Trends in advanced oral drug delivery system for curcumin: A systematic review. J. Control. Release 2022, 348, 335–345. [Google Scholar] [CrossRef]
  28. Prasad, S.; Tyagi, A.K.; Aggarwal, B.B. Recent developments in delivery, bioavailability, absorption and metabolism of curcumin: The golden pigment from golden spice. Cancer Res. Treat. Off. J. Korean Cancer Assoc. 2014, 46, 2–18. [Google Scholar] [CrossRef]
  29. Dei Cas, M.; Ghidoni, R. Dietary curcumin: Correlation between bioavailability and health potential. Nutrients 2019, 11, 2147. [Google Scholar] [CrossRef]
  30. Mohi-Ud-Din, R.; Mir, R.H.; Wani, T.U.; Shah, A.J.; Mohi-Ud-Din, I.; Dar, M.A.; Pottoo, F.H. Novel drug delivery system for curcumin: Implementation to improve therapeutic efficacy against neurological disorders. Comb. Chem. High. Throughput Screen. 2022, 25, 607–615. [Google Scholar] [CrossRef]
  31. Aminnezhad, S.; Zonobian, M.A.; Moradi Douki, M.; Mohammadi, M.R.; Azarakhsh, Y. Curcumin and their derivatives with anti-inflammatory, neuroprotective, anticancer, and antimicrobial activities: A review. Micro Nano Bio Asp. 2023, 2, 25–34. [Google Scholar]
  32. Hunyadi, A. The mechanism (s) of action of antioxidants: From scavenging reactive oxygen/nitrogen species to redox signaling and the generation of bioactive secondary metabolites. Med. Res. Rev. 2019, 39, 2505–2533. [Google Scholar] [CrossRef]
  33. Wei, Q.-Y.; Chen, W.-F.; Zhou, B.; Yang, L.; Liu, Z.-L. Inhibition of lipid peroxidation and protein oxidation in rat liver mitochondria by curcumin and its analogues. Biochim. Biophys. Acta (BBA)-Gen. Subj. 2006, 1760, 70–77. [Google Scholar] [CrossRef] [PubMed]
  34. Surh, Y.-J.; Chun, K.-S.; Cha, H.-H.; Han, S.S.; Keum, Y.-S.; Park, K.-K.; Lee, S.S. Molecular mechanisms underlying chemopreventive activities of anti-inflammatory phytochemicals: Down-regulation of COX-2 and iNOS through suppression of NF-κB activation. Mutat. Res. Fundam. Mol. Mech. Mutagen. 2001, 480, 243–268. [Google Scholar] [CrossRef]
  35. Reuter, S.; Gupta, S.C.; Park, B.; Goel, A.; Aggarwal, B.B. Epigenetic changes induced by curcumin and other natural compounds. Genes Nutr. 2011, 6, 93–108. [Google Scholar] [CrossRef]
  36. Teiten, M.H.; Dicato, M.; Diederich, M. Curcumin as a regulator of epigenetic events. Mol. Nutr. Food Res. 2013, 57, 1619–1629. [Google Scholar] [CrossRef]
  37. Boyanapalli, S.S.; Kong, A.-N.T. “Curcumin, the king of spices”: Epigenetic regulatory mechanisms in the prevention of cancer, neurological, and inflammatory diseases. Curr. Pharmacol. Rep. 2015, 1, 129–139. [Google Scholar] [CrossRef] [PubMed]
  38. Azzini, E.; Peña-Corona, S.I.; Hernández-Parra, H.; Chandran, D.; Saleena, L.A.K.; Sawikr, Y.; Peluso, I.; Dhumal, S.; Kumar, M.; Leyva-Gómez, G. Neuroprotective and anti-inflammatory effects of curcumin in Alzheimer’s disease: Targeting neuroinflammation strategies. Phytother. Res. 2024, 38, 3169–3189. [Google Scholar] [CrossRef]
  39. Moldoveanu, C.-A.; Tomoaia-Cotisel, M.; Sevastre-Berghian, A.; Tomoaia, G.; Mocanu, A.; Pal-Racz, C.; Toma, V.-A.; Roman, I.; Ujica, M.-A.; Pop, L.-C. A Review on Current Aspects of Curcumin-Based Effects in Relation to Neurodegenerative, Neuroinflammatory and Cerebrovascular Diseases. Molecules 2024, 30, 43. [Google Scholar] [CrossRef]
  40. Mukherjee, S.; Mishra, A.K.; Peer, G.D.G.; Bagabir, S.A.; Haque, S.; Pandey, R.P.; Raj, V.S.; Jain, N.; Pandey, A.; Kar, S.K. The Interplay of the Unfolded Protein Response in Neurodegenerative Diseases: A Therapeutic Role of Curcumin. Front. Aging Neurosci. 2021, 13, 767493. [Google Scholar] [CrossRef]
  41. Ng, T.P.; Nyunt, S.Z.; Gao, Q.; Gwee, X.; Chua, D.Q.L.; Yap, K.B. Curcumin-rich curry consumption and life expectancy: Singapore longitudinal ageing study. Geroscience 2024, 46, 969–980. [Google Scholar] [CrossRef]
  42. Izadi, M.; Sadri, N.; Abdi, A.; Zadeh, M.M.R.; Jalaei, D.; Ghazimoradi, M.M.; Shouri, S.; Tahmasebi, S. Longevity and anti-aging effects of curcumin supplementation. Geroscience 2024, 46, 2933–2950. [Google Scholar] [CrossRef] [PubMed]
  43. Strong, R.; Miller, R.A.; Astle, C.M.; Baur, J.A.; de Cabo, R.; Fernandez, E.; Guo, W.; Javors, M.; Kirkland, J.L.; Nelson, J.F.; et al. Evaluation of resveratrol, green tea extract, curcumin, oxaloacetic acid, and medium-chain triglyceride oil on life span of genetically heterogeneous mice. J. Gerontol. A Biol. Sci. Med. Sci. 2013, 68, 6–16. [Google Scholar] [CrossRef] [PubMed]
  44. Pospiech, E.; Bar, A.; Pisarek-Pacek, A.; Karas, A.; Branicki, W.; Chlopicki, S. Epigenetic clock in the aorta and age-related endothelial dysfunction in mice. Geroscience 2024, 46, 3993–4002. [Google Scholar] [CrossRef]
  45. Sun, Z.Z.; Li, X.Y.; Wang, S.; Shen, L.; Ji, H.F. Bidirectional interactions between curcumin and gut microbiota in transgenic mice with Alzheimer’s disease. Appl. Microbiol. Biotechnol. 2020, 104, 3507–3515. [Google Scholar] [CrossRef] [PubMed]
  46. Zhu, J.; He, L. The modulatory effects of curcumin on the gut microbiota: A potential strategy for disease treatment and health promotion. Microorganisms 2024, 12, 642. [Google Scholar] [CrossRef]
  47. Balaji, S.; Jeyaraman, N.; Jeyaraman, M.; Ramasubramanian, S.; Muthu, S.; Santos, G.S.; da Fonseca, L.F.; Lana, J.F. Impact of curcumin on gut microbiome. World J. Exp. Med. 2025, 15, 100275. [Google Scholar] [CrossRef] [PubMed]
  48. Zam, W. Gut microbiota as a prospective therapeutic target for curcumin: A review of mutual influence. J. Nutr. Metab. 2018, 2018, 1367984. [Google Scholar] [CrossRef]
  49. Cerullo, M.; Armeli, F.; Mengoni, B.; Menin, M.; Crudeli, M.L.; Businaro, R. Curcumin Modulation of the Gut–Brain Axis for Neuroinflammation and Metabolic Disorders Prevention and Treatment. Nutrients 2025, 17, 1430. [Google Scholar] [CrossRef]
  50. Ghosh, S.S.; He, H.; Wang, J.; Gehr, T.W.; Ghosh, S. Curcumin-mediated regulation of intestinal barrier function: The mechanism underlying its beneficial effects. Tissue Barriers 2018, 6, e1425085. [Google Scholar] [CrossRef]
  51. Wang, J.; Ghosh, S.S.; Ghosh, S. Curcumin improves intestinal barrier function: Modulation of intracellular signaling, and organization of tight junctions. Am. J. Physiol.-Cell Physiol. 2017, 312, C438–C445. [Google Scholar] [CrossRef]
  52. Di Meo, F.; Margarucci, S.; Galderisi, U.; Crispi, S.; Peluso, G. Curcumin, gut microbiota, and neuroprotection. Nutrients 2019, 11, 2426. [Google Scholar] [CrossRef]
  53. Ramalingam, M.; Kim, S.-J. Reactive oxygen/nitrogen species and their functional correlations in neurodegenerative diseases. J. Neural Transm. 2012, 119, 891–910. [Google Scholar] [CrossRef] [PubMed]
  54. Tabner, B.J.; Turnbull, S.; El-Aganf, O.; Allsop, D. Production of reactive oxygen species from aggregating proteins implicated in Alzheimer’s disease, Parkinson’s disease and other neurodegenerative diseases. Curr. Top. Med. Chem. 2001, 1, 507–517. [Google Scholar] [CrossRef]
  55. Bhat, A.H.; Dar, K.B.; Anees, S.; Zargar, M.A.; Masood, A.; Sofi, M.A.; Ganie, S.A. Oxidative stress, mitochondrial dysfunction and neurodegenerative diseases; a mechanistic insight. Biomed. Pharmacother. 2015, 74, 101–110. [Google Scholar] [CrossRef]
  56. Borra, S.K.; Mahendra, J.; Gurumurthy, P.; Iqbal, S.S.; Mahendra, L. Effect of curcumin against oxidation of biomolecules by hydroxyl radicals. J. Clin. Diagn. Res. JCDR 2014, 8, CC01. [Google Scholar] [CrossRef] [PubMed]
  57. Gupta, N.; Verma, K.; Nalla, S.; Kulshreshtha, A.; Lall, R.; Prasad, S. Free radicals as a double-edged sword: The cancer preventive and therapeutic roles of curcumin. Molecules 2020, 25, 5390. [Google Scholar] [CrossRef] [PubMed]
  58. Wolnicka-Glubisz, A.; Wisniewska-Becker, A. Dual action of curcumin as an anti-and pro-oxidant from a biophysical perspective. Antioxidants 2023, 12, 1725. [Google Scholar] [CrossRef]
  59. Liu, Y.; Wu, Y.-M.; Zhang, P.-Y. Protective effects of curcumin and quercetin during benzo (a) pyrene induced lung carcinogenesis in mice. Eur. Rev. Med. Pharmacol. Sci. 2015, 19, 1736–1743. [Google Scholar]
  60. Biswas, J.; Sinha, D.; Mukherjee, S.; Roy, S.; Siddiqi, M.; Roy, M. Curcumin protects DNA damage in a chronically arsenic-exposed population of West Bengal. Hum. Exp. Toxicol. 2010, 29, 513–524. [Google Scholar] [CrossRef]
  61. Iqbal, M.; Sharma, S.D.; Okazaki, Y.; Fujisawa, M.; Okada, S. Dietary supplementation of curcumin enhances antioxidant and phase II metabolizing enzymes in ddY male mice: Possible role in protection against chemical carcinogenesis and toxicity. Pharmacol. Toxicol. 2003, 92, 33–38. [Google Scholar] [CrossRef]
  62. Piper, J.T.; Singhal, S.S.; Salameh, M.S.; Torman, R.T.; Awasthi, Y.C.; Awasthi, S. Mechanisms of anticarcinogenic properties of curcumin: The effect of curcumin on glutathione linked detoxification enzymes in rat liver. Int. J. Biochem. Cell Biol. 1998, 30, 445–456. [Google Scholar] [CrossRef]
  63. Shahcheraghi, S.H.; Salemi, F.; Peirovi, N.; Ayatollahi, J.; Alam, W.; Khan, H.; Saso, L. Nrf2 regulation by curcumin: Molecular aspects for therapeutic prospects. Molecules 2021, 27, 167. [Google Scholar] [CrossRef]
  64. Shen, G.; Xu, C.; Hu, R.; Jain, M.R.; Gopalkrishnan, A.; Nair, S.; Huang, M.T.; Chan, J.Y.; Kong, A.N. Modulation of nuclear factor E2-related factor 2-mediated gene expression in mice liver and small intestine by cancer chemopreventive agent curcumin. Mol. Cancer Ther. 2006, 5, 39–51. [Google Scholar] [CrossRef]
  65. Balogun, E.; Hoque, M.; Gong, P.; Killeen, E.; Green, C.J.; Foresti, R.; Alam, J.; Motterlini, R. Curcumin activates the haem oxygenase-1 gene via regulation of Nrf2 and the antioxidant-responsive element. Biochem. J. 2003, 371, 887–895. [Google Scholar] [CrossRef]
  66. González-Reyes, S.; Guzmán-Beltrán, S.; Medina-Campos, O.N.; Pedraza-Chaverri, J. Curcumin pretreatment induces Nrf2 and an antioxidant response and prevents hemin-induced toxicity in primary cultures of cerebellar granule neurons of rats. Oxidative Med. Cell. Longev. 2013, 2013, 801418. [Google Scholar] [CrossRef]
  67. Kocyigit, A.; Guler, E.M. Curcumin induce DNA damage and apoptosis through generation of reactive oxygen species and reducing mitochondrial membrane potential in melanoma cancer cells. Cell. Mol. Biol. 2017, 63, 97–105. [Google Scholar] [CrossRef] [PubMed]
  68. Naserzadeh, P.; Mehr, S.N.; Sadabadi, Z.; Seydi, E.; Salimi, A.; Pourahmad, J. Curcumin protects mitochondria and cardiomyocytes from oxidative damage and apoptosis induced by hemiscorpius lepturus venom. Drug Res. 2018, 68, 113–120. [Google Scholar] [CrossRef] [PubMed]
  69. Ghosh, D.; Choudhury, S.T.; Ghosh, S.; Mandal, A.K.; Sarkar, S.; Ghosh, A.; Saha, K.D.; Das, N. Nanocapsulated curcumin: Oral chemopreventive formulation against diethylnitrosamine induced hepatocellular carcinoma in rat. Chem.-Biol. Interact. 2012, 195, 206–214. [Google Scholar] [CrossRef]
  70. Shishodia, S.; Potdar, P.; Gairola, C.G.; Aggarwal, B.B. Curcumin (diferuloylmethane) down-regulates cigarette smoke-induced NF-kappaB activation through inhibition of IkappaBalpha kinase in human lung epithelial cells: Correlation with suppression of COX-2, MMP-9 and cyclin D1. Carcinogenesis 2003, 24, 1269–1279. [Google Scholar] [CrossRef] [PubMed]
  71. Kao, N.-J.; Hu, J.-Y.; Wu, C.-S.; Kong, Z.-L. Curcumin represses the activity of inhibitor-κB kinase in dextran sulfate sodium-induced colitis by S-nitrosylation. Int. Immunopharmacol. 2016, 38, 1–7. [Google Scholar] [CrossRef]
  72. Yu, Y.; Shen, Q.; Lai, Y.; Park, S.Y.; Ou, X.; Lin, D.; Jin, M.; Zhang, W. Anti-inflammatory effects of curcumin in microglial cells. Front. Pharmacol. 2018, 9, 386. [Google Scholar] [CrossRef]
  73. Liang, X.; Wu, L.; Wang, Q.; Hand, T.; Bilak, M.; McCullough, L.; Andreasson, K. Function of COX-2 and prostaglandins in neurological disease. J. Mol. Neurosci. 2007, 33, 94–99. [Google Scholar] [CrossRef] [PubMed]
  74. Coleman, J.W. Nitric oxide in immunity and inflammation. Int. Immunopharmacol. 2001, 1, 1397–1406. [Google Scholar] [CrossRef]
  75. Habib, S.; Ali, R. Acquired antigenicity of DNA after modification with peroxynitrite. Int. J. Biol. Macromol. 2005, 35, 221–225. [Google Scholar] [CrossRef]
  76. Shishodia, S. Molecular mechanisms of curcumin action: Gene expression. Biofactors 2013, 39, 37–55. [Google Scholar] [CrossRef]
  77. Chan, M.M.-Y.; Huang, H.-I.; Fenton, M.R.; Fong, D. In vivo inhibition of nitric oxide synthase gene expression by curcumin, a cancer preventive natural product with anti-inflammatory properties. Biochem. Pharmacol. 1998, 55, 1955–1962. [Google Scholar] [CrossRef] [PubMed]
  78. Jung, K.K.; Lee, H.S.; Cho, J.Y.; Shin, W.C.; Rhee, M.H.; Kim, T.G.; Kang, J.H.; Kim, S.H.; Hong, S.; Kang, S.Y. Inhibitory effect of curcumin on nitric oxide production from lipopolysaccharide-activated primary microglia. Life Sci. 2006, 79, 2022–2031. [Google Scholar] [CrossRef]
  79. Liu, C.; Takada, K.; Zhu, D. Targeting Wnt/β-catenin pathway for drug therapy. Med. Drug Discov. 2020, 8, 100066. [Google Scholar]
  80. Martin, D.A.; Bolling, B.W. A review of the efficacy of dietary polyphenols in experimental models of inflammatory bowel diseases. Food Funct. 2015, 6, 1773–1786. [Google Scholar] [CrossRef]
  81. Kannan, K.; Jain, S.K. Oxidative stress and apoptosis. Pathophysiology 2000, 7, 153–163. [Google Scholar] [CrossRef]
  82. Yu, J.; Zhou, X.; He, X.; Dai, M.; Zhang, Q. Curcumin induces apoptosis involving bax/bcl-2 in human hepatoma SMMC-7721 cells. Asian Pac. J. Cancer Prev. 2011, 12, 1925–1929. [Google Scholar] [PubMed]
  83. Kuttikrishnan, S.; Siveen, K.S.; Prabhu, K.S.; Khan, A.Q.; Ahmed, E.I.; Akhtar, S.; Ali, T.A.; Merhi, M.; Dermime, S.; Steinhoff, M. Curcumin induces apoptotic cell death via inhibition of PI3-kinase/AKT pathway in B-precursor acute lymphoblastic leukemia. Front. Oncol. 2019, 9, 484. [Google Scholar] [CrossRef]
  84. Bai, C.; Zhao, J.; Su, J.; Chen, J.; Cui, X.; Sun, M.; Zhang, X. Curcumin induces mitochondrial apoptosis in human hepatoma cells through BCLAF1-mediated modulation of PI3K/AKT/GSK-3β signaling. Life Sci. 2022, 306, 120804. [Google Scholar] [CrossRef]
  85. Chang, R.; Sun, L.; Webster, T.J. Selective cytotoxicity of curcumin on osteosarcoma cells compared to healthy osteoblasts. Int. J. Nanomed. 2014, 9, 461–465. [Google Scholar] [CrossRef][Green Version]
  86. Syng-Ai, C.; Kumari, A.L.; Khar, A. Effect of curcumin on normal and tumor cells: Role of glutathione and bcl-2. Mol. Cancer Ther. 2004, 3, 1101–1108. [Google Scholar] [CrossRef]
  87. Ismail, N.I.; Othman, I.; Abas, F.; H Lajis, N.; Naidu, R. Mechanism of apoptosis induced by curcumin in colorectal cancer. Int. J. Mol. Sci. 2019, 20, 2454. [Google Scholar] [CrossRef] [PubMed]
  88. Jiang, A.J.; Jiang, G.; Li, L.T.; Zheng, J.N. Curcumin induces apoptosis through mitochondrial pathway and caspases activation in human melanoma cells. Mol. Biol. Rep. 2015, 42, 267–275. [Google Scholar] [PubMed]
  89. Bush, J.A.; Cheung, K.J., Jr.; Li, G. Curcumin induces apoptosis in human melanoma cells through a Fas receptor/caspase-8 pathway independent of p53. Exp. Cell Res. 2001, 271, 305–314. [Google Scholar] [CrossRef]
  90. Wahl, H.; Tan, L.; Griffith, K.; Choi, M.; Liu, J.R. Curcumin enhances Apo2L/TRAIL-induced apoptosis in chemoresistant ovarian cancer cells. Gynecol. Oncol. 2007, 105, 104–112. [Google Scholar] [CrossRef]
  91. Prakobwong, S.; Gupta, S.C.; Kim, J.H.; Sung, B.; Pinlaor, P.; Hiraku, Y.; Wongkham, S.; Sripa, B.; Pinlaor, S.; Aggarwal, B.B. Curcumin suppresses proliferation and induces apoptosis in human biliary cancer cells through modulation of multiple cell signaling pathways. Carcinogenesis 2011, 32, 1372–1380. [Google Scholar]
  92. Choudhuri, T.; Pal, S.; Agwarwal, M.L.; Das, T.; Sa, G. Curcumin induces apoptosis in human breast cancer cells through p53-dependent Bax induction. FEBS Lett. 2002, 512, 334–340. [Google Scholar] [CrossRef] [PubMed]
  93. Wu, M.F.; Huang, Y.H.; Chiu, L.Y.; Cherng, S.H.; Sheu, G.T.; Yang, T.Y. Curcumin Induces Apoptosis of Chemoresistant Lung Cancer Cells via ROS-Regulated p38 MAPK Phosphorylation. Int. J. Mol. Sci. 2022, 23, 8248. [Google Scholar] [CrossRef]
  94. Zhu, L.; Han, M.B.; Gao, Y.; Wang, H.; Dai, L.; Wen, Y.; Na, L.X. Curcumin triggers apoptosis via upregulation of Bax/Bcl-2 ratio and caspase activation in SW872 human adipocytes. Mol. Med. Rep. 2015, 12, 1151–1156. [Google Scholar] [CrossRef]
  95. Yang, F.; Lim, G.P.; Begum, A.N.; Ubeda, O.J.; Simmons, M.R.; Ambegaokar, S.S.; Chen, P.P.; Kayed, R.; Glabe, C.G.; Frautschy, S.A. Curcumin inhibits formation of amyloid β oligomers and fibrils, binds plaques, and reduces amyloid in vivo. J. Biol. Chem. 2005, 280, 5892–5901. [Google Scholar] [CrossRef]
  96. Reinke, A.A.; Gestwicki, J.E. Structure–activity Relationships of amyloid beta-aggregation inhibitors based on curcumin: Influence of linker length and flexibility. Chem. Biol. Drug Des. 2007, 70, 206–215. [Google Scholar] [CrossRef] [PubMed]
  97. Lakey-Beitia, J.; González, Y.; Doens, D.; Stephens, D.E.; Santamaría, R.; Murillo, E.; Gutiérrez, M.; Fernández, P.L.; Rao, K.; Larionov, O.V. Assessment of novel curcumin derivatives as potent inhibitors of inflammation and amyloid-β aggregation in Alzheimer’s disease. J. Alzheimer’s Dis. 2017, 60, S59–S68. [Google Scholar] [CrossRef]
  98. Orteca, G.; Tavanti, F.; Bednarikova, Z.; Gazova, Z.; Rigillo, G.; Imbriano, C.; Basile, V.; Asti, M.; Rigamonti, L.; Saladini, M. Curcumin derivatives and Aβ-fibrillar aggregates: An interactions’ study for diagnostic/therapeutic purposes in neurodegenerative diseases. Bioorganic Med. Chem. 2018, 26, 4288–4300. [Google Scholar] [CrossRef]
  99. Baum, L.; Ng, A. Curcumin interaction with copper and iron suggests one possible mechanism of action in Alzheimer’s disease animal models. J. Alzheimer’s Dis. 2004, 6, 367–377. [Google Scholar] [CrossRef]
  100. Ferrari, E. Curcumin derivatives as metal-chelating agents: Implications for potential therapeutic agents for neurological disorders. In Curcumin for Neurological and Psychiatric Disorders: Neurochemical and Pharmacological Properties; Elsevier: Amsterdam, The Netherlands, 2019; pp. 275–299. [Google Scholar]
  101. Schulz-Schaeffer, W.J. The synaptic pathology of α-synuclein aggregation in dementia with Lewy bodies, Parkinson’s disease and Parkinson’s disease dementia. Acta Neuropathol. 2010, 120, 131–143. [Google Scholar] [CrossRef] [PubMed]
  102. Genchi, G.; Lauria, G.; Catalano, A.; Carocci, A.; Sinicropi, M.S. Neuroprotective effects of curcumin in neurodegenerative diseases. Foods 2024, 13, 1774. [Google Scholar] [CrossRef] [PubMed]
  103. Javed, H.; Nagoor Meeran, M.F.; Azimullah, S.; Adem, A.; Sadek, B.; Ojha, S.K. Plant extracts and phytochemicals targeting α-synuclein aggregation in Parkinson’s disease models. Front. Pharmacol. 2019, 9, 1555. [Google Scholar] [CrossRef]
  104. Sajjad, R.; Arif, R.; Shah, A.; Manzoor, I.; Mustafa, G. Pathogenesis of Alzheimer’s Disease: Role of Amyloid-β and Hyperphosphorylated Tau Protein. Indian. J. Pharm. Sci. 2018, 80, 581–591. [Google Scholar] [CrossRef]
  105. Dubey, T.; Sonawane, S.K.; Mannava, M.C.; Nangia, A.K.; Chandrashekar, M.; Chinnathambi, S. The inhibitory effect of Curcumin-Artemisinin co-amorphous on Tau aggregation and Tau phosphorylation. Colloids Surf. B Biointerfaces 2023, 221, 112970. [Google Scholar] [CrossRef]
  106. Rane, J.S.; Bhaumik, P.; Panda, D. Curcumin inhibits tau aggregation and disintegrates preformed tau filaments in vitro. J. Alzheimer’s Dis. 2017, 60, 999–1014. [Google Scholar] [CrossRef]
  107. Sivanantharajah, L.; Mudher, A. Curcumin as a holistic treatment for tau pathology. Front. Pharmacol. 2022, 13, 903119. [Google Scholar] [CrossRef] [PubMed]
  108. Ma, Q.-L.; Zuo, X.; Yang, F.; Ubeda, O.J.; Gant, D.J.; Alaverdyan, M.; Teng, E.; Hu, S.; Chen, P.-P.; Maiti, P.; et al. Curcumin suppresses soluble tau dimers and corrects molecular chaperone, synaptic, and behavioral deficits in aged human tau transgenic mice. J. Biol. Chem. 2013, 288, 4056–4065. [Google Scholar] [CrossRef] [PubMed]
  109. Ullah, F.; Asgarov, R.; Venigalla, M.; Liang, H.; Niedermayer, G.; Münch, G.; Gyengesi, E. Effects of a solid lipid curcumin particle formulation on chronic activation of microglia and astroglia in the GFAP-IL6 mouse model. Sci. Rep. 2020, 10, 2365. [Google Scholar] [CrossRef]
  110. Zhou, H.; S Beevers, C.; Huang, S. The targets of curcumin. Curr. Drug Targets 2011, 12, 332–347. [Google Scholar] [CrossRef] [PubMed]
  111. Zhu, J.; Zhao, B.; Xiong, P.; Wang, C.; Zhang, J.; Tian, X.; Huang, Y. Curcumin induces autophagy via inhibition of yes-associated protein (YAP) in human colon cancer cells. Med. Sci. Monit. Int. Med. J. Exp. Clin. Res. 2018, 24, 7035. [Google Scholar] [CrossRef]
  112. Cao, S.; Wang, C.; Yan, J.; Li, X.; Wen, J.; Hu, C. Curcumin ameliorates oxidative stress-induced intestinal barrier injury and mitochondrial damage by promoting Parkin dependent mitophagy through AMPK-TFEB signal pathway. Free Radic. Biol. Med. 2020, 147, 8–22. [Google Scholar] [CrossRef]
  113. Kim, J.Y.; Cho, T.J.; Woo, B.H.; Choi, K.U.; Lee, C.H.; Ryu, M.H.; Park, H.R. Curcumin-induced autophagy contributes to the decreased survival of oral cancer cells. Arch. Oral. Biol. 2012, 57, 1018–1025. [Google Scholar] [CrossRef]
  114. Takei, Y.; Amagase, Y.; Goto, A.; Kambayashi, R.; Izumi-Nakaseko, H.; Hirasawa, A.; Sugiyama, A. Adipose chemokine ligand CX3CL1 contributes to maintaining the hippocampal BDNF level, and the effect is attenuated in advanced age. Geroscience 2025, 47, 5969–5983. [Google Scholar] [CrossRef]
  115. Tait, J.L.; Duckham, R.L.; Rantalainen, T.; Milte, C.M.; Main, L.C.; Nowson, C.A.; Sanders, K.M.; Taaffe, D.R.; Hill, K.D.; Abbott, G.; et al. Effects of a 6-month dual-task, power-based exercise program on cognitive function, neurological and inflammatory markers in older adults: Secondary analysis of a cluster randomised controlled trial. Geroscience 2024, 47, 1251–1268. [Google Scholar] [CrossRef]
  116. Bali, Z.K.; Nagy, L.V.; Bruszt, N.; Bodo, K.; Engelmann, P.; Hernadi, Z.; Gonter, K.; Tadepalli, S.A.; Hernadi, I. Increased brain cytokine level associated impairment of vigilance and memory in aged rats can be alleviated by alpha7 nicotinic acetylcholine receptor agonist treatment. Geroscience 2024, 46, 645–664. [Google Scholar] [CrossRef]
  117. Eroglu, B.; Isales, C.; Eroglu, A. Age and duration of obesity modulate the inflammatory response and expression of neuroprotective factors in mammalian female brain. Aging Cell 2024, 23, e14313. [Google Scholar] [CrossRef]
  118. Lautrup, S.; Myrup Holst, C.; Yde, A.; Asmussen, S.; Thinggaard, V.; Larsen, K.; Laursen, L.S.; Richner, M.; Vaegter, C.B.; Prieto, G.A.; et al. The role of aging and brain-derived neurotrophic factor signaling in expression of base excision repair genes in the human brain. Aging Cell 2023, 22, e13905. [Google Scholar] [CrossRef] [PubMed]
  119. Jia, P.; Wang, J.; Ren, X.; He, J.; Wang, S.; Xing, Y.; Chen, D.; Zhang, X.; Zhou, S.; Liu, X.; et al. An enriched environment improves long-term functional outcomes in mice after intracerebral hemorrhage by mechanisms that involve the Nrf2/BDNF/glutaminase pathway. J. Cereb. Blood Flow. Metab. 2023, 43, 694–711. [Google Scholar] [CrossRef] [PubMed]
  120. Park, H.; Poo, M.-m. Neurotrophin regulation of neural circuit development and function. Nat. Rev. Neurosci. 2013, 14, 7–23. [Google Scholar] [CrossRef]
  121. Xu, Y.; Ku, B.; Tie, L.; Yao, H.; Jiang, W.; Ma, X.; Li, X. Curcumin reverses the effects of chronic stress on behavior, the HPA axis, BDNF expression and phosphorylation of CREB. Brain Res. 2006, 1122, 56–64. [Google Scholar] [CrossRef]
  122. Wang, R.; Li, Y.B.; Li, Y.H.; Xu, Y.; Wu, H.L.; Li, X.J. Curcumin protects against glutamate excitotoxicity in rat cerebral cortical neurons by increasing brain-derived neurotrophic factor level and activating TrkB. Brain Res. 2008, 1210, 84–91. [Google Scholar] [CrossRef]
  123. Zachariou, V.; Pappas, C.; Bauer, C.E.; Shao, X.; Liu, P.; Lu, H.; Wang, D.J.J.; Gold, B.T. Regional differences in the link between water exchange rate across the blood-brain barrier and cognitive performance in normal aging. Geroscience 2024, 46, 265–282. [Google Scholar] [CrossRef] [PubMed]
  124. van Dinther, M.; Voorter, P.H.M.; Zhang, E.; van Kuijk, S.M.J.; Jansen, J.F.A.; van Oostenbrugge, R.J.; Backes, W.H.; Staals, J. The neurovascular unit and its correlation with cognitive performance in patients with cerebral small vessel disease: A canonical correlation analysis approach. Geroscience 2024, 46, 5061–5073. [Google Scholar] [CrossRef]
  125. Stankovics, L.; Ungvari, A.; Fekete, M.; Nyul-Toth, A.; Mukli, P.; Patai, R.; Csik, B.; Gulej, R.; Conley, S.; Csiszar, A.; et al. The vasoprotective role of IGF-1 signaling in the cerebral microcirculation: Prevention of cerebral microhemorrhages in aging. Geroscience 2024, 47, 445. [Google Scholar] [CrossRef]
  126. Salwierz, P.; Thapa, S.; Taghdiri, F.; Vasilevskaya, A.; Anastassiadis, C.; Tang-Wai, D.F.; Golas, A.C.; Tartaglia, M.C. Investigating the association between a history of depression and biomarkers of Alzheimer’s disease, cerebrovascular disease, and neurodegeneration in patients with dementia. Geroscience 2024, 46, 783–793. [Google Scholar] [CrossRef]
  127. Nyul-Toth, A.; Patai, R.; Csiszar, A.; Ungvari, A.; Gulej, R.; Mukli, P.; Yabluchanskiy, A.; Benyo, Z.; Sotonyi, P.; Prodan, C.I.; et al. Linking peripheral atherosclerosis to blood-brain barrier disruption: Elucidating its role as a manifestation of cerebral small vessel disease in vascular cognitive impairment. Geroscience 2024, 46, 6511–6536. [Google Scholar] [CrossRef]
  128. Gulej, R.; Nyul-Toth, A.; Csik, B.; Patai, R.; Petersen, B.; Negri, S.; Chandragiri, S.S.; Shanmugarama, S.; Mukli, P.; Yabluchanskiy, A.; et al. Young blood-mediated cerebromicrovascular rejuvenation through heterochronic parabiosis: Enhancing blood-brain barrier integrity and capillarization in the aged mouse brain. Geroscience 2024, 46, 4415–4442. [Google Scholar] [CrossRef] [PubMed]
  129. Csiszar, A.; Ungvari, A.; Patai, R.; Gulej, R.; Yabluchanskiy, A.; Benyo, Z.; Kovacs, I.; Sotonyi, P.; Kirkpartrick, A.C.; Prodan, C.I.; et al. Atherosclerotic burden and cerebral small vessel disease: Exploring the link through microvascular aging and cerebral microhemorrhages. Geroscience 2024, 46, 5103–5132. [Google Scholar] [CrossRef]
  130. Weijs, R.W.J.; Oudegeest-Sander, M.H.; Vloet, J.I.A.; Hopman, M.T.E.; Claassen, J.; Thijssen, D.H.J. A decade of aging in healthy older adults: Longitudinal findings on cerebrovascular and cognitive health. Geroscience 2023, 45, 2629–2641. [Google Scholar] [CrossRef] [PubMed]
  131. Gulej, R.; Nyul-Toth, A.; Ahire, C.; DelFavero, J.; Balasubramanian, P.; Kiss, T.; Tarantini, S.; Benyo, Z.; Pacher, P.; Csik, B.; et al. Elimination of senescent cells by treatment with Navitoclax/ABT263 reverses whole brain irradiation-induced blood-brain barrier disruption in the mouse brain. Geroscience 2023, 45, 2983–3002. [Google Scholar] [CrossRef]
  132. Zhang, H.; Roman, R.J.; Fan, F. Hippocampus is more susceptible to hypoxic injury: Has the Rosetta Stone of regional variation in neurovascular coupling been deciphered? Geroscience 2022, 44, 127–130. [Google Scholar] [CrossRef] [PubMed]
  133. Nyul-Toth, A.; Fulop, G.A.; Tarantini, S.; Kiss, T.; Ahire, C.; Faakye, J.A.; Ungvari, A.; Toth, P.; Toth, A.; Csiszar, A.; et al. Cerebral venous congestion exacerbates cerebral microhemorrhages in mice. Geroscience 2022, 44, 805–816. [Google Scholar] [CrossRef]
  134. Sabayan, B.; Westendorp, R.G.J. Neurovascular-glymphatic dysfunction and white matter lesions. Geroscience 2021, 43, 1635–1642. [Google Scholar] [CrossRef]
  135. Wang, S.; Lv, W.; Zhang, H.; Liu, Y.; Li, L.; Jefferson, J.R.; Guo, Y.; Li, M.; Gao, W.; Fang, X.; et al. Aging exacerbates impairments of cerebral blood flow autoregulation and cognition in diabetic rats. Geroscience 2020, 42, 1387–1410. [Google Scholar] [CrossRef]
  136. Verheggen, I.C.M.; de Jong, J.J.A.; van Boxtel, M.P.J.; Postma, A.A.; Jansen, J.F.A.; Verhey, F.R.J.; Backes, W.H. Imaging the role of blood-brain barrier disruption in normal cognitive ageing. Geroscience 2020, 42, 1751–1764. [Google Scholar] [CrossRef]
  137. Sorond, F.A.; Whitehead, S.; Arai, K.; Arnold, D.; Carmichael, S.T.; De Carli, C.; Duering, M.; Fornage, M.; Flores-Obando, R.E.; Graff-Radford, J.; et al. Proceedings from the Albert Charitable Trust Inaugural Workshop on white matter and cognition in aging. Geroscience 2020, 42, 81–96. [Google Scholar] [CrossRef]
  138. Levit, A.; Hachinski, V.; Whitehead, S.N. Neurovascular unit dysregulation, white matter disease, and executive dysfunction: The shared triad of vascular cognitive impairment and Alzheimer disease. Geroscience 2020, 42, 445–465. [Google Scholar] [CrossRef] [PubMed]
  139. Jor’dan, A.J.; Manor, B.; Iloputaife, I.; Habtemariam, D.A.; Bean, J.F.; Sorond, F.A.; Lipsitz, L.A. Diminished Locomotor Control Is Associated With Reduced Neurovascular Coupling in Older Adults. J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, 1516–1522. [Google Scholar] [CrossRef] [PubMed]
  140. Cermakova, P.; Ding, J.; Meirelles, O.; Reis, J.; Religa, D.; Schreiner, P.J.; Jacobs, D.R.; Bryan, R.N.; Launer, L.J. Carotid Intima-Media Thickness and Markers of Brain Health in a Biracial Middle-Aged Cohort: CARDIA Brain MRI Sub-study. J. Gerontol. A Biol. Sci. Med. Sci. 2020, 75, 380–386. [Google Scholar] [CrossRef] [PubMed]
  141. Gulej, R.; Nyul-Toth, A.; Csik, B.; Petersen, B.; Faakye, J.; Negri, S.; Chandragiri, S.S.; Mukli, P.; Yabluchanskiy, A.; Conley, S.; et al. Rejuvenation of cerebromicrovascular function in aged mice through heterochronic parabiosis: Insights into neurovascular coupling and the impact of young blood factors. Geroscience 2024, 46, 327–347. [Google Scholar] [CrossRef]
  142. Li, B.; Yabluchanskiy, A.; Tarantini, S.; Allu, S.R.; Sencan-Egilmez, I.; Leng, J.; Alfadhel, M.A.H.; Porter, J.E.; Fu, B.; Ran, C.; et al. Measurements of cerebral microvascular blood flow, oxygenation, and morphology in a mouse model of whole-brain irradiation-induced cognitive impairment by two-photon microscopy and optical coherence tomography: Evidence for microvascular injury in the cerebral white matter. Geroscience 2023, 45, 1491–1510. [Google Scholar]
  143. Fang, X.; Border, J.J.; Rivers, P.L.; Zhang, H.; Williams, J.M.; Fan, F.; Roman, R.J. Amyloid beta accumulation in TgF344-AD rats is associated with reduced cerebral capillary endothelial Kir2.1 expression and neurovascular uncoupling. Geroscience 2023, 45, 2909–2926. [Google Scholar] [CrossRef]
  144. Vestergaard, M.B.; Lindberg, U.; Knudsen, M.H.; Urdanibia-Centelles, O.; Bakhtiari, A.; Mortensen, E.L.; Osler, M.; Fagerlund, B.; Benedek, K.; Lauritzen, M.; et al. Subclinical cognitive deficits are associated with reduced cerebrovascular response to visual stimulation in mid-sixties men. Geroscience 2022, 44, 1905–1923. [Google Scholar] [CrossRef]
  145. Toth, L.; Czigler, A.; Hegedus, E.; Komaromy, H.; Amrein, K.; Czeiter, E.; Yabluchanskiy, A.; Koller, A.; Orsi, G.; Perlaki, G.; et al. Age-related decline in circulating IGF-1 associates with impaired neurovascular coupling responses in older adults. Geroscience 2022, 44, 2771–2783. [Google Scholar] [CrossRef]
  146. Montagne, A.; Barnes, S.R.; Nation, D.A.; Kisler, K.; Toga, A.W.; Zlokovic, B.V. Imaging subtle leaks in the blood-brain barrier in the aging human brain: Potential pitfalls, challenges, and possible solutions. Geroscience 2022, 44, 1339–1351. [Google Scholar] [CrossRef]
  147. Tarantini, S.; Nyul-Toth, A.; Yabluchanskiy, A.; Csipo, T.; Mukli, P.; Balasubramanian, P.; Ungvari, A.; Toth, P.; Benyo, Z.; Sonntag, W.E.; et al. Endothelial deficiency of insulin-like growth factor-1 receptor (IGF1R) impairs neurovascular coupling responses in mice, mimicking aspects of the brain aging phenotype. Geroscience 2021, 43, 2387–2394. [Google Scholar] [CrossRef]
  148. Istvan, L.; Czako, C.; Elo, A.; Mihaly, Z.; Sotonyi, P.; Varga, A.; Ungvari, Z.; Csiszar, A.; Yabluchanskiy, A.; Conley, S.; et al. Imaging retinal microvascular manifestations of carotid artery disease in older adults: From diagnosis of ocular complications to understanding microvascular contributions to cognitive impairment. Geroscience 2021, 43, 1703–1723. [Google Scholar] [CrossRef]
  149. Fan, F.; Roman, R.J. Reversal of cerebral hypoperfusion: A novel therapeutic target for the treatment of AD/ADRD? Geroscience 2021, 43, 1065–1067. [Google Scholar] [CrossRef] [PubMed]
  150. Kiss, T.; Nyul-Toth, A.; Balasubramanian, P.; Tarantini, S.; Ahire, C.; Yabluchanskiy, A.; Csipo, T.; Farkas, E.; Wren, J.D.; Garman, L.; et al. Nicotinamide mononucleotide (NMN) supplementation promotes neurovascular rejuvenation in aged mice: Transcriptional footprint of SIRT1 activation, mitochondrial protection, anti-inflammatory, and anti-apoptotic effects. Geroscience 2020, 42, 527–546. [Google Scholar] [CrossRef]
  151. Cao, Y.; Xu, W.; Liu, Q. Alterations of the blood-brain barrier during aging. J. Cereb. Blood Flow. Metab. 2024, 44, 881–895. [Google Scholar] [CrossRef] [PubMed]
  152. Moon, Y.; Jeon, H.J.; Han, S.H.; Min-Young, N.; Kim, H.J.; Kwon, K.J.; Moon, W.J.; Kim, S.H. Blood-brain barrier breakdown is linked to tau pathology and neuronal injury in a differential manner according to amyloid deposition. J. Cereb. Blood Flow. Metab. 2023, 43, 1813–1825. [Google Scholar] [CrossRef] [PubMed]
  153. Wang, Y.; Du, W.; Sun, Y.; Zhang, J.; Ma, C.; Jin, X. CRTC1 is a potential target to delay aging-induced cognitive deficit by protecting the integrity of the blood-brain barrier via inhibiting inflammation. J. Cereb. Blood Flow. Metab. 2023, 43, 1042–1059. [Google Scholar] [CrossRef]
  154. Preininger, M.K.; Zaytseva, D.; Lin, J.M.; Kaufer, D. Blood-brain barrier dysfunction promotes astrocyte senescence through albumin-induced TGFbeta signaling activation. Aging Cell 2023, 22, e13747. [Google Scholar] [CrossRef]
  155. Udo, M.S.B.; Zaccarelli-Magalhaes, J.; Clemons, G.A.; Citadin, C.T.; Langman, J.; Smith, D.J.; Matuguma, L.H.; Tesic, V.; Lin, H.W. Blockade of A(2A)R improved brain perfusion and cognitive function in a mouse model of Alzheimer’s disease. Geroscience 2025, 47, 4153–4167. [Google Scholar] [CrossRef]
  156. Patai, R.; Csik, B.; Nyul-Toth, A.; Gulej, R.; Vali Kordestan, K.; Chandragiri, S.S.; Shanmugarama, S.; Tarantini, S.; Mukli, P.; Ungvari, A.; et al. Persisting blood-brain barrier disruption following cisplatin treatment in a mouse model of chemotherapy-associated cognitive impairment. Geroscience 2025, 47, 3835–3847. [Google Scholar] [CrossRef]
  157. Cummins, M.J.; Cresswell, E.T.; Bevege, R.J.; Smith, D.W. Aging disrupts blood-brain and blood-spinal cord barrier homeostasis, but does not increase paracellular permeability. Geroscience 2024, 47, 263–285. [Google Scholar] [CrossRef] [PubMed]
  158. Chandra, P.K.; Panner Selvam, M.K.; Castorena-Gonzalez, J.A.; Rutkai, I.; Sikka, S.C.; Mostany, R.; Busija, D.W. Fibrinogen in mice cerebral microvessels induces blood-brain barrier dysregulation with aging via a dynamin-related protein 1-dependent pathway. Geroscience 2024, 46, 395–415. [Google Scholar] [CrossRef] [PubMed]
  159. Ayyanar, M.P.; Vijayan, M. A review on gut microbiota and miRNA crosstalk: Implications for Alzheimer’s disease. Geroscience 2024, 47, 339–385. [Google Scholar] [CrossRef]
  160. Ting, K.K.; Coleman, P.; Kim, H.J.; Zhao, Y.; Mulangala, J.; Cheng, N.C.; Li, W.; Gunatilake, D.; Johnstone, D.M.; Loo, L.; et al. Vascular senescence and leak are features of the early breakdown of the blood-brain barrier in Alzheimer’s disease models. Geroscience 2023, 45, 3307–3331. [Google Scholar] [CrossRef] [PubMed]
  161. Ruggiero, A.D.; Vemuri, R.; Blawas, M.; Long, M.; DeStephanis, D.; Williams, A.G.; Chen, H.; Justice, J.N.; Macauley, S.L.; Day, S.M.; et al. Long-term dasatinib plus quercetin effects on aging outcomes and inflammation in nonhuman primates: Implications for senolytic clinical trial design. Geroscience 2023, 45, 2785–2803. [Google Scholar] [CrossRef]
  162. Pandics, T.; Major, D.; Fazekas-Pongor, V.; Szarvas, Z.; Peterfi, A.; Mukli, P.; Gulej, R.; Ungvari, A.; Fekete, M.; Tompa, A.; et al. Exposome and unhealthy aging: Environmental drivers from air pollution to occupational exposures. Geroscience 2023, 45, 3381–3408. [Google Scholar] [CrossRef]
  163. Bakhtiari, A.; Vestergaard, M.B.; Benedek, K.; Fagerlund, B.; Mortensen, E.L.; Osler, M.; Lauritzen, M.; Larsson, H.B.W.; Lindberg, U. Changes in hippocampal volume during a preceding 10-year period do not correlate with cognitive performance and hippocampal blood–brain barrier permeability in cognitively normal late-middle-aged men. Geroscience 2023, 45, 1161–1175. [Google Scholar] [CrossRef]
  164. Towner, R.A.; Gulej, R.; Zalles, M.; Saunders, D.; Smith, N.; Lerner, M.; Morton, K.A.; Richardson, A. Rapamycin restores brain vasculature, metabolism, and blood-brain barrier in an inflammaging model. Geroscience 2021, 43, 563–578. [Google Scholar] [CrossRef]
  165. Kerkhofs, D.; Wong, S.M.; Zhang, E.; Uiterwijk, R.; Hoff, E.I.; Jansen, J.F.A.; Staals, J.; Backes, W.H.; van Oostenbrugge, R.J. Blood-brain barrier leakage at baseline and cognitive decline in cerebral small vessel disease: A 2-year follow-up study. Geroscience 2021, 43, 1643–1652. [Google Scholar] [CrossRef] [PubMed]
  166. Verheggen, I.C.M.; de Jong, J.J.A.; van Boxtel, M.P.J.; Gronenschild, E.; Palm, W.M.; Postma, A.A.; Jansen, J.F.A.; Verhey, F.R.J.; Backes, W.H. Increase in blood-brain barrier leakage in healthy, older adults. Geroscience 2020, 42, 1183–1193. [Google Scholar] [CrossRef]
  167. Islam, M.T.; Hall, S.A.; Dutson, T.; Bloom, S.I.; Bramwell, R.C.; Kim, J.; Tucker, J.R.; Machin, D.R.; Donato, A.J.; Lesniewski, L.A. Endothelial cell-specific reduction in mTOR ameliorates age-related arterial and metabolic dysfunction. Aging Cell 2024, 23, e14040. [Google Scholar] [CrossRef]
  168. Kiss, T.; Nyul-Toth, A.; Balasubramanian, P.; Tarantini, S.; Ahire, C.; DelFavero, J.; Yabluchanskiy, A.; Csipo, T.; Farkas, E.; Wiley, G.; et al. Single-cell RNA sequencing identifies senescent cerebromicrovascular endothelial cells in the aged mouse brain. Geroscience 2020, 42, 429–444. [Google Scholar] [CrossRef]
  169. Tarantini, S.; Balasubramanian, P.; Delfavero, J.; Csipo, T.; Yabluchanskiy, A.; Kiss, T.; Nyul-Toth, A.; Mukli, P.; Toth, P.; Ahire, C.; et al. Treatment with the BCL-2/BCL-xL inhibitor senolytic drug ABT263/Navitoclax improves functional hyperemia in aged mice. Geroscience 2021, 43, 2427–2440. [Google Scholar] [CrossRef]
  170. Faakye, J.; Nyul-Toth, A.; Muranyi, M.; Gulej, R.; Csik, B.; Shanmugarama, S.; Tarantini, S.; Negri, S.; Prodan, C.; Mukli, P.; et al. Preventing spontaneous cerebral microhemorrhages in aging mice: A novel approach targeting cellular senescence with ABT263/navitoclax. Geroscience 2024, 46, 21–37. [Google Scholar] [CrossRef]
  171. Cummings, S.R.; Lui, L.Y.; Zaira, A.; Mau, T.; Fielding, R.A.; Atkinson, E.J.; Patel, S.; LeBrasseur, N. Biomarkers of cellular senescence and major health outcomes in older adults. Geroscience 2024, 47, 3407–3415. [Google Scholar] [CrossRef]
  172. Mahoney, S.A.; Venkatasubramanian, R.; Darrah, M.A.; Ludwig, K.R.; VanDongen, N.S.; Greenberg, N.T.; Longtine, A.G.; Hutton, D.A.; Brunt, V.E.; Campisi, J.; et al. Intermittent supplementation with fisetin improves arterial function in old mice by decreasing cellular senescence. Aging Cell 2024, 23, e14060. [Google Scholar] [CrossRef] [PubMed]
  173. Novo, J.P.; Gee, L.; Caetano, C.A.; Tome, I.; Vilaca, A.; von Zglinicki, T.; Moreira, I.S.; Jurk, D.; Rosa, S.; Ferreira, L. Blood-brain barrier dysfunction in aging is mediated by brain endothelial senescence. Aging Cell 2024, 23, e14270. [Google Scholar] [CrossRef] [PubMed]
  174. Budamagunta, V.; Kumar, A.; Rani, A.; Bean, L.; Manohar-Sindhu, S.; Yang, Y.; Zhou, D.; Foster, T.C. Effect of peripheral cellular senescence on brain aging and cognitive decline. Aging Cell 2023, 22, e13817. [Google Scholar] [CrossRef] [PubMed]
  175. Ahire, C.; Nyul-Toth, A.; DelFavero, J.; Gulej, R.; Faakye, J.A.; Tarantini, S.; Kiss, T.; Kuan-Celarier, A.; Balasubramanian, P.; Ungvari, A.; et al. Accelerated cerebromicrovascular senescence contributes to cognitive decline in a mouse model of paclitaxel (Taxol)-induced chemobrain. Aging Cell 2023, 22, e13832. [Google Scholar] [CrossRef]
  176. Bloom, S.I.; Liu, Y.; Tucker, J.R.; Islam, M.T.; Machin, D.R.; Abdeahad, H.; Thomas, T.G.; Bramwell, R.C.; Lesniewski, L.A.; Donato, A.J. Endothelial cell telomere dysfunction induces senescence and results in vascular and metabolic impairments. Aging Cell 2023, 22, e13875. [Google Scholar] [CrossRef]
  177. Nyul-Toth, A.; Negri, S.; Sanford, M.; Jiang, R.; Patai, R.; Budda, M.; Petersen, B.; Pinckard, J.; Chandragiri, S.S.; Shi, H.; et al. Novel intravital approaches to quantify deep vascular structure and perfusion in the aging mouse brain using ultrasound localization microscopy (ULM). J. Cereb. Blood Flow. Metab. 2024, 44, 1378–1396. [Google Scholar] [CrossRef]
  178. Andonian, B.J.; Hippensteel, J.A.; Abuabara, K.; Boyle, E.M.; Colbert, J.F.; Devinney, M.J.; Faye, A.S.; Kochar, B.; Lee, J.; Litke, R.; et al. Inflammation and aging-related disease: A transdisciplinary inflammaging framework. Geroscience 2024, 47, 515–542. [Google Scholar] [CrossRef]
  179. Gotz, L.; Rueckschloss, U.; Reimer, A.; Bommel, H.; Beilhack, A.; Ergun, S.; Kleefeldt, F. Vascular inflammaging: Endothelial CEACAM1 expression is upregulated by TNF-alpha via independent activation of NF-kappaB and beta-catenin signaling. Aging Cell 2025, 24, e14384. [Google Scholar] [CrossRef] [PubMed]
  180. Shirzadi, Z.; Boyle, R.; Yau, W.W.; Coughlan, G.; Fu, J.F.; Properzi, M.J.; Buckley, R.F.; Yang, H.S.; Scanlon, C.E.; Hsieh, S.; et al. Vascular contributions to cognitive decline: Beyond amyloid and tau in the Harvard Aging Brain Study. J. Cereb. Blood Flow. Metab. 2024, 44, 1319–1328. [Google Scholar] [CrossRef]
  181. van Dijk, S.E.; Drenth, N.; Hafkemeijer, A.; Labadie, G.; Witjes-Ane, M.W.; Blauw, G.J.; Rombouts, S.A.; van der Grond, J.; van Rooden, S. Neurovascular coupling in early stage dementia—A case-control study. J. Cereb. Blood Flow. Metab. 2024, 44, 1013–1023. [Google Scholar] [CrossRef]
  182. van der Horn, H.J.; Vakhtin, A.A.; Julio, K.; Nitschke, S.; Shaff, N.; Dodd, A.B.; Erhardt, E.; Phillips, J.P.; Pirio Richardson, S.; Deligtisch, A.; et al. Parkinson’s disease cerebrovascular reactivity pattern: A feasibility study. J. Cereb. Blood Flow. Metab. 2024, 44, 1774–1786. [Google Scholar] [CrossRef]
  183. Jones, O.A.; Mohamed, S.; Hinz, R.; Paterson, A.; Sobowale, O.A.; Dickie, B.R.; Parkes, L.M.; Parry-Jones, A.R. Neuroinflammation and blood-brain barrier breakdown in acute, clinical intracerebral hemorrhage. J. Cereb. Blood Flow. Metab. 2025, 45, 233–243. [Google Scholar] [CrossRef] [PubMed]
  184. Zhou, S.; Liu, C.; Wang, J.; Ye, J.; Lian, Q.; Gan, L.; Deng, S.; Xu, T.; Guo, Y.; Li, W.; et al. CCL5 mediated astrocyte-T cell interaction disrupts blood-brain barrier in mice after hemorrhagic stroke. J. Cereb. Blood Flow. Metab. 2024, 44, 367–383. [Google Scholar] [CrossRef]
  185. Zhang, S.; Zhao, J.; Sha, W.M.; Zhang, X.P.; Mai, J.Y.; Bartlett, P.F.; Hou, S.T. Inhibition of EphA4 reduces vasogenic edema after experimental stroke in mice by protecting the blood-brain barrier integrity. J. Cereb. Blood Flow. Metab. 2024, 44, 419–433. [Google Scholar] [CrossRef]
  186. Wu, S.; Ren, X.S.; Shi, Y. Early and enduring: Targeting the endothelium for blood-brain barrier protection. J. Cereb. Blood Flow. Metab. 2024, 44, 1674–1676. [Google Scholar] [CrossRef] [PubMed]
  187. Mandeville, E.T.; Buchan, A.M.; Tiedt, S. Not open and shut: Complex and prolonged blood-brain barrier responses after stroke. J. Cereb. Blood Flow. Metab. 2024, 44, 446–448. [Google Scholar] [CrossRef] [PubMed]
  188. Liu, C.; Guo, Y.; Deng, S.; Zhou, S.; Wu, S.; Chen, T.; Shi, X.; Mamtilahun, M.; Xu, T.; Liu, Z.; et al. Hemorrhagic stroke-induced subtype of inflammatory reactive astrocytes disrupts blood-brain barrier. J. Cereb. Blood Flow. Metab. 2024, 44, 1102–1116. [Google Scholar] [CrossRef] [PubMed]
  189. Roth, M.; Carlsson, R.; Buizza, C.; Enstrom, A.; Paul, G. Pericyte response to ischemic stroke precedes endothelial cell death and blood-brain barrier breakdown. J. Cereb. Blood Flow. Metab. 2025, 45, 617–629. [Google Scholar] [CrossRef]
  190. Tudor, T.; Spinazzi, E.F.; Alexander, J.E.; Mandigo, G.K.; Lavine, S.D.; Grinband, J.; Connolly, E.S., Jr. Progressive microvascular failure in acute ischemic stroke: A systematic review, meta-analysis, and time-course analysis. J. Cereb. Blood Flow. Metab. 2024, 44, 192–208. [Google Scholar] [CrossRef]
  191. Schmitzer, L.; Kaczmarz, S.; Gottler, J.; Hoffmann, G.; Kallmayer, M.; Eckstein, H.H.; Hedderich, D.M.; Kufer, J.; Zimmer, C.; Preibisch, C.; et al. Macro- and microvascular contributions to cerebral structural alterations in patients with asymptomatic carotid artery stenosis. J. Cereb. Blood Flow. Metab. 2024, 44, 1629–1642. [Google Scholar] [CrossRef]
  192. Li, S.; Tian, X.; Ip, B.; Feng, X.; Ip, H.L.; Abrigo, J.; Lan, L.; Liu, H.; Zheng, L.; Liu, Y.; et al. Cerebral hemodynamics and stroke risks in symptomatic intracranial atherosclerotic stenosis with internal versus cortical borderzone infarcts: A computational fluid dynamics study. J. Cereb. Blood Flow. Metab. 2024, 44, 516–526. [Google Scholar] [CrossRef]
  193. Bagi, Z.; Kroenke, C.D.; Fopiano, K.A.; Tian, Y.; Filosa, J.A.; Sherman, L.S.; Larson, E.B.; Keene, C.D.; Degener O’Brien, K.; Adeniyi, P.A.; et al. Association of cerebral microvascular dysfunction and white matter injury in Alzheimer’s disease. Geroscience 2022, 44, 1–14. [Google Scholar] [CrossRef]
  194. Czakó, C.; Kovács, T.; Ungvari, Z.; Csiszar, A.; Yabluchanskiy, A.; Conley, S.; Csipo, T.; Lipecz, A.; Horváth, H.; Sándor, G.L.; et al. Retinal biomarkers for Alzheimer’s disease and vascular cognitive impairment and dementia (VCID): Implication for early diagnosis and prognosis. Geroscience 2020, 42, 1499–1525. [Google Scholar] [CrossRef]
  195. Fekete, M.; Lehoczki, A.; Szappanos, A.; Toth, A.; Mahdi, M.; Sotonyi, P.; Benyo, Z.; Yabluchanskiy, A.; Tarantini, S.; Ungvari, Z. Cerebromicrovascular mechanisms contributing to long COVID: Implications for neurocognitive health. Geroscience 2025, 47, 745–779. [Google Scholar] [CrossRef]
  196. Elmansi, A.M.; Kassem, A.; Castilla, R.M.; Miller, R.A. Downregulation of the NF-kappaB protein p65 is a shared phenotype among most anti-aging interventions. Geroscience 2024, 47, 3077–3094. [Google Scholar] [CrossRef]
  197. Riordan, R.; Rong, W.; Yu, Z.; Ross, G.; Valerio, J.; Dimas-Munoz, J.; Heredia, V.; Magnusson, K.; Galvan, V.; Perez, V.I. Effect of Nrf2 loss on senescence and cognition of tau-based P301S mice. Geroscience 2023, 45, 1451–1469. [Google Scholar] [CrossRef]
  198. Shin, J.W.; Chun, K.S.; Kim, D.H.; Kim, S.J.; Kim, S.H.; Cho, N.C.; Na, H.K.; Surh, Y.J. Curcumin induces stabilization of Nrf2 protein through Keap1 cysteine modification. Biochem. Pharmacol. 2020, 173, 113820. [Google Scholar] [CrossRef]
  199. Yang, C.; Zhang, X.; Fan, H.; Liu, Y. Curcumin upregulates transcription factor Nrf2, HO-1 expression and protects rat brains against focal ischemia. Brain Res. 2009, 1282, 133–141. [Google Scholar] [CrossRef] [PubMed]
  200. Khor, T.O.; Huang, Y.; Wu, T.Y.; Shu, L.; Lee, J.; Kong, A.N. Pharmacodynamics of curcumin as DNA hypomethylation agent in restoring the expression of Nrf2 via promoter CpGs demethylation. Biochem. Pharmacol. 2011, 82, 1073–1078. [Google Scholar] [CrossRef] [PubMed]
  201. Kobutree, P.; Tothonglor, A.; Roumwong, A.; Jindatip, D.; Agthong, S. Curcumin reduces blood-nerve barrier abnormalities and cytotoxicity to endothelial cells and pericytes induced by cisplatin. Folia Morphol. 2023, 82, 533–542. [Google Scholar] [CrossRef]
  202. Moore, T.L.; Bowley, B.; Shultz, P.; Calderazzo, S.; Shobin, E.; Killiany, R.J.; Rosene, D.L.; Moss, M.B. Chronic curcumin treatment improves spatial working memory but not recognition memory in middle-aged rhesus monkeys. Geroscience 2017, 39, 571–584. [Google Scholar] [CrossRef]
  203. Kaufmann, F.N.; Gazal, M.; Bastos, C.R.; Kaster, M.P.; Ghisleni, G. Curcumin in depressive disorders: An overview of potential mechanisms, preclinical and clinical findings. Eur. J. Pharmacol. 2016, 784, 192–198. [Google Scholar] [CrossRef]
  204. Basu, P.; Maier, C.; Basu, A. Effects of Curcumin and Its Different Formulations in Preclinical and Clinical Studies of Peripheral Neuropathic and Postoperative Pain: A Comprehensive Review. Int. J. Mol. Sci. 2021, 22, 4666. [Google Scholar] [CrossRef]
  205. Marques, M.; Marinho, M.; Vian, C.; Horn, A. The action of curcumin against damage resulting from cerebral stroke: A systematic review. Pharmacol. Res. 2022, 183, 106369. [Google Scholar] [CrossRef] [PubMed]
  206. Lin, L.; Li, C.; Zhang, D.; Yuan, M.; Chen, C.-h.; Li, M. Synergic effects of berberine and curcumin on improving cognitive function in an Alzheimer’s disease mouse model. Neurochem. Res. 2020, 45, 1130–1141. [Google Scholar] [CrossRef] [PubMed]
  207. Su, I.J.; Chang, H.Y.; Wang, H.C.; Tsai, K.J. A Curcumin Analog Exhibits Multiple Biologic Effects on the Pathogenesis of Alzheimer’s Disease and Improves Behavior, Inflammation, and β-Amyloid Accumulation in a Mouse Model. Int. J. Mol. Sci. 2020, 21, 5459. [Google Scholar] [CrossRef] [PubMed]
  208. ELBini-Dhouib, I.; Doghri, R.; Ellefi, A.; Degrach, I.; Srairi-Abid, N.; Gati, A. Curcumin Attenuated Neurotoxicity in Sporadic Animal Model of Alzheimer’s Disease. Molecules 2021, 26, 3001. [Google Scholar] [CrossRef]
  209. Reddy, P.H.; Manczak, M.; Yin, X.; Grady, M.C.; Mitchell, A.; Kandimalla, R.; Kuruva, C.S. Protective effects of a natural product, curcumin, against amyloid β induced mitochondrial and synaptic toxicities in Alzheimer’s disease. J. Investig. Med. 2016, 64, 1220–1234. [Google Scholar] [CrossRef] [PubMed]
  210. Wang, Y.; Yin, H.; Wang, L.; Shuboy, A.; Lou, J.; Han, B.; Zhang, X.; Li, J. Curcumin as a potential treatment for Alzheimer’s disease: A study of the effects of curcumin on hippocampal expression of glial fibrillary acidic protein. Am. J. Chin. Med. 2013, 41, 59–70. [Google Scholar] [CrossRef] [PubMed]
  211. Cheng, K.K.; Yeung, C.F.; Ho, S.W.; Chow, S.F.; Chow, A.H.; Baum, L. Highly stabilized curcumin nanoparticles tested in an in vitro blood-brain barrier model and in Alzheimer’s disease Tg2576 mice. Aaps J. 2013, 15, 324–336. [Google Scholar] [CrossRef]
  212. Ray, B.; Bisht, S.; Maitra, A.; Maitra, A.; Lahiri, D.K. Neuroprotective and neurorescue effects of a novel polymeric nanoparticle formulation of curcumin (NanoCurc™) in the neuronal cell culture and animal model: Implications for Alzheimer’s disease. J. Alzheimers Dis. 2011, 23, 61–77. [Google Scholar] [CrossRef]
  213. Begum, A.N.; Jones, M.R.; Lim, G.P.; Morihara, T.; Kim, P.; Heath, D.D.; Rock, C.L.; Pruitt, M.A.; Yang, F.; Hudspeth, B.; et al. Curcumin structure-function, bioavailability, and efficacy in models of neuroinflammation and Alzheimer’s disease. J. Pharmacol. Exp. Ther. 2008, 326, 196–208. [Google Scholar] [CrossRef]
  214. Xiong, Z.; Hongmei, Z.; Lu, S.; Yu, L. Curcumin mediates presenilin-1 activity to reduce β-amyloid production in a model of Alzheimer’s Disease. Pharmacol. Rep. 2011, 63, 1101–1108. [Google Scholar] [CrossRef]
  215. Ishrat, T.; Hoda, M.N.; Khan, M.B.; Yousuf, S.; Ahmad, M.; Khan, M.M.; Ahmad, A.; Islam, F. Amelioration of cognitive deficits and neurodegeneration by curcumin in rat model of sporadic dementia of Alzheimer’s type (SDAT). Eur. Neuropsychopharmacol. 2009, 19, 636–647. [Google Scholar] [CrossRef]
  216. Nam, S.M.; Choi, J.H.; Yoo, D.Y.; Kim, W.; Jung, H.Y.; Kim, J.W.; Yoo, M.; Lee, S.; Kim, C.J.; Yoon, Y.S.; et al. Effects of curcumin (Curcuma longa) on learning and spatial memory as well as cell proliferation and neuroblast differentiation in adult and aged mice by upregulating brain-derived neurotrophic factor and CREB signaling. J. Med. Food 2014, 17, 641–649. [Google Scholar] [CrossRef]
  217. Bassani, T.B.; Turnes, J.M.; Moura, E.L.R.; Bonato, J.M.; Cóppola-Segovia, V.; Zanata, S.M.; Oliveira, R.; Vital, M. Effects of curcumin on short-term spatial and recognition memory, adult neurogenesis and neuroinflammation in a streptozotocin-induced rat model of dementia of Alzheimer’s type. Behav. Brain Res. 2017, 335, 41–54. [Google Scholar] [CrossRef]
  218. McClure, R.; Ong, H.; Janve, V.; Barton, S.; Zhu, M.; Li, B.; Dawes, M.; Jerome, W.G.; Anderson, A.; Massion, P. Aerosol delivery of curcumin reduced amyloid-β deposition and improved cognitive performance in a transgenic model of Alzheimer’s disease. J. Alzheimer’s Dis. 2016, 55, 797–811. [Google Scholar] [CrossRef] [PubMed]
  219. Alamro, A.A.; Alsulami, E.A.; Almutlaq, M.; Alghamedi, A.; Alokail, M.; Haq, S.H. Therapeutic Potential of Vitamin D and Curcumin in an In Vitro Model of Alzheimer Disease. J. Cent. Nerv. Syst. Dis. 2020, 12, 1179573520924311. [Google Scholar] [CrossRef] [PubMed]
  220. Lim, G.P.; Chu, T.; Yang, F.; Beech, W.; Frautschy, S.A.; Cole, G.M. The curry spice curcumin reduces oxidative damage and amyloid pathology in an Alzheimer transgenic mouse. J. Neurosci. 2001, 21, 8370–8377. [Google Scholar] [CrossRef]
  221. Savall, A.S.P.; de Mello, J.D.; Fidelis, E.M.; Comis-Neto, A.A.; Nepomuceno, M.R.; Pacheco, C.d.O.; Haas, S.E.; Pinton, S. Nanoencapsulated curcumin: Enhanced efficacy in reversing memory loss in an Alzheimer disease model. Brain Sci. 2024, 14, 130. [Google Scholar] [CrossRef]
  222. Abbaoui, A.; Chatoui, H.; El Hiba, O.; Gamrani, H. Neuroprotective effect of curcumin-I in copper-induced dopaminergic neurotoxicity in rats: A possible link with Parkinson’s disease. Neurosci. Lett. 2017, 660, 103–108. [Google Scholar] [CrossRef]
  223. Rajeswari, A.; Sabesan, M. Inhibition of monoamine oxidase-B by the polyphenolic compound, curcumin and its metabolite tetrahydrocurcumin, in a model of Parkinson’s disease induced by MPTP neurodegeneration in mice. Inflammopharmacology 2008, 16, 96–99. [Google Scholar] [CrossRef]
  224. Pan, J.; Li, H.; Ma, J.F.; Tan, Y.Y.; Xiao, Q.; Ding, J.Q.; Chen, S.D. Curcumin inhibition of JNKs prevents dopaminergic neuronal loss in a mouse model of Parkinson’s disease through suppressing mitochondria dysfunction. Transl. Neurodegener. 2012, 1, 16. [Google Scholar] [CrossRef] [PubMed]
  225. Jiang, T.F.; Zhang, Y.J.; Zhou, H.Y.; Wang, H.M.; Tian, L.P.; Liu, J.; Ding, J.Q.; Chen, S.D. Curcumin ameliorates the neurodegenerative pathology in A53T α-synuclein cell model of Parkinson’s disease through the downregulation of mTOR/p70S6K signaling and the recovery of macroautophagy. J. Neuroimmune Pharmacol. 2013, 8, 356–369. [Google Scholar] [CrossRef] [PubMed]
  226. Wang, M.S.; Boddapati, S.; Emadi, S.; Sierks, M.R. Curcumin reduces alpha-synuclein induced cytotoxicity in Parkinson’s disease cell model. BMC Neurosci. 2010, 11, 57. [Google Scholar] [CrossRef]
  227. Ramires Júnior, O.V.; Alves, B.D.S.; Barros, P.A.B.; Rodrigues, J.L.; Ferreira, S.P.; Monteiro, L.K.S.; Araújo, G.M.S.; Fernandes, S.S.; Vaz, G.R.; Dora, C.L.; et al. Nanoemulsion Improves the Neuroprotective Effects of Curcumin in an Experimental Model of Parkinson’s Disease. Neurotox. Res. 2021, 39, 787–799. [Google Scholar] [CrossRef] [PubMed]
  228. Yang, J.; Song, S.; Li, J.; Liang, T. Neuroprotective effect of curcumin on hippocampal injury in 6-OHDA-induced Parkinson’s disease rat. Pathol. Res. Pract. 2014, 210, 357–362. [Google Scholar] [CrossRef]
  229. Fikry, H.; Saleh, L.A.; Abdel Gawad, S. Neuroprotective effects of curcumin on the cerebellum in a rotenone-induced Parkinson’s Disease Model. CNS Neurosci. Ther. 2022, 28, 732–748. [Google Scholar] [CrossRef]
  230. Jagatha, B.; Mythri, R.B.; Vali, S.; Bharath, M.M. Curcumin treatment alleviates the effects of glutathione depletion in vitro and in vivo: Therapeutic implications for Parkinson’s disease explained via in silico studies. Free Radic. Biol. Med. 2008, 44, 907–917. [Google Scholar] [CrossRef]
  231. Sharma, N.; Nehru, B. Curcumin affords neuroprotection and inhibits α-synuclein aggregation in lipopolysaccharide-induced Parkinson’s disease model. Inflammopharmacology 2018, 26, 349–360. [Google Scholar] [CrossRef]
  232. Wang, X.-S.; Zhang, Z.-R.; Zhang, M.-M.; Sun, M.-X.; Wang, W.-W.; Xie, C.-L. Neuroprotective properties of curcumin in toxin-base animal models of Parkinson’s disease: A systematic experiment literatures review. BMC Complement. Altern. Med. 2017, 17, 412. [Google Scholar] [CrossRef]
  233. Liu, X.; Zhang, H.; Li, C.; Chen, Z.; Gao, Q.; Han, M.; Zhao, F.; Chen, D.; Chen, Q.; Hu, M. The dosage of curcumin to alleviate movement symptoms in a 6-hydroxydopamine-induced Parkinson’s disease rat model. Heliyon 2023, 9, e16921. [Google Scholar] [CrossRef]
  234. Geng, X.; Zhang, H.; Hu, M.; Liu, X.; Han, M.; Xie, J.; Li, Z.; Zhao, F.; Liu, W.; Wei, S. A novel curcumin oil solution can better alleviate the motor activity defects and neuropathological damage of a Parkinson’s disease mouse model. Front. Aging Neurosci. 2022, 14, 984895. [Google Scholar] [CrossRef]
  235. Li, Y.; Huang, J.; Wang, J.; Xia, S.; Ran, H.; Gao, L.; Feng, C.; Gui, L.; Zhou, Z.; Yuan, J. Human umbilical cord-derived mesenchymal stem cell transplantation supplemented with curcumin improves the outcomes of ischemic stroke via AKT/GSK-3β/β-TrCP/Nrf2 axis. J. Neuroinflamm. 2023, 20, 49. [Google Scholar] [CrossRef] [PubMed]
  236. Jia, G.; Tan, B.; Ma, J.; Zhang, L.; Jin, X.; Li, C. Prdx6 Upregulation by Curcumin Attenuates Ischemic Oxidative Damage via SP1 in Rats after Stroke. Biomed. Res. Int. 2017, 2017, 6597401. [Google Scholar] [CrossRef] [PubMed]
  237. Altinay, S.; Cabalar, M.; Isler, C.; Yildirim, F.; Celik, D.S.; Zengi, O.; Tas, A.; Gulcubuk, A. Is Chronic Curcumin Supplementation Neuroprotective Against Ischemia for Antioxidant Activity, Neurological Deficit, or Neuronal Apoptosis in an Experimental Stroke Model? Turk. Neurosurg. 2017, 27, 537–545. [Google Scholar] [PubMed][Green Version]
  238. Marques, M.S.; Cordeiro, M.F.; Marinho, M.A.G.; Vian, C.O.; Vaz, G.R.; Alves, B.S.; Jardim, R.D.; Hort, M.A.; Dora, C.L.; Horn, A.P. Curcumin-loaded nanoemulsion improves haemorrhagic stroke recovery in wistar rats. Brain Res. 2020, 1746, 147007. [Google Scholar] [CrossRef] [PubMed]
  239. Ran, Y.; Su, W.; Gao, F.; Ding, Z.; Yang, S.; Ye, L.; Chen, X.; Tian, G.; Xi, J.; Liu, Z. Curcumin Ameliorates White Matter Injury after Ischemic Stroke by Inhibiting Microglia/Macrophage Pyroptosis through NF-κB Suppression and NLRP3 Inflammasome Inhibition. Oxid. Med. Cell Longev. 2021, 2021, 1552127. [Google Scholar] [CrossRef]
  240. Xie, C.J.; Gu, A.P.; Cai, J.; Wu, Y.; Chen, R.C. Curcumin protects neural cells against ischemic injury in N2a cells and mouse brain with ischemic stroke. Brain Behav. 2018, 8, e00921. [Google Scholar] [CrossRef]
  241. Li, W.; Suwanwela, N.C.; Patumraj, S. Curcumin prevents reperfusion injury following ischemic stroke in rats via inhibition of NF-κB, ICAM-1, MMP-9 and caspase-3 expression. Mol. Med. Rep. 2017, 16, 4710–4720. [Google Scholar] [CrossRef]
  242. Liu, S.; Cao, Y.; Qu, M.; Zhang, Z.; Feng, L.; Ye, Z.; Xiao, M.; Hou, S.T.; Zheng, R.; Han, Z. Curcumin protects against stroke and increases levels of Notch intracellular domain. Neurol. Res. 2016, 38, 553–559. [Google Scholar] [CrossRef]
  243. Wu, S.; Guo, T.; Qi, W.; Li, Y.; Gu, J.; Liu, C.; Sha, Y.; Yang, B.; Hu, S.; Zong, X. Curcumin ameliorates ischemic stroke injury in rats by protecting the integrity of the blood-brain barrier. Exp. Ther. Med. 2021, 22, 783. [Google Scholar] [CrossRef]
  244. Liu, Z.; Ran, Y.; Huang, S.; Wen, S.; Zhang, W.; Liu, X.; Ji, Z.; Geng, X.; Ji, X.; Du, H.; et al. Curcumin Protects against Ischemic Stroke by Titrating Microglia/Macrophage Polarization. Front. Aging Neurosci. 2017, 9, 233. [Google Scholar] [CrossRef]
  245. Wu, J.; Li, Q.; Wang, X.; Yu, S.; Li, L.; Wu, X.; Chen, Y.; Zhao, J.; Zhao, Y. Neuroprotection by curcumin in ischemic brain injury involves the Akt/Nrf2 pathway. PLoS ONE 2013, 8, e59843. [Google Scholar] [CrossRef] [PubMed]
  246. Jiang, J.; Wang, W.; Sun, Y.J.; Hu, M.; Li, F.; Zhu, D.Y. Neuroprotective effect of curcumin on focal cerebral ischemic rats by preventing blood-brain barrier damage. Eur. J. Pharmacol. 2007, 561, 54–62. [Google Scholar] [CrossRef]
  247. Zhang, Y.; Yan, Y.; Cao, Y.; Yang, Y.; Zhao, Q.; Jing, R.; Hu, J.; Bao, J. Potential therapeutic and protective effect of curcumin against stroke in the male albino stroke-induced model rats. Life Sci. 2017, 183, 45–49. [Google Scholar] [CrossRef]
  248. He, Y.; Liu, Y.; Zhang, M. The beneficial effects of curcumin on aging and age-related diseases: From oxidative stress to antioxidant mechanisms, brain health and apoptosis. Front. Aging Neurosci. 2025, 17, 1533963. [Google Scholar] [CrossRef]
  249. Dutta, A.; Patil, R.; Pati, H.l.L. Curcumin: Its bioavailability and nanoparticle formulation: A review. Int. J. Health Sci. Res. 2021, 11, 228–238. [Google Scholar] [CrossRef]
  250. Sharifi-Rad, J.; Rayess, Y.E.; Rizk, A.A.; Sadaka, C.; Zgheib, R.; Zam, W.; Sestito, S.; Rapposelli, S.; Neffe-Skocińska, K.; Zielińska, D. Turmeric and its major compound curcumin on health: Bioactive effects and safety profiles for food, pharmaceutical, biotechnological and medicinal applications. Front. Pharmacol. 2020, 11, 1021. [Google Scholar] [CrossRef] [PubMed]
  251. Gagliardi, S.; Morasso, C.; Stivaktakis, P.; Pandini, C.; Tinelli, V.; Tsatsakis, A.; Prosperi, D.; Hickey, M.; Corsi, F.; Cereda, C. Curcumin formulations and trials: What’s new in neurological diseases. Molecules 2020, 25, 5389. [Google Scholar] [CrossRef]
  252. Wang, W.; Zhao, R.; Liu, B.; Li, K. The effect of curcumin supplementation on cognitive function: An updated systematic review and meta-analysis. Front. Nutr. 2025, 12, 1549509. [Google Scholar] [CrossRef]
  253. Francis, A.J.; Sreenivasan, C.; Parikh, A.; AlQassab, O.; Kanthajan, T.; Pandey, M.; Nwosu, M. Curcumin and Cognitive Function: A Systematic Review of the Effects of Curcumin on Adults With and Without Neurocognitive Disorders. Cureus 2024, 16, e67706. [Google Scholar] [CrossRef]
  254. Zhang, Y.; Li, L.; Zhang, J. Curcumin in antidepressant treatments: An overview of potential mechanisms, pre-clinical/clinical trials and ongoing challenges. Basic. Clin. Pharmacol. Toxicol. 2020, 127, 243–253. [Google Scholar] [CrossRef] [PubMed]
  255. Small, G.W.; Siddarth, P.; Li, Z.; Miller, K.J.; Ercoli, L.; Emerson, N.D.; Martinez, J.; Wong, K.P.; Liu, J.; Merrill, D.A.; et al. Memory and Brain Amyloid and Tau Effects of a Bioavailable Form of Curcumin in Non-Demented Adults: A Double-Blind, Placebo-Controlled 18-Month Trial. Am. J. Geriatr. Psychiatry 2018, 26, 266–277. [Google Scholar] [CrossRef] [PubMed]
  256. Cox, K.H.; Pipingas, A.; Scholey, A.B. Investigation of the effects of solid lipid curcumin on cognition and mood in a healthy older population. J. Psychopharmacol. 2015, 29, 642–651. [Google Scholar] [CrossRef]
  257. Baum, L.; Lam, C.W.; Cheung, S.K.; Kwok, T.; Lui, V.; Tsoh, J.; Lam, L.; Leung, V.; Hui, E.; Ng, C.; et al. Six-month randomized, placebo-controlled, double-blind, pilot clinical trial of curcumin in patients with Alzheimer disease. J. Clin. Psychopharmacol. 2008, 28, 110–113. [Google Scholar] [CrossRef]
  258. Ringman, J.M.; Frautschy, S.A.; Teng, E.; Begum, A.N.; Bardens, J.; Beigi, M.; Gylys, K.H.; Badmaev, V.; Heath, D.D.; Apostolova, L.G.; et al. Oral curcumin for Alzheimer’s disease: Tolerability and efficacy in a 24-week randomized, double blind, placebo-controlled study. Alzheimers Res. Ther. 2012, 4, 43. [Google Scholar] [CrossRef]
  259. Rainey-Smith, S.R.; Brown, B.M.; Sohrabi, H.R.; Shah, T.; Goozee, K.G.; Gupta, V.B.; Martins, R.N. Curcumin and cognition: A randomised, placebo-controlled, double-blind study of community-dwelling older adults. Br. J. Nutr. 2016, 115, 2106–2113. [Google Scholar] [CrossRef]
  260. Ghodsi, H.; Rahimi, H.R.; Aghili, S.M.; Saberi, A.; Shoeibi, A. Evaluation of curcumin as add-on therapy in patients with Parkinson’s disease: A pilot randomized, triple-blind, placebo-controlled trial. Clin. Neurol. Neurosurg. 2022, 218, 107300. [Google Scholar] [CrossRef]
  261. Maghbooli, M.; Safarnejad, B.; Mostafavi, H.; Mazloomzadeh, S.; Ghoreishi, A. Effect of nanomicelle curcumin on quality of life and sleep in patients with Parkinson’s disease: A double-blind, randomized, and PlaceboControlled trial. Int. Clin. Neurosci. J. 2019, 6, 140–145. [Google Scholar] [CrossRef]
  262. Donadio, V.; Incensi, A.; Rizzo, G.; Fileccia, E.; Ventruto, F.; Riva, A.; Tiso, D.; Recchia, M.; Vacchiano, V.; Infante, R.; et al. The Effect of Curcumin on Idiopathic Parkinson Disease: A Clinical and Skin Biopsy Study. J. Neuropathol. Exp. Neurol. 2022, 81, 545–552. [Google Scholar] [CrossRef]
  263. Boshagh, K.; Khorvash, F.; Sahebkar, A.; Majeed, M.; Bahreini, N.; Askari, G.; Bagherniya, M. The effects of curcumin-piperine supplementation on inflammatory, oxidative stress and metabolic indices in patients with ischemic stroke in the rehabilitation phase: A randomized controlled trial. Nutr. J. 2023, 22, 69. [Google Scholar] [CrossRef] [PubMed]
  264. Yakubu, J.; Pandey, A.V. Innovative delivery systems for curcumin: Exploring nanosized and conventional formulations. Pharmaceutics 2024, 16, 637. [Google Scholar] [CrossRef] [PubMed]
  265. Gota, V.S.; Maru, G.B.; Soni, T.G.; Gandhi, T.R.; Kochar, N.; Agarwal, M.G. Safety and pharmacokinetics of a solid lipid curcumin particle formulation in osteosarcoma patients and healthy volunteers. J. Agric. Food Chem. 2010, 58, 2095–2099. [Google Scholar] [CrossRef]
  266. Nahar, P.P.; Slitt, A.L.; Seeram, N.P. Anti-Inflammatory Effects of Novel Standardized Solid Lipid Curcumin Formulations. J. Med. Food 2015, 18, 786–792. [Google Scholar] [CrossRef]
  267. Gupte, P.A.; Giramkar, S.A.; Harke, S.M.; Kulkarni, S.K.; Deshmukh, A.P.; Hingorani, L.L.; Mahajan, M.P.; Bhalerao, S.S. Evaluation of the efficacy and safety of Capsule Longvida(®) Optimized Curcumin (solid lipid curcumin particles) in knee osteoarthritis: A pilot clinical study. J. Inflamm. Res. 2019, 12, 145–152. [Google Scholar] [CrossRef]
  268. Pancholi, V.; Smina, T.P.; Kunnumakkara, A.B.; Maliakel, B.; Krishnakumar, I.M. Safety assessment of a highly bioavailable curcumin-galactomannoside complex (CurQfen) in healthy volunteers, with a special reference to the recent hepatotoxic reports of curcumin supplements: A 90-days prospective study. Toxicol. Rep. 2021, 8, 1255–1264. [Google Scholar] [CrossRef]
  269. Cicero, A.F.G.; Sahebkar, A.; Fogacci, F.; Bove, M.; Giovannini, M.; Borghi, C. Effects of phytosomal curcumin on anthropometric parameters, insulin resistance, cortisolemia and non-alcoholic fatty liver disease indices: A double-blind, placebo-controlled clinical trial. Eur. J. Nutr. 2020, 59, 477–483. [Google Scholar] [CrossRef]
  270. Nakagawa, Y.; Mukai, S.; Yamada, S.; Murata, S.; Yabumoto, H.; Maeda, T.; Akamatsu, S. The Efficacy and Safety of Highly-Bioavailable Curcumin for Treating Knee Osteoarthritis: A 6-Month Open-Labeled Prospective Study. Clin. Med. Insights Arthritis Musculoskelet. Disord. 2020, 13, 1179544120948471. [Google Scholar] [CrossRef]
  271. Sunagawa, Y.; Miyazaki, Y.; Funamoto, M.; Shimizu, K.; Shimizu, S.; Nurmila, S.; Katanasaka, Y.; Ito, M.; Ogawa, T.; Ozawa-Umeta, H. A novel amorphous preparation improved curcumin bioavailability in healthy volunteers: A single-dose, double-blind, two-way crossover study. J. Funct. Foods 2021, 81, 104443. [Google Scholar] [CrossRef]
  272. Tagde, P.; Tagde, P.; Islam, F.; Tagde, S.; Shah, M.; Hussain, Z.D.; Rahman, M.H.; Najda, A.; Alanazi, I.S.; Germoush, M.O. The multifaceted role of curcumin in advanced nanocurcumin form in the treatment and management of chronic disorders. Molecules 2021, 26, 7109. [Google Scholar] [CrossRef] [PubMed]
  273. Mandal, M.; Jaiswal, P.; Mishra, A. Role of curcumin and its nanoformulations in neurotherapeutics: A comprehensive review. J. Biochem. Mol. Toxicol. 2020, 34, e22478. [Google Scholar] [CrossRef] [PubMed]
  274. Zhao, L.; Ciallella, H.L.; Aleksunes, L.M.; Zhu, H. Advancing computer-aided drug discovery (CADD) by big data and data-driven machine learning modeling. Drug Discov. Today 2020, 25, 1624–1638. [Google Scholar] [CrossRef] [PubMed]
  275. Vinicki, M.; Pribic, T.; Vuckovic, F.; Frkatovic-Hodzic, A.; Plaza-Andrades, I.; Tinahones, F.; Raffaele, J.; Fernandez-Garcia, J.C.; Lauc, G. Effects of testosterone and metformin on the GlycanAge index of biological age and the composition of the IgG glycome. Geroscience 2024, 47, 1777–1788. [Google Scholar] [CrossRef] [PubMed]
  276. Malavolta, M. Anti-aging interventions in geriatric mice: Insights into the timing of treatment, benefits, and limitations. Geroscience 2024, 47, 109–119. [Google Scholar] [CrossRef]
  277. Landsberger, T.; Amit, I.; Alon, U. Geroprotective interventions converge on gene expression programs of reduced inflammation and restored fatty acid metabolism. Geroscience 2024, 46, 1627–1639. [Google Scholar] [CrossRef]
  278. Shintani, T.; Shintani, H.; Sato, M.; Ashida, H. Calorie restriction mimetic drugs could favorably influence gut microbiota leading to lifespan extension. Geroscience 2023, 45, 3475–3490. [Google Scholar] [CrossRef]
  279. Li, X.; McPherson, M.; Hager, M.; Lee, M.; Chang, P.; Miller, R.A. Four anti-aging drugs and calorie-restricted diet produce parallel effects in fat, brain, muscle, macrophages, and plasma of young mice. Geroscience 2023, 45, 2495–2510. [Google Scholar] [CrossRef]
  280. Katsube, M.; Ishimoto, T.; Fukushima, Y.; Kagami, A.; Shuto, T.; Kato, Y. Ergothioneine promotes longevity and healthy aging in male mice. Geroscience 2024, 46, 3889–3909. [Google Scholar] [CrossRef]
  281. Sonsalla, M.M.; Lamming, D.W. Geroprotective interventions in the 3xTg mouse model of Alzheimer’s disease. Geroscience 2023, 45, 1343–1381. [Google Scholar] [CrossRef] [PubMed]
  282. Sandalova, E.; Goh, J.; Lim, Z.X.; Lim, Z.M.; Barardo, D.; Dorajoo, R.; Kennedy, B.K.; Maier, A.B. Alpha-ketoglutarate supplementation and BiologicaL agE in middle-aged adults (ABLE)-intervention study protocol. Geroscience 2023, 45, 2897–2907. [Google Scholar] [CrossRef]
  283. McGee, K.C.; Sullivan, J.; Hazeldine, J.; Schmunk, L.J.; Martin-Herranz, D.E.; Jackson, T.; Lord, J.M. A combination nutritional supplement reduces DNA methylation age only in older adults with a raised epigenetic age. Geroscience 2024, 46, 4333–4347. [Google Scholar] [CrossRef]
  284. Dai, Z.; Lee, S.Y.; Sharma, S.; Ullah, S.; Tan, E.C.K.; Brodaty, H.; Schutte, A.E.; Sachdev, P.S. A systematic review of diet and medication use among centenarians and near-centenarians worldwide. Geroscience 2024, 46, 6625–6639. [Google Scholar] [CrossRef] [PubMed]
  285. Bizzozero-Peroni, B.; Diaz-Goni, V.; Beneit, N.; Oliveira, A.; Jimenez-Lopez, E.; Martinez-Vizcaino, V.; Mesas, A.E. Nut consumption is associated with a lower risk of all-cause dementia in adults: A community-based cohort study from the UK Biobank. Geroscience 2024, 47, 1721–1733. [Google Scholar] [CrossRef]
  286. Zupo, R.; Donghia, R.; Castellana, F.; Bortone, I.; De Nucci, S.; Sila, A.; Tatoli, R.; Lampignano, L.; Sborgia, G.; Panza, F.; et al. Ultra-processed food consumption and nutritional frailty in older age. Geroscience 2023, 45, 2229–2243. [Google Scholar] [CrossRef]
  287. Zhou, Z.; Kim, K.; Ramsey, J.J.; Rutkowsky, J.M. Ketogenic diets initiated in late mid-life improved measures of spatial memory in male mice. Geroscience 2023, 45, 2481–2494. [Google Scholar] [CrossRef]
  288. Aversa, Z.; White, T.A.; Heeren, A.A.; Hulshizer, C.A.; Saul, D.; Zhang, X.; Molina, A.J.A.; Redman, L.M.; Martin, C.K.; Racette, S.B.; et al. Calorie restriction reduces biomarkers of cellular senescence in humans. Aging Cell 2024, 23, e14038. [Google Scholar] [CrossRef]
  289. Das, S.K.; Silver, R.E.; Senior, A.; Gilhooly, C.H.; Bhapkar, M.; Le Couteur, D. Diet composition, adherence to calorie restriction, and cardiometabolic disease risk modification. Aging Cell 2023, 22, e14018. [Google Scholar] [CrossRef] [PubMed]
  290. Fiorito, G.; Tosti, V.; Polidoro, S.; Bertozzi, B.; Veronese, N.; Cava, E.; Spelta, F.; Piccio, L.; Early, D.S.; Raftery, D.; et al. Multi-omic analysis of biological aging biomarkers in long-term calorie restriction and endurance exercise practitioners: A cross-sectional study. Aging Cell 2025, 24, e14442. [Google Scholar] [CrossRef]
  291. Han, W.; Zhang, B.; Zhao, W.; Zhao, W.; He, J.; Qiu, X.; Zhang, L.; Wang, X.; Wang, Y.; Lu, H.; et al. Ketogenic beta-hydroxybutyrate regulates beta-hydroxybutyrylation of TCA cycle-associated enzymes and attenuates disease-associated pathologies in Alzheimer’s mice. Aging Cell 2025, 24, e14368. [Google Scholar] [CrossRef] [PubMed]
  292. Ribeiro, R.V.; Senior, A.M.; Simpson, S.J.; Tan, J.; Raubenheimer, D.; Le Couteur, D.; Macia, L.; Holmes, A.; Eberhard, J.; O’Sullivan, J.; et al. Rapid benefits in older age from transition to whole food diet regardless of protein source or fat to carbohydrate ratio: Arandomised control trial. Aging Cell 2024, 23, e14276. [Google Scholar] [CrossRef]
  293. Ungvari, Z.; Fazekas-Pongor, V.; Csiszar, A.; Kunutsor, S.K. The multifaceted benefits of walking for healthy aging: From Blue Zones to molecular mechanisms. Geroscience 2023, 45, 3211–3239. [Google Scholar] [CrossRef]
  294. Ungvari, Z.; Fekete, M.; Varga, P.; Fekete, J.T.; Buda, A.; Szappanos, A.; Lehoczki, A.; Mozes, N.; Grosso, G.; Menyhart, O.; et al. Impact of adherence to the Mediterranean diet on stroke risk. Geroscience 2025, 47, 3565–3581. [Google Scholar] [CrossRef]
  295. Fekete, M.; Varga, P.; Ungvari, Z.; Fekete, J.T.; Buda, A.; Szappanos, A.; Lehoczki, A.; Mozes, N.; Grosso, G.; Godos, J.; et al. The role of the Mediterranean diet in reducing the risk of cognitive impairement, dementia, and Alzheimer’s disease: A meta-analysis. Geroscience 2025, 47, 3111–3130. [Google Scholar] [CrossRef]
  296. Talavera-Rodriguez, I.; Banegas, J.R.; de la Cruz, J.J.; Martinez-Gomez, D.; Ruiz-Canela, M.; Ortola, R.; Hershey, M.S.; Artalejo, F.R.; Sotos-Prieto, M. Mediterranean lifestyle index and 24-h systolic blood pressure and heart rate in community-dwelling older adults. Geroscience 2024, 46, 1357–1369. [Google Scholar] [CrossRef]
  297. Dobreva, I.; Marston, L.; Mukadam, N. Which components of the Mediterranean diet are associated with dementia? A UK Biobank cohort study. Geroscience 2022, 44, 2541–2554. [Google Scholar] [CrossRef]
  298. Gensous, N.; Garagnani, P.; Santoro, A.; Giuliani, C.; Ostan, R.; Fabbri, C.; Milazzo, M.; Gentilini, D.; di Blasio, A.M.; Pietruszka, B.; et al. One-year Mediterranean diet promotes epigenetic rejuvenation with country- and sex-specific effects: A pilot study from the NU-AGE project. Geroscience 2020, 42, 687–701. [Google Scholar] [CrossRef] [PubMed]
  299. Martens, C.R.; Rossman, M.J.; Mazzo, M.R.; Jankowski, L.R.; Nagy, E.E.; Denman, B.A.; Richey, J.J.; Johnson, S.A.; Ziemba, B.P.; Wang, Y.; et al. Short-term time-restricted feeding is safe and feasible in non-obese healthy midlife and older adults. Geroscience 2020, 42, 667–686. [Google Scholar] [CrossRef] [PubMed]
  300. Gonzalez-Armenta, J.L.; Bergstrom, J.; Lee, J.; Furdui, C.M.; Nicklas, B.J.; Molina, A.J.A. Serum factors mediate changes in mitochondrial bioenergetics associated with diet and exercise interventions. Geroscience 2024, 46, 349–365. [Google Scholar] [CrossRef] [PubMed]
  301. Szilagyi, A.; Takacs, B.; Szekeres, R.; Tarjanyi, V.; Nagy, D.; Priksz, D.; Bombicz, M.; Kiss, R.; Szabo, A.M.; Lehoczki, A.; et al. Effects of voluntary and forced physical exercise on the retinal health of aging Wistar rats. Geroscience 2024, 46, 4707–4728. [Google Scholar] [CrossRef]
  302. Roslund, K.J.; Ramsey, J.J.; Rutkowsky, J.M.; Zhou, Z.; Slupsky, C.M. Two-month ketogenic diet alters systemic and brain metabolism in middle-aged female mice. Geroscience 2024, 47, 935–952. [Google Scholar] [CrossRef]
  303. Frohlich, S.; Kutz, D.F.; Muller, K.; Voelcker-Rehage, C. Cardiorespiratory fitness is associated with cognitive performance in 80+-year-olds: Detangling processing levels. Geroscience 2024, 46, 3297–3310. [Google Scholar] [CrossRef]
  304. Chauquet, S.; Willis, E.F.; Grice, L.; Harley, S.B.R.; Powell, J.E.; Wray, N.R.; Nguyen, Q.; Ruitenberg, M.J.; Shah, S.; Vukovic, J. Exercise rejuvenates microglia and reverses T cell accumulation in the aged female mouse brain. Aging Cell 2024, 23, e14172. [Google Scholar] [CrossRef]
  305. Fox, F.A.U.; Liu, D.; Breteler, M.M.B.; Aziz, N.A. Physical activity is associated with slower epigenetic ageing-Findings from the Rhineland study. Aging Cell 2023, 22, e13828. [Google Scholar] [CrossRef]
  306. Kawamura, T.; Radak, Z.; Tabata, H.; Akiyama, H.; Nakamura, N.; Kawakami, R.; Ito, T.; Usui, C.; Jokai, M.; Torma, F.; et al. Associations between cardiorespiratory fitness and lifestyle-related factors with DNA methylation-based ageing clocks in older men: WASEDA’S Health Study. Aging Cell 2024, 23, e13960. [Google Scholar] [CrossRef] [PubMed]
  307. Lohman, T.; Bains, G.; Cole, S.; Gharibvand, L.; Berk, L.; Lohman, E. High-Intensity interval training reduces transcriptomic age: A randomized controlled trial. Aging Cell 2023, 22, e13841. [Google Scholar] [CrossRef] [PubMed]
  308. Malin, S.K.; Battillo, D.J.; Beeri, M.S.; Mustapic, M.; Delgado-Peraza, F.; Kapogiannis, D. Two weeks of exercise alters neuronal extracellular vesicle insulin signaling proteins and pro-BDNF in older adults with prediabetes. Aging Cell 2025, 24, e14369. [Google Scholar] [CrossRef] [PubMed]
  309. Manning, K.M.; Hall, K.S.; Sloane, R.; Magistro, D.; Rabaglietti, E.; Lee, C.C.; Castle, S.; Kopp, T.; Giffuni, J.; Katzel, L.; et al. Longitudinal analysis of physical function in older adults: The effects of physical inactivity and exercise training. Aging Cell 2024, 23, e13987. [Google Scholar] [CrossRef]
  310. Moore, A.Z.; Simonsick, E.M.; Landman, B.; Schrack, J.; Wanigatunga, A.A.; Ferrucci, L. Correlates of life course physical activity in participants of the Baltimore longitudinal study of aging. Aging Cell 2024, 23, e14078. [Google Scholar] [CrossRef]
  311. Voisin, S.; Seale, K.; Jacques, M.; Landen, S.; Harvey, N.R.; Haupt, L.M.; Griffiths, L.R.; Ashton, K.J.; Coffey, V.G.; Thompson, J.M.; et al. Exercise is associated with younger methylome and transcriptome profiles in human skeletal muscle. Aging Cell 2024, 23, e13859. [Google Scholar] [CrossRef]
  312. Kunnumakkara, A.B.; Bordoloi, D.; Padmavathi, G.; Monisha, J.; Roy, N.K.; Prasad, S.; Aggarwal, B.B. Curcumin, the golden nutraceutical: Multitargeting for multiple chronic diseases. Br. J. Pharmacol. 2017, 174, 1325–1348. [Google Scholar] [CrossRef]
  313. Far, B.F.; Safaei, M.; Pourmolaei, A.; Adibamini, S.; Shirdel, S.; Shirdel, S.; Emadi, R.; Kaushik, A.K. Exploring curcumin-loaded lipid-based nanomedicine as efficient targeted therapy for Alzheimer’s diseases. ACS Appl. Bio Mater. 2024, 7, 3535–3555. [Google Scholar] [CrossRef]
  314. Naoi, M.; Maruyama, W.; Shamoto-Nagai, M. Disease-modifying treatment of Parkinson’s disease by phytochemicals: Targeting multiple pathogenic factors. J. Neural Transm. 2022, 129, 737–753. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Pharmacokinetic challenges of oral curcumin.
Figure 1. Pharmacokinetic challenges of oral curcumin.
Nutrients 17 02884 g001
Figure 2. Anti-Inflammatory Mechanisms of Curcumin: Inhibition of NF-κB, COX-2, and iNOS: Curcumin reduces inflammation by inhibiting the NF-κB (nuclear factor kappa B) signaling pathway, as well as COX-2 (cyclooxygenase-2) and iNOS (inducible nitric oxide synthase) activity, thereby lowering pro-inflammatory cytokine and nitric oxide production. Arrow indicates the direction of change: ↓—decrease.
Figure 2. Anti-Inflammatory Mechanisms of Curcumin: Inhibition of NF-κB, COX-2, and iNOS: Curcumin reduces inflammation by inhibiting the NF-κB (nuclear factor kappa B) signaling pathway, as well as COX-2 (cyclooxygenase-2) and iNOS (inducible nitric oxide synthase) activity, thereby lowering pro-inflammatory cytokine and nitric oxide production. Arrow indicates the direction of change: ↓—decrease.
Nutrients 17 02884 g002
Table 1. Preclinical studies on curcumin in Alzheimer’s disease models.
Table 1. Preclinical studies on curcumin in Alzheimer’s disease models.
Ref.ModelDoseDurationRouteKey EffectsMechanisms
Sun et al. [45]APP/PS1 male mice50 or 200 mg/kg3 monthsOral gavage (daily)Improved spatial memory; reduced hippocampal Aβ plaquesModulated gut microbiota (↑ Lactobacillaceae, ↓ Bacteroidaceae); curcumin biotransformation to neuroprotective metabolites
Lin et al. [206]APPswe/PSEN1dE9 mice (both sexes)Curcumin 200 mg/kg; Berberine 100 mg/kg3 monthsOral gavage (daily)Improved cognition; ↓ Aβ1–42, inflammation (IL-1β, TNF-α, IL-6), oxidative stress; ↑ AMPKα phosphorylation and autophagySynergistic anti-amyloid, anti-inflammatory, antioxidative effects; modulation of APP processing and AMPK–autophagy pathway
Su et al. [207]Triple-transgenic Alzheimer’s disease mice150 mg/kg (TML-6)4 monthsOralImproved learning; reduced brain Aβ and microglial activationEnhanced bioavailability; reduced Aβ synthesis; ↑ ApoE; NF-κB/mTOR suppression; Nrf2 pathway activation
ELBini-Dhouib et al. [208]AlCl3-induced sporadic AD rats100 mg/kg (CUR1 co-treatment; CUR2 post-treatment)90 daysOralImproved cognition; reduced anxiety; restored AChE activity; improved oxidative markers; ↓ apoptosis and inflammationAntioxidant and anti-inflammatory actions; neuroprotection; restored cholinergic function
Reddy et al. [209]SH-SY5Y cells + Aβ25–3566.3 μM24 h (pretreatment)In vitroImproved cell viability; ↑ MFN1/2, OPA1; ↓ DRP1, FIS1; ↑ PGC1α, Nrf1/2, TFAM; ↑ synaptic markersEnhanced mitochondrial biogenesis and dynamics; synaptic protection; reduced oxidative stress
Wang et al. [210]1–40-induced AD in Sprague–Dawley rats300 mg/kg7 daysIntraperitonealImproved spatial memory; reduced hippocampal GFAP mRNA; less reactive astrocytosisSuppression of astrogliosis and neuroinflammation via GFAP downregulation
Cheng et al. [211]Tg2576 AD mice100 mg/kg3 monthsOral gavage (weekly)Improved cue memory; trend to improved working memory; ↓ amyloid plaques; ↑ brain/plasma curcuminNanoparticle formulation enhanced BBB penetration, pharmacokinetics, and bioavailability
Ray et al. [212]SK-N-SH cells and athymic mice25 mg/kg NanoCurc™ i.p. twice daily; 1 nM–5 μM in vitroIn vitro: 24 h; in vivo: ≤16 h post-doseIn vitro; i.p.Neuroprotection from ROS; rescued injured cells; ↑ brain curcumin; ↓ ROS and caspase 3/7; ↑ GSHPolymeric nanoparticles improved bioavailability; enhanced redox balance; antioxidant and anti-apoptotic effects
Begum et al. [213]Tg2576 APPsw mice; LPS-injected WT mice; neuron/microglia culturesChronic: ~83 mg/kg diet (~500 ppm); Acute: 0.4 μmolChronic: ~4 months; Acute: 2 days pre-LPSOral (diet/gavage), i.p., i.m.Both regimens reduced neuroinflammation (iNOS, IL-1β); Cur reduced plaques/insoluble Aβ, TC reduced soluble Aβ and JNK phosphorylationCur dienone bridge essential for plaque reduction; TC selectively inhibits JNK phosphorylation; both antioxidant/anti-inflammatory
Xiong et al. [214]SH-SY5Y neuroblastoma cells (APP-transfected)0–20 μM (24 h), or 5 μM (12–48 h)24–48 hIn vitroDose- and time-dependent reduction in extracellular Aβ40/42Inhibition of Aβ generation via PS1 and GSK-3β suppression; ↑ GSK-3β Ser9 phosphorylation; downregulation of APP-processing enzymes
Ishrat et al. [215]ICV-STZ-infused male Wistar rats80 mg/kg3 weeksOral gavageImproved cognition; ↓ oxidative stress (↓ 4-HNE, MDA, TBARS, H2O2, PC, GSSG; ↑ GSH, GPx, GR); ↑ choline acetyltransferaseAntioxidant activity; restored glutathione system; reduced lipid peroxidation; enhanced cholinergic function
Nam et al. [216]Adult and D-galactose-induced aged male C57BL/6 mice300 mg/kg Curcuma longa extract; 100 mg/kg D-galactose10 weeks D-gal; 3 weeks C. longaOral gavage; subcutaneous (D-gal)Reduced escape latency; ↑ cell proliferation (Ki67), neuroblast differentiation (DCX); ↑ pCREB, ↑ BDNF in hippocampusCREB pathway activation; enhanced neurogenesis; increased neurotrophic factors
Bassani et al. [217]Male Wistar rats25, 50, or 100 mg/kg30 daysOral gavageImproved object recognition (50/100 mg/kg) in ICV-STZ dementia; mild anti-inflammatory effectMemory improvement unrelated to hippocampal neurogenesis; mild anti-inflammatory action
Ma et al. [108]hTau transgenic mice (15–16 months) and controls500 ppm curcumin (Longvida SLN)4–5 monthsDiet (ad libitum)Reduced behavioral, synaptic, and chaperone deficits; ↓ soluble Tau dimers; ↑ HSP90/HSC70 in synaptic fractionsIncreased HSP proteins; modulation of HSP90 client kinases; enhanced Tau clearance
McClure et al. [218]5XFAD transgenic mice5 mg/kg18 weeksInhalation (aerosol)Reduced Aβ plaques in hippocampus/subiculum; improved memory; no toxicity↓ Aβ and COX-2; prevention of neuritic dystrophy and autophagic vesicle formation
Alamro et al. [219]Primary cortical neurons (Wistar rat pups)1 μM Aβ1–42 + 1 nM vitamin D3 + 5 μM curcumin≤72 hIn vitroReduced lipid peroxidation; ↑ GST, catalase, SOD; improved viability; ↑ NGFReduced oxidative stress; enhanced antioxidant defense; increased NGF
Lim et al. [220]APPSw Tg+ mice160 or 5000 ppm curcumin6 monthsDiet↓ oxidized proteins, IL-1β, Aβ plaques; ↓ GFAP (low dose); ↓ microglial activationAnti-inflammatory (IL-1β); reduced oxidative stress and plaque formation; microglial suppression
Savall et al. [221]STZ-induced AD male Wistar rats6 mg/kg14 days treatment (total 36 days)Oral gavage; STZ i.c.v.Nanoencapsulated curcumin improved memory; both forms ↓ AChE activity; NC reduced oxidative stressReduced neuroinflammation (GFAP); antioxidant effects; cognitive improvement
4-HNE—4-hydroxynonenal, Aβ—amyloid-beta, AChE—acetylcholinesterase, AMPKα—AMP-activated protein kinase alpha, APP—amyloid precursor protein, APPswe—Swedish mutant of APP, BDNF—brain-derived neurotrophic factor, BBB—blood–brain barrier, COX-2—cyclooxygenase-2, CREB—cAMP response element-binding protein, DCX—doublecortin, DRP1—dynamin-related protein 1, FIS1—mitochondrial fission 1 protein, GFAP—glial fibrillary acidic protein, GSH—glutathione, GSSG—glutathione disulfide; GPx—glutathione peroxidase, GR—glutathione reductase, GSK-3β—glycogen synthase kinase-3 beta, HSP—heat shock protein, IL-1β—interleukin-1 beta, iNOS—inducible nitric oxide synthase, JNK—c-Jun N-terminal kinase, LPS—lipopolysaccharide, MDA—malondialdehyde, MFN1/2—mitofusin 1/2, mTOR—mechanistic target of rapamycin, NC—nanoencapsulated curcumin, Nrf2—nuclear factor erythroid 2-related factor 2, NGF—nerve growth factor, NF-κB—nuclear factor kappa-light-chain-enhancer of activated B cells, OPA1—optic atrophy 1, PC—protein carbonyl, PGC1α—peroxisome proliferator-activated receptor gamma coactivator 1-alpha, PS1—presenilin-1, pCREB—phosphorylated CREB, ROS—reactive oxygen species, SLN—solid lipid nanoparticles, STZ—streptozotocin, TC—Tetrahydrocurcumin, TBARS—thiobarbituric acid reactive substances, Tg—transgenic, TFAM—mitochondrial transcription factor A, WT—wild type. Arrows indicate the direction of change: ↑ increase, ↓ decrease.
Table 2. Preclinical studies on curcumin in experimental Parkinson’s disease models.
Table 2. Preclinical studies on curcumin in experimental Parkinson’s disease models.
ReferenceModelDoseDurationRouteKey EffectsProposed Mechanisms
Abbaoui et al. [222]Wistar rats (copper-induced Parkinson’s disease)Cu: 10 mg/kg i.p.; Curcumin: 30 mg/kg oral3 daysi.p., oralRestored tyrosine hydroxylase (TH) expression and locomotor functionAntioxidant and anti-inflammatory activity; protection of TH+ neurons from copper-induced neurotoxicity
Rajeswari et al. [223] Mice (MPTP model)MPTP: 40 mg/kg; Curcumin: 80 mg/kg; ThC: 60 mg/kg7 daysi.p.Restored dopamine (DA) and DOPAC levels; reduced MAO-B activityDopaminergic neuron preservation via MAO-B inhibition and dopamine metabolism protection
Pan et al. [224]C57BL/6 mice (MPTP model)50 mg/kg5 daysi.p.Reduced dopaminergic neuron loss; improved TH levelsInhibition of JNK-mediated mitochondrial apoptosis
Jiang et al. [225]SH-SY5Y cells (A53T α-syn overexpression)6 μM24–48 hIn vitroReduced α-synuclein accumulation; restored autophagic flux; improved neuronal viabilityDownregulation of mTOR/p70S6K signaling; increased LC3-II; enhanced autophagosome formation
Ramires Júnior et al. [227]Mice (rotenone model)25–50 mg/kg30 daysOralImproved motor function; preserved mitochondrial complex I activityOxidative stress reduction; mitochondrial protection; enhanced efficacy via nanoemulsion formulation
Yang et al. [228]Rats (6-OHDA model)5–20 mg/kg37 daysOralImproved behavior, cognition, DA and NE levelsActivation of BDNF/TrkB/PI3K pathway; neurogenesis; monoaminergic system support
Fikry et al. [229]Rats (rotenone model)30 mg/kg60 daysi.p.Preserved cerebellar structure and motor function; reduced oxidative stressAntioxidant defense restoration (↑GSH, ↑SOD, ↓MDA); reduced gliosis; improved acetylcholinesterase (AChE) activity
Wang et al. [226]SH-SY5Y cells (α-syn oligomers)4 μM48–72 hIn vitroReduced ROS and apoptosis; stabilized α-syn aggregatesCaspase-3 inhibition; ROS suppression; attenuation of α-syn toxicity
Jagatha et al. [230]N27 cells (in vitro) and GSH-deficient mice (BSO-induced)10 μM (in vitro); 50 mg/kg i.p. (in vivo)1–3 daysIn vitro, i.p.Restored GSH; protected mitochondrial complex I; reduced protein oxidationγ-GCL activation; mitochondrial stabilization; redox balance normalization
Sharma et al. [231]Rats (LPS-induced PD)40 mg/kg21 daysi.p.Reduced astrocytosis, cytokines, iron deposition, and α-syn aggregationNF-κB inhibition; antioxidant effects; apoptosis regulation; iron metabolism normalization
Liu et al. [233]Wistar rats (6-OHDA model)40, 80, or 160 mg/kg/day14 daysOral (gavage)High dose improved motor performance; preserved TH+ neurons in substantia nigraDose- and time-dependent neuroprotection; dopaminergic preservation; reduced oxidative stress and inflammation
Geng et al. [234]Male C57BL/6J mice (MPTP model)MPTP: 30 mg/kg/day i.p.; Curcumin or Curoil: 120 mg/kg/day oral7 daysi.p., oralCuroil improved motor function; protected TH+ neurons in substantia nigraEnhanced curcumin bioavailability; antioxidant and neuroprotective effects; TH+ neuron preservation
6-OHDA—6-hydroxydopamine, AChE—acetylcholinesterase, α-syn—alpha-synuclein, BDNF—brain-derived neurotrophic factor, BDNF/TrkB/PI3K: brain-derived neurotrophic factor/tropomyosin receptor kinase B/phosphoinositide 3-kinase; BSO—buthionine sulfoximine, Curcumin—diferuloylmethane, DA—dopamine, DOPAC—3,4-dihydroxyphenylacetic acid, GSH—glutathione, γ-GCL—gamma-glutamylcysteine ligase, i.p.—intraperitoneal, JNK—c-Jun N-terminal kinase, LC3-II—microtubule-associated proteins 1A/1B light chain 3B (lipidated form), LPS—lipopolysaccharide, MPTP—1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine, MAO-B—monoamine oxidase B, mTOR—mechanistic target of rapamycin, NE—norepinephrine, NF-κB—nuclear factor kappa-light-chain-enhancer of activated B cells, ROS—reactive oxygen species, TH—tyrosine hydroxylase. Arrows indicate the direction of change: ↑ increase, ↓ decrease.
Table 3. Preclinical studies on curcumin in experimental stroke models.
Table 3. Preclinical studies on curcumin in experimental stroke models.
Ref.ModelDoseDurationRouteKey EffectsMechanisms
Li et al. [235]MCAO in mice; neuronal cultures100 mg/kg i.p. (in vivo); 4 μM (in vitro)3 days (in vivo); 24 h (in vitro)i.p., i.v.Improved neurological function; reduced brain edema and infarct volumeActivation of AKT/GSK-3β/Nrf2 pathway; M2 microglia polarization; reduced oxidative stress
Jia et al. [236]Transient MCAO in rats + MTM300 mg/kg i.p. + 250 μg/kg MTM i.c.v.24 hi.p., i.c.v.Reduced infarct size and oxidative stress; improved neurological outcomesSP1-mediated upregulation of Prdx6; antioxidant protection
Altinay et al. [237]Global ischemia (BCCAO)300 mg/kg oral or i.p.21 days (oral); 72 h (i.p.)Oral, i.p.Increased antioxidant enzymes; reduced inflammation; improved outcomeIncreased SOD, CAT, GPx; reduced IL-6, TNF-α, apoptosis
Marques et al. [238]ICH in rats30 mg/kg3 doses/48 hi.p.Improved motor recovery; reduced hematoma volumeEnhanced antioxidant activity; improved bioavailability
Ran et al. [239]MCAO in mice150 mg/kg7 daysi.p.Reduced white matter damage; improved sensorimotor functionInhibition of NF-κB/NLRP3; reduced pyroptosis
Xie et al. [240]MCAO in mice; OGD/R in cells100–400 mg/kg (in vivo); 5–35 μM (in vitro)Single dosei.p., in vitroReduced apoptosis; improved mitochondrial functionDownregulation of Bax, upregulation of Bcl-2; preserved mitochondrial membrane potential
Li et al. [241]MCAO/reperfusion in rats300 mg/kgSingle dose, pre-reperfusioni.p.Reduced infarct size, brain edema, and inflammationInhibition of NF-κB, MMP-9, and caspase-3; BBB protection
Liu et al. [242]MCAO in rats300 mg/kg7 daysi.p.Increased neurogenesis; reduced neurological deficitsActivation of Notch signaling
Wu et al. [243]MCAO/reperfusion in rats300 mg/kgSingle dose, pre-MCAOi.p.Reduced BBB leakage; increased tight junction protein expressionUpregulation of ZO-1 and occludin; inhibition of NF-κB, MMP-9
Liu et al. [244]Distal MCAO in mice150 mg/kg10 daysi.p.Reduced infarct size; improved sensorimotor recovery; M2 microglia shiftAnti-inflammatory activity; enhanced M2 polarization
Wu et al. [245]Primary cortical neurons (OGD/R); MCAO in rats2.5–25 μM (in vitro); dose not specified (in vivo)1 h OGD + 24 h reoxygenation (in vitro); 60 min MCAO (in vivo)In vitro; likely i.p. (in vivo)Reduced neuronal injury and oxidative stress; improved cell viability; reduced infarct volumeActivation of Akt/Nrf2 pathway; NQO1 upregulation; effect blocked by PI3K inhibitor LY294002
Jiang et al. [246]MCAO in rats; astrocytes; brain capillary endothelial cells (BCECs)0.5–2.0 mg/kgSingle i.v. injection, 30 min post-reperfusion (effects at 48 h)i.v.Reduced infarct size, BBB permeability, brain edema, and mortality; protected BCECsBBB preservation via ONOO inhibition; suppression of iNOS/NO pathway
Zhang et al. [247]MCAO in rats25 mg/kgSingle dosei.p.Reduced brain edema and infarct size; decreased TNF-α, IL-6, p53, Bax; increased Bcl-2 and Sirt1; improved mitochondrial functionAnti-inflammatory and anti-apoptotic activity; mitochondrial protection
AKT—protein kinase B, Bax—Bcl-2-associated X protein, Bcl-2—B-cell lymphoma 2, BBB—blood–brain barrier, BCCAO—bilateral common carotid artery occlusion, BCECs—brain capillary endothelial cells, CAT—catalase, GPx—glutathione peroxidase, ICH—intracerebral hemorrhage, IL-6—interleukin-6, i.c.v.—intracerebroventricular, iNOS—inducible nitric oxide synthase, i.p.—intraperitoneal, i.v.—intravenous, MCAO—middle cerebral artery occlusion, MMP-9—matrix metalloproteinase-9, MTM—metformin, NF-κB—nuclear factor kappa-light-chain-enhancer of activated B cells, NLRP3—NOD-like receptor pyrin domain-containing protein 3, NQO1—NAD(P)H quinone oxidoreductase 1, Nrf2—nuclear factor erythroid 2–related factor 2, OGD/R—oxygen–glucose deprivation/reoxygenation, ONOO—peroxynitrite, p53—tumor suppressor protein p53, PI3K—phosphoinositide 3-kinase, Prdx6—peroxiredoxin 6, Sirt1—sirtuin 1, SOD—superoxide dismutase, SP1—specificity protein 1, TNF-α—tumor necrosis factor alpha, ZO-1—zonula occludens-1.
Table 4. Summary of human clinical trials investigating the effects of curcumin in Alzheimer’s disease, Parkinson’s disease, and stroke.
Table 4. Summary of human clinical trials investigating the effects of curcumin in Alzheimer’s disease, Parkinson’s disease, and stroke.
Ref.PopulationDoseDurationRouteKey EffectsMechanisms
Small et al. [255]40 non-demented adults (51–84 y)180 mg/day Theracurmin® (90 mg BID)18 monthsOralImproved verbal and visual memory; enhanced attention; reduced amyloid and tau deposition (PET)Anti-inflammatory; anti-amyloidogenic; Akt/Nrf2 activation
Cox et al. [256]60 healthy adults (60–85 y)400 mg/day Longvida®4 weeksOralAcute: improved attention and working memory (1 h); Chronic: improved mood, reduced fatigue and LDLAntioxidant; anti-inflammatory; modulation of lipid metabolism
Baum et al. [257]34 probable/possible AD patients1 g or 4 g/day6 monthsOralNo cognitive improvement (MMSE); safe; curcuminoids detected in plasmaPossible Aβ disaggregation; antioxidant; low bioavailability; improved capsule absorption
Ringman et al. [258]36 mild-to-moderate AD patients2 g or 4 g/day Curcumin C3 Complex®24 weeks double-blind + 24 weeks open-labelOralNo cognitive benefit (ADAS-Cog); well toleratedAnti-inflammatory; antioxidant; inhibition of amyloid-β aggregation
Rainey-Smith et al. [259]96 older adults (40–90 y)1500 mg/day Biocurcumax™ (BCM-95®)12 monthsOralPrevention of cognitive decline (MoCA) at 6 months; no other changesAntioxidant; anti-inflammatory
Ghodsi et al. [260]60 idiopathic PD patients (≥30 y)80 mg/day nanomicellar curcumin9 monthsOralNo motor or QoL improvement; mild GI adverse effectsAnti-inflammatory; anti-apoptotic
Maghbooli et al. [261]50 idiopathic PD patients (≥35 y)160 mg/day nanomicellar curcumin (80 mg BID)3 monthsOralImproved sleep quality and QoL (PDQ-39); no effect on fatigueNeuroprotective; antioxidant; anti-inflammatory
Donadio et al. [262]19 PD patients + 14 PD controls2 g/day curcumin–phospholipid (Meriva®)12 monthsOralReduced autonomic and non-motor symptoms; slowed progression; reduced phosphorylated α-synuclein depositionBBB penetration; anti-amyloidogenic; antioxidant; anti-inflammatory
Boshagh et al. [263]56 ischemic stroke patients in rehabilitation500 mg curcumin + 5 mg piperine/day12 weeksOralReduced hs-CRP, cholesterol, TG, CIMT, BP, weight, waist circumference; increased TACAnti-inflammatory; antioxidant; vascular plaque reduction; BP lowering; lipid profile improvement
Aβ—amyloid-β, ADAS-Cog—Alzheimer’s Disease Assessment Scale–Cognitive Subscale, AD—Alzheimer’s disease, Akt—protein kinase B, α-synuclein—alpha-synuclein, BBB—blood–brain barrier, BID—twice daily, BP—blood pressure, BCM-95®—Biocurcumax™ curcumin formulation, CIMT—carotid intima–media thickness, GI—gastrointestinal, hs-CRP—high-sensitivity C-reactive protein, LDL—low-density lipoprotein, Longvida®—lipidated curcumin formulation, Meriva®—curcumin–phospholipid complex, MMSE—Mini-Mental State Examination, MoCA—Montreal Cognitive Assessment, Nrf2—nuclear factor erythroid 2–related factor 2, PD—Parkinson’s disease, PDQ-39—Parkinson’s Disease Questionnaire-39, PET—positron emission tomography, QoL—quality of life, TAC—total antioxidant capacity, TG—triglycerides, Theracurmin®—highly bioavailable curcumin formulation, y—years.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lehoczki, A.; Fekete, M.; Jarecsny, T.; Zábó, V.; Szappanos, Á.; Csípő, T.; Lipécz, Á.; Major, D.; Fazekas-Pongor, V.; Varga, P.; et al. The Neuroprotective Role of Curcumin: From Molecular Pathways to Clinical Translation—A Narrative Review. Nutrients 2025, 17, 2884. https://doi.org/10.3390/nu17172884

AMA Style

Lehoczki A, Fekete M, Jarecsny T, Zábó V, Szappanos Á, Csípő T, Lipécz Á, Major D, Fazekas-Pongor V, Varga P, et al. The Neuroprotective Role of Curcumin: From Molecular Pathways to Clinical Translation—A Narrative Review. Nutrients. 2025; 17(17):2884. https://doi.org/10.3390/nu17172884

Chicago/Turabian Style

Lehoczki, Andrea, Mónika Fekete, Tamás Jarecsny, Virág Zábó, Ágnes Szappanos, Tamás Csípő, Ágnes Lipécz, Dávid Major, Vince Fazekas-Pongor, Péter Varga, and et al. 2025. "The Neuroprotective Role of Curcumin: From Molecular Pathways to Clinical Translation—A Narrative Review" Nutrients 17, no. 17: 2884. https://doi.org/10.3390/nu17172884

APA Style

Lehoczki, A., Fekete, M., Jarecsny, T., Zábó, V., Szappanos, Á., Csípő, T., Lipécz, Á., Major, D., Fazekas-Pongor, V., Varga, P., & Varga, J. T. (2025). The Neuroprotective Role of Curcumin: From Molecular Pathways to Clinical Translation—A Narrative Review. Nutrients, 17(17), 2884. https://doi.org/10.3390/nu17172884

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop