Skip to main content
Erschienen in: Journal of Neuroinflammation 1/2022

Open Access 01.12.2022 | Research

TLR4 mutation protects neurovascular function and cognitive decline in high-fat diet-fed mice

verfasst von: Nathalie Obadia, Giulia Andrade, Marina Leardini-Tristão, Letícia Albuquerque, Celina Garcia, Flavia Lima, Júlio Daleprane, Hugo C. Castro-Faria-Neto, Eduardo Tibiriçá, Vanessa Estato

Erschienen in: Journal of Neuroinflammation | Ausgabe 1/2022

download
DOWNLOAD
print
DRUCKEN
insite
SUCHEN

Abstract

Background

Metabolic syndrome (MS) is defined as a low-grade proinflammatory state in which abnormal metabolic and cardiovascular factors increase the risk of developing cardiovascular disease and neuroinflammation. Events, such as the accumulation of visceral adipose tissue, increased plasma concentrations of free fatty acids, tissue hypoxia, and sympathetic hyperactivity in MS may contribute to the direct or indirect activation of Toll-like receptors (TLRs), specifically TLR4, which is thought to be a major component of this syndrome. Activation of the innate immune response via TLR4 may contribute to this state of chronic inflammation and may be related to the neuroinflammation and neurodegeneration observed in MS. In this study, we investigated the role of TLR4 in the brain microcirculation and in the cognitive performance of high-fat diet (HFD)-induced MS mice.

Methods

Wild-type (C3H/He) and TLR4 mutant (C3H/HeJ) mice were maintained under a normal diet (ND) or a HFD for 24 weeks. Intravital video-microscopy was used to investigate the functional capillary density, endothelial function, and endothelial–leukocyte interactions in the brain microcirculation. Plasma concentrations of monocyte chemoattractant protein-1 (MCP-1), adipokines and metabolic hormones were measured with a multiplex immunoassay. Brain postsynaptic density protein-95 and synaptophysin were evaluated by western blotting; astrocytic coverage of the vessels, microglial activation and structural capillary density were evaluated by immunohistochemistry.

Results

The HFD-induced MS model leads to metabolic, hemodynamic, and microcirculatory alterations, as evidenced by capillary rarefaction, increased rolling and leukocyte adhesion in postcapillary venules, endothelial dysfunction, and less coverage of astrocytes in the vessels, which are directly related to cognitive decline and neuroinflammation. The same model of MS reproduced in mice deficient for TLR4 because of a genetic mutation does not generate such changes. Furthermore, the comparison of wild-type mice fed a HFD and a normolipid diet revealed differences in inflammation in the cerebral microcirculation, possibly related to lower TLR4 activation.

Conclusions

Our results demonstrate that TLR4 is involved in the microvascular dysfunction and neuroinflammation associated with HFD-induced MS and possibly has a causal role in the development of cognitive decline.
Hinweise

Supplementary Information

The online version contains supplementary material available at https://​doi.​org/​10.​1186/​s12974-022-02465-3.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Abkürzungen
ACh
Acetylcholine
BBB
Blood–brain barrier
BDNF
Neurotrophic factors derived from the brain
DEXA
Dual-energy X-ray absorptiometry
eNOS
Endothelial nitric oxide synthase
FITC
Fluorescein isothiocyanate
GFAP
Glial fibrillary acidic protein
GIP
Glucose-dependent insulinotropic polypeptide
HFD
High-fat diet
HR
Heart rate
IB4
Biotinylated isolectin B4
Iba1
Ionized calcium-binding adaptor molecule 1
LPS
Lipopolysaccharides
MCP-1
Monocyte chemoattractant protein-1
MS
Metabolic syndrome
MWM test
Morris water maze test
ND
Normal diet
NVU
Neurovascular unit
PSD-95
Postsynaptic density protein 95
SAP
Systolic arterial pressure
SFAs
Saturated fatty acids
TLR4
Toll-like 4 receptor
TLRs
Toll-like receptors

Background

Over the past two decades, studies to determine a mechanism that links the pathogenesis of obesity to insulin resistance and diabetes have revealed a close association between nutrient excess and innate immune system activation [1]. Understanding the pathophysiology and risk factors for a range of chronic diseases with the precursor concept of inflammation could help treat associated comorbidities. Metabolic syndrome (MS) is defined as a combination of risk factors that culminate in adverse outcomes [2, 3], including type 2 diabetes mellitus, cardiovascular disease [4, 5], and neurodegenerative diseases [6]. At least three risk factors are associated with MS: obesity in conjunction with hyperglycemia, high arterial blood pressure, and dyslipidemia, usually associated with hyperinsulinemia and insulin resistance [710]. Overnutrition is an environmental stimulus that can activate Toll-like receptor (TLR) pathways to mediate the development of MS-related disorders [11, 12].
There is evidence that the combination of obesity and hypertension can impair performance across various cognitive domains [13]. In addition, type 2 diabetes mellitus constitutes a strong risk factor for cognitive decline and Alzheimer’s disease, as well as generally poorer memory performance in adults [1416]. Animal studies have shown that a high-calorie diet impairs the structure and function of the hippocampus and cortex, critical brain regions for learning and memory [17]. The main factor conferring a risk of cognitive decline in obese and hyperglycemic individuals is neuroinflammation, which is characterized by the activation of the resident population of macrophages and microglia, initiating/priming an inflammatory state that leads to the establishment of low-grade neuroinflammation [18, 19].
The neurovascular unit (NVU), formed by neurons, interneurons, astrocytes, and basal lamina covered with smooth muscle cells, pericytes, endothelial cells, and extracellular matrix, enables cross-talk between neurons and microcirculation [20]. Each component is closely and reciprocally linked, establishing an anatomical and functional role, which results in a highly efficient system for regulating cerebral blood flow [21].
It has been proposed that changes in the NVU lead to brain dysfunction and damage, and there is a growing interest in exploring the potential contribution of neurovascular dysfunction to neurodegeneration [22, 23]. However, the mechanistic link between changes in the NVU components and cognitive decline in obese experimental models is not yet fully understood.
In the present study, we investigated the role of TLR4 in the brain microcirculation and cognitive performance of diet-induced MS mice. Our objectives were to 1) determine whether mice lacking functional TLR4 would be protected against microvascular brain disorders during high-fat diet (HFD)-induced MS and 2) identify the underlying NVU components, such as astrocytes, blood vessels, neurons, and microglia, that were altered during HFD-induced MS.

Methods

Animals and experimental protocol

We used C3H/HeJ and C3H/He mice. The animals were supplied from the central animal facilities of the Oswaldo Cruz Foundation, Brazil, and were housed in standard cages in a temperature-controlled room (22 ± 1 °C) with a 12 h light/dark cycle and water ad libitum. All experimental procedures were conducted in accordance with the internationally accepted principles for the Care and Use of Laboratory Animals and were approved by the Oswaldo Cruz Foundation Animal Welfare Committee (license # 038/2015). C3H/HeJ mice are used in many areas of research, including cancer, immunology, and inflammation, as well as sensory and cardiovascular biology; they have a mutation in the third exon of TLR4, which compromises TLR4 signaling [24, 25]. C3H/He mice were used as background control mice (WT).
At eight weeks of age, the wild-type (C3H/He) and TLR4 mutant (C3H/HeJ) mice were randomly divided into two groups for each mouse strain (n = 12 per group) and fed either standard chow (AIN-93 M; normolipid diet, ND) or a high-fat diet (HFD) for twenty-four weeks. The ND contained 14% protein, 77% carbohydrate, and 9% lipids, and the HFD contained 14% protein, 33% carbohydrate, 53% lipids, and 0.5% NaCl. Both diets were manufactured by Prag Soluções (Jaú, SP, Brazil). All animals were weighed at the beginning and end of the diet period. The groups are referred to as WT-ND, WT-HFD, TLR4-mut-ND, and TLR4-mut-HFD.

Hemodynamic measurements

Systolic arterial pressure (SAP) and heart rate (HR) were assessed by a noninvasive measurement of tail pressure in conscious animals at the beginning and in the 24th week of the experimental protocol using a computerized tail-cuff plethysmography system (BP-2000 Blood Pressure Analysis System; Visitech Systems, Apex, NC, USA). The animals were adapted to the apparatus for three consecutive days prior to undergoing their baseline measurements.

Morris water maze

One week before the end of the diet period, cognitive function was investigated by the Morris water maze (MWM) test, which consisted of a black circular pool (100 cm in diameter) that was conceptually divided into four equal, but imaginary, quadrants for the purpose of data analysis. The water temperature was set at 25 °C. A platform (10 cm2), which was hidden from the mouse’s view, was located 2 cm beneath the surface of the water, allowing the mouse to easily climb onto it once its presence was detected. The MWM was placed in a well-lit white room with several posters and other distal visual stimuli hanging on the walls to provide spatial cues. The training on the hidden platform (spatial) version of the MWM was carried out on 4 consecutive days [26]. On each day, the mice (WT-ND, n = 8; WT-HFD, n = 10; TLR4-mut-ND, n = 10; TLR4-mut-HFD, n = 10) were allowed to swim for 60 s until they found the hidden platform. The time required to find the platform was recorded. On Day 5, a probe trial was conducted immediately after the four-day period to evaluate memory retention. The probe trial involved removing the submerged platform from the pool and allowing mice to swim for 60 s in any of the four quadrants of the pool. The time spent in the quadrant where the platform had been previously located was recorded. An observer blinded to the animals’ treatment group performed all the analyses.

Analysis of body composition

Changes in body composition were determined by dual-energy X-ray absorptiometry (DEXA) using a densitometer (Lunar PIXImus; GE Medical Systems, Madison, WI, USA). This system employs a cone beam X-ray source generating energies of 35 and 80 keV and a flat 100 × 80 mm detector with individual pixel dimensions of 0.18 × 0.18 mm. The ratio of energy attenuation in the luminescent panel separates bone, lean tissue, and fat mass. A quality control procedure was routinely performed with a calibration phantom before imaging. Mice were anesthetized by intraperitoneal (IP) injection of 100 mg/kg ketamine and 10 mg/kg xylazine (Cristália, SP, Brazil) to ensure good immobilization and positioning during a five-minute acquisition [27]. Analyses were performed at baseline and at the end of the 24th week of diet, and the results were expressed as a percentage of body fat.

Cerebral intravital microscopy with epi-illumination and fluorescence

At the end of the diet period, all the animals were randomly assigned for microcirculatory analysis. Brain microcirculation of all groups of animals was observed via cranial window imaging, and the functional capillary density, leukocyte–endothelium interactions, and microvascular endothelial function were evaluated by fluorescence intravital video microscopy as previously described [28]. Briefly, the animals were anesthetized with ketamine (90 mg/kg, i.p.) and xylazine (10 mg/kg, i.p.) and fixed in a stereotaxic frame. Then, their left parietal bones were exposed via midline skin incisions, and a 5-mm-diameter cranial window (between the coronal and lambdoid sutures) was created with a high-speed drill. The dura mater and arachnoid membranes were subsequently excised and withdrawn to expose the cerebral microcirculation, after which the cranial windows were suffused with artificial cerebrospinal fluid (132 mM NaCl, 2.95 mM KCl, 1.71 mM CaCl2, 0.64 mM MgCl2, 24.6 mM NaHCO3, 3.71 mM dextrose, and 6.7 mM urea, pH 7.4, at 37 °C). After the intravenous administration of 0.1 ml of 1% fluorescein isothiocyanate (FITC)-labeled dextran (molecular weight of 150,000 kDa; Sigma Chemical Co., St. Louis, MO, USA), the brain capillaries were observed with epi-illumination at 460 to 490 nm using a 520-nm emission filter, and the images were acquired under a 10 × ocular and 10 × objective lens on a Zeiss Axioscope A1 microscope (Zeiss, Oberkochen, Germany), producing a final magnification of 100 × on the monitor. The capillary count was performed using cellSens Standard 1.9 Zen blue software (Zeiss, Oberkochen, Germany), was used for one min per field. Only the continuously perfused capillaries were counted to determine the mean functional capillary density, expressed as the number of capillaries/mm2.
In addition, this cranial window enabled visualization of in vivo leukocyte recruitment. To label circulating leukocytes, 0.3 mg/kg rhodamine 6G was injected intravenously. Fluorescent leukocytes were visualized by intravital microscopy on postcapillary venules via epi-illumination at 510 to 550 nm using a 590-nm emission filter. The leukocyte–endothelial interaction was evaluated by counting the number of leukocytes adhering to the venular wall (100 µm long) over 30 s and was expressed as the number of cells/min/100 µm. Leukocyte rolling was defined as the movement of white blood cells into the vessel at a speed lower than the circulating red blood cells and was expressed as the number of cells/min. These parameters were determined in brain surface venules with diameters ranging from 50 to 120 cm. Microvascular endothelial function was evaluated by the pial arteriolar responses to the topical application of the endothelium-dependent vasodilator acetylcholine (ACh) in all animal groups (10−6 mol L−1). The cranial window was suffused with the vasoactive substance over 5 min, and the arteriolar diameters were measured before and after the application of ACh. Vascular responses were expressed as percent changes from the baseline.

Tissue processing

After intravital microscopy, mice were euthanized by pentobarbital overdose (150 mg/kg, i.p.). The brain was dissected, and the left hemisphere cortex regions were stored in a freezer at − 80 °C for subsequent Western blot analysis.
For immunohistochemical and adipocyte analyses, 5 animals per group were anesthetized with ketamine and xylazine (90 and 10 mg/kg i.p., respectively), whole blood was collected for subsequent metabolic analyses, and the mice were transcardially perfused with 0.9% saline followed by 4% paraformaldehyde. The brains and epididymal white adipose tissue were collected and maintained in paraformaldehyde solution.

Biochemical assays to evaluate metabolic function

The glucose test was carried out for whole blood using an automatic glucometer (One Touch Ultra 2®; Johnson & Johnson Medical S.A., Buenos Aires, Argentina). The blood was then centrifuged (15,000 rpm), and the serum was collected for determination of the insulin level by enzyme-linked immunosorbent assay (ELISA) using a commercial kit (Millipore, Burlington, MA, USA) following the manufacturer’s protocol. Insulin resistance was determined using the index of homeostasis model assessment (HOMA-IR) using the following formula [29]: HOMA-IR = fasting insulin (μU/ml) × fasting glucose (mmol/l)/22.5.

Measurement of plasma concentrations of adipokines and metabolic hormones

Adipokines and metabolic hormones were measured in plasma that had been stored at − 80 °C after collection. The plasma concentrations of monocyte chemoattractant protein-1 (MCP-1), ghrelin, resistin, leptin, gastric inhibitory peptide (GIP) and glucagon were measured with a microbead-based Luminex multiplex immunoassay, the MILLIPLEX® MAP Kit for the Mouse Metabolic Hormone Expanded Panel (Millipore, Darmstadt, Germany), following the manufacturer’s protocol. These specific analytes were chosen for their effects on different (patho)physiological mechanisms: inflammation, glucose metabolism, and obesity-related hormones. All adipokine concentrations were reported in pg/ml.

Adipocyte area quantification

Epididymal white adipose tissue sections were H&E stained and imaged at 40 × magnification (ZEISS Primo Star light microscopy Oberkochen, Germany). White adipocyte area was calculated using ImageJ software (NIH, Bethesda, MD, USA) by drawing ellipses circumscribing white adipocytes. The scale was set to 8 pixels per micron based on the pixel length of a 100-μm scale bar at 40 × magnification. Two to three images, each from a different area of a given sample, were captured per animal. Adipocyte area was measured for 10–20 adipocytes per image (25–40 adipocytes per mouse) and averaged on a per mouse basis.

Western blot analysis

Protein levels of synaptophysin and postsynaptic density protein 95 (PSD-95) in the whole brain were determined by Western blot analysis. The brain was homogenized in 50 mM Tris, pH 8.0, containing 150 mM NaCl, 1 mM EDTA, 1% Triton X-100, 0.5% sodium deoxycholate, 0.1% SDS, and protease inhibitor cocktail (Complete Mini Protease Inhibitor Cocktail; Roche Diagnostics, Indianapolis, IN, USA). Homogenates were incubated for 30 min on ice, sonicated for 3 min, and centrifuged for 10 min at 10,000 × g and 4 °C. The supernatant was collected, and the protein concentration was measured using the bicinchoninic acid method with a BCA Protein Assay Kit (Thermo Scientific, Waltham, MA, USA), with bovine serum albumin (BSA) as the standard. Then, 50 μg of protein per lane was separated on a 12% SDS–PAGE gel and transferred to a nitrocellulose membrane (Bio-Rad Laboratories, Munich, Germany). After blocking in 5% nonfat milk in TBS-Tween-0.05% (blocking solution) for 2 h at room temperature, the membranes were incubated overnight at 4 °C with antibodies to synaptophysin (Abcam, ab8049, 1:000) and PSD-95 (Abcam, ab18258, 1:1000) in blocking solution. Then, TBS-Tween solution was used to wash the membrane in a shaker three times for 5 min per wash. After that, the membranes were incubated with the following secondary antibodies: horse antimouse IgG antibody (H + L), peroxidase (Vector Laboratories, PI-2000, 1:10,000), goat antirabbit IgG antibody (H + L), and peroxidase (Vector Laboratories, PI-1000, 1:10,000) in blocking solution for 1 h at room temperature. Then, the membranes were washed in a shaker three times for 5 min per wash followed by three times for 10 min per wash in TBS-Tween. Finally, using a hypercassette™ (Amersham Pharmacia Biotech, UK), the nitrocellulose membrane was exposed to the luminescent agent (A:B = 1:1) in a chemiluminescence film (Amersham Hyperfilm™, GE Healthcare Limited).
Beta-tubulin (Proteintech, 66240-1-Ig, 1:20,000) was used as a loading control. Densitometry analysis was performed by Odyssey software (Li-Cor Biosciences, Lincoln, NE, USA).

Immunohistochemical analysis

For immunohistochemistry analysis, the brains were quickly excised after perfusion fixation as described above and serially sectioned at 50 μm. The sections were washed with PBS and incubated with 10% NGS diluted in PBS with 0.3% Triton X-100 for 90 min. The sections were then incubated with antibodies to glial fibrillary acidic protein (GFAP, Dako Glostrup, Denmark; 1:400), to ionized calcium-binding adaptor molecule 1 (Iba1, Wako; Tokyo, Japan 1:200), to TLR4 (MTS510; Santa Cruz Biotechnology, INC) or with biotinylated isolectin B4 (IB4; Vector Laboratories California, USA 1:100) overnight at 4 °C, washed again with PBS and incubated with secondary antibodies conjugated with Alexa Fluor 488 or 546 (Sigma–Aldrich Missouri, USA 1:400) or streptavidin-Cy3 (Sigma–Aldrich Missouri, USA 1:400) for 2 h. The sections were counterstained with DAPI, and coverslips were mounted with fluoromount (Vectashield—Vector Laboratories California, USA). Negative controls were treated with nonimmune rabbit IgG (Sigma–Aldrich Missouri, USA). Slices were imaged using a confocal microscope (Zeiss LSM 510 META; Zeiss, Oberkochen, Germany) equipped with a 63 × NA 1.40 oil immersion objective. The same selection criteria were carefully applied to all slices. The CA1 region of the hippocampus was first located followed by the parietal cortex.
For microglial morphology analysis, the ImageJ Sholl Analysis tool was used. This is a quantitative analysis method commonly used in neuronal studies to characterize the morphological characteristics of an imaged neuron. Using Sholl analysis, a mathematical algorithm of the program called the “branching index” is used to analyze neuronal morphology [30]. This index compares the difference in the number of intersections by making consecutive circles in relation to the distance of the neuronal sum. In this study, microglial processes were quantified as the number of intersections from 2 to 40 μm from the sum and the total number of intersections within the analysis circles, representing the total branches of the microglia.
To quantify morphological changes in the IBA1+ cells, consecutive Z-stack images were converted to a maximum-intensity projection image in ImageJ software (NIH, Bethesda, MD, USA). Using the Sholl analysis plugin, concentric circles were created, centered on the soma, beginning at 5.5-μm radii and increasing 2 μm with every circle. The number of intersections made by microglial branching processes with each successive increasing circle and the maximum number of intersections for the cell (Nm) were determined, as well as the critical value at which Nm occurred and the maximum length at which a branch intersection was observed [31].
For the analysis of the cerebral microvascular angioarchitecture, the AngioTool program was used (https://​ccrod.​cancer.​gov/​confluence/​display/​ROB2/​Downloads) to determine the total length of the brain capillaries in each image obtained by confocal microscopy. AngioTool is a validated source for measuring vascular networks [32] and has already been described in analyses of murine brain and retinal angiogenesis [33].
Confocal images of vessels immunostained with IB4 isolectin were analyzed using the AngioTool program and were processed as described [34]. Images for quantitative analysis were acquired with a confocal microscope using a 20 × objective (oil) zoom (200 ×). The image was immediately segmented and skeletonized, and the analysis of the vasculature was performed automatically. The resulting image shows the overlay of the area of all vessels and the computed branching points inside the imaged area.

Statistical analysis

The results are expressed as the mean ± standard error of the mean (SEM) for each group. Between-group statistical comparisons were performed using one-way analysis of variance (ANOVA), followed by Bonferroni’s post hoc test. When appropriate, the results were analyzed using unpaired Student’s t tests. Differences with p values less than 0.05 were considered significant. All calculations were performed using commercially available statistical software (GraphPad Prism, La Jolla, CA, USA).

Results

TLR4-mutant mice do not develop metabolic syndrome

Table 1 summarizes the average measures of the mice in the WT and TLR4-mut groups fed either a normolipid diet, ND, or a high fat diet, HFD. We observed marked increases in abdominal fat and plasma fasting glucose, as well as increased blood pressure, insulin levels, HOMA-IR, and cholesterol parameters, in the WT-HFD group compared to the WT-ND group. However, the TLR4-mut-HFD group did not present any significant difference in these parameters compared to the TLR4-mut-ND group. There were increases in cardiometabolic parameters in the TLR4 mutant mice fed a normal diet as compared to their WT controls; however, these increases were not statistically significant. In addition, the measures were also higher in the WT-HFD group than in the TLR4-mut-HFD group.
Table 1
Effects of high fat diet consumption on the cardiometabolic parameters of WT or TLR4-mut mice
Parameters
WT
TLR4-mut
ND
HFD
ND
HFD
Body weight (g)
38 ± 1
41 ± 0.8
36 ± 0.7
39 ± 0.6
Abdominal fat (g)
1.53 ± 0.1
2.47 ± 0.2**
1.7 ± 0.1
1.9 ± 0.1#
Body fat (%)
35.6 ± 1.4
47.2 ± 1.5**
37.5 ± 2.9
43.3 ± 1.2
SBP (mmHg)
94.5 ± 1.3
151.4 ± 2.4*
94.1 ± 2.5
98.7 ± 2.9#
Fasting glucose (mg/dl)
101.3 ± 13.5
170.9 ± 10.7**
122.7 ± 8.2
129.6 ± 7.5#
Fasting insulin (ng/ml)
0.21 ± 0.1
0.67 ± 0.3*
0.29 ± 0.7
0.26 ± 0.7#
HOMA-IR
1.51 ± 0.7
7.82 ± 3.7*
2.57 ± 0.7
2.43 ± 0.7#
Total cholesterol (mg/dl)
120.1 ± 18.1
180.2 ± 12.2*
140.4 ± 15.6
139.9 ± 6.3#
Triglycerides (mg/dl)
59.2 ± 7.9
264.6 ± 29.2***
47.2 ± 6.4
58.4 ± 4.7###
SBP systolic blood pressure, HOMA-IR homeostatic model assessment. Values represent the mean ± SEM of 8–12 animals per group. *p < 0.05, **p < 0.01, ***p < 0.001 vs. WT-ND, and #p < 0.05 and ###p < 0.001 vs. WT-HFD

TLR4 mutation mitigates adipocyte hypertrophy in the HFD-induced metabolic syndrome mouse model

The changes in adipocyte size in HFD-fed mice are shown in Fig. 1. WT-HFD mice had significantly larger adipocytes from the epididymal adipose tissue than all other groups. While the adipocytes of TLR4-mut-HFD mice were smaller than those of the WT-HFD group, there was no difference in the size compared to those of the WT-ND and TLR4-mut-ND mice (Fig. 1A, B).
In addition to significant differences in adipocyte size, histological examination indicated macrophage infiltration in the adipose tissue of WT-HFD mice but not in control mice (WT-ND) (Fig. 1A, B). Thus, we next decided to evaluate the levels of metabolic hormones in the plasma of these animals. Table 2 shows increased levels of ghrelin, and gastric inhibitory peptide (GIP), the leptin levels were increased but this were not statistically relevant. The glucagon levels decreased in the WT-HFD group compared with the WT-ND group. The TLR4-mut-HFD mice did not show any difference in the levels of these hormones when compared with the TLR4-mut-ND group.
Table 2
Effects of high fat diet consumption on the metabolic hormones of WT or TLR4-mut mice
Metabolic hormones (pg/mL)
WT
TLR4-mut
ND
HFD
ND
HFD
Leptin
2323 ± 166
3302 ± 407
1868 ± 777
2535 ± 312
Resistin
4730 ± 499
8649 ± 145
3221 ± 499
4495 ± 517#
Glucagon
11.9 ± 0.2
7.3 ± 0.9
6.5 ± 1.0
5.1 ± 0.7
Ghrelin
6.9 ± 0.3
11.2 ± 1.9*
8.5 ± 0.5
8.20 ± 0.8
GIP
220 ± 19
430 ± 134*
211 ± 44
237 ± 45
GIP glucose-dependent insulinotropic polypeptide. Values represent the mean ± SEM of 8–12 animals per group. *p < 0.05 vs. WT-ND and #p < 0.05 versus WT-HFD

TLR4 mutation prevents cerebral microcirculatory rarefaction and endothelial dysfunction in the brains of mice with HFD-induced metabolic syndrome

To investigate the potential protective effects of TLR4 mutation in an experimental model of MS, we performed intravital microscopic examinations at the end of the diet period. Figure 2 shows the influence of 24 weeks of HFD ingestion on the brain microcirculation of WT and TLR4-mut mice. The TLR4-mut mice fed the HFD did not show differences in the numbers of spontaneously perfused capillaries in their brain cortices compared to those fed the ND. However, the WT mice with MS (WT-HFD) exhibited a lower number of spontaneously perfused capillaries than the WT control mice (WT-ND) (Fig. 2A and C).
Intravital assessments of cerebrovascular function revealed that microvascular vasodilator responses to ACh were blunted in the WT-HFD group compared with the WT-ND group, whereas vasodilator responses to ACh were restored in the TLR4-mut-HFD and TLR4-mut-ND groups (Fig. 2B).

TLR4 mutation mitigates the recruitment of adherent leukocytes into the cerebral microvasculature of HFD-induced metabolic syndrome mice

Intravital fluorescence microscopic examinations of the cerebral venules of WT-HFD mice showed marked increases in the numbers of rolling and firmly adherent leukocytes in comparison with those of the WT-ND group, while in the TLR4-mut-HFD group, this inflammatory event was not observed in the cerebral microvasculature (Fig. 3A, B and D). In addition, the levels of MCP-1, a potent chemotactic factor for monocytes, were significantly reduced in TLR4-mut-HFD mice compared to WT-HFD mice (Fig. 3C).

TLR4 mutation protects animals from metabolic syndrome-associated cognitive decline

Figure 4 shows the effect of HFD-induced MS on the cognitive performance of WT and TLR4-mut mice, as assessed by the MWM test. WT-HFD mice took longer to find the platform from the third day (Fig. 4A) and spent a shorter time in the quadrant from which the platform was removed (Fig. 4B) than mice from the WT-ND group. TLR4-mut mice fed both the HFD and ND performed similarly in both tasks (Fig. 4), indicating that the TLR4 mutation protected animals from MS-associated cognitive decline.

TLR4 mutation restored synaptic protein expression in the brains of HFD-induced metabolic syndrome mice

The expression levels of proteins related to brain plasticity, synaptophysin and PSD-95, were evaluated by Western blot analysis of the whole brain. The HFD group of TLR4-mut mice showed a significant increase in synaptophysin compared to the WT-HFD group. However, there was no difference in the expression of PSD-95 in the TLR4-mut HFD group compared to the WT-HFD group (Fig. 5).

TLR4 mutation improves astrocyte interaction with the hippocampal vasculature in HFD-induced metabolic syndrome mice

Next, we examined brain vessel morphology, including astrocytes, microglia, and neurons. Figure 6 shows the quantitative results, with the corresponding segmented cortex images obtained from AngioToll software. Measurements of the following parameters were obtained: the total vessel length and the lacunarity. We observed a significant decrease in the total vessel length (p < 0.05) in mice of the WT-HFD group, while the lacunarity was higher than that in the WT-ND group. However, in the TLR4-mut-HFD group, there was no difference observed in either of these parameters.
We also examined the hippocampus, a region of the brain responsible for consolidating short- and long-term memory [35], which has been shown to be particularly prone to inflammation and damage during metabolic disease [36]. However, no alterations in the total vessel length or lacunarity were noticed in the WT-HFD and TLR4-mut-HFD groups when compared with their respective ND controls (data not shown).
Astrocytes and microglia and their interaction with the hippocampal vasculature are important features of NVU and blood–brain barrier (BBB) integrity (21). Using immunofluorescence, we quantified the numbers of Iba1+ microglia and assessed the area of astrocytes as denoted by GFAP+ staining. In analysis of colocalization of astrocytes and vessels in cerebral microcirculation (Fig. 7A), the coverage with astrocyte endfeet in the WT-HFD group was markedly decreased compared to that in the WT-ND group; however, in the TLR4-mut-HFD group we observed the same colocalization values as compared to the TLR4-mut-ND group (Fig. 7B).

HFD-induced metabolic syndrome changes the microglial morphology

We decided to examine microglial morphology in cortex and hippocampus of all groups of animals. We observed that the microglia of WT-HFD mice presented an amoeboid phenotype, with retracted processes as shown in Fig. 8 (in cortex: B and F; in hippocampus: M and Q), while the WT-ND group presented more ramified microglial morphology in sections of the cortex (Fig. 8A and E) and the hippocampus (Fig. 8L and P). This is supported by our quantitative data, as the maximum number of hippocampal microglia increased in WT-HFD (Fig. 8; T), and the number of branches decreased in both cortical (Fig. 8J and K) and hippocampal sections (Fig. 8U and V) compared to the WT-ND group. No difference was observed in the TLR4-mut-HFD group in either the cortex (Fig. 8D and H) or the hippocampus (Fig. 8O and S) compared to the TLR4-mut ND group in the cortex (Fig. 8C and G) or hippocampus (Fig. 8N and R), suggesting that microglial activation is an indicator of inflammation after MS and it is mitigated by TLR4 mutation (Fig. 8T–V).

Discussion

Studies of human and animal models provide strong evidence that alterations in glucose metabolism, such as hyperglycemia or MS, could have a negative effect on brain function [29, 37, 38]. In this study, we explored the specific contributions of a HFD to neuroinflammation in an experimental mouse model of MS. We demonstrate that long-term ingestion of a HFD induces region-specific effects on the brain, elevating microglial numbers, and decreasing the coverage of astrocytes in the cerebral vessels of the hippocampus. Moreover, we provide evidence that the ingestion of a HFD led to increased fasting blood glucose and abdominal fat, which are features of MS also observed in patients [39, 40]. Visceral adiposity plays a key role in the development of human MS and is associated with all MS criteria [41]. The increase in visceral adipose tissue leads to adipocyte hypertrophy, which is considered a key event in insulin resistance in obesity and MS, as it not only elevates proinflammatory cytokine levels but also leads to leptin resistance [4244]. We demonstrated that HFD intake in the WT group led to higher visceral fat accumulation that was not observed in TLR4 mutant mice; we also demonstrated in this study that a mutation in the TLR4 receptor reduces leukocyte–endothelium interactions, particularly the numbers of rolling and adherent leukocytes in the cerebral microvasculature. Moreover, TLR4 appears to contribute to the development of regional neuroinflammation in animal models of MS. We confirmed the upregulation expression of TLR4 in MS by immunostaining TLR4 in the brain cortex of the WT-HFD group as compared to the WT-ND group, and it was possible to correlate this increased expression with the observed metabolic syndrome-related changes. These results are shown in the Additional file 1.
An attenuation of obesity-induced adipose tissue inflammation has already been demonstrated in C3H/HeJ mice (which have a mutation in TLR4) when compared to their controls after ingestion of a HFD [24, 45]. TLR4 dysfunction in these animals suggests a fundamental role of this receptor in the diet-induced hyperlipidic inflammatory response [11]. In fact, there were increases in cardiometabolic parameters in the TLR4 mutant mice fed a normal diet compared to their WT controls; however, these increases were not statistically significant. In addition, it is important to emphasize that even with the nonsignificant increase in cardiometabolic parameters, this was not able to generate an inflammatory effect, reinforcing the importance in TLR4 as mediator of this process.
Saturated fatty acids (SFAs) can trigger inflammatory pathways, including those mediated by TLR4, similar to lipopolysaccharides (LPS) [46, 47] [48]. We believe that the source of fatty acids from the HFD used in the present study may be an important proinflammatory factor in the induction of MS. Indeed, some studies have shown that TLR4 activation in visceral adipose tissue is important in the genesis of the low-grade chronic inflammatory response observed in MS [11, 49, 50]. Once activated, these pathways can increase the synthesis and secretion of many chemokines (e.g., MCP-1) in adipocytes, contributing to insulin resistance and proinflammatory macrophage infiltration [50].
In the present work, we evaluated adipocytokines (insulin, leptin, resistin) and gut hormones (ghrelin and GIP). We observed increased levels of adipocytokines in the WT-HFD group compared to the WT-ND group but not in the TLR4-mut-HFD group, suggesting that this receptor may have an important role in insulin and leptin resistance in MS. Consistent with these findings, other studies have demonstrated that the deletion, knockdown or pharmacological inhibition of TLR4 protects mice from leptin and insulin resistance [45, 51].
An increase in resistin levels has been observed in obese patients compared to those with normal body mass. However, ghrelin, an appetite-regulating hormone secreted during fasting by gastrointestinal endocrine cells, plays an important role in weight regulation [52]. Increased levels of serum ghrelin concentration have been observed in pathological conditions, including insulin resistance and diabetes, whereas physical exercise and weight loss can reduce blood levels of this hormone [53]. We also observed increased levels of ghrelin and GIP in the WT-HFD group compared with the WT-ND group but not in the TLR4-mut-HFD group, again suggesting that TLR4 triggers metabolic and hormonal alterations leading to MS. The increase in resistin levels observed in TLR4-mut HFD mice was significantly inferior when compared to WT HFD mice. This finding demonstrates that TLR4 is involved in the inflammation process caused by resistin release by adipose tissue. This result corroborates previous studies showing the role of resistin on the activation of TLR4 and may explain the spectrum of resistin effects beyond inflammation and a loss of TLR4 function preventing diet-induced obesity and insulin resistance [32]
Endothelial dysfunction is an important component of MS and is caused by an imbalance between vasodilatory substances such as nitric oxide (˚NO), vasoconstricting substances such as endothelin, and prothrombotic factors such as plasminogen-1 activator inhibitors [54]. The decrease in ˚NO levels in individuals with MS is already well established [55]. In the physiological state, insulin stimulates the production of ˚NO by endothelial cells through an increase in endothelial nitric oxide synthase (eNOS) activity. This action is reduced or overturned in the case of insulin resistance, as is seen in MS [56]. In the present study, the vasodilatory response to acetylcholine evaluated in superficial arterioles of the cerebral cortex was abolished after the HFD period in the WT group but not in the TLR-mut group. Studies have already shown that 12 weeks of eating a HFD is sufficient to induce endothelial dysfunction in the cerebral microvascular bed [57]. Moreover, TLR4 contributes to endothelial dysfunction observed in hypertension, and anti-TLR4 treatment improves endothelium-dependent relaxation [58]. Our study also shows that HFD consumption induces functional capillary rarefaction in the brain, accompanied by increased vascular inflammation characterized by significant amounts of leukocyte rolling and adhesion to venular walls.
Several studies have reported that MS is increasingly related to cognitive decline [5962]. The reduction in cerebral blood flow is responsible for lower cognitive performance due to neuronal damage in brain regions related to memory and learning [63]. We employed the Morris water maze test, a complex task that assesses not only hippocampus-dependent reference memory changes but also spatial working memory and working memory load. Our results demonstrate that TLR4-mut mice, regardless of diet, learned the task, exhibiting a learning curve and a decreasing number of errors during the test. Conversely, the WT-HFD group exhibited worse performance on the Morris water maze test, indicating that the decline in cognitive function is prevented in TLR4 mutant mice.
The NVU includes neurons, astrocytes, pericytes, and microglia, as well as the blood vessels themselves, and enables tight regulation of blood flow through the brain vasculature [20]. Astrocytes support neuronal function by providing essential structural and nutritional support. As astrocytes are involved in neurotransmitter trafficking, they are also sensitive to circulating endocrine signals, such as the hormones ghrelin, glucagon‐like peptide‐1, and leptin, which have major impacts on central nervous system (CNS) mechanisms, controlling food intake and energy balance [64, 65]. Therefore, any disruption to astrocyte endfeet can have dramatic consequences on the integrity of the BBB, with consequent development of brain disorders [65]. In the present work, we used immunohistochemistry to assess the coverage of astrocytes in brain microvessels. Our results showed that the coverage of cerebral microvessels by astrocyte endfeet was markedly decreased in WT animals with MS compared to their controls; however, the TLR4 mutation appears to prevent this interruption in the endfeet coverage after chronic ingestion of the HFD.
To investigate whether differences in the neuronal and glial components could account for the diet-related changes found in the cognitive tests, brain cortical and hippocampal microglia morphology were analyzed. Microglia appear to play a key role in the inflammatory response that triggers neuronal death and synaptic dysfunction. Changes in microglial phenotype were observed, suggesting activated microglia in the cortex and hippocampus of WT-HFD mice that might indicate a greater degree of inflammation [32]. However, neither the HFD nor the ND TLR4-mut mice presented changes in microglial morphology, reinforcing our interpretation that TLR4 is a key player in the inflammatory component in MS. Under steady-state conditions, the number and function of microglia are tightly regulated, comprising the immune dominant cell type in the brain. In response to an acute immunological stimulus, microglia proliferate, increase phagocytosis, and eliminate pathogenic insult and later undergo apoptosis after the resolution of inflammation. In the case of a chronic stimulus, such as low-grade inflammation, characteristic of metabolic syndrome, microglia can become activated and remarkably long lasting, releasing cytokines and contributing to the perpetuation of a harmful environment [66]. In our study, microglia activation could be inferred by sholl analysis, which is considered the gold standard for studying microglial morphology and a good marker for an inflammatory response [67]; Furthermore, Intravital microscopy analysis in the brain was also performed to assess leukocyte–endothelium interactions, particularly the number of rolling and adherent leukocytes in the brain microvasculature. Rolling and adhesion of leukocytes were both increased in HFD animals and are regarded as classic and well characterized early events of the inflammatory process associated with both endothelial (with expression of adhesion molecules) and leukocyte activation. Finally, secretion of chemical mediators such as cytokines and chemokines are another hallmark of an inflammatory response, and in that sense, we showed elevated levels of monocyte chemoattractant protein-1 (MCP-1) HFD animals. However, a full inflammatory process is not undoubtedly demonstrated in our study. In addition, it is reasonable to assume that inflammation under the conditions of our study shall be low grade, not exuberant, and long lasting.
It has been suggested that prolonged feeding with a high-fat and carbohydrate-rich diet reduces neuroplasticity processes, as well as hippocampal innervation, which is also associated with cognitive impairment [68]. Neuronal plasticity reflects the ability of synapses to be modified in number and strength in response to physiological and pathological stimuli. We analyzed the expression levels of neurotrophic factors, such as synaptophysin, a presynaptic marker, and PSD-95, a postsynaptic marker. We wanted to understand the putative protective effect of the TLR4 mutation on neuronal plasticity in an HFD MS-induced model. Synapse-associated proteins, such as the postsynaptic density 95 (PSD-95) and synaptophysin, are essential for synaptic maturation, plasticity, learning, and cognition. [6971]. PSD-95 is a member of a family of proteins known as membrane-associated guanylate kinases [72] and the most abundant protein at the postsynaptic density [73]. In the same way, synaptophysin is another synaptic marker protein, located in the synaptic vesicle membrane and its increase indicates greater synaptic activity [74]. An increase in synaptophysin and PSD-95 occurs during the first decade of life, highlighting the critical role of these proteins in synaptogenesis [75]. It is worth noting that synaptic dysregulation is observed in neuroinflammation and neurological disorders, demonstrating a correlation between synaptic loss and cognitive impairments [76, 77]. In addition, therapeutic approaches that are able to induce and improve the expression of these proteins are associated with neuroprotection and improvement of cognitive function and memory [78, 79].
We observed that mice in the TLR4-mut group showed an increase in synaptophysin protein content, but this did not occur in the WT-HFD group. However, no significant alterations were observed in the PSD-95 protein content in either group. Several studies have already demonstrated that long-term consumption of HFD may be activating signaling pathways with deleterious effects in various regions of the brain [80, 81]. It is known that differences may occur in the synaptic activity of brain regions, such as the cortex and the hippocampus, the latter being the memory storage center and an important region for the memory of habituation to a new environment. We believe that no differences were observed in PSD-95, as the analyses were performed in the whole brain and not in specific brain regions, and we consider this one limitation of this study. Therefore, the results suggest that although the expression of synaptic proteins are not altered, the reduction in the astrocytic–vessel communication can greatly compromise the integrity of the NVU and, consequently, the cognitive function of the animals.
The relationship between MS and neurodegeneration has many unanswered questions. There is consensus that inflammation is an inherent problem in this syndrome, but it is still unclear whether a component of the high-fat diet is primarily responsible for inducing inflammation or whether the inflammatory profile of visceral adipose tissue is a critical trigger of this process. Furthermore, evaluating the local modulation of TLR4 would be an interesting pharmacological approach to prevent or treat cognitive decline in MS and could be the subject of future studies.

Conclusions

Our data suggest that TLR4 signaling activation is involved in microvascular dysfunction and neuroinflammation associated with HFD-induced MS and that this receptor possibly plays a causative role in the development of cognitive decline.

Acknowledgements

We would like to thank the Confocal Platform from Oswald Cruz Foundation, the Confocal Microscopy Unit from the Graduate Program in Morphological Sciences at the Federal University of Rio de Janeiro for their support with the immunohistochemistry images and Edson Fernandes de Assis for performing the biochemical assays in the Multiplex platform of the Instituto Oswaldo Cruz, FIOCRUZ, Rio de Janeiro.

Declarations

All experiments were conducted in accordance with internationally accepted principles for the care and use of laboratory animals and were approved by the Animal Ethics Committee of the Oswaldo Cruz Foundation (protocol number L-012/2021).
Not applicable.

Competing interests

The authors declare no competing interests.
Open AccessThis article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​. The Creative Commons Public Domain Dedication waiver (http://​creativecommons.​org/​publicdomain/​zero/​1.​0/​) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Literatur
1.
Zurück zum Zitat Castro-Barquero S, Ruiz-León AM, Sierra-Pérez M, Estruch R, Casas R. Dietary strategies for metabolic syndrome: a comprehensive review. Nutrients. 2020;12:2983.PubMedCentralCrossRef Castro-Barquero S, Ruiz-León AM, Sierra-Pérez M, Estruch R, Casas R. Dietary strategies for metabolic syndrome: a comprehensive review. Nutrients. 2020;12:2983.PubMedCentralCrossRef
2.
Zurück zum Zitat Wang HH, Lee DK, Liu M, Portincasa P, Wang DQ. Novel insights into the pathogenesis and management of the metabolic syndrome. Pediatr Gastroenterol Hepatol Nutr. 2020;23:189–230.PubMedPubMedCentralCrossRef Wang HH, Lee DK, Liu M, Portincasa P, Wang DQ. Novel insights into the pathogenesis and management of the metabolic syndrome. Pediatr Gastroenterol Hepatol Nutr. 2020;23:189–230.PubMedPubMedCentralCrossRef
4.
Zurück zum Zitat Grundy SM, Brewer HB Jr, Cleeman JI, Smith SC Jr, Lenfant C, American Heart A, National Heart L, Blood I. Definition of metabolic syndrome: report of the national heart, lung, and blood Institute/American Heart Association conference on scientific issues related to definition. Circulation. 2004;109:433–8.PubMedCrossRef Grundy SM, Brewer HB Jr, Cleeman JI, Smith SC Jr, Lenfant C, American Heart A, National Heart L, Blood I. Definition of metabolic syndrome: report of the national heart, lung, and blood Institute/American Heart Association conference on scientific issues related to definition. Circulation. 2004;109:433–8.PubMedCrossRef
5.
Zurück zum Zitat Punthakee Z, Goldenberg R, Katz P. Definition, classification and diagnosis of diabetes, prediabetes and metabolic syndrome. Can J Diabetes. 2018;42(Suppl 1):S10-s15.PubMedCrossRef Punthakee Z, Goldenberg R, Katz P. Definition, classification and diagnosis of diabetes, prediabetes and metabolic syndrome. Can J Diabetes. 2018;42(Suppl 1):S10-s15.PubMedCrossRef
6.
Zurück zum Zitat Campos-Pena V, Toral-Rios D, Becerril-Perez F, Sanchez-Torres C, Delgado-Namorado Y, Torres-Ossorio E, Franco-Bocanegra D, Carvajal K. Metabolic syndrome as a risk factor for Alzheimer’s disease: is abeta a crucial factor in both pathologies? Antioxid Redox Signal. 2017;26:542–60.PubMedCrossRef Campos-Pena V, Toral-Rios D, Becerril-Perez F, Sanchez-Torres C, Delgado-Namorado Y, Torres-Ossorio E, Franco-Bocanegra D, Carvajal K. Metabolic syndrome as a risk factor for Alzheimer’s disease: is abeta a crucial factor in both pathologies? Antioxid Redox Signal. 2017;26:542–60.PubMedCrossRef
7.
Zurück zum Zitat Nilsson PM, Tuomilehto J, Rydén L. The metabolic syndrome—what is it and how should it be managed? Eur J Prev Cardiol. 2019;26:33–46.PubMedCrossRef Nilsson PM, Tuomilehto J, Rydén L. The metabolic syndrome—what is it and how should it be managed? Eur J Prev Cardiol. 2019;26:33–46.PubMedCrossRef
8.
Zurück zum Zitat Engin A. The definition and prevalence of obesity and metabolic syndrome. Adv Exp Med Biol. 2017;960:1–17.PubMedCrossRef Engin A. The definition and prevalence of obesity and metabolic syndrome. Adv Exp Med Biol. 2017;960:1–17.PubMedCrossRef
9.
Zurück zum Zitat Gluvic Z, Zaric B, Resanovic I, Obradovic M, Mitrovic A, Radak D, Isenovic ER. Link between metabolic syndrome and insulin resistance. Curr Vasc Pharmacol. 2017;15:30–9.PubMedCrossRef Gluvic Z, Zaric B, Resanovic I, Obradovic M, Mitrovic A, Radak D, Isenovic ER. Link between metabolic syndrome and insulin resistance. Curr Vasc Pharmacol. 2017;15:30–9.PubMedCrossRef
10.
Zurück zum Zitat Rask Larsen J, Dima L, Correll CU, Manu P. The pharmacological management of metabolic syndrome. Expert Rev Clin Pharmacol. 2018;11:397–410.PubMedCrossRef Rask Larsen J, Dima L, Correll CU, Manu P. The pharmacological management of metabolic syndrome. Expert Rev Clin Pharmacol. 2018;11:397–410.PubMedCrossRef
12.
Zurück zum Zitat Wong SK, Chin KY, Ima-Nirwana S. Toll-like receptor as a molecular link between metabolic syndrome and inflammation: a review. Curr Drug Targets. 2019;20:1264–80.PubMedCrossRef Wong SK, Chin KY, Ima-Nirwana S. Toll-like receptor as a molecular link between metabolic syndrome and inflammation: a review. Curr Drug Targets. 2019;20:1264–80.PubMedCrossRef
13.
Zurück zum Zitat de Vries R, Morquecho-Campos P, de Vet E, de Rijk M, Postma E, de Graaf K, Engel B, Boesveldt S. Human spatial memory implicitly prioritizes high-calorie foods. Sci Rep. 2020;10:15174.PubMedPubMedCentralCrossRef de Vries R, Morquecho-Campos P, de Vet E, de Rijk M, Postma E, de Graaf K, Engel B, Boesveldt S. Human spatial memory implicitly prioritizes high-calorie foods. Sci Rep. 2020;10:15174.PubMedPubMedCentralCrossRef
14.
Zurück zum Zitat Pugazhenthi S, Qin L, Reddy PH. Common neurodegenerative pathways in obesity, diabetes, and Alzheimer’s disease. Biochim Biophys Acta Mol Basis Dis. 2017;1863:1037–45.PubMedCrossRef Pugazhenthi S, Qin L, Reddy PH. Common neurodegenerative pathways in obesity, diabetes, and Alzheimer’s disease. Biochim Biophys Acta Mol Basis Dis. 2017;1863:1037–45.PubMedCrossRef
15.
Zurück zum Zitat Fiore V, De Rosa A, Falasca P, Marci M, Guastamacchia E, Licchelli B, Giagulli VA, De Pergola G, Poggi A, Triggiani V. Focus on the Correlations between Alzheimer’s Disease and Type 2 Diabetes. Endocr Metab Immune Disord Drug Targets. 2019;19:571–9.PubMedCrossRef Fiore V, De Rosa A, Falasca P, Marci M, Guastamacchia E, Licchelli B, Giagulli VA, De Pergola G, Poggi A, Triggiani V. Focus on the Correlations between Alzheimer’s Disease and Type 2 Diabetes. Endocr Metab Immune Disord Drug Targets. 2019;19:571–9.PubMedCrossRef
16.
Zurück zum Zitat Srikanth V, Sinclair AJ, Hill-Briggs F, Moran C, Biessels GJ. Type 2 diabetes and cognitive dysfunction-towards effective management of both comorbidities. Lancet Diabetes Endocrinol. 2020;8:535–45.PubMedCrossRef Srikanth V, Sinclair AJ, Hill-Briggs F, Moran C, Biessels GJ. Type 2 diabetes and cognitive dysfunction-towards effective management of both comorbidities. Lancet Diabetes Endocrinol. 2020;8:535–45.PubMedCrossRef
17.
Zurück zum Zitat Tomassoni D, Martinelli I, Moruzzi M, Di MicioniBonaventura MV, Cifani C, Amenta F, Tayebati SK. Obesity and age-related changes in the brain of the Zucker Lepr (fa/fa) rats. Nutrients. 2020;12:1356.PubMedCentralCrossRef Tomassoni D, Martinelli I, Moruzzi M, Di MicioniBonaventura MV, Cifani C, Amenta F, Tayebati SK. Obesity and age-related changes in the brain of the Zucker Lepr (fa/fa) rats. Nutrients. 2020;12:1356.PubMedCentralCrossRef
18.
Zurück zum Zitat Chunchai T, Chattipakorn N, Chattipakorn SC. The possible factors affecting microglial activation in cases of obesity with cognitive dysfunction. Metab Brain Dis. 2018;33:615–35.PubMedCrossRef Chunchai T, Chattipakorn N, Chattipakorn SC. The possible factors affecting microglial activation in cases of obesity with cognitive dysfunction. Metab Brain Dis. 2018;33:615–35.PubMedCrossRef
19.
Zurück zum Zitat Kacířová M, Zmeškalová A, Kořínková L, Železná B, Kuneš J, Maletínská L. Inflammation: major denominator of obesity, Type 2 diabetes and Alzheimer’s disease-like pathology? Clin Sci (Lond). 2020;134:547–70.CrossRef Kacířová M, Zmeškalová A, Kořínková L, Železná B, Kuneš J, Maletínská L. Inflammation: major denominator of obesity, Type 2 diabetes and Alzheimer’s disease-like pathology? Clin Sci (Lond). 2020;134:547–70.CrossRef
21.
Zurück zum Zitat Hayden MR. Type 2 diabetes mellitus increases the risk of late-onset Alzheimer’s disease: ultrastructural remodeling of the neurovascular unit and diabetic Gliopathy. Brain Sci. 2019;9:262.PubMedCentralCrossRef Hayden MR. Type 2 diabetes mellitus increases the risk of late-onset Alzheimer’s disease: ultrastructural remodeling of the neurovascular unit and diabetic Gliopathy. Brain Sci. 2019;9:262.PubMedCentralCrossRef
22.
23.
Zurück zum Zitat Ahmad A, Patel V, Xiao J, Khan MM. The role of neurovascular system in neurodegenerative diseases. Mol Neurobiol. 2020;57:4373–93.PubMedCrossRef Ahmad A, Patel V, Xiao J, Khan MM. The role of neurovascular system in neurodegenerative diseases. Mol Neurobiol. 2020;57:4373–93.PubMedCrossRef
24.
Zurück zum Zitat Poggi M, Bastelica D, Gual P, Iglesias MA, Gremeaux T, Knauf C, Peiretti F, Verdier M, Juhan-Vague I, Tanti JF, et al. C3H/HeJ mice carrying a toll-like receptor 4 mutation are protected against the development of insulin resistance in white adipose tissue in response to a high-fat diet. Diabetologia. 2007;50:1267–76.PubMedCrossRef Poggi M, Bastelica D, Gual P, Iglesias MA, Gremeaux T, Knauf C, Peiretti F, Verdier M, Juhan-Vague I, Tanti JF, et al. C3H/HeJ mice carrying a toll-like receptor 4 mutation are protected against the development of insulin resistance in white adipose tissue in response to a high-fat diet. Diabetologia. 2007;50:1267–76.PubMedCrossRef
25.
Zurück zum Zitat Poltorak A, He X, Smirnova I, Liu M-Y, Van Huffel C, Du X, Birdwell D, Alejos E, Silva M, Galanos C. Defective LPS signaling in C3H/HeJ and C57BL/10ScCr mice: mutations in Tlr4 gene. Science. 1998;282:2085–8.PubMedCrossRef Poltorak A, He X, Smirnova I, Liu M-Y, Van Huffel C, Du X, Birdwell D, Alejos E, Silva M, Galanos C. Defective LPS signaling in C3H/HeJ and C57BL/10ScCr mice: mutations in Tlr4 gene. Science. 1998;282:2085–8.PubMedCrossRef
26.
Zurück zum Zitat Morris R. Developments of a water-maze procedure for studying spatial learning in the rat. J Neurosci Methods. 1984;11:47–60.PubMedCrossRef Morris R. Developments of a water-maze procedure for studying spatial learning in the rat. J Neurosci Methods. 1984;11:47–60.PubMedCrossRef
27.
Zurück zum Zitat Gargiulo S, Gramanzini M, Megna R, Greco A, Albanese S, Manfredi C, Brunetti A. Evaluation of growth patterns and body composition in C57Bl/6J mice using dual energy X-ray absorptiometry. Biomed Res Int. 2014;2014: 253067.PubMedPubMedCentralCrossRef Gargiulo S, Gramanzini M, Megna R, Greco A, Albanese S, Manfredi C, Brunetti A. Evaluation of growth patterns and body composition in C57Bl/6J mice using dual energy X-ray absorptiometry. Biomed Res Int. 2014;2014: 253067.PubMedPubMedCentralCrossRef
28.
Zurück zum Zitat Carvalho-Tavares J, Hickey MJ, Hutchison J, Michaud J, Sutcliffe IT, Kubes P. A role for platelets and endothelial selectins in tumor necrosis factor-alpha-induced leukocyte recruitment in the brain microvasculature. Circ Res. 2000;87:1141–8.PubMedCrossRef Carvalho-Tavares J, Hickey MJ, Hutchison J, Michaud J, Sutcliffe IT, Kubes P. A role for platelets and endothelial selectins in tumor necrosis factor-alpha-induced leukocyte recruitment in the brain microvasculature. Circ Res. 2000;87:1141–8.PubMedCrossRef
29.
Zurück zum Zitat Treviño S, Vázquez-Roque RA, López-López G, Perez-Cruz C, Moran C, Handal-Silva A, González-Vergara E, Flores G, Guevara J, Díaz A. Metabolic syndrome causes recognition impairments and reduced hippocampal neuronal plasticity in rats. J Chem Neuroanat. 2017;82:65–75.PubMedCrossRef Treviño S, Vázquez-Roque RA, López-López G, Perez-Cruz C, Moran C, Handal-Silva A, González-Vergara E, Flores G, Guevara J, Díaz A. Metabolic syndrome causes recognition impairments and reduced hippocampal neuronal plasticity in rats. J Chem Neuroanat. 2017;82:65–75.PubMedCrossRef
31.
32.
Zurück zum Zitat Agrawal R, Gomez-Pinilla F. “Metabolic syndrome” in the brain: deficiency in omega-3 fatty acid exacerbates dysfunctions in insulin receptor signalling and cognition. J Physiol. 2012;590:2485–99.PubMedPubMedCentralCrossRef Agrawal R, Gomez-Pinilla F. “Metabolic syndrome” in the brain: deficiency in omega-3 fatty acid exacerbates dysfunctions in insulin receptor signalling and cognition. J Physiol. 2012;590:2485–99.PubMedPubMedCentralCrossRef
33.
Zurück zum Zitat Arshad N, Lin TS, Yahaya MF. Metabolic syndrome and its effect on the brain: possible mechanism. CNS Neurol Disord Drug Targets. 2018;17:595–603.CrossRef Arshad N, Lin TS, Yahaya MF. Metabolic syndrome and its effect on the brain: possible mechanism. CNS Neurol Disord Drug Targets. 2018;17:595–603.CrossRef
34.
36.
Zurück zum Zitat Ivanova N, Liu Q, Agca C, Agca Y, Noble EG, Whitehead SN, Cechetto DF. White matter inflammation and cognitive function in a co-morbid metabolic syndrome and prodromal Alzheimer’s disease rat model. J Neuroinflammation. 2020;17:29.PubMedPubMedCentralCrossRef Ivanova N, Liu Q, Agca C, Agca Y, Noble EG, Whitehead SN, Cechetto DF. White matter inflammation and cognitive function in a co-morbid metabolic syndrome and prodromal Alzheimer’s disease rat model. J Neuroinflammation. 2020;17:29.PubMedPubMedCentralCrossRef
37.
Zurück zum Zitat Lima LCF, Saliba SW, Andrade JMO, Cunha ML, Cassini-Vieira P, Feltenberger JD, Barcelos LS, Guimarães ALS, de Paula AMB, de Oliveira ACP. Neurodegeneration alters metabolic profile and Sirt 1 signaling in high-fat-induced obese mice. Mol Neurobiol. 2017;54:3465–75.PubMedCrossRef Lima LCF, Saliba SW, Andrade JMO, Cunha ML, Cassini-Vieira P, Feltenberger JD, Barcelos LS, Guimarães ALS, de Paula AMB, de Oliveira ACP. Neurodegeneration alters metabolic profile and Sirt 1 signaling in high-fat-induced obese mice. Mol Neurobiol. 2017;54:3465–75.PubMedCrossRef
38.
Zurück zum Zitat Lemos C, Rial D, Gonçalves F, Pires J, Silva H, Matheus F, Da Silva A, Marques J, Rodrigues R, Jarak I. High sucrose consumption induces memory impairment in rats associated with electrophysiological modifications but not with metabolic changes in the hippocampus. Neuroscience. 2016;315:196–205.PubMedCrossRef Lemos C, Rial D, Gonçalves F, Pires J, Silva H, Matheus F, Da Silva A, Marques J, Rodrigues R, Jarak I. High sucrose consumption induces memory impairment in rats associated with electrophysiological modifications but not with metabolic changes in the hippocampus. Neuroscience. 2016;315:196–205.PubMedCrossRef
40.
Zurück zum Zitat Wali JA, Jarzebska N, Raubenheimer D, Simpson SJ, Rodionov RN, O’Sullivan JF. Cardio-metabolic effects of high-fat diets and their underlying mechanisms-a narrative review. Nutrients. 2020;12:1505.PubMedCentralCrossRef Wali JA, Jarzebska N, Raubenheimer D, Simpson SJ, Rodionov RN, O’Sullivan JF. Cardio-metabolic effects of high-fat diets and their underlying mechanisms-a narrative review. Nutrients. 2020;12:1505.PubMedCentralCrossRef
41.
Zurück zum Zitat Chait A, den Hartigh LJ. Adipose tissue distribution, inflammation and its metabolic consequences, including diabetes and cardiovascular disease. Front Cardiovasc Med. 2020;7:22.PubMedPubMedCentralCrossRef Chait A, den Hartigh LJ. Adipose tissue distribution, inflammation and its metabolic consequences, including diabetes and cardiovascular disease. Front Cardiovasc Med. 2020;7:22.PubMedPubMedCentralCrossRef
42.
Zurück zum Zitat Kloting N, Fasshauer M, Dietrich A, Kovacs P, Schon MR, Kern M, Stumvoll M, Bluher M. Insulin-sensitive obesity. Am J Physiol Endocrinol Metab. 2010;299:E506-515.PubMedCrossRef Kloting N, Fasshauer M, Dietrich A, Kovacs P, Schon MR, Kern M, Stumvoll M, Bluher M. Insulin-sensitive obesity. Am J Physiol Endocrinol Metab. 2010;299:E506-515.PubMedCrossRef
43.
Zurück zum Zitat Cotillard A, Poitou C, Torcivia A, Bouillot JL, Dietrich A, Kloting N, Gregoire C, Lolmede K, Bluher M, Clement K. Adipocyte size threshold matters: link with risk of type 2 diabetes and improved insulin resistance after gastric bypass. J Clin Endocrinol Metab. 2014;99:E1466-1470.PubMedCrossRef Cotillard A, Poitou C, Torcivia A, Bouillot JL, Dietrich A, Kloting N, Gregoire C, Lolmede K, Bluher M, Clement K. Adipocyte size threshold matters: link with risk of type 2 diabetes and improved insulin resistance after gastric bypass. J Clin Endocrinol Metab. 2014;99:E1466-1470.PubMedCrossRef
44.
Zurück zum Zitat de Git KC, Adan RA. Leptin resistance in diet-induced obesity: the role of hypothalamic inflammation. Obes Rev. 2015;16:207–24.PubMedCrossRef de Git KC, Adan RA. Leptin resistance in diet-induced obesity: the role of hypothalamic inflammation. Obes Rev. 2015;16:207–24.PubMedCrossRef
45.
Zurück zum Zitat Suganami T, Mieda T, Itoh M, Shimoda Y, Kamei Y, Ogawa Y. Attenuation of obesity-induced adipose tissue inflammation in C3H/HeJ mice carrying a Toll-like receptor 4 mutation. Biochem Biophys Res Commun. 2007;354:45–9. PubMedCrossRef Suganami T, Mieda T, Itoh M, Shimoda Y, Kamei Y, Ogawa Y. Attenuation of obesity-induced adipose tissue inflammation in C3H/HeJ mice carrying a Toll-like receptor 4 mutation. Biochem Biophys Res Commun. 2007;354:45–9. PubMedCrossRef
46.
Zurück zum Zitat Kim SJ, Choi Y, Choi YH, Park T. Obesity activates toll-like receptor-mediated proinflammatory signaling cascades in the adipose tissue of mice. J Nutr Biochem. 2012;23:113–22.PubMedCrossRef Kim SJ, Choi Y, Choi YH, Park T. Obesity activates toll-like receptor-mediated proinflammatory signaling cascades in the adipose tissue of mice. J Nutr Biochem. 2012;23:113–22.PubMedCrossRef
47.
Zurück zum Zitat Rocha DM, Caldas AP, Oliveira LL, Bressan J, Hermsdorff HH. Saturated fatty acids trigger TLR4-mediated inflammatory response. Atherosclerosis. 2016;244:211–5.PubMedCrossRef Rocha DM, Caldas AP, Oliveira LL, Bressan J, Hermsdorff HH. Saturated fatty acids trigger TLR4-mediated inflammatory response. Atherosclerosis. 2016;244:211–5.PubMedCrossRef
48.
Zurück zum Zitat AlZaim I, Hammoud SH, Al-Koussa H, Ghazi A, Eid AH, El-Yazbi AF. Adipose tissue immunomodulation: a novel therapeutic approach in cardiovascular and metabolic diseases. Front Cardiovasc Med. 2020;7: 602088.PubMedPubMedCentralCrossRef AlZaim I, Hammoud SH, Al-Koussa H, Ghazi A, Eid AH, El-Yazbi AF. Adipose tissue immunomodulation: a novel therapeutic approach in cardiovascular and metabolic diseases. Front Cardiovasc Med. 2020;7: 602088.PubMedPubMedCentralCrossRef
49.
Zurück zum Zitat Kopp A, Buechler C, Bala M, Neumeier M, Schölmerich J, Schäffler A. Toll-like receptor ligands cause proinflammatory and prodiabetic activation of adipocytes via phosphorylation of extracellular signal-regulated kinase and c-Jun N-terminal kinase but not interferon regulatory factor-3. Endocrinology. 2010;151:1097–108.PubMedCrossRef Kopp A, Buechler C, Bala M, Neumeier M, Schölmerich J, Schäffler A. Toll-like receptor ligands cause proinflammatory and prodiabetic activation of adipocytes via phosphorylation of extracellular signal-regulated kinase and c-Jun N-terminal kinase but not interferon regulatory factor-3. Endocrinology. 2010;151:1097–108.PubMedCrossRef
50.
Zurück zum Zitat Longo M, Zatterale F, Naderi J, Parrillo L, Formisano P, Raciti GA, Beguinot F, Miele C. Adipose tissue dysfunction as determinant of obesity-associated metabolic complications. Int J Mol Sci. 2019;20:2358.PubMedCentralCrossRef Longo M, Zatterale F, Naderi J, Parrillo L, Formisano P, Raciti GA, Beguinot F, Miele C. Adipose tissue dysfunction as determinant of obesity-associated metabolic complications. Int J Mol Sci. 2019;20:2358.PubMedCentralCrossRef
51.
Zurück zum Zitat Campolim CM, Weissmann L, Ferreira CKO, Zordão OP, Dornellas APS, de Castro G, Zanotto TM, Boico VF, Quaresma PGF, Lima RPA, et al. Short-term exposure to air pollution (PM(2.5)) induces hypothalamic inflammation, and long-term leads to leptin resistance and obesity via Tlr4/Ikbke in mice. Sci Rep. 2020;10:10160.PubMedPubMedCentralCrossRef Campolim CM, Weissmann L, Ferreira CKO, Zordão OP, Dornellas APS, de Castro G, Zanotto TM, Boico VF, Quaresma PGF, Lima RPA, et al. Short-term exposure to air pollution (PM(2.5)) induces hypothalamic inflammation, and long-term leads to leptin resistance and obesity via Tlr4/Ikbke in mice. Sci Rep. 2020;10:10160.PubMedPubMedCentralCrossRef
52.
Zurück zum Zitat Elamin MA, Youseif SM, Mohammed HA. Ghrelin, resistin and insulin in obese diabetic women in Wad-Madani, Sudan. Afr Health Sci. 2020;20:266–76.PubMedPubMedCentralCrossRef Elamin MA, Youseif SM, Mohammed HA. Ghrelin, resistin and insulin in obese diabetic women in Wad-Madani, Sudan. Afr Health Sci. 2020;20:266–76.PubMedPubMedCentralCrossRef
53.
Zurück zum Zitat Al Qarni AA, Joatar FE, Das N, Awad M, Eltayeb M, Al-Zubair AG, Ali ME, Al Masaud A, Shire AM, Gumaa K, Giha HA. Association of plasma ghrelin levels with insulin resistance in type 2 diabetes mellitus among saudi subjects. Endocrinol Metab (Seoul). 2017;32:230–40.CrossRef Al Qarni AA, Joatar FE, Das N, Awad M, Eltayeb M, Al-Zubair AG, Ali ME, Al Masaud A, Shire AM, Gumaa K, Giha HA. Association of plasma ghrelin levels with insulin resistance in type 2 diabetes mellitus among saudi subjects. Endocrinol Metab (Seoul). 2017;32:230–40.CrossRef
54.
Zurück zum Zitat Ahirwar AK, Jain A, Singh A, Goswami B, Bhatnagar MK, Bhatacharjee J. The study of markers of endothelial dysfunction in metabolic syndrome. Horm Mol Biol Clin Investig. 2015;24:131–6.PubMed Ahirwar AK, Jain A, Singh A, Goswami B, Bhatnagar MK, Bhatacharjee J. The study of markers of endothelial dysfunction in metabolic syndrome. Horm Mol Biol Clin Investig. 2015;24:131–6.PubMed
55.
Zurück zum Zitat Kowalski J, Sliwczyńska-Rodziewicz D, Kowalczyk E, Ciećwierz J, Irzmański R, Pawlicki L, Mejer A, Barylski M. Plasma nitric oxide and vascular endothelial growth factor levels in patients with metabolic syndrome and co-existing vascular complications. Pol Merkur Lekarski. 2011;30:249–52.PubMed Kowalski J, Sliwczyńska-Rodziewicz D, Kowalczyk E, Ciećwierz J, Irzmański R, Pawlicki L, Mejer A, Barylski M. Plasma nitric oxide and vascular endothelial growth factor levels in patients with metabolic syndrome and co-existing vascular complications. Pol Merkur Lekarski. 2011;30:249–52.PubMed
56.
Zurück zum Zitat Sarafidis PA, Bakris GL. Review: Insulin and endothelin: an interplay contributing to hypertension development? J Clin Endocrinol Metab. 2007;92:379–85.PubMedCrossRef Sarafidis PA, Bakris GL. Review: Insulin and endothelin: an interplay contributing to hypertension development? J Clin Endocrinol Metab. 2007;92:379–85.PubMedCrossRef
57.
Zurück zum Zitat Lynch CM, Kinzenbaw DA, Chen X, Zhan S, Mezzetti E, Filosa J, Ergul A, Faulkner JL, Faraci FM, Didion SP. Nox2-derived superoxide contributes to cerebral vascular dysfunction in diet-induced obesity. Stroke. 2013;44:3195–201.PubMedPubMedCentralCrossRef Lynch CM, Kinzenbaw DA, Chen X, Zhan S, Mezzetti E, Filosa J, Ergul A, Faulkner JL, Faraci FM, Didion SP. Nox2-derived superoxide contributes to cerebral vascular dysfunction in diet-induced obesity. Stroke. 2013;44:3195–201.PubMedPubMedCentralCrossRef
58.
Zurück zum Zitat De Batista PR, Palacios R, Martin A, Hernanz R, Medici CT, Silva MA, Rossi EM, Aguado A, Vassallo DV, Salaices M, Alonso MJ. Toll-like receptor 4 upregulation by angiotensin II contributes to hypertension and vascular dysfunction through reactive oxygen species production. PLoS ONE. 2014;9: e104020.PubMedPubMedCentralCrossRef De Batista PR, Palacios R, Martin A, Hernanz R, Medici CT, Silva MA, Rossi EM, Aguado A, Vassallo DV, Salaices M, Alonso MJ. Toll-like receptor 4 upregulation by angiotensin II contributes to hypertension and vascular dysfunction through reactive oxygen species production. PLoS ONE. 2014;9: e104020.PubMedPubMedCentralCrossRef
59.
Zurück zum Zitat Trevino S, Aguilar-Alonso P, Flores Hernandez JA, Brambila E, Guevara J, Flores G, Lopez-Lopez G, Munoz-Arenas G, Morales-Medina JC, Toxqui V, et al. A high calorie diet causes memory loss, metabolic syndrome and oxidative stress into hippocampus and temporal cortex of rats. Synapse. 2015;69:421–33.PubMedCrossRef Trevino S, Aguilar-Alonso P, Flores Hernandez JA, Brambila E, Guevara J, Flores G, Lopez-Lopez G, Munoz-Arenas G, Morales-Medina JC, Toxqui V, et al. A high calorie diet causes memory loss, metabolic syndrome and oxidative stress into hippocampus and temporal cortex of rats. Synapse. 2015;69:421–33.PubMedCrossRef
60.
Zurück zum Zitat Assuncao N, Sudo FK, Drummond C, de Felice FG, Mattos P. Metabolic syndrome and cognitive decline in the elderly: a systematic review. PLoS ONE. 2018;13: e0194990.PubMedPubMedCentralCrossRef Assuncao N, Sudo FK, Drummond C, de Felice FG, Mattos P. Metabolic syndrome and cognitive decline in the elderly: a systematic review. PLoS ONE. 2018;13: e0194990.PubMedPubMedCentralCrossRef
61.
Zurück zum Zitat Gómez-Gómez ME, Zapico SC. Frailty, cognitive decline, neurodegenerative diseases and nutrition interventions. Int J Mol Sci. 2019;20:2842.PubMedCentralCrossRef Gómez-Gómez ME, Zapico SC. Frailty, cognitive decline, neurodegenerative diseases and nutrition interventions. Int J Mol Sci. 2019;20:2842.PubMedCentralCrossRef
62.
Zurück zum Zitat Borshchev YY, Uspensky YP, Galagudza MM. Pathogenetic pathways of cognitive dysfunction and dementia in metabolic syndrome. Life Sci. 2019;237: 116932.PubMedCrossRef Borshchev YY, Uspensky YP, Galagudza MM. Pathogenetic pathways of cognitive dysfunction and dementia in metabolic syndrome. Life Sci. 2019;237: 116932.PubMedCrossRef
63.
Zurück zum Zitat Birdsill AC, Carlsson CM, Willette AA, Okonkwo OC, Johnson SC, Xu G, Oh JM, Gallagher CL, Koscik RL, Jonaitis EM, et al. Low cerebral blood flow is associated with lower memory function in metabolic syndrome. Obesity (Silver Spring). 2013;21:1313–20.CrossRef Birdsill AC, Carlsson CM, Willette AA, Okonkwo OC, Johnson SC, Xu G, Oh JM, Gallagher CL, Koscik RL, Jonaitis EM, et al. Low cerebral blood flow is associated with lower memory function in metabolic syndrome. Obesity (Silver Spring). 2013;21:1313–20.CrossRef
64.
Zurück zum Zitat Heithoff BP, George KK, Phares AN, Zuidhoek IA, Munoz-Ballester C, Robel S. Astrocytes are necessary for blood-brain barrier maintenance in the adult mouse brain. Glia. 2021;69:436–72.PubMedCrossRef Heithoff BP, George KK, Phares AN, Zuidhoek IA, Munoz-Ballester C, Robel S. Astrocytes are necessary for blood-brain barrier maintenance in the adult mouse brain. Glia. 2021;69:436–72.PubMedCrossRef
65.
Zurück zum Zitat Hallof-Büstrich H, Di Benedetto B. Examining the coverage of blood vessels by astrocyte endfeet in an animal model of major depressive disorder. Methods Mol Biol. 2019;1938:255–63.PubMedCrossRef Hallof-Büstrich H, Di Benedetto B. Examining the coverage of blood vessels by astrocyte endfeet in an animal model of major depressive disorder. Methods Mol Biol. 2019;1938:255–63.PubMedCrossRef
66.
Zurück zum Zitat Ledreux A, Wang X, Schultzberg M, Granholm A-C, Freeman LR. Detrimental effects of a high fat/high cholesterol diet on memory and hippocampal markers in aged rats. Behav Brain Res. 2016;312:294–304.PubMedPubMedCentralCrossRef Ledreux A, Wang X, Schultzberg M, Granholm A-C, Freeman LR. Detrimental effects of a high fat/high cholesterol diet on memory and hippocampal markers in aged rats. Behav Brain Res. 2016;312:294–304.PubMedPubMedCentralCrossRef
67.
Zurück zum Zitat Campagno KE, Lu W, Jassim AH, Albalawi F, Cenaj A, Tso H-Y, Clark SP, Sripinun P, Gómez NM, Mitchell CH. Rapid morphologic changes to microglial cells and upregulation of mixed microglial activation state markers induced by P2X7 receptor stimulation and increased intraocular pressure. J Neuroinflammation. 2021;18:1–18.CrossRef Campagno KE, Lu W, Jassim AH, Albalawi F, Cenaj A, Tso H-Y, Clark SP, Sripinun P, Gómez NM, Mitchell CH. Rapid morphologic changes to microglial cells and upregulation of mixed microglial activation state markers induced by P2X7 receptor stimulation and increased intraocular pressure. J Neuroinflammation. 2021;18:1–18.CrossRef
68.
Zurück zum Zitat Del Olmo N, Ruiz-Gayo M. Influence of high-fat diets consumed during the juvenile period on hippocampal morphology and function. Front Cel Neurosci. 2018;12:439.CrossRef Del Olmo N, Ruiz-Gayo M. Influence of high-fat diets consumed during the juvenile period on hippocampal morphology and function. Front Cel Neurosci. 2018;12:439.CrossRef
69.
Zurück zum Zitat Béïque JC, Andrade R. PSD-95 regulates synaptic transmission and plasticity in rat cerebral cortex. J Physiol. 2003;546:859–67.PubMedCrossRef Béïque JC, Andrade R. PSD-95 regulates synaptic transmission and plasticity in rat cerebral cortex. J Physiol. 2003;546:859–67.PubMedCrossRef
70.
Zurück zum Zitat Coley AA, Gao W-J. PSD-95 deficiency disrupts PFC-associated function and behavior during neurodevelopment. Sci Rep. 2019;9:1–13.CrossRef Coley AA, Gao W-J. PSD-95 deficiency disrupts PFC-associated function and behavior during neurodevelopment. Sci Rep. 2019;9:1–13.CrossRef
71.
Zurück zum Zitat Schmitt U, Tanimoto N, Seeliger M, Schaeffel F, Leube R. Detection of behavioral alterations and learning deficits in mice lacking synaptophysin. Neuroscience. 2009;162:234–43.PubMedCrossRef Schmitt U, Tanimoto N, Seeliger M, Schaeffel F, Leube R. Detection of behavioral alterations and learning deficits in mice lacking synaptophysin. Neuroscience. 2009;162:234–43.PubMedCrossRef
72.
Zurück zum Zitat Niethammer M, Kim E, Sheng M. Interaction between the C terminus of NMDA receptor subunits and multiple members of the PSD-95 family of membrane-associated guanylate kinases. J Neurosci. 1996;16:2157–63.PubMedPubMedCentralCrossRef Niethammer M, Kim E, Sheng M. Interaction between the C terminus of NMDA receptor subunits and multiple members of the PSD-95 family of membrane-associated guanylate kinases. J Neurosci. 1996;16:2157–63.PubMedPubMedCentralCrossRef
73.
74.
Zurück zum Zitat Becher A, Drenckhahn A, Pahner I, Margittai M, Jahn R, Ahnert-Hilger G. The synaptophysin–synaptobrevin complex: a hallmark of synaptic vesicle maturation. J Neurosci. 1999;19:1922–31.PubMedPubMedCentralCrossRef Becher A, Drenckhahn A, Pahner I, Margittai M, Jahn R, Ahnert-Hilger G. The synaptophysin–synaptobrevin complex: a hallmark of synaptic vesicle maturation. J Neurosci. 1999;19:1922–31.PubMedPubMedCentralCrossRef
75.
Zurück zum Zitat Glantz LA, Gilmore JH, Hamer RM, Lieberman JA, Jarskog LF. Synaptophysin and postsynaptic density protein 95 in the human prefrontal cortex from mid-gestation into early adulthood. Neuroscience. 2007;149:582–91.PubMedCrossRef Glantz LA, Gilmore JH, Hamer RM, Lieberman JA, Jarskog LF. Synaptophysin and postsynaptic density protein 95 in the human prefrontal cortex from mid-gestation into early adulthood. Neuroscience. 2007;149:582–91.PubMedCrossRef
76.
Zurück zum Zitat Liu Y, Zhang Y, Zheng X, Fang T, Yang X, Luo X, Guo A, Newell KA, Huang X-F, Yu Y. Galantamine improves cognition, hippocampal inflammation, and synaptic plasticity impairments induced by lipopolysaccharide in mice. J Neuroinflammation. 2018;15:1–15.CrossRef Liu Y, Zhang Y, Zheng X, Fang T, Yang X, Luo X, Guo A, Newell KA, Huang X-F, Yu Y. Galantamine improves cognition, hippocampal inflammation, and synaptic plasticity impairments induced by lipopolysaccharide in mice. J Neuroinflammation. 2018;15:1–15.CrossRef
77.
Zurück zum Zitat Shao CY, Mirra SS, Sait HB, Sacktor TC, Sigurdsson EM. Postsynaptic degeneration as revealed by PSD-95 reduction occurs after advanced Aβ and tau pathology in transgenic mouse models of Alzheimer’s disease. Acta Neuropathol. 2011;122:285–92.PubMedPubMedCentralCrossRef Shao CY, Mirra SS, Sait HB, Sacktor TC, Sigurdsson EM. Postsynaptic degeneration as revealed by PSD-95 reduction occurs after advanced Aβ and tau pathology in transgenic mouse models of Alzheimer’s disease. Acta Neuropathol. 2011;122:285–92.PubMedPubMedCentralCrossRef
78.
79.
Zurück zum Zitat Revilla S, Suñol C, García-Mesa Y, Giménez-Llort L, Sanfeliu C, Cristòfol R. Physical exercise improves synaptic dysfunction and recovers the loss of survival factors in 3xTg-AD mouse brain. Neuropharmacology. 2014;81:55–63.PubMedCrossRef Revilla S, Suñol C, García-Mesa Y, Giménez-Llort L, Sanfeliu C, Cristòfol R. Physical exercise improves synaptic dysfunction and recovers the loss of survival factors in 3xTg-AD mouse brain. Neuropharmacology. 2014;81:55–63.PubMedCrossRef
80.
Zurück zum Zitat Dutheil S, Ota KT, Wohleb ES, Rasmussen K, Duman RS. High-Fat Diet Induced Anxiety and Anhedonia: Impact on Brain Homeostasis and Inflammation. Neuropsychopharmacology. 2016;41:1874–87.PubMedPubMedCentralCrossRef Dutheil S, Ota KT, Wohleb ES, Rasmussen K, Duman RS. High-Fat Diet Induced Anxiety and Anhedonia: Impact on Brain Homeostasis and Inflammation. Neuropsychopharmacology. 2016;41:1874–87.PubMedPubMedCentralCrossRef
81.
Zurück zum Zitat Chen TJ, Chen SS, Wang DC, Hung HS. High-fat diet reduces novelty-induced expression of activity-regulated cytoskeleton-associated protein. J Cell Physiol. 2020;235:1065–75.PubMedCrossRef Chen TJ, Chen SS, Wang DC, Hung HS. High-fat diet reduces novelty-induced expression of activity-regulated cytoskeleton-associated protein. J Cell Physiol. 2020;235:1065–75.PubMedCrossRef
Metadaten
Titel
TLR4 mutation protects neurovascular function and cognitive decline in high-fat diet-fed mice
verfasst von
Nathalie Obadia
Giulia Andrade
Marina Leardini-Tristão
Letícia Albuquerque
Celina Garcia
Flavia Lima
Júlio Daleprane
Hugo C. Castro-Faria-Neto
Eduardo Tibiriçá
Vanessa Estato
Publikationsdatum
01.12.2022
Verlag
BioMed Central
Erschienen in
Journal of Neuroinflammation / Ausgabe 1/2022
Elektronische ISSN: 1742-2094
DOI
https://doi.org/10.1186/s12974-022-02465-3

Weitere Artikel der Ausgabe 1/2022

Journal of Neuroinflammation 1/2022 Zur Ausgabe

Neu in den Fachgebieten Neurologie und Psychiatrie

Prämenstruelle Beschwerden mit Suizidrisiko assoziiert

04.06.2024 Suizidalität Nachrichten

Manche Frauen, die regelmäßig psychische und körperliche Symptome vor ihrer Menstruation erleben, haben ein deutlich erhöhtes Suizidrisiko. Jüngere Frauen sind besonders gefährdet.

Bei seelischem Stress sind Checkpoint-Hemmer weniger wirksam

03.06.2024 NSCLC Nachrichten

Wie stark Menschen mit fortgeschrittenem NSCLC von einer Therapie mit Immun-Checkpoint-Hemmern profitieren, hängt offenbar auch davon ab, wie sehr die Diagnose ihre psychische Verfassung erschüttert

Demenzkranke durch Antipsychotika vielfach gefährdet

Demenz Nachrichten

Der Einsatz von Antipsychotika gegen psychische und Verhaltenssymptome in Zusammenhang mit Demenzerkrankungen erfordert eine sorgfältige Nutzen-Risiken-Abwägung. Neuen Erkenntnissen zufolge sind auf der Risikoseite weitere schwerwiegende Ereignisse zu berücksichtigen.

Schlaganfall: frühzeitige Blutdrucksenkung im Krankenwagen ohne Nutzen

31.05.2024 Apoplex Nachrichten

Der optimale Ansatz für die Blutdruckkontrolle bei Patientinnen und Patienten mit akutem Schlaganfall ist noch nicht gefunden. Ob sich eine frühzeitige Therapie der Hypertonie noch während des Transports in die Klinik lohnt, hat jetzt eine Studie aus China untersucht.