Skip to main content

Altered substrate metabolism in neurodegenerative disease: new insights from metabolic imaging

Abstract

Neurodegenerative diseases (NDs), such as Alzheimer’s disease (AD), Parkinson’s disease (PD) and multiple sclerosis (MS), are relatively common and devastating neurological disorders. For example, there are 6 million individuals living with AD in the United States, a number that is projected to grow to 14 million by the year 2030. Importantly, AD, PD and MS are all characterized by the lack of a true disease-modifying therapy that is able to reverse or halt disease progression. In addition, the existing standard of care for most NDs only addresses the symptoms of the disease. Therefore, alternative strategies that target mechanisms underlying the neuropathogenesis of disease are much needed. Recent studies have indicated that metabolic alterations in neurons and glia are commonly observed in AD, PD and MS and lead to changes in cell function that can either precede or protect against disease onset and progression. Specifically, single-cell RNAseq studies have shown that AD progression is tightly linked to the metabolic phenotype of microglia, the key immune effector cells of the brain. However, these analyses involve removing cells from their native environment and performing measurements in vitro, influencing metabolic status. Therefore, technical approaches that can accurately assess cell-specific metabolism in situ have the potential to be transformative to our understanding of the mechanisms driving AD. Here, we review our current understanding of metabolism in both neurons and glia during homeostasis and disease. We also evaluate recent advances in metabolic imaging, and discuss how emerging modalities, such as fluorescence lifetime imaging microscopy (FLIM) have the potential to determine how metabolic perturbations may drive the progression of NDs. Finally, we propose that the temporal, regional, and cell-specific characterization of brain metabolism afforded by FLIM will be a critical first step in the rational design of metabolism-focused interventions that delay or even prevent NDs.

Background

Despite being only 2% of the body’s mass, the brain consumes 20% of the body’s oxygen intake and 25% of its glucose intake [1]. This energy utilization is not uniform, with 70–80% of total energy consumed via neurons, and microglia, astrocytes and oligodendrocytes using the remaining portion [2,3,4]. Although the brain primarily consumes glucose to meet its energy demands, in disease states or starvation energy utilization can shift to other substrates [5,6,7,8,9]. During aging, glucose metabolism in the brain is impaired, a phenomenon that is exacerbated in neurodegenerative (ND) and neuroinflammatory diseases (NID) such as Alzheimer’s disease (AD), Parkinson’s disease (PD) and multiple sclerosis (MS). Specifically in AD, several changes to overall brain metabolism have been observed, including dysregulated oxidative metabolism, amino acid metabolism and creatine degradation [10], and impaired glucose transport [11]. In particular, the pathogenesis of AD is complex, and despite major research efforts, successful therapeutics that can prevent, or delay AD are lacking. Since perturbed brain metabolism has been repeatedly implicated in AD, understanding changes in substrate utilization prior to and during the development and progression of AD may highlight critical mechanisms that can be exploited to diagnose or treat AD.

Despite changes in brain metabolism being a potential mechanism driving the development of NDs, the field has been stymied by the lack of techniques that can accurately measure brain metabolism with sufficient temporal, spatial, and cell-specific resolution. On one hand, it is becoming widely appreciated that both neurons and glial cells, which have very different metabolic needs, contribute to disease pathology. Recent methodological advances have facilitated the identification of microglial subpopulations with defined metabolic phenotypes that may protect against or precede AD [12, 13]. However, these measurements have been made in isolated cells that have been subjected to mechanical or enzymatic stress, which likely modifies their metabolic state, underscoring the need for methodologies that can determine cell-specific metabolic changes in situ.

Specific brain regions have different cellular composition, functions, and therefore metabolic needs. Hence metabolic perturbations in certain regions may have a greater impact on disease state and severity. Nonetheless, existing “metabolic imaging” modalities, such as magnetic resonance imaging (MRI), magnetic resonance spectroscopy (MRS), and positron emission tomography (PET), have provided important insights regarding regional changes in metabolic processing that occur preceding and during NDs, such as AD [14]. Although these techniques are becoming increasingly advanced, they cannot discriminate between regional and cell-specific metabolic aberrations. In contrast, the development of two-photon imaging and fluorescence lifetime imaging microscopy (FLIM), enables direct measurement of the endogenous metabolic fluorophores, such as reduced nicotinamide adenine dinucleotide (NADH) and flavin adenine dinucleotide (FAD), which have very distinct lifetimes in free and bound-to-enzyme forms [15]. The ratio of free and bound-to-enzyme NADH or FAD can be determined within an individual cell or on a cell cluster in tissue, and therefore has the potential to measure cell and region-specific changes in brain metabolism that may precede or drive AD onset and progression. Here, we revisit our understanding of brain metabolism and review recent findings suggesting that changes in brain metabolism may be a potential target for the treatment of ND. We also highlight recent advances in metabolic imaging that enable precise monitoring of brain metabolism, guiding the design of interventions that may prevent or improve outcomes for individuals with NDs such as AD.

Metabolism in the homeostatic brain

Overview of metabolic processes in the brain

Energy usage is dependent on a given cell’s microenvironment, the substrates available for metabolism, and the role the cell has in maintaining homeostasis or fighting off disease. Generally, the pathways available for energy production are glycolysis, the tricarboxylic acid cycle (TCA cycle), and oxidative phosphorylation (OXPHOS). Glycolysis takes 1 glucose molecule, 2 nicotinamide adenine dinucleotide (NAD+) molecules, 2 adenine diphosphate (ADP), and 2 phosphate groups to yield 2 pyruvate molecules, 2 reduced nicotinamide adenine dinucleotide (NADH) molecules, 2 water (H2O) molecules, and 4 adenine triphosphate (ATP) molecules. These pyruvate molecules are converted to two acetyl coenzyme A (acetyl-CoA) before entering TCA cycle, where they undergo a series of reactions yielding either 2 guanosine triphosphate (GTP) or ATP molecules, 6 NADH molecules, 2 ubiquinone (QH2) molecules, 2 FADH2 molecules, and 4 carbon dioxide (CO2) molecules. The reducing agents generated by these pathways are used in oxidative phosphorylation to break covalent bonds in molecular oxygen (O2), converting chemical energy into a proton gradient. ATP synthase, an enzyme found in the mitochondrial inner membrane, then uses the proton gradient to phosphorylate ADP to ATP, a form of chemical energy that can be used throughout the cell.

Another pathway important in homeostasis and cell repair is the pentose phosphate pathway (PPP). This pathway is a parallel of glycolysis and converts glucose-6-phosphate to pentoses, and ribulose-5-phosphate which can be used in the synthesis of nucleotides. The PPP can be broken up into an oxidative phase where 2 nicotinamide adenine dinucleotide phosphate (NADP+) are reduced to NADPH via the oxidation of glucose-6-phosphate, and a non-oxidative phase where pentoses are generated. Unlike glycolysis, the function of the PPP is anabolic rather than catabolic.

The brain is composed of neurons and glial cells. Glial cells can be further categorized as ependymal cells, oligodendrocytes, microglia, and astrocytes. Below we will examine the preferred substrate utilization within neurons, astrocytes, microglia, and oligodendrocytes and metabolic cross-talk between them. We will also evaluate how alterations in these processes may contribute to ND.

Astrocytes and neurons

Neuronal activity, which involves action potentials and synaptic transmission, accounts for the majority of neuronal energy usage [3, 16]. At baseline, neurons are capable of generating ATP through complete oxidation of glucose to CO2 via glycolysis, but in a healthy state, were found to favor oxidative pathways, like the TCA cycle and OXPHOS [8, 17,18,19]. This stems from their inability to increase their glycolytic activity in response to cellular stress, which would lead to the generation and buildup of toxic glycation byproducts, ultimately causing apoptosis [8, 20, 21]. The substrate for neuronal OXPHOS is a source of controversy, with some reports pointing towards the existence of an astrocyte-neuronal lactate shuttle (ANLS) where astrocytes provide neurons with lactate to fuel the TCA cycle [8, 16, 17, 22,23,24,25] (see Fig. 1). Opponents of this theory base their opposition on theoretical and modeling studies that found neurons to have a larger capacity for glucose transport than astrocytes, supporting the view that glucose and not lactate is the primary substrate for neuronal oxidative metabolism [26,27,28]. Magistretti et al. rebuts these claims citing that glucose phosphorylation by hexokinase, not transport of glucose into the cell, is the rate-limiting step in glycolysis [16]. In addition, haploinsufficiency in neuron-specific glucose transporters (GLUT3) does not result in a pathological phenotype, whereas haploinsufficiency in astrocyte-specific glucose transporters (GLUT1) does [16, 29, 30]. Cultured neurons have also been found to prefer lactate over glucose when both are available [6].

Fig. 1
figure 1

The metabolism of astrocytes (yellow), microglia (blue), neurons (pink), and oligodendrocytes (green) are intricately linked and change depending on substrate availability. The neuron relies primarily on glycolysis, the TCA cycle, and OXPHOS to meet metabolic needs. In times of low glucose, it receives lactate from astrocytes and oligodendrocytes to fuel its TCA cycle and generate reducing agents for OXPHOS via the ANLS. Neurons and astrocytes can also turn to ketone bodies produced by the liver and secreted into the blood in times of hypoglycemia. A key role of astrocytes is removal and conversion of glutamate from the synaptic cleft into glutamine which can be used in the TCA cycle by microglia and neurons

Astrocyte end-feet cover much of the capillary surface and reside at every synapse, allowing them to both control the neuronal uptake of blood-borne nutrients and take in more glucose than other brain cells [31]. A key role of astrocytes in maintaining brain homeostasis is the removal of neurotransmitters from the synaptic cleft. The removal of neurotransmitters like glutamate is crucial for neurons as it can cause excitotoxicity. Glutamate is then converted to glutamine before being shuttled back to the neuron. This recycling and neuroprotective role of astrocytes is energetically expensive and requires a large amount of ATP [23, 32]. Like neurons, astrocytes can oxidize glucose to CO2; however, they have high glycolytic activity and low oxygen metabolism [7, 8, 32,33,34]. Increased glycolytic rates along with increased lactate export in astrocytes have been found to be markedly upregulated in response to glutamate stimulation. The lactate released into the extracellular environment can then be taken in by neurons and used in oxidative metabolism, as per the ANLS hypothesis. Astrocytes also possess a glyoxalase system that provides them with a pathway to detoxify the otherwise toxic byproducts of glycolysis [35]. With this, astrocytes are able to significantly upregulate and rely on glycolysis as a means of ATP production in a way that is inaccessible to neurons, while simultaneously providing substrates for the TCA cycle in neurons.

In hypoglycemia, astrocytes and neurons are forced to turn to alternative substrates. In the early stages of hypoglycemia in a healthy brain, astrocytes can utilize stored glycogen to produce lactate via glycogenolysis, which in turn can fuel neuronal metabolism [36]. This pathway is only available to astrocytes as they are the only location of glycogen storage in the brain, with total brain stores ranging from 3 to 12 μmol/g of tissue [37]. When these stores run out, astrocytes and neurons must then turn to ketone bodies (KBs), such as acetoacetate, acetone, and β-hydroxybutyrate. KBs are made in the adult liver in states of hypoglycemia and cross the blood–brain barrier (BBB), where they are converted to acetyl-CoA and used in the TCA in neurons and glial cells [7] (see Fig. 1). In various disorders including some neurodegenerative diseases like Lafora disease, there is aberrant glycogen metabolism that can lead to neuronal damage [38]. Recent studies have also found evidence of glycogen accumulation in the brain in glycogen storage disease type II, also known as Pompe disease [39].

Fatty acid (FA) metabolism is used for energetic and signaling purposes in glial cells and plays a signaling role in neurons. However, the contribution of fatty acids to substrate metabolism in neurons and glia is debatable and remains an emerging area of investigation. Recent research on neuronal fatty acid metabolism has focused on a brain-specific carnitine palmitoyltransferase 1 C (CPT1C). Unlike other CPT1 isoforms found predominantly in peripheral metabolic tissues, CPT1C is not involved in the β-oxidation of fatty acids and is not found in mitochondria [40]. Instead, CPT1C appears to play a regulatory role via its interaction with malonyl-coenzyme A (malonyl-CoA) [9, 41, 42]. Malonyl-CoA is the first product in the process of de novo fatty acid synthesis [9]. In hypothalamic neurons of the arcuate nucleus, the inhibition of malonyl-CoA by CPTC1C can regulate food intake [9].

In addition to the liver, astrocytes are also capable of making KBs from fatty acids [7, 43]. This process starts with β-oxidation of the fatty acid to make acetyl-CoA which can enter either the TCA cycle or ketogenesis. The role that KB production has in brain metabolism is not fully understood, but it has been hypothesized that it may provide metabolic support in states of starvation or ischemia and provide substrates for anaplerotic replenishment of neurotransmitters [7, 44].

Oligodendrocytes

Oligodendrocytes (OL) are chiefly associated with the synthesis of lipid-rich myelin to form the myelin sheath that supports neuronal health and signal transmission. Oligodendrocytes have been demonstrated to use glycolysis and the TCA cycle to generate ATP to fuel biological processes, as well as the PPP for pyruvate carboxylation [45] (see Fig. 1). They have been also found to have the highest rate of oxidative metabolism of all brain cells [46]. It is likely that their metabolic requirements are tightly linked to the need for myelin synthesis. For example, recent studies using fatty acid synthase (FASN)-depleted OLs, have shown that endogenous FA synthesis is required for myelination during development and for efficient remyelination in adulthood [47]. Although it remains to be experimentally determined, this would suggest that increased NADPH recycling via the PPP and citrate production via the TCA cycle in OLs may be required to provide intermediates for FASN to drive myelin synthesis. While OL substrate requirements require further study, it is of note that provision of a high-fat diet can partially rescue myelin deficits OL-specific FASN KD mice [47].

Microglia

Microglia, the brain resident macrophages have pleiotropic roles in the central nervous system (CNS). To perform a variety of functions, they express numerous cell surface receptors, including but not limited to toll-like receptors (TLRs), CD11b, and triggering receptor expressed by myeloid cells 2 (TREM2) [48,49,50]. The roles of these receptors in microglial function and how they could be harnessed for clinical diagnosis and treatment of CNS disorders continues to be a subject of intense research. However, it is becoming fairly well established that in response to invasion or injury microglia can phenotypically switch in response to specific environmental cues, and take on several changes including cell surface receptor expression, release of inflammatory cytokines, generation of reactive oxygen species (ROS), and changes in substrate utilization and metabolic requirements [48, 51, 52].

Transcriptomic studies suggest that microglia can use glucose, FAs, lipoproteins and glutamine as substrates for metabolism [51, 52]. Recent studies using CPT1 inhibitors to block mitochondrial fatty acid oxidation (FAO) in microglia showed increased microglial ramification and reduced motile surveillance and damage sensing capabilities, suggesting that FAO occurs in microglia and may drive function [53]. However, these studies are complicated by the fact that CPT1a is only expressed in astrocytes, implicating that mitochondrial β-oxidation is unlikely in microglia. It remains to be determined whether FAO could occur in microglial peroxisomes, but recent findings showing that inactivation of multifunctional protein-2 (MFP2), a pivotal enzyme in peroxisomal β-oxidation, leads to fatal neurological disorder characterized by microglial dysfunction suggest that peroxisomal β-oxidation may be an important and understudied facet of microglial metabolism [54]. Functional metabolic studies, particularly to determine whether FAs are utilized for fuel in microglia, either by mitochondrial or possibility peroxisomal β-oxidation, or rather intermediates for structural components or signaling molecules are much needed.

Under homeostatic, anti-inflammatory conditions microglia are believed to rely primarily on glucose and OXPHOS to meet their energy demands whereas pro-inflammatory ‘activated’ microglia have been found to rely more heavily on glycolysis [51, 52, 55]. In support, in homeostatic conditions microglia only express the glucose transporters GLUT3 and GLUT5. However, upon stimulation by lipopolysaccharide (LPS) and interferon gamma (IFNγ), microglia upregulate the expression of GLUT1 and GLUT4 as well as hexokinase 2 (HK2) a key glycolytic enzyme [51, 56, 57]. This shift in GLUT expression may represent the greater glycolytic need associated with pro-inflammatory activation and is reminiscent of the Warburg effect, which is often been reported in cancer cells [42, 58]. In light of recent findings suggesting that microglia in fact exist in a plethora of “activation” states, it is likely that changes to substrate utilization following “microglial activation” is more complex than previously thought. Since microglial activation plays a major role in the pathogenesis of all neurodegenerative diseases, and particularly AD, further studies aimed at defining the metabolic characteristics microglia are much needed. It is worth noting that recent transcriptomic studies that have highlighted the importance of distinct metabolic phenotypes in disease have used in vitro cell cultures or cells that have been isolated from the brain [12, 13, 59,60,61]. The ability of microglia to rapidly respond to their environment could affect the validity of these studies. For example, it has also been noted that microglia grown in cultures tend to appear more “activated” than microglia in their native environment [62]. This calls for the development of new methodologies that assess the metabolic profile of microglia removing them from their native environment.

Metabolism in the brain during damage and disease

Neurodegenerative diseases affect millions of people worldwide and can be devastating for the afflicted and their families [63]. Of these diseases AD is the most common [2, 64], and will become more prevalent as the world’s population gets older [65]. Importantly AD, PD and MS share metabolic characteristics such as neuroinflammation and mitochondrial dysfunction. For example, during neuropathogenesis metabolic homeostasis becomes lost, which is associated with glial activation, neuronal damage, and cell death [66, 67]. Existing treatments for these diseases are not curative and at best slow the progression of the disease, highlighting the need to further understand metabolic aberrations underlying disease onset to develop new metabolism-focused intervention strategies with the potential to reverse or halt disease progression. Here, we will review our current understanding of the metabolic changes involved in AD, PD and MS.

Alzheimer’s disease

AD is the most common cause of dementia. It was first described by Alois Alzheimer in 1907 who characterized it as a disease of progressive dementia with fibrils that were chemically distinct from the surrounding tissue [68, 69]. With the advent of advanced microscopy techniques the disease has been further characterized by formation of plaques in the brain consisting of amyloid beta (Aβ), neurofibrillary tangles (NFTs), and loss of neuronal synapses [70]. It is estimated that around 44 million people live with dementia worldwide, a number that is expected to triple by 2050 [71]. While many cases present sporadically, some genetic risk factors have been identified of which the Apoprotein E (APOE) ε4 allele presents the largest risk. Heterozygous APOE4 carriers have a reported odds ratio for AD of 3, whereas homozygous APOE4 carriers have an odds ratio of 12 [68, 72].

The APOE protein is a major component in glia-derived high-density lipoprotein (HDL)-like lipoproteins and is the primary apolipoprotein in CNS lipid metabolism. It is also involved in cholesterol and phospholipid efflux and clearance of lipids from the brain [73,74,75]. The APOE4 isoform has been found to be a significantly less potent acceptor of cholesterol and lipids than the APOE2 and APOE3 isoforms [73, 74, 76, 77]. This could be due to APOE4 lacking cysteine residues and leading to a propensity for posttranslational C-terminus cleavage preventing it from forming homodimers that allow it to accept lipids and to form lipoproteins [78,79,80,81]. This decreased lipid carrying capacity may lead to decreased lipid clearance from the brain, reduced lipid transport, reduced neuronal protection, and increased lipid accumulation, particular in microglia where lipid droplet accumulation may lead to cellular dysfunction [82]. Therefore, it is plausible to suggest APOE4-mediated defects in lipid and lipoprotein processing are central to the neuropathogenesis of AD. In support, astrocytes expressing the APOE4 isoform were found to accumulate more and smaller lipid droplets (LDs) than those with the APOE3 isoform [81]. Accumulation of LDs is indicative of cell dysfunction and a pro-inflammatory state, suggesting that cells expressing APOE4 isoform are more prone to activation and dysfunction that exacerbates the development of AD [81,82,83,84].

Recent studies have also found sex-dependent differences in APOE4 influences on metabolism. Female APOE4 carriers were found to have higher cerebral glucose metabolism than non-carriers while male APOE4 carriers displayed no significant difference [85]. This could be due to an interaction between estrogen receptors, estradiol, and APOE4 as suggested by literature, but clinical studies yield inconsistent results, and further studies are much needed [85,86,87,88,89].

Microglial activation and astrogliosis in response to amyloid deposition has been implicated in disease progression [70, 72, 90]. These phenotypic changes are coupled to metabolic alterations. Astrocytes and microglia have been found to switch to glycolysis in mouse models of AD, producing more lactate and exhibiting less TCA cycle activity [64]. This seems to contradict findings from fluorodeoxyglucose positron emission topography studies (FDG-PET) suggesting that the brains of AD and PD patients exhibit glucose hypometabolism [46, 91,92,93]. However, it is plausible that relative hypoglycemia may be representative of aberrations in systemic metabolism that are becoming increasingly associated with AD risk [94, 95]. It should also be noted that blood glucose levels also affect FDG-PET measurements, as the FDG competes with glucose to bind hexokinase [96]. In hypoglycemia, FDG will bind hexokinase more readily and not have as much competition due to relatively lower glucose levels. Similarly, in hyperglycemia, FDG will bind its target less readily as there is more relative competition. This confounds FDG-PET findings in neurodegenerative disease and should be considered during study design and data interpretation.

It has been proposed that inflammatory cells in the AD brain consume enough glucose to mask the metabolic deficits in neurons [97]. For example, metabolic tracer studies in a rodent model of AD have shown that AD neurons exhibit less TCA cycle activity and decreased expression of key mitochondrial enzymes [98], leading to poorly maintained ionic gradients and neuronal dysfunction. In addition, the loss of astrocytic TCA cycle activity cuts off neuronal supply of neurotransmitters, leading to an inability to maintain neurotransmitter homeostasis [98]. The interaction between neurons and astrocytes that both exhibit metabolic alterations in AD, leads to a cascade of events that culminates in neuronal death and cognitive decline, further supporting the need to understand this sequence of events in a temporal and cell-specific manner.

Fatty acid metabolism is associated with the pathogenesis of AD. Current research has focused on omega-3 and omega-6 poly unsaturated fatty acids (PUFAs) for the anti-inflammatory and pro-inflammatory effects of their breakdown products. The breakdown products of omega-6 PUFAs, like arachidonic acid (AA), promote inflammation and have been associated with increased AD pathology [99, 100]. On the other hand, omega-3 PUFAs are believed to have an anti-inflammatory effect [101]. However, this notion is often inconsistent with experimental findings. For example, increased levels of tissue omega-3 PUFAs have been associated with worse cognitive function in AD patients, suggesting that perturbation of fatty acid metabolism is implicated in AD pathogenesis [99]. Although it has been long postulated that supplementation with omega-3 PUFAs may be beneficial for patients with AD, clinical studies are also contradictory. While several studies have found that omega-3 PUFAs supplementation can improve cognitive function in both humans and animal models of AD [102,103,104], supplementation with the omega-3 PUFA docosahexaenoic acid (DHA) was not effective in preventing AD progression [105]. There are likely several reasons for these conflicting findings, including insufficient dosing regimens, the method of omega-3 delivery to the brain, and inherent defects in lipid processing that may simultaneously contribute to disease and unresponsiveness to omega-3 supplementation. Further research is needed to determine how changes in lipid metabolism may contribute or protect against the development of AD, particularly since dietary regimens that target lipid processing in the brain are a promising strategy to improve outcomes in patients with AD.

Parkinson’s disease

PD is the second most common progressive neurodegenerative disorder in the United States. It is clinically characterized by a resting tremor, bradykinesia, stooping posture and rigidity. It is associated with cognitive impairment, neurobehavioral disorders like depression and anxiety, and autonomic dysfunction [106]. Pathologically, PD is characterized by dopaminergic neuronal loss in the substantia nigra and intracellular α-synuclein deposition and accumulation in cholinergic and monoaminergic brainstem neurons [106,107,108]. Unlike with AD, most cases of PD are sporadic although familial forms have allowed for the identification pathways involved in the disease. Some of these pathways include oxidative stress, axonal transport, calcium homeostasis, α-synuclein proteostasis, mitochondrial function, and neuroinflammation. Like AD, magnetic resonance imaging and PET studies have found glucose hypometabolism in PD patients [91, 93, 109]. Decreased levels of PPP enzymes have been identified in PD patients along with mutations in genes encoding α-synuclein, Parkin, PTEN Induced Kinase 1 (PINK1), Leucine Rich Repeat Kinase 2 (LRRK2), and DJ-1 in familial forms of PD [110, 111]. Decreased levels of PPP enzymes while not directly deleterious, impair the cells’ ability to make pentoses necessary for ATP generation and DNA repair. Mutations in the genes found in familial PD have been implicated in mitochondrial dysfunction, inhibiting OXPHOS and forcing cells to rely on glycolysis to generate ATP [110, 111]. While astrocytes are less dependent on OXPHOS, neurons rely heavily on mitochondrial respiration to meet energy demands. Therefore, impaired oxidative metabolism in neurons will disrupt neuronal functions and continue to the progression of PD.

Amyloid deposition is also seen in PD and has been hypothesized as a cause for microglial activation [93]. Several studies have shown that microglial activation is increased in PD patients and may exacerbate neurodegeneration in PD brains [64, 112]. Edison et al., used [(11)C](R)PK11195-PET to measure the upregulation of translocator protein, a marker associated with microglial activation in patients with PD and Parkinson’s disease dementia (PDD) [93]. They found that patients with PD showed a significant increase in microglial activation, and that this activation was even more pronounced in patients with Parkinson’s disease dementia (PDD), but that microglial activation was an early event in the neuropathogenesis of PD and was independent of amyloid pathology [93]. Moreover, they found concurrent reductions in glucose metabolism in PD patients, supporting the notion that metabolic permutation may drive disease pathology.

These activated microglia are believed to be key to pathogenesis of PD and generate oxidative species that damage neurons [64]. It has also been suggested that neuronal damage could be due to increased secretion of glutamate from activated microglia, since increased extracellular glutamate leads to impaired neuronal mitochondrial respiration and an inability to generate sufficient levels of ATP [113]. This cascade of events ultimately leads to cell death [114], and highlights the importance of metabolic cross-talk between different cells of the brain in the development of PD.

Multiple sclerosis

MS is a chronic inflammatory and neurodegenerative disease in which immune cells cause demyelination and neuronal damage leading to cognitive decline. The disease has multiple clinical, radiologic, and histologic presentations. It is currently not known whether the inflammatory or neurodegenerative process is primary or secondary, and it is likely that pathogenesis is variable among patients, thus giving rise to different presentations [46, 115,116,117,118]. Although the disease progression is variable, most patients experience relapsing remitting multiple sclerosis (RRMS) characterized by periods of muscle weakness, blindness, double vision, trouble with sensation, and trouble with coordination, followed by total or partial recovery. In roughly 50% of patients, RRMS proceeds to secondary progressive multiple sclerosis (SPMS) where patients experience progressive deterioration. SPMS is typically diagnosed based on a history of progressive worsening of symptoms following a relapsing course of the disease. In some patients, the disease causes progressive damage from the onset with no periods of relapse. This form is termed primary progressive multiple sclerosis (PPMS). The speed of deterioration occurs at a similar rate in PPMS as it does in SPMS [119].

Four types of lesions have been characterized in MS. 20% of MS patients have immunopattern I lesions have T-cell inflammation, activated microglia, myelin-laden macrophages, and active demyelination suggesting that activated macrophages release toxic products that drive the demyelination process [120]. Approximately 50% of patients have immunopattern II lesions. These lesions are characterized by T-lymphocyte and macrophage infiltration, immunoglobulin deposition and complement activation on myelin. Immunopattern III lesions are characterized by oligodendrocyte apoptosis, microglial activation, T-cell inflammation and loss of myelin-associated glycoprotein, 2,3-cyclic nucleotide-3-phosphodiestarase [120]. This form describes about 29% of MS patients. The final form, immunopattern IV lesions are defined by non-apoptotic oligodendrocyte death and comprise only 1% of MS patients.

Regardless of the form, metabolic disturbances are central to the pathogenesis of MS. For example there is a growing body of evidence to suggest that mitochondrial dysfunction may underlie the neurodegeneration in MS, and is often present at the onset of clinical symptoms [121]. Dysfunctional mitochondria over-produce ROS, which can cause damage to proteins, lipids, DNA, and mitochondrial DNA (mtDNA) leading to cell damage and death [122,123,124]. OLs are particularly susceptible to damage by ROS due to their high metabolic rate and reliance on mitochondrial respiration to fuel the production of myelin proteins. Neurons are also susceptible to damage by ROS as they rely heavily on the TCA cycle to generate ATP. Astrocytes on the other hand are more glycolysis-dependent and therefore less affected by mitochondrial dysfunction and damage. Microglia become activated in response to the release of inflammatory mediators that are released by other microglia or infiltrating macrophages leading to further ROS generation and damage to nearby cells [125]. Microglia switch increased glucose utilization and increased fatty acid synthesis due to activation of mTOR and disruption of the TCA cycle [125]. Intracellular metabolites like malonyl-CoA in activated microglia also prevent fatty acids from associating with carnitine acyltransferase, preventing their entry into mitochondria, and thus inhibiting FAO [125].

Like AD and PD, metabolic alterations appear to be a critical component, if not causative, to the development and progression of MS. Importantly, AD, PD and MS are all characterized by the lack of a true disease-modifying therapy that is able to reverse or halt disease progression. This can be due to many factors like drug half-life, penetration of the active form of a drug into the BBB, the timing of the intervention, or ability of the drug to reach its target, inadequate bioavailability of the drug, or patient intolerance of therapy to name a few. Regardless of the reason for the lack of disease-modifying therapy, it is plausible to suggest that therapeutics aimed at restoring mitochondrial function may be an improved strategy that may rescue an increased susceptibility to develop ND in some individuals [126, 127]. To validate this hypothesis, further research is needed to define the time, region, and cell-specific changes in oxidative metabolism. Although current imaging modalities can measure changes in brain metabolism, the resolution is often lacking to determine cell-specific metabolic alterations. Going forward we will review the current range of methods used to visualize brain metabolism and propose FLIM as an emerging modality with sufficient resolution to empirically determine metabolic changes leading to ND.

Metabolic imaging of the brain

It is clear by now that neurodegenerative diseases are accompanied by metabolic alterations in one or more cell types of the brain. Therefore, examining these alterations is a crucial step in understanding the etiology and progression of these diseases. There are many ways to study metabolism in cells and tissues such as mass spectrometry, activity and colorimetric assays, seahorse XF analyzers, and imaging [128]. Specifically, metabolic imaging is advantageous in a sense that it provides the spatial and temporal information that other techniques lack. Such spatiotemporal measurements are essential to resolve the metabolic changes in the different regions of the brain, as well as at different time points for longitudinal studies of disease progression.

Metabolic imaging modalities can be primarily divided into either ionizing radiation methods or non-ionizing radiation methods. Ionizing radiation methods include nuclear imaging techniques such as positron emission tomography (PET) and single-photon emission computed tomography (SPECT). Whereas non-ionizing radiation methods include magnetic resonance imaging (MRI) techniques such as magnetic resonance spectroscopy (MRS), hyperpolarized magnetic resonance, chemical exchange saturation transfer (CEST), optoacoustic (photoacoustic) imaging, as well as fluorescence microscopy imaging methods [129]. Coherent Raman scattering (CRS) microscopy [130] and mass spectrometry imaging (MSI) [131, 132] are other non-ionizing radiation methods which have been used to image metabolism in cells and tissues; although the latter does involve desorption/ionization of analytes by means of absorbing short ultraviolet (UV) wavelengths that are non-ionizing per se. To be clinically relevant, metabolic imaging techniques need to be noninvasive or minimally invasive (i.e., no biopsies and tissue sampling are required), and preferably without the use of exogenous tracers. Moreover, they need to provide enough resolution to differentiate between different cell types at least, if not at a subcellular level.

In vivo changes in brain metabolism have been assessed using the following noninvasive or minimally invasive neuroimaging techniques: MRI, functional magnetic resonance imaging (fMRI), MRS, diffusion weighted imaging (DWI), diffusion tensor imaging (DTI), PET and SPECT [133], in addition to functional photoacoustic microscopy (PAM) [134] (a focused review on advanced MRI and MRS methods used in brain metabolic imaging can be found in [135]). On the other hand, postmortem ex vivo examination of these changes has been generally performed using microscopy techniques [133] such as immunohistochemistry (IHC) [136], fluorescence microscopy with a variety of biosensors [137] (although such biosensors have also been used in in vivo brain metabolic imaging [138]), stimulated Raman scattering (SRS) [139]—which is a CRS technique, MSI techniques [140,141,142], and more recently fluorescence lifetime imaging microscopy (FLIM) [143,144,145] (though FLIM has also been used in in vivo brain metabolic imaging [53]). Autoradiography is another technique that has been used to observe changes in brain cell receptor expression or architecture [146,147,148]. Here, a radioactive tracer is injected into the sample prior to brain isolation and exposure of a radiation sensitive imager plate [149]. This technique provides good data on the regional localization of the target of a given radioactive tracer. Depending on the tracer, inferences can then be made on changes in metabolism. This technique has been applied to analyze changes in PD, AD, and TBI [146,147,148,149]. This technique is particularly useful to determine drug or ligand occupancy of specific-receptors and is therefore a gold-standard technique in pharmacological studies evaluating novel drugs, and their targets [150, 151]. However, the use of radiation, and the ex vivo nature of this approach is not always desirable.

Additionally, brain imaging methods can be broadly divided into three categories based on the information they provide: structural imaging, functional imaging, and molecular imaging. Structural imaging modalities include techniques such as computed tomography (CT) and MRI. CT images are created by exposing a patient to a rotating beam of X-ray radiation, whose signal is then collected and reconstructed to form a 3D image [152]. This can be used in conjunction with intravenously administered contrast agents, e.g., CT angiography, to allow for better imaging of vasculature [152, 153]. In the clinic, CT is useful for its quick acquisition time and is often used to rule out acute pathology such as hemorrhage, ischemia, or lesions. However, magnetic resonance images can provide better structural resolution especially when it comes to ND-associated structural changes. These images are created by exciting hydrogen ions and detecting energy release to construct an image [153]. Variations of MRI such as structural MRI and DTI have also been developed to study brain volume and tissue characteristics, and the integrity of white matter tracts [154].

Many MRI-based modalities have been developed to provide functional and molecular information. The most prevailing functional neuroimaging modality is blood oxygen level-dependent (BOLD) fMRI. This method relies on differences in the magnetic resonance signal created by changes in blood flow to brain regions, and differences in the magnetic properties of oxygen-rich and oxygen-poor blood. This provides an indirect measure of neural activity by associating metabolic activity with hemodynamic parameters that imply increased delivery or availability of nutrients to a specific brain region [155]. BOLD fMRI can be used for mapping brain networks associated with tasks or stimuli [156]. A similar technique is arterial spin labeling (ASL) wherein arterial blood is magnetically labeled and acts as a tracer to then measure differences in cerebral blood flow [157]. Rather than directly measure metabolism, these techniques provide good information on cerebral blood flow that can then be used to make inferences on regional cerebral metabolism.

PET is a well-established method of imaging that can provide metabolic information; the metabolic activity of brain regions can be imaged and analyzed using radioactive forms of glucose, dopamine, or radioligands for the 5-HT2 receptor, serotonin transporter, or D2 receptor [158, 159]. 18F-fluorodeoxyglucose (FDG) is one such tracer that has been developed to study regions of altered glucose metabolism in the brain. This is now widely used to study regions of infection, inflammation, or tumor growth [158, 160]. It should be noted, however, that localization of FDG is a measure of glucose uptake and does not provide a precise measure of glucose metabolism [96]. As previously mentioned, FDG will bind hexokinase but not be metabolized.

Each of the aforementioned methodologies has its merits. Nonetheless, there exist limitations to their usability and suitability for the metabolic research in question. PET technique, which is one of the most commonly used metabolic imaging methods for the brain- in addition to MRS—is limited by the high cost accompanying the production of its radiopharmaceuticals as well as the inability to distinguish anatomic details [161]. MRS limitations include potential tissue heating from the applied radio frequency (RF) energy, lengthy infusion and sampling schemes of the isotopically labeled substrates, and more significantly the metabolic composition retrieved from a given voxel may belong to different tissue types [161], and thereby depreciating the importance of brain tissue-dependent metabolite heterogeneity. Additionally, some MRS techniques (carbon spectroscopy) require either long acquisition time or infusion of 13C-labeled substrates, while others (hyperpolarization-based) are limited by their hyperpolarized metabolic probe’s short relaxation time [162]. Since many of these also rely on cerebral vasculature to either determine differences in brain regions or to transport a substrate, making these measurements susceptible to confounding by anatomical variation, as well as pathology that alters pericytes, endothelial cells, and smooth muscle cells which play a large role in controlling cerebral blood flow [163, 164]. Admittedly, many diseases also affect cerebral microvasculature, but this complicates separating metabolic differences from vascular differences [165].

The previously mentioned methods provide good structural, functional, and molecular information especially when used in conjunction with one another. However, they only provide regional resolution and are dependent on the detectable metabolites. On the other hand, optical imaging modalities provide cellular resolution. In light of these differences, an optical imaging modality such as FLIM, may be a better tool to study metabolic changes in the brain [15].

Fluorescence lifetime imaging microscopy as a tool to measure in situ brain metabolism

Reduced nicotinamide adenine dinucleotide (NADH) and flavin adenine dinucleotide (FAD) are endogenous molecules in cells and tissues that have weak natural fluorescence [166]. Using specific wavelength of a light source, they can be transformed into an excited state with subsequent relaxation back to the ground state followed by simultaneous emission of a light signal of a different wavelength. The first NADH fluorescence imaging on live cells was performed in 1957 by Duysens and Amesz [167]. Since then, it has been widely used to study metabolic state of cells and tissues. NADH and FAD have divergent emission spectra with a little spectral overlap, which allows them to be imaged simultaneously. Fluorescence microscopy of NADH and FAD molecules helps to visualize where metabolic alterations are taking place in biological samples, including cells and tissue sections.

While the fluorescence intensity signal of NADH and FAD is highly informative and represents the activity of metabolic pathways, it is proportional to their concentration in the imaged samples. This can become problematic when reproducing measurements or when a fluorophore is only present at small concentrations in a sample. One way to circumvent this reliance on fluorophore concentration is to look at fluorescence lifetime instead of fluorescence intensity. With fluorescence lifetime the amount of time a fluorophore spends in the excited state is measured. This property of a fluorophore is dependent only on the molecule’s microenvironment and not on its concentration [168,169,170]. This can be particularly advantageous in the field of biology as some traditional fluorophores are cytotoxic or not native to a cell’s environment. As previously mentioned, FLIM allows for the use of endogenous fluorophores that are already present in biological systems [167, 168, 171]. Looking at a fluorophore’s fluorescence lifetime as opposed to its fluorescence signal intensity also eliminates the requirement that a fluorophore displays spectral shift in response to a change in environment [168]. This also allows for the use of fluorophores with large wavelength absorption spectra, potentially lowering the cost of instrumentation [168].

Development of laser scanning microscopy methods in conjunction with fluorescent molecules lifetime measurement marked the advent of FLIM. With FLIM, contrast in an image is generated by differences in fluorescence lifetimes [168]. This new method allows for the study of a fluorophore’s microenvironment, shedding light on the different environments in a cell or tissue. Examples of factors that can change a fluorophore’s fluorescence lifetime include the presence of oxygen, calcium, pH, energy transfer, as well as many other factors like nearby proteins or quenchers [172,173,174,175,176,177,178,179,180,181,182,183,184]. As the field grows, more fluorophores will enter the field. NADH and FAD, which have intrinsic fluorescence and their lifetime is tied to the state of the molecule [185, 186]. Given the subcellular resolution fluorescence microcopy provides, examination of these fluorophores allows for a direct analysis of the metabolic profile of individual cells.

Advances in the field of microscopy over the last century have been key in the development of FLIM. Two-photon excitation microscopy allowed for deeper imaging of tissues and reduced photodamage even with longer imaging times and allowed for the interrogation of femtoliter volumes in a sample [185, 187,188,189,190,191]. The ability to capture spatial and temporal resolution simultaneously in microscopy was also key in the development of FLIM as this allowed for fluorescence lifetime data acquisition in volumes small enough to analyze individual subcellular compartments in real time [189, 191].

The fluorescence lifetime of NADH changes depending on whether it is free or in a bound-to-enzyme form [186, 192]. For free NADH, the lifetime is around 0.4 ns versus around 3.4 ns when it is in a bound-to-enzyme form [15]. Other factors that affect the fluorescence lifetime include pH, temperature, and redox state. Since NADH is intricately involved as a cofactor in multiple metabolic pathways, differences in its lifetime can be used to determine which pathways a cell is using to meet its metabolic demands. FAD is another cofactor whose autofluorescence can and has been used to investigate metabolic processes in situ [193,194,195,196]. Specifically, FLIM has found a niche in analyzing tissue for markers of cancer, since a hallmark of cancerous cells is the reprogramming of metabolic pathways [197,198,199,200,201,202,203]. It should be noted that there are many other molecules in cells and tissues that exhibit autofluorescence and whose fluorescence spectra overlap with these probes, and techniques have been developed to bypass this problem [186, 192, 204, 205].

The phasor approach for FLIM data analysis helps to avoid multiexponential analysis used in typical fluorescence decay fitting [187, 189]. In phasor analysis each pixel of an image has its fluorescence decay plotted as a vector on a universal semicircle plot with vertices at (0,0) and (0,1). These points correspond to an infinite lifetime and a lifetime of zero, respectively. Pixels with contribution of multiple different lifetimes lie inside the circle while pixels with more homogenous lifetime composition lie closer to the periphery of the circle [187, 189]. The resultant phasor plot provides insight on the state of various fluorophores within a sample, which is not intended to determine exact fluorescence lifetimes (Fig. 2). Notably, there are caveats to this fit free analysis of FLIM data. Phasor analysis is not sensitive to low photon counts [206], and cannot reliably distinguish small changes in the redox state in the sample [186]. Moreover, the metabolic analysis does not have a single pathway resolution and response is reported by the ratio of free-to-bound fractions, where both numerator and denominator can change, while maintaining similar ratio [207]. Further, if the calibration decay is not acquired properly, this can lead to improper conversions of data to phasors [208]. Additionally, if the phasor transformation is not done with data acquired in entire period of the laser repetition rate, advantages of phasor plot analysis can be diminished [209]. The consequences of such errors as well as ways to avoid them are outlined elsewhere [185, 209].

Fig. 2
figure 2

FLIM imaging and analysis. Mouse hypothalamus (A) is excited (B) with 2 photon laser to collect NADH and FAD fluorescence signal on multiple smaller fields of view (C) and measure their lifetimes. Phasor plots for NADH (D) and FAD (E) and corresponding lifetime maps distributions (F and G, respectively) are created

As previously discussed, many neurodegenerative diseases involve metabolic reprogramming. By using FLIM, these changes in brain metabolism can be characterized with the possibility of identifying metabolic derangements prior to the onset of symptoms. For example, we have recently shown that neurons lacking the lipid processing enzyme lipoprotein lipase (LPL) show a shift in the NADH phasor plot indicative (Fig. 2) of increased free NADH compared to wild-type neurons, suggesting that LPL depletion results in metabolic shifts towards glucose utilization and away from oxidative metabolism (Fig. 2) [144]. These findings were the first to combine electrophysiological measurements of neuronal function with metabolic changes and highlight how FLIM may be used to measure metabolism alterations in fresh-frozen brain sections [144]. Recent studies have also employed FLIM to successfully determine metabolic changes in primary cultures of neurons and astrocytes in response to toxins that mimic the effects of PD [210]. Specifically, manganese (Mn) treatment leads to increased bound NAD(P)H in neurons consistent with enhanced apoptosis, but leads to decreased bound NADH in astrocytes, possibly due to a shift towards glycolytic metabolism and impaired respiration [210]. FLIM has also been used to study chronic neuroinflammation in the brains of mouse models of MS and in venous blood of human subjects [211,212,213]. Here, NAD(P)H was used as a fluorophore to detect active NADPH oxidase and as an analog for oxidative stress [211,212,213]. NAD(P)H can be used as an analog for oxidative stress because it serves as a source of electrons to generate ROS via NADPH oxidase [214,215,216]. In the brain, overactivation of NADPH oxidase has been linked to neurodegenerative disease [214,215,216]. This technique has also been applied to studying the pathophysiology of cerebral amyloid angiopathy and glioblastoma [217, 218].

With enough characterization of a given fluorophore, and its possible intracellular bound states and respective fluorescence lifetimes, one can generate a database that allows for near complete analysis of a cell’s metabolic state in various compartments [219]. Leben et al. undertook such an endeavor, generating a database of the various fluorescence lifetimes that NAD(P)H has when bound to various enzymes. With the help of RNA-seq data, they identified metabolic enzymes like malate dehydrogenase or lactate dehydrogenase, inducible enzymes like nitric oxide synthase and NAD(P)H oxidases, as well as other abundant small enzymes like sorbitol dehydrogenase that were most likely to bind to NAD(P)H and change its fluorescence lifetime [219]. By performing measurements of homogenous mixtures of NAD(P)H and single enzymes they were able to generate a database listing nearly all possible fluorescence lifetimes for this fluorophore [219]. This demonstrates that with enough fluorophore characterization, FLIM has the power to identify the state that a given fluorophore is in, what enzymes it is bound to, as well as what its microenvironment looks like in various intracellular or extracellular compartments. Although FLIM characterization is somewhat new, and often require validation studies to support the findings, it is clear even from the limited number of published studies that FLIM analysis can resolve cell-specific changes in brain metabolism that would otherwise be missed by alternative imaging modalities.

Microglia are of particular interest since altered microglial metabolism coincides with changes to phenotype and function that may contribute to, or protect against, the pathogenesis of many neurodegenerative diseases. However, existing methods for studying them rely on removing microglia from their native environment to study their gene expression and phenotype.

Recently, several publications have begun to utilize FLIM to understand inherent characteristics of microglial metabolism that may differentiate them from other cells in the brain. Sagar et al., measured the lifetime of NADH in both primary microglial cultures and in 100-μm-thick coronal sections from mice that had been treated with increasing doses of LPS in order to modulate the inflammatory status of the brain and “activate” the microglia [145]. Interestingly, they found that microglia in their activated state surveillant or inactivated state had a shorter NADH mean lifetime compared to other glial cells, and that they had a greater fractional contribution from free NADH [145]. In contrast, LPS treatment increased their mean NADH lifetime and decreased the fractional contribution of free NADH (compared to vehicle treatment) [145]. In support, Bernier et al., 2020 recently reported that resting microglia in situ have a much shorter NAD(P)H lifetime than surrounding non-microglial cells [53], and therefore a more glycolytic state. Importantly this study performed glycolytic inhibition with iodoacetate to confirm that glycolytic activity correlates with measurements of free short lifetime NADH [53]. Together, these studies suggest that microglia are much more glycolytic than previously thought, particularly in comparison to non-microglial cells and that activated microglia in fact may metabolically shift away towards oxidative metabolism to support the greater bioenergetic needs of activation-associated functions such as phagocytosis. Since these findings are somewhat in contrast to previous literature, and in particular findings from (albeit peripheral) inflammatory and macrophage models, it is imperative to further validate these findings using models of neurodegenerative disease such as AD, PD and MS. These discrepancies also highlight the importance of further work to validate our understanding of cell-and-disease-state specific brain metabolism, and to determine whether our findings corroborate with those from existing metabolic imaging approaches. In other words, will our findings using FLIM challenge our previous notions? For example, FDG-PET, described above, is predominantly used to measure brain glucose metabolism, which is interpreted as neuronal metabolism and neuronal activity. Since we know that non-neuronal cells majorly contribute to glucose uptake and metabolism, could glial cells also contribute to FDG-PET readouts? Although further method development is required, FLIM analysis of brain tissue from various metabolic and disease state paradigms, in conjunction with cell-specific markers, may eventually help answer these questions. Nonetheless, it is clear from the few studies employing FLIM, that this technique will enable major advancements in our understanding of microglial metabolism in situ. With more widespread use of FLIM and better metabolic characterization of microglia, FLIM could be used to identify mechanisms that promote or protect against the development of ND.

Conclusion

A striking similarity between ND is the immunometabolic switching of glial cells. Specifically, in AD, PD and MS oxidative-to-glycolytic, and shifts are often observed. It is plausible to suggest that targeting glycolytic shifts may be a potential therapeutic strategy for the treatment of ND. This is particularly promising in the case of AD, since disease-modifying treatments are mostly lacking. Here we suggest that imaging modalities such as FLIM, that can discriminate between subcellular, cell-type and regional differences in substrate utilization may be instrumental in the development of strategies that target metabolic processes to modify cellular function and improve outcomes in patients with AD and beyond.

Availability of data and materials

Not applicable.

Abbreviations

ND:

Neurodegenerative diseases

AD:

Alzheimer’s disease

PD:

Parkinson’s disease

MS:

Multiple sclerosis

FLIM:

Fluorescence lifetime imaging microscopy

MRI:

Magnetic resonance imaging

MRS:

Magnetic resonance spectroscopy

PET:

Positron emission tomography

NADH:

Nicotinamide adenine dinucleotide

FAD:

Flavin adenine dinucleotide

TCA cycle:

Tricarboxylic acid cycle

OXPHOS:

Oxidative phosphorylation

ATP:

Adenine triphosphate

Acetyl-CoA:

Acetyl coenzyme A

CO2 :

Carbon dioxide

O2 :

Oxygen

PPP:

Pentose phosphate pathway

NADP+ :

Nicotinamide adenine dinucleotide phosphate

ANLS:

Astrocyte-neuronal lactate shuttle

GLUT3:

Glucose transporter 3

KBs:

Ketone bodies

FA:

Fatty acid

CPT1C:

Carnitine palmitoyltransferase 1 C

malonyl-CoA:

Malonyl-coenzyme A

OL:

Oligodendrocytes

FASN:

Fatty acid synthase

CNS:

Central nervous system

TLRs:

Toll-like receptors

TREM2:

Triggering receptor expressed by myeloid cells 2

FAO:

Fatty acid oxidation

LPS:

Lipopolysaccharide

HK2:

Hexokinase 2

Aβ:

Amyloid beta

NFTs:

Neurofibrillary tangles

APOE:

Apoprotein E

HDL:

High-density lipoprotein

PUFAs:

Unsaturated fatty acids

PINK1:

PTEN induced kinase 1

LRRK2:

Leucine rich repeat kinase 2

RRMS:

Remitting multiple sclerosis

SPMS:

Progressive multiple sclerosis

mtDNA:

Mitochondrial DNA

CEST:

Chemical exchange saturation transfer

CRS:

Coherent Raman scattering

UV:

Ultraviolet

fMRI:

Functional magnetic resonance imaging

PAM:

Photoacoustic microscopy

IHC:

Immunohistochemistry

CT:

Computed tomography

BOLD:

Blood oxygen level-dependent

FDG:

18F-fluorodeoxyglucose

References

  1. Attwell D, Laughlin SB. An energy budget for signaling in the grey matter of the brain. J Cereb Blood Flow Metab. 2001;21:1133–45.

    Article  CAS  PubMed  Google Scholar 

  2. Camandola S, Mattson MP. Brain metabolism in health, aging, and neurodegeneration. EMBO J. 2017;36:1474–92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Harris JJ, Jolivet R, Attwell D. Synaptic energy use and supply. Neuron. 2012;75:762–77.

    Article  CAS  PubMed  Google Scholar 

  4. Hyder F, Rothman DL, Bennett MR. Cortical energy demands of signaling and nonsignaling components in brain are conserved across mammalian species and activity levels. Proc Natl Acad Sci USA. 2013;110:3549–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Nehlig A. Brain uptake and metabolism of ketone bodies in animal models. Prostaglandins Leukot Essent Fatty Acids. 2004;70:265–75.

    Article  CAS  PubMed  Google Scholar 

  6. Wyss MT, Jolivet R, Buck A, Magistretti PJ, Weber B. In vivo evidence for lactate as a neuronal energy source. J Neurosci. 2011;31:7477–85.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Takahashi S. Metabolic compartmentalization between astroglia and neurons in physiological and pathophysiological conditions of the neurovascular unit. Neuropathology. 2020;40:121–37.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Belanger M, Allaman I, Magistretti PJ. Brain energy metabolism: focus on astrocyte-neuron metabolic cooperation. Cell Metab. 2011;14:724–38.

    Article  CAS  PubMed  Google Scholar 

  9. Romano A, Koczwara JB, Gallelli CA, Vergara D, Micioni Di Bonaventura MV, Gaetani S, Giudetti AM. Fats for thoughts: an update on brain fatty acid metabolism. Int J Biochem Cell Biol. 2017;84:40–5.

    Article  CAS  PubMed  Google Scholar 

  10. van der Velpen V, Teav T, Gallart-Ayala H, Mehl F, Konz I, Clark C, Oikonomidi A, Peyratout G, Henry H, Delorenzi M, Ivanisevic J, Popp J. Systemic and central nervous system metabolic alterations in Alzheimer’s disease. Alzheimers Res Ther. 2019;11:93.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  11. Kyrtata N, Emsley HCA, Sparasci O, Parkes LM, Dickie BR. A systematic review of glucose transport alterations in Alzheimer’s disease. Front Neurosci. 2021;15:626636.

    Article  PubMed  PubMed Central  Google Scholar 

  12. Hammond TR, Dufort C, Dissing-Olesen L, Giera S, Young A, Wysoker A, Walker AJ, Gergits F, Segel M, Nemesh J, Marsh SE, Saunders A, Macosko E, Ginhoux F, Chen J, Franklin RJM, Piao X, McCarroll SA, Stevens B. Single-cell RNA sequencing of microglia throughout the mouse lifespan and in the injured brain reveals complex cell-state changes. Immunity. 2019;50:253-271 e6.

    Article  CAS  PubMed  Google Scholar 

  13. Keren-Shaul H, Spinrad A, Weiner A, Matcovitch-Natan O, Dvir-Szternfeld R, Ulland TK, David E, Baruch K, Lara-Astaiso D, Toth B, Itzkovitz S, Colonna M, Schwartz M, Amit I. A unique microglia type associated with restricting development of Alzheimer’s disease. Cell. 2017;169:1276-1290 e17.

    Article  CAS  PubMed  Google Scholar 

  14. Shokouhi S, Claassen D, Riddle W. Imaging brain metabolism and pathology in Alzheimer’s disease with positron emission tomography. J Alzheimers Dis Parkinsonism. 2014. https://doi.org/10.4172/2161-0460.1000143.

    Article  PubMed  PubMed Central  Google Scholar 

  15. Berezin MY, Achilefu S. Fluorescence lifetime measurements and biological imaging. Chem Rev. 2010;110:2641–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Magistretti PJ, Allaman I. A cellular perspective on brain energy metabolism and functional imaging. Neuron. 2015;86:883–901.

    Article  CAS  PubMed  Google Scholar 

  17. Itoh Y, Esaki T, Shimoji K, Cook M, Law MJ, Kaufman E, Sokoloff L. Dichloroacetate effects on glucose and lactate oxidation by neurons and astroglia in vitro and on glucose utilization by brain in vivo. Proc Natl Acad Sci USA. 2003;100:4879–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Zhang Y, Chen K, Sloan SA, Bennett ML, Scholze AR, O’Keeffe S, Phatnani HP, Guarnieri P, Caneda C, Ruderisch N, Deng S, Liddelow SA, Zhang C, Daneman R, Maniatis T, Barres BA, Wu JQ. An RNA-sequencing transcriptome and splicing database of glia, neurons, and vascular cells of the cerebral cortex. J Neurosci. 2014;34:11929–47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Mergenthaler P, Lindauer U, Dienel GA, Meisel A. Sugar for the brain: the role of glucose in physiological and pathological brain function. Trends Neurosci. 2013;36:587–97.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Herrero-Mendez A, Almeida A, Fernández E, Maestre C, Moncada S, Bolaños JP. The bioenergetic and antioxidant status of neurons is controlled by continuous degradation of a key glycolytic enzyme by APC/C-Cdh1. Nat Cell Biol. 2009;11:747–52.

    Article  CAS  PubMed  Google Scholar 

  21. Almeida A, Moncada S, Bolaños JP. Nitric oxide switches on glycolysis through the AMP protein kinase and 6-phosphofructo-2-kinase pathway. Nat Cell Biol. 2004;6:45–51.

    Article  CAS  PubMed  Google Scholar 

  22. Pellerin L, Magistretti PJ. Glutamate uptake into astrocytes stimulates aerobic glycolysis: a mechanism coupling neuronal activity to glucose utilization. Proc Natl Acad Sci USA. 1994;91:10625–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Figley CR, Stroman PW. The role(s) of astrocytes and astrocyte activity in neurometabolism, neurovascular coupling, and the production of functional neuroimaging signals. Eur J Neurosci. 2011;33:577–88.

    Article  PubMed  Google Scholar 

  24. Bittner CX, Loaiza A, Ruminot I, Larenas V, Sotelo-Hitschfeld T, Gutierrez R, Cordova A, Valdebenito R, Frommer WB, Barros LF. High resolution measurement of the glycolytic rate. Front Neuroenerg. 2010. https://doi.org/10.3389/fnene.2010.00026.

    Article  Google Scholar 

  25. Hashimoto T, Hussien R, Cho HS, Kaufer D, Brooks GA. Evidence for the mitochondrial lactate oxidation complex in rat neurons: demonstration of an essential component of brain lactate shuttles. PLoS ONE. 2008;3:e2915.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  26. Dienel GA. Brain lactate metabolism: the discoveries and the controversies. J Cereb Blood Flow Metab. 2012;32:1107–38.

    Article  CAS  PubMed  Google Scholar 

  27. Hertz L, Peng L, Dienel GA. Energy metabolism in astrocytes: high rate of oxidative metabolism and spatiotemporal dependence on glycolysis/glycogenolysis. J Cereb Blood Flow Metab. 2007;27:219–49.

    Article  CAS  PubMed  Google Scholar 

  28. DiNuzzo M, Giove F, Maraviglia B, Mangia S. Computational flux balance analysis predicts that stimulation of energy metabolism in astrocytes and their metabolic interactions with neurons depend on uptake of K(+) rather than glutamate. Neurochem Res. 2017;42:202–16.

    Article  CAS  PubMed  Google Scholar 

  29. Wang D, Pascual JM, Yang H, Engelstad K, Mao X, Cheng J, Yoo J, Noebels JL, De Vivo DC. A mouse model for Glut-1 haploinsufficiency. Hum Mol Genet. 2006;15:1169–79.

    Article  CAS  PubMed  Google Scholar 

  30. Stuart CA, Ross IR, Howell ME, McCurry MP, Wood TG, Ceci JD, Kennel SJ, Wall J. Brain glucose transporter (Glut3) haploinsufficiency does not impair mouse brain glucose uptake. Brain Res. 2011;1384:15–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Mathiisen TM, Lehre KP, Danbolt NC, Ottersen OP. The perivascular astroglial sheath provides a complete covering of the brain microvessels: an electron microscopic 3D reconstruction. Glia. 2010;58:1094–103.

    Article  PubMed  Google Scholar 

  32. Falkowska A, Gutowska I, Goschorska M, Nowacki P, Chlubek D, Baranowska-Bosiacka I. Energy metabolism of the brain, including the cooperation between astrocytes and neurons, especially in the context of glycogen metabolism. Int J Mol Sci. 2015;16:25959–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Supplie LM, Duking T, Campbell G, Diaz F, Moraes CT, Gotz M, Hamprecht B, Boretius S, Mahad D, Nave KA. Respiration-deficient astrocytes survive as glycolytic cells in vivo. J Neurosci. 2017;37:4231–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Jakoby P, Schmidt E, Ruminot I, Gutierrez R, Barros LF, Deitmer JW. Higher transport and metabolism of glucose in astrocytes compared with neurons: a multiphoton study of hippocampal and cerebellar tissue slices. Cereb Cortex. 2014;24:222–31.

    Article  PubMed  Google Scholar 

  35. Belanger M, Yang J, Petit JM, Laroche T, Magistretti PJ, Allaman I. Role of the glyoxalase system in astrocyte-mediated neuroprotection. J Neurosci. 2011;31:18338–52.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Yellen G. Fueling thought: management of glycolysis and oxidative phosphorylation in neuronal metabolism. J Cell Biol. 2018;217:2235–46.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Schonfeld P, Reiser G. Why does brain metabolism not favor burning of fatty acids to provide energy? Reflections on disadvantages of the use of free fatty acids as fuel for brain. J Cereb Blood Flow Metab. 2013;33:1493–9.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  38. Dienel GA, Carlson GM. Major advances in brain glycogen research: understanding of the roles of glycogen have evolved from emergency fuel reserve to dynamic, regulated participant in diverse brain functions. Adv Neurobiol. 2019;23:1–16.

    Article  PubMed  Google Scholar 

  39. Musumeci O, Marino S, Granata F, Morabito R, Bonanno L, Brizzi T, Lo Buono V, Corallo F, Longo M, Toscano A. Central nervous system involvement in late-onset Pompe disease: clues from neuroimaging and neuropsychological analysis. Eur J Neurol. 2019;26:442-e35.

    Article  PubMed  Google Scholar 

  40. Sierra AY, Gratacós E, Carrasco P, Clotet J, Ureña J, Serra D, Asins G, Hegardt FG, Casals N. CPT1c is localized in endoplasmic reticulum of neurons and has carnitine palmitoyltransferase activity. J Biol Chem. 2008;283:6878–85.

    Article  CAS  PubMed  Google Scholar 

  41. Wolfgang MJ, Lane MD. Control of energy homeostasis: role of enzymes and intermediates of fatty acid metabolism in the central nervous system. Annu Rev Nutr. 2006;26:23–44.

    Article  CAS  PubMed  Google Scholar 

  42. Zaugg K, Yao Y, Reilly PT, Kannan K, Kiarash R, Mason J, Huang P, Sawyer SK, Fuerth B, Faubert B, Kalliomäki T, Elia A, Luo X, Nadeem V, Bungard D, Yalavarthi S, Growney JD, Wakeham A, Moolani Y, Silvester J, Ten AY, Bakker W, Tsuchihara K, Berger SL, Hill RP, Jones RG, Tsao M, Robinson MO, Thompson CB, Pan G, Mak TW. Carnitine palmitoyltransferase 1C promotes cell survival and tumor growth under conditions of metabolic stress. Genes Dev. 2011;25:1041–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Bruce KD, Zsombok A, Eckel RH. Lipid processing in the brain: a key regulator of systemic metabolism. Front Endocrinol (Lausanne). 2017;8:60.

    Article  Google Scholar 

  44. Panov A, Orynbayeva Z, Vavilin V, Lyakhovich V. Fatty acids in energy metabolism of the central nervous system. Biomed Res Int. 2014;2014:472459.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  45. Amaral AI, Hadera MG, Tavares JM, Kotter MR, Sonnewald U. Characterization of glucose-related metabolic pathways in differentiated rat oligodendrocyte lineage cells. Glia. 2016;64:21–34.

    Article  PubMed  Google Scholar 

  46. Adiele RC, Adiele CA. Metabolic defects in multiple sclerosis. Mitochondrion. 2019;44:7–14.

    Article  CAS  PubMed  Google Scholar 

  47. Dimas P, Montani L, Pereira JA, Moreno D, Trotzmuller M, Gerber J, Semenkovich CF, Kofeler HC, Suter U. CNS myelination and remyelination depend on fatty acid synthesis by oligodendrocytes. Elife. 2019. https://doi.org/10.7554/eLife.44702.

    Article  PubMed  PubMed Central  Google Scholar 

  48. Ghosh S, Castillo E, Frias ES, Swanson RA. Bioenergetic regulation of microglia. Glia. 2018;66:1200–12.

    Article  PubMed  Google Scholar 

  49. Hsieh CL, Koike M, Spusta SC, Niemi EC, Yenari M, Nakamura MC, Seaman WE. A role for TREM2 ligands in the phagocytosis of apoptotic neuronal cells by microglia. J Neurochem. 2009;109:1144–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Marinelli S, Basilico B, Marrone MC, Ragozzino D. Microglia-neuron crosstalk: signaling mechanism and control of synaptic transmission. Semin Cell Dev Biol. 2019;94:138–51.

    Article  PubMed  Google Scholar 

  51. Kalsbeek MJ, Mulder L, Yi CX. Microglia energy metabolism in metabolic disorder. Mol Cell Endocrinol. 2016;438:27–35.

    Article  CAS  PubMed  Google Scholar 

  52. van der Poel M, Ulas T, Mizee MR, Hsiao CC, Miedema SSM, Adelia, Schuurman KG, Helder B, Tas SW, Schultze JL, Hamann J, Huitinga I. Transcriptional profiling of human microglia reveals grey-white matter heterogeneity and multiple sclerosis-associated changes. Nat Commun. 2019;10:1139.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  53. Bernier LP, York EM, Kamyabi A, Choi HB, Weilinger NL, MacVicar BA. Microglial metabolic flexibility supports immune surveillance of the brain parenchyma. Nat Commun. 2020;11:1559.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Beckers L, Geric I, Stroobants S, Beel S, Van Damme P, D’Hooge R, Baes M. Microglia lacking a peroxisomal beta-oxidation enzyme chronically alter their inflammatory profile without evoking neuronal and behavioral deficits. J Neuroinflamm. 2019;16:61.

    Article  Google Scholar 

  55. Voloboueva LA, Emery JF, Sun X, Giffard RG. Inflammatory response of microglial BV-2 cells includes a glycolytic shift and is modulated by mitochondrial glucose-regulated protein 75/mortalin. FEBS Lett. 2013;587:756–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Gimeno-Bayon J, Lopez-Lopez A, Rodriguez MJ, Mahy N. Glucose pathways adaptation supports acquisition of activated microglia phenotype. J Neurosci Res. 2014;92:723–31.

    Article  CAS  PubMed  Google Scholar 

  57. Wang L, Pavlou S, Du X, Bhuckory M, Xu H, Chen M. Glucose transporter 1 critically controls microglial activation through facilitating glycolysis. Mol Neurodegener. 2019;14:2.

    Article  PubMed  PubMed Central  Google Scholar 

  58. Schwartz L, Supuran CT, Alfarouk KO. The Warburg effect and the hallmarks of cancer. Anticancer Agents Med Chem. 2017;17:164–70.

    Article  CAS  PubMed  Google Scholar 

  59. Bachiller S, Jiménez-Ferrer I, Paulus A, Yang Y, Swanberg M, Deierborg T, Boza-Serrano A. Microglia in neurological diseases: a road map to brain-disease dependent-inflammatory response. Front Cell Neurosci. 2018. https://doi.org/10.3389/fncel.2018.00488.

    Article  PubMed  PubMed Central  Google Scholar 

  60. De Biase LM, Schuebel KE, Fusfeld ZH, Jair K, Hawes IA, Cimbro R, Zhang HY, Liu QR, Shen H, Xi ZX, Goldman D, Bonci A. Local cues establish and maintain region-specific phenotypes of basal ganglia microglia. Neuron. 2017;95:341-356.e6.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  61. Kettenmann H, Hanisch UK, Noda M, Verkhratsky A. Physiology of microglia. Physiol Rev. 2011;91:461–553.

    Article  CAS  PubMed  Google Scholar 

  62. Glenn JA, Ward SA, Stone CR, Booth PL, Thomas WE. Characterisation of ramified microglial cells: detailed morphology, morphological plasticity and proliferative capability. J Anat. 1992;180(Pt 1):109–18.

    PubMed  PubMed Central  Google Scholar 

  63. Brown RC, Lockwood AH, Sonawane BR. Neurodegenerative diseases: an overview of environmental risk factors. Environ Health Perspect. 2005;113:1250–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Aldana BI. Microglia-specific metabolic changes in neurodegeneration. J Mol Biol. 2019;431:1830–42.

    Article  CAS  PubMed  Google Scholar 

  65. Erkkinen MG, Kim MO, Geschwind MD. Clinical neurology and epidemiology of the major neurodegenerative diseases. Cold Spring Harb Perspect Biol. 2018. https://doi.org/10.1101/cshperspect.a033118.

    Article  PubMed  PubMed Central  Google Scholar 

  66. Prinz M, Priller J, Sisodia SS, Ransohoff RM. Heterogeneity of CNS myeloid cells and their roles in neurodegeneration. Nat Neurosci. 2011;14:1227–35.

    Article  CAS  PubMed  Google Scholar 

  67. Kipnis J, Filiano AJ. The central nervous system: privileged by immune connections. Nat Rev Immunol. 2018;18:83–4.

    Article  CAS  PubMed  Google Scholar 

  68. Lane CA, Hardy J, Schott JM. Alzheimer’s disease. Eur J Neurol. 2018;25:59–70.

    Article  CAS  PubMed  Google Scholar 

  69. Alzheimer A, Stelzmann RA, Schnitzlein HN, Murtagh FR. An English translation of Alzheimer’s 1907 paper, “Uber eine eigenartige Erkankung der Hirnrinde.” Clin Anat. 1995;8:429–31.

    Article  CAS  PubMed  Google Scholar 

  70. Butterfield DA, Boyd-Kimball D. Oxidative stress, amyloid-beta peptide, and altered key molecular pathways in the pathogenesis and progression of Alzheimer’s disease. J Alzheimers Dis. 2018;62:1345–67.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Prince M, Wimo A, Guerchet M, Ali G-C, Wu Y-T, Prina A. World Alzheimer report 2015: the global impact of dementia: an analysis of prevalence, incidence, cost and trends. London: Alzheimer’s Disease International; 2015.

    Google Scholar 

  72. Serrano-Pozo A, Frosch MP, Masliah E, Hyman BT. Neuropathological alterations in Alzheimer disease. Cold Spring Harb Perspect Med. 2011;1:a006189.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  73. Michikawa M, Fan QW, Isobe I, Yanagisawa K. Apolipoprotein E exhibits isoform-specific promotion of lipid efflux from astrocytes and neurons in culture. J Neurochem. 2000;74:1008–16.

    Article  CAS  PubMed  Google Scholar 

  74. Minagawa H, Gong JS, Jung CG, Watanabe A, Lund-Katz S, Phillips MC, Saito H, Michikawa M. Mechanism underlying apolipoprotein E (ApoE) isoform-dependent lipid efflux from neural cells in culture. J Neurosci Res. 2009;87:2498–508.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Rebeck GW. The role of APOE on lipid homeostasis and inflammation in normal brains. J Lipid Res. 2017;58:1493–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Fu Y, Zhao J, Atagi Y, Nielsen HM, Liu CC, Zheng H, Shinohara M, Kanekiyo T, Bu G. Apolipoprotein E lipoprotein particles inhibit amyloid-β uptake through cell surface heparan sulphate proteoglycan. Mol Neurodegener. 2016;11:37.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  77. Bar R, Boehm-Cagan A, Luz I, Kleper-Wall Y, Michaelson DM. The effects of apolipoprotein E genotype, α-synuclein deficiency, and sex on brain synaptic and Alzheimer’s disease-related pathology. Alzheimers Dement (Amst). 2018;10:1–11.

    Article  Google Scholar 

  78. Nagata KO, Nakada C, Kasai RS, Kusumi A, Ueda K. ABCA1 dimer-monomer interconversion during HDL generation revealed by single-molecule imaging. Proc Natl Acad Sci USA. 2013;110:5034–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Rebeck GW, Alonzo NC, Berezovska O, Harr SD, Knowles RB, Growdon JH, Hyman BT, Mendez AJ. Structure and functions of human cerebrospinal fluid lipoproteins from individuals of different APOE genotypes. Exp Neurol. 1998;149:175–82.

    Article  CAS  PubMed  Google Scholar 

  80. Harris FM, Brecht WJ, Xu Q, Tesseur I, Kekonius L, Wyss-Coray T, Fish JD, Masliah E, Hopkins PC, Scearce-Levie K, Weisgraber KH, Mucke L, Mahley RW, Huang Y. Carboxyl-terminal-truncated apolipoprotein E4 causes Alzheimer’s disease-like neurodegeneration and behavioral deficits in transgenic mice. Proc Natl Acad Sci USA. 2003;100:10966–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Farmer BC, Kluemper J, Johnson LA. Apolipoprotein E4 alters astrocyte fatty acid metabolism and lipid droplet formation. Cells. 2019;8:182.

    Article  CAS  PubMed Central  Google Scholar 

  82. Marschallinger J, Iram T, Zardeneta M, Lee SE, Lehallier B, Haney MS, Pluvinage JV, Mathur V, Hahn O, Morgens DW, Kim J, Tevini J, Felder TK, Wolinski H, Bertozzi CR, Bassik MC, Aigner L, Wyss-Coray T. Lipid-droplet-accumulating microglia represent a dysfunctional and proinflammatory state in the aging brain. Nat Neurosci. 2020;23:194–208.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Lalancette-Hébert M, Moquin A, Choi AO, Kriz J, Maysinger D. Lipopolysaccharide-QD micelles induce marked induction of TLR2 and lipid droplet accumulation in olfactory bulb microglia. Mol Pharm. 2010;7:1183–94.

    Article  PubMed  CAS  Google Scholar 

  84. Khatchadourian A, Bourque SD, Richard VR, Titorenko VI, Maysinger D. Dynamics and regulation of lipid droplet formation in lipopolysaccharide (LPS)-stimulated microglia. Biochim Biophys Acta. 1821;2012:607–17.

    Google Scholar 

  85. Jiang L, Lin H, Alzheimer's Disease Neuroimaging Initiative, Chen Y. Sex difference in the association of APOE4 with cerebral glucose metabolism in older adults reporting significant memory concern. Neurosci Lett. 2020; 722:134824.

  86. Kang JH, Weuve J, Grodstein F. Postmenopausal hormone therapy and risk of cognitive decline in community-dwelling aging women. Neurology. 2004;63:101–7.

    Article  PubMed  Google Scholar 

  87. Wang JM, Irwin RW, Brinton RD. Activation of estrogen receptor alpha increases and estrogen receptor beta decreases apolipoprotein E expression in hippocampus in vitro and in vivo. Proc Natl Acad Sci USA. 2006;103:16983–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Beydoun MA, Boueiz A, Abougergi MS, Kitner-Triolo MH, Beydoun HA, Resnick SM, O’Brien R, Zonderman AB. Sex differences in the association of the apolipoprotein E epsilon 4 allele with incidence of dementia, cognitive impairment, and decline. Neurobiol Aging. 2012;33:720-731.e4.

    Article  CAS  PubMed  Google Scholar 

  89. Ryan J, Carrière I, Scali J, Dartigues JF, Tzourio C, Poncet M, Ritchie K, Ancelin ML. Characteristics of hormone therapy, cognitive function, and dementia: the prospective 3C Study. Neurology. 2009;73:1729–37.

    Article  CAS  PubMed  Google Scholar 

  90. Burda JE, Sofroniew MV. Reactive gliosis and the multicellular response to CNS damage and disease. Neuron. 2014;81:229–48.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Hoyer S. Brain glucose and energy metabolism abnormalities in sporadic Alzheimer disease. Causes and consequences: an update. Exp Gerontol. 2000;35:1363–72.

    Article  CAS  PubMed  Google Scholar 

  92. Sperling RA, Aisen PS, Beckett LA, Bennett DA, Craft S, Fagan AM, Iwatsubo T, Jack CR Jr, Kaye J, Montine TJ, Park DC, Reiman EM, Rowe CC, Siemers E, Stern Y, Yaffe K, Carrillo MC, Thies B, Morrison-Bogorad M, Wagster MV, Phelps CH. Toward defining the preclinical stages of Alzheimer’s disease: recommendations from the National Institute on Aging-Alzheimer’s Association workgroups on diagnostic guidelines for Alzheimer’s disease. Alzheimers Dement. 2011;7:280–92.

    Article  PubMed  PubMed Central  Google Scholar 

  93. Edison P, Ahmed I, Fan Z, Hinz R, Gelosa G, Ray Chaudhuri K, Walker Z, Turkheimer FE, Brooks DJ. Microglia, amyloid, and glucose metabolism in Parkinson’s disease with and without dementia. Neuropsychopharmacology. 2013;38:938–49.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Lee HJ, Seo HI, Cha HY, Yang YJ, Kwon SH, Yang SJ. Diabetes and Alzheimer’s disease: mechanisms and nutritional aspects. Clin Nutr Res. 2018;7:229–40.

    Article  PubMed  PubMed Central  Google Scholar 

  95. Vagelatos NT, Eslick GD. Type 2 diabetes as a risk factor for Alzheimer’s disease: the confounders, interactions, and neuropathology associated with this relationship. Epidemiol Rev. 2013;35:152–60.

    Article  PubMed  Google Scholar 

  96. Sarikaya I, Sarikaya A, Sharma P. Assessing the effect of various blood glucose levels on 18F-FDG activity in the brain, liver, and blood pool. J Nucl Med Technol. 2019;47:313.

    Article  PubMed  Google Scholar 

  97. Backes H, Walberer M, Ladwig A, Rueger MA, Neumaier B, Endepols H, Hoehn M, Fink GR, Schroeter M, Graf R. Glucose consumption of inflammatory cells masks metabolic deficits in the brain. Neuroimage. 2016;128:54–62.

    Article  CAS  PubMed  Google Scholar 

  98. Nilsen LH, Witter MP, Sonnewald U. Neuronal and astrocytic metabolism in a transgenic rat model of Alzheimer’s disease. J Cereb Blood Flow Metab. 2014;34:906–14.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Snowden SG, Ebshiana AA, Hye A, An Y, Pletnikova O, O’Brien R, Troncoso J, Legido-Quigley C, Thambisetty M. Association between fatty acid metabolism in the brain and Alzheimer disease neuropathology and cognitive performance: a nontargeted metabolomic study. PLoS Med. 2017;14:e1002266.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  100. Amtul Z, Uhrig M, Wang L, Rozmahel RF, Beyreuther K. Detrimental effects of arachidonic acid and its metabolites in cellular and mouse models of Alzheimer’s disease: structural insight. Neurobiol Aging. 2012;33(831):e21-31.

    Google Scholar 

  101. Thomas J, Thomas CJ, Radcliffe J, Itsiopoulos C. Omega-3 fatty acids in early prevention of inflammatory neurodegenerative disease: a focus on Alzheimer’s disease. Biomed Res Int. 2015;2015:172801.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Martínez-Lapiscina EH, Clavero P, Toledo E, San Julián B, Sanchez-Tainta A, Corella D, Lamuela-Raventós RM, Martínez JA, Martínez-Gonzalez M. Virgin olive oil supplementation and long-term cognition: the PREDIMED-NAVARRA randomized, trial. J Nutr Health Aging. 2013;17:544–52.

    Article  PubMed  Google Scholar 

  103. Cole GM, Frautschy SA. Docosahexaenoic acid protects from amyloid and dendritic pathology in an Alzheimer’s disease mouse model. Nutr Health. 2006;18:249–59.

    Article  CAS  PubMed  Google Scholar 

  104. Hashimoto M, Hossain S, Shimada T, Sugioka K, Yamasaki H, Fujii Y, Ishibashi Y, Oka J-I, Shido O. Docosahexaenoic acid provides protection from impairment of learning ability in Alzheimer’s disease model rats. J Neurochem. 2002;81:1084–91.

    Article  CAS  PubMed  Google Scholar 

  105. Arellanes IC, Choe N, Solomon V, He X, Kavin B, Martinez AE, Kono N, Buennagel DP, Hazra N, Kim G, D’Orazio LM, McCleary C, Sagare A, Zlokovic BV, Hodis HN, Mack WJ, Chui HC, Harrington MG, Braskie MN, Schneider LS, Yassine HN. Brain delivery of supplemental docosahexaenoic acid (DHA): a randomized placebo-controlled clinical trial. EBioMedicine. 2020;59:102883.

    Article  PubMed  PubMed Central  Google Scholar 

  106. Poewe W, Seppi K, Tanner CM, Halliday GM, Brundin P, Volkmann J, Schrag AE, Lang AE. Parkinson disease. Nat Rev Dis Primers. 2017;3:17013.

    Article  PubMed  Google Scholar 

  107. Damier P, Hirsch EC, Agid Y, Graybiel AM. The substantia nigra of the human brain. II. Patterns of loss of dopamine-containing neurons in Parkinson’s disease. Brain. 1999;122(Pt 8):1437–48.

    Article  PubMed  Google Scholar 

  108. Fearnley JM, Lees AJ. Ageing and Parkinson’s disease: substantia nigra regional selectivity. Brain. 1991;114(Pt 5):2283–301.

    Article  PubMed  Google Scholar 

  109. Borghammer P, Chakravarty M, Jonsdottir KY, Sato N, Matsuda H, Ito K, Arahata Y, Kato T, Gjedde A. Cortical hypometabolism and hypoperfusion in Parkinson’s disease is extensive: probably even at early disease stages. Brain Struct Funct. 2010;214:303–17.

    Article  PubMed  Google Scholar 

  110. Dunn L, Allen GF, Mamais A, Ling H, Li A, Duberley KE, Hargreaves IP, Pope S, Holton JL, Lees A, Heales SJ, Bandopadhyay R. Dysregulation of glucose metabolism is an early event in sporadic Parkinson’s disease. Neurobiol Aging. 2014;35:1111–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Knight AL, Yan X, Hamamichi S, Ajjuri RR, Mazzulli JR, Zhang MW, Daigle JG, Zhang S, Borom AR, Roberts LR, Lee SK, DeLeon SM, Viollet-Djelassi C, Krainc D, O’Donnell JM, Caldwell KA, Caldwell GA. The glycolytic enzyme, GPI, is a functionally conserved modifier of dopaminergic neurodegeneration in Parkinson’s models. Cell Metab. 2014;20:145–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. McGeer PL, Itagaki S, Boyes BE, McGeer EG. Reactive microglia are positive for HLA-DR in the substantia nigra of Parkinson’s and Alzheimer’s disease brains. Neurology. 1988;38:1285–91.

    Article  CAS  PubMed  Google Scholar 

  113. Reynolds AD, Stone DK, Mosley RL, Gendelman HE. Nitrated {alpha}-synuclein-induced alterations in microglial immunity are regulated by CD4+ T cell subsets. J Immunol. 2009;182:4137–49.

    Article  CAS  PubMed  Google Scholar 

  114. Takeuchi H, Jin S, Wang J, Zhang G, Kawanokuchi J, Kuno R, Sonobe Y, Mizuno T, Suzumura A. Tumor necrosis factor-induces neurotoxicity via glutamate release from hemichannels of activated microglia in an autocrine manner. J Biol Chem. 2006;281:21362–8.

    Article  CAS  PubMed  Google Scholar 

  115. Correale J, Gaitán MI, Ysrraelit MC, Fiol MP. Progressive multiple sclerosis: from pathogenic mechanisms to treatment. Brain. 2017;140:527–46.

    PubMed  Google Scholar 

  116. Popescu BF, Lucchinetti CF. Meningeal and cortical grey matter pathology in multiple sclerosis. BMC Neurol. 2012;12:11.

    Article  PubMed  Google Scholar 

  117. Mahad DJ, Ziabreva I, Campbell G, Lax N, White K, Hanson PS, Lassmann H, Turnbull DM. Mitochondrial changes within axons in multiple sclerosis. Brain. 2009;132:1161–74.

    Article  PubMed  Google Scholar 

  118. Hutchinson M. Neurodegeneration in multiple sclerosis is a process separate from inflammation: no. Mult Scler. 2015;21:1628–31.

    Article  PubMed  Google Scholar 

  119. Lublin FD, Reingold SC, Cohen JA, Cutter GR, Sørensen PS, Thompson AJ, Wolinsky JS, Balcer LJ, Banwell B, Barkhof F, Bebo B Jr, Calabresi PA, Clanet M, Comi G, Fox RJ, Freedman MS, Goodman AD, Inglese M, Kappos L, Kieseier BC, Lincoln JA, Lubetzki C, Miller AE, Montalban X, O’Connor PW, Petkau J, Pozzilli C, Rudick RA, Sormani MP, Stüve O, Waubant E, Polman CH. Defining the clinical course of multiple sclerosis: the 2013 revisions. Neurology. 2013;83(2014):278–86.

    Google Scholar 

  120. Lucchinetti C, Brück W, Parisi J, Scheithauer B, Rodriguez M, Lassmann H. Heterogeneity of multiple sclerosis lesions: implications for the pathogenesis of demyelination. Ann Neurol. 2000;47:707–17.

    Article  CAS  PubMed  Google Scholar 

  121. Barcelos IP, Troxell RM, Graves JS. Mitochondrial dysfunction and multiple sclerosis. Biology (Basel). 2019. https://doi.org/10.3390/biology8020037.

    Article  Google Scholar 

  122. Glass CK, Saijo K, Winner B, Marchetto MC, Gage FH. Mechanisms underlying inflammation in neurodegeneration. Cell. 2010;140:918–34.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Campbell GR, Ziabreva I, Reeve AK, Krishnan KJ, Reynolds R, Howell O, Lassmann H, Turnbull DM, Mahad DJ. Mitochondrial DNA deletions and neurodegeneration in multiple sclerosis. Ann Neurol. 2011;69:481–92.

    Article  CAS  PubMed  Google Scholar 

  124. Mahad D, Ziabreva I, Lassmann H, Turnbull D. Mitochondrial defects in acute multiple sclerosis lesions. Brain. 2008;131:1722–35.

    Article  PubMed  PubMed Central  Google Scholar 

  125. Peruzzotti-Jametti L, Pluchino S. Targeting mitochondrial metabolism in neuroinflammation: towards a therapy for progressive multiple sclerosis. Trends Mol Med. 2018;24:838–55.

    Article  CAS  PubMed  Google Scholar 

  126. Kalia LV, Kalia SK, Lang AE. Disease-modifying strategies for Parkinson’s disease. Mov Disord. 2015;30:1442–50.

    Article  CAS  PubMed  Google Scholar 

  127. Kulshreshtha A, Piplani P. Current pharmacotherapy and putative disease-modifying therapy for Alzheimer’s disease. Neurol Sci. 2016;37:1403–35.

    Article  PubMed  Google Scholar 

  128. Bernier LP, York EM, MacVicar BA. Immunometabolism in the brain: how metabolism shapes microglial function. Trends Neurosci. 2020;43:854–69.

    Article  CAS  PubMed  Google Scholar 

  129. Ntziachristos V, Pleitez MA, Aime S, Brindle KM. Emerging technologies to image tissue metabolism. Cell Metab. 2019;29:518–38.

    Article  CAS  PubMed  Google Scholar 

  130. Yue S, Cheng JX. Deciphering single cell metabolism by coherent Raman scattering microscopy. Curr Opin Chem Biol. 2016;33:46–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Fujimura Y, Miura D. MALDI mass spectrometry imaging for visualizing in situ metabolism of endogenous metabolites and dietary phytochemicals. Metabolites. 2014;4:319–46.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  132. Buchberger AR, DeLaney K, Johnson J, Li L. Mass spectrometry imaging: a review of emerging advancements and future insights. Anal Chem. 2018;90:240–65.

    Article  CAS  PubMed  Google Scholar 

  133. Van Horn JD, Bhattrai A, Irimia A. Multimodal imaging of neurometabolic pathology due to traumatic brain injury. Trends Neurosci. 2017;40:39–59.

    Article  PubMed  CAS  Google Scholar 

  134. Yao J, Wang L, Yang JM, Maslov KI, Wong TT, Li L, Huang CH, Zou J, Wang LV. High-speed label-free functional photoacoustic microscopy of mouse brain in action. Nat Methods. 2015;12:407–10.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Hyder F, Rothman DL. Advances in imaging brain metabolism. Annu Rev Biomed Eng. 2017;19:485–515.

    Article  CAS  PubMed  Google Scholar 

  136. Oe Y, Akther S, Hirase H. Regional distribution of glycogen in the mouse brain visualized by immunohistochemistry. Adv Neurobiol. 2019;23:147–68.

    Article  PubMed  Google Scholar 

  137. Zhang Z, Chen W, Zhao Y, Yang Y. Spatiotemporal imaging of cellular energy metabolism with genetically-encoded fluorescent sensors in brain. Neurosci Bull. 2018;34:875–86.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  138. Benson S, Fernandez A, Barth ND, de Moliner F, Horrocks MH, Herrington CS, Abad JL, Delgado A, Kelly L, Chang Z, Feng Y, Nishiura M, Hori Y, Kikuchi K, Vendrell M. SCOTfluors: small, conjugatable, orthogonal, and tunable fluorophores for in vivo imaging of cell metabolism. Angew Chem Int Ed Engl. 2019;58:6911–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. Hu F, Chen Z, Zhang L, Shen Y, Wei L, Min W. Vibrational imaging of glucose uptake activity in live cells and tissues by stimulated Raman scattering. Angew Chem Int Ed Engl. 2015;54:9821–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Benabdellah F, Touboul D, Brunelle A, Laprevote O. In situ primary metabolites localization on a rat brain section by chemical mass spectrometry imaging. Anal Chem. 2009;81:5557–60.

    Article  CAS  PubMed  Google Scholar 

  141. Miura D, Fujimura Y, Yamato M, Hyodo F, Utsumi H, Tachibana H, Wariishi H. Ultrahighly sensitive in situ metabolomic imaging for visualizing spatiotemporal metabolic behaviors. Anal Chem. 2010;82:9789–96.

    Article  CAS  PubMed  Google Scholar 

  142. Sugiura Y, Zaima N, Setou M, Ito S, Yao I. Visualization of acetylcholine distribution in central nervous system tissue sections by tandem imaging mass spectrometry. Anal Bioanal Chem. 2012;403:1851–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. York EM, Zhang J, Choi HB, MacVicar BA. Neuroinflammatory inhibition of synaptic long-term potentiation requires immunometabolic reprogramming of microglia. Glia. 2021;69:567–78.

    Article  CAS  PubMed  Google Scholar 

  144. Bruce KD, Dobrinskikh E, Wang H, Rudenko I, Gao H, Libby AE, Gorkhali S, Yu T, Zsombok A, Eckel RH. Neuronal lipoprotein lipase deficiency alters neuronal function and hepatic metabolism. Metabolites. 2020;10:385.

    Article  CAS  PubMed Central  Google Scholar 

  145. Sagar MAK, Ouellette JN, Cheng KP, Williams JC, Watters JJ, Eliceiri KW. Microglia activation visualization via fluorescence lifetime imaging microscopy of intrinsically fluorescent metabolic cofactors. Neurophotonics. 2020;7:035003.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Li H, Yang P, Knight W, Guo Y, Perlmutter JS, Benzinger TLS, Morris JC, Xu J. The interactions of dopamine and oxidative damage in the striatum of patients with neurodegenerative diseases. J Neurochem. 2020;152:235–51.

    Article  CAS  PubMed  Google Scholar 

  147. López-Picón F, Snellman A, Shatillo O, Lehtiniemi P, Grönroos TJ, Marjamäki P, Trigg W, Jones PA, Solin O, Pitkänen A, Haaparanta-Solin M. Ex vivo tracing of NMDA and GABA-A receptors in rat brain after traumatic brain injury using 18F-GE-179 and 18F-GE-194 autoradiography. J Nucl Med. 2016;57:1442–7.

    Article  PubMed  CAS  Google Scholar 

  148. Hachem M, Géloën A, Van AL, Foumaux B, Fenart L, Gosselet F, Da Silva P, Breton G, Lagarde M, Picq M, Bernoud-Hubac N. Efficient docosahexaenoic acid uptake by the brain from a structured phospholipid. Mol Neurobiol. 2016;53:3205–15.

    Article  CAS  PubMed  Google Scholar 

  149. Barthe N, Maîtrejean S, Carvou N, Cardona A. Chapter 9—high-resolution beta imaging. In: L’Annunziata MF, editor. Handbook of radioactivity analysis: volume 2 (fourth edition). London: Academic Press; 2020. p. 669–727.

    Chapter  Google Scholar 

  150. Gackenheimer SL, Gehlert DR. In vitro and ex vivo autoradiography of the NK-1 antagonist [(3)H]-LY686017 in Guinea pig brain. Neuropeptides. 2011;45:157–64.

    Article  CAS  PubMed  Google Scholar 

  151. Ishiwata K, Ogi N, Tanaka A, Senda M. Quantitative ex vivo and in vitro receptor autoradiography using 11C-labeled ligands and an imaging plate: a study with a dopamine D2-like receptor ligand [11C]nemonapride. Nucl Med Biol. 1999;26:291–6.

    Article  CAS  PubMed  Google Scholar 

  152. Mikla VI, Mikla VV. 2—computed tomography. In: Mikla VI, Mikla VV, editors. Medical imaging technology. Oxford: Elsevier; 2014. p. 23–38.

    Chapter  Google Scholar 

  153. Risacher SL, Saykin AJ. Chapter 12—neuroimaging in aging and neurologic diseases. In: Dekosky ST, Asthana S, editors. Handbook of clinical neurology. Amsterdam: Elsevier; 2019. p. 191–227.

    Google Scholar 

  154. Wong YQ, Tan LK, Seow P, Tan MP, Abd Kadir KA, Vijayananthan A, Ramli N. Microstructural integrity of white matter tracts amongst older fallers: a DTI study. PLoS ONE. 2017;12:e0179895.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  155. Buchbinder BR. Chapter 4—functional magnetic resonance imaging. In: Masdeu JC, González RG, editors. Handbook of clinical neurology. Amsterdam: Elsevier; 2016. p. 61–92.

    Google Scholar 

  156. Jackson GD, Badawy R, Gotman J. Chapter 23—functional magnetic resonance imaging: focus localization. In: Stefan H, Theodore WH, editors. Handbook of clinical neurology. Amsterdam: Elsevier; 2012. p. 369–85.

    Google Scholar 

  157. Haller S, Zaharchuk G, Thomas DL, Lovblad K-O, Barkhof F, Golay X. Arterial spin labeling perfusion of the brain: emerging clinical applications. Radiology. 2016;281:337–56.

    Article  PubMed  Google Scholar 

  158. Sarikaya I, Sarikaya A, Elgazzar AH. Current status of (18)F-FDG PET brain imaging in patients with dementia. J Nucl Med Technol. 2018;46:362–7.

    Article  PubMed  Google Scholar 

  159. Kropotov JD. Chapter 1.4—positron emission tomography. In: Kropotov JD, editor. Functional neuromarkers for psychiatry. San Diego: Academic Press; 2016. p. 27–30.

    Chapter  Google Scholar 

  160. Lameka K, Farwell MD, Ichise M. Chapter 11—positron emission tomography. In: Masdeu JC, González RG, editors. Handbook of clinical neurology. Amsterdam: Elsevier; 2016. p. 209–27.

    Google Scholar 

  161. Lin AL, Rothman DL. What have novel imaging techniques revealed about metabolism in the aging brain? Future Neurol. 2014;9:341–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  162. Le Page LM, Guglielmetti C, Taglang C, Chaumeil MM. Imaging brain metabolism using hyperpolarized (13)C magnetic resonance spectroscopy. Trends Neurosci. 2020;43:343–54.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  163. Armstead WM. Cerebral blood flow autoregulation and dysautoregulation. Anesthesiol Clin. 2016;34:465–77.

    Article  PubMed  PubMed Central  Google Scholar 

  164. Brown LS, Foster CG, Courtney JM, King NE, Howells DW, Sutherland BA. Pericytes and neurovascular function in the healthy and diseased brain. Front Cell Neurosci. 2019;13:282.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  165. Macdonald JA, Murugesan N, Pachter JS. Endothelial cell heterogeneity of blood–brain barrier gene expression along the cerebral microvasculature. J Neurosci Res. 2010;88:1457–74.

    CAS  PubMed  Google Scholar 

  166. Chance B, Schoener B, Oshino R, Itshak F, Nakase Y. Oxidation-reduction ratio studies of mitochondria in freeze-trapped samples. NADH and flavoprotein fluorescence signals. J Biol Chem. 1979;254:4764–71.

    Article  CAS  PubMed  Google Scholar 

  167. Duysens LN, Amesz J. Fluorescence spectrophotometry of reduced phosphopyridine nucleotide in intact cells in the near-ultraviolet and visible region. Biochim Biophys Acta. 1957;24:19–26.

    Article  CAS  PubMed  Google Scholar 

  168. Lakowicz JR, Szmacinski H, Nowaczyk K, Berndt KW, Johnson M. Fluorescence lifetime imaging. Anal Biochem. 1992;202:316–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  169. McQuarrie DA, Simon JD. Physical chemistry: a molecular approach. Sausalito: University Science Books; 1997.

    Google Scholar 

  170. Tsai HM, Souris JS, Kim HJ, Cheng SH, Chen L, Lo LW, Chen CT, Kao CM. Note: rapid measurement of fluorescence lifetimes using SiPM detection and waveform sampling. Rev Sci Instrum. 2017;88:096107.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  171. Lakowicz JR, Szmacinski H, Nowaczyk K, Johnson ML. Fluorescence lifetime imaging of free and protein-bound NADH. Proc Natl Acad Sci USA. 1992;89:1271–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  172. Hammler D, Marx A, Zumbusch A. Fluorescence-lifetime-sensitive probes for monitoring ATP cleavage. Chemistry (Easton). 2018;24:15329–35.

    CAS  Google Scholar 

  173. Sicchieri LB, de Andrade Natal R, Courrol LC. Fluorescent lifetime imaging microscopy using Europium complexes improves atherosclerotic plaques discrimination. Int J Cardiovasc Imaging. 2016;32:1595–604.

    Article  PubMed  Google Scholar 

  174. Thulborn KR, Sawyer WH. Properties and the locations of a set of fluorescent probes sensitive to the fluidity gradient of the lipid bilayer. Biochim Biophys Acta. 1978;511:125–40.

    Article  CAS  PubMed  Google Scholar 

  175. Leto TL, Roseman MA, Holloway PW. Mechanism of exchange of cytochrome b5 between phosphatidylcholine vesicles. Biochemistry. 1980;19:1911–6.

    Article  CAS  PubMed  Google Scholar 

  176. Lehrer SS. Solute perturbation of protein fluorescence. The quenching of the tryptophyl fluorescence of model compounds and of lysozyme by iodide ion. Biochemistry. 1971;10:3254–63.

    Article  CAS  PubMed  Google Scholar 

  177. Eftink MR, Ghiron CA. Fluorescence quenching studies with proteins. Anal Biochem. 1981;114:199–227.

    Article  CAS  PubMed  Google Scholar 

  178. Steinberg IZ. Long-range nonradiative transfer of electronic excitation energy in proteins and polypeptides. Annu Rev Biochem. 1971;40:83–114.

    Article  CAS  PubMed  Google Scholar 

  179. Stryer L. Fluorescence energy transfer as a spectroscopic ruler. Annu Rev Biochem. 1978;47:819–46.

    Article  CAS  PubMed  Google Scholar 

  180. Miyoshi N, Hara K, Yokoyama I, Tomita G, Fukuda M. fluorescence lifetime of acridine orange in sodium dodecyl sulfate premicellar solutions. Photochem Photobiol. 1988;47:685–8.

    Article  CAS  PubMed  Google Scholar 

  181. Gafni A, Brand L. Excited state proton transfer reactions of acridine studied by nanosecond fluorometry. Chem Phys Lett. 1978;58:346–50.

    Article  CAS  Google Scholar 

  182. Lakowicz JR, Weber G. Quenching of fluorescence by oxygen. A probe for structural fluctuations in macromolecules. Biochemistry. 1973;12:4161–70.

    Article  CAS  PubMed  Google Scholar 

  183. Lakowicz JR, Joshi NB, Johnson ML, Szmacinski H, Gryczynski I. Diffusion coefficients of quenchers in proteins from transient effects in the intensity decays. J Biol Chem. 1987;262:10907–10.

    Article  CAS  PubMed  Google Scholar 

  184. Kautsky H. Quenching of luminescence by oxygen. Trans Faraday Soc. 1939;35:216–9.

    Article  CAS  Google Scholar 

  185. Ranjit S, Lanzanò L, Libby AE, Gratton E, Levi M. Advances in fluorescence microscopy techniques to study kidney function. Nat Rev Nephrol. 2020;17:128–44.

    Article  PubMed  CAS  Google Scholar 

  186. Schaefer PM, Kalinina S, Rueck A, von Arnim CAF, von Einem B. NADH autofluorescence—a marker on its way to boost bioenergetic research. Cytometry A. 2019;95:34–46.

    Article  PubMed  Google Scholar 

  187. Digman MA, Caiolfa VR, Zamai M, Gratton E. The phasor approach to fluorescence lifetime imaging analysis. Biophys J. 2008;94:L14–6.

    Article  CAS  PubMed  Google Scholar 

  188. Gratton E, Breusegem S, Sutin J, Ruan Q, Barry N. Fluorescence lifetime imaging for the two-photon microscope: time-domain and frequency-domain methods. J Biomed Opt. 2003;8:381.

    Article  PubMed  Google Scholar 

  189. Stringari C, Cinquin A, Cinquin O, Digman MA, Donovan PJ, Gratton E. Phasor approach to fluorescence lifetime microscopy distinguishes different metabolic states of germ cells in a live tissue. Proc Natl Acad Sci USA. 2011;108:13582–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  190. Sheppard CJR. Multiphoton microscopy: a personal historical review, with some future predictions. J Biomed Opt. 2020;25:1–11.

    Article  PubMed  Google Scholar 

  191. Helmchen F, Denk W. Deep tissue two-photon microscopy. Nat Methods. 2005;2:932–40.

    Article  CAS  PubMed  Google Scholar 

  192. Xu C, Zipfel WR. Multiphoton excitation of fluorescent probes. Cold Spring Harb Protoc. 2015;2015:250–8.

    Article  PubMed  Google Scholar 

  193. Blacker TS, Mann ZF, Gale JE, Ziegler M, Bain AJ, Szabadkai G, Duchen MR. Separating NADH and NADPH fluorescence in live cells and tissues using FLIM. Nat Commun. 2014;5:3936.

    Article  CAS  PubMed  Google Scholar 

  194. Butte PV, Fang Q, Jo JA, Yong WH, Pikul BK, Black KL, Marcu L. Intraoperative delineation of primary brain tumors using time-resolved fluorescence spectroscopy. J Biomed Opt. 2010;15:027008.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  195. De Ruyck J, Famerée M, Wouters J, Perpète EA, Preat J, Jacquemin D. Towards the understanding of the absorption spectra of NAD(P)H/NAD(P)+ as a common indicator of dehydrogenase enzymatic activity. Chem Phys Lett. 2007;450:119–22.

    Article  CAS  Google Scholar 

  196. Evans ND, Gnudi L, Rolinski OJ, Birch DJS, Pickup JC. Glucose-dependent changes in NAD(P)H-related fluorescence lifetime of adipocytes and fibroblasts in vitro: potential for non-invasive glucose sensing in diabetes mellitus. J Photochem Photobiol B. 2005;80:122–9.

    Article  CAS  PubMed  Google Scholar 

  197. Walsh A, Cook RS, Rexer B, Arteaga CL, Skala MC. Optical imaging of metabolism in HER2 overexpressing breast cancer cells. Biomed Opt Express. 2012;3:75–85.

    Article  CAS  PubMed  Google Scholar 

  198. Walsh AJ, Cook RS, Sanders ME, Aurisicchio L, Ciliberto G, Arteaga CL, Skala MC. Quantitative optical imaging of primary tumor organoid metabolism predicts drug response in breast cancer. Cancer Res. 2014;74:5184–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  199. Walsh AJ, Castellanos JA, Nagathihalli NS, Merchant NB, Skala MC. Optical imaging of drug-induced metabolism changes in murine and human pancreatic cancer organoids reveals heterogeneous drug response. Pancreas. 2016;45:863–9.

    Article  CAS  PubMed  Google Scholar 

  200. Acevedo-Acevedo S, Millar DC, Simmons AD, Favreau P, Cobra PF, Skala M, Palecek SP. Metabolomics revealed the influence of breast cancer on lymphatic endothelial cell metabolism, metabolic crosstalk, and lymphangiogenic signaling in co-culture. Sci Rep. 2020;10:21244.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  201. Conklin MW, Provenzano PP, Eliceiri KW, Sullivan R, Keely PJ. Fluorescence lifetime imaging of endogenous fluorophores in histopathology sections reveals differences between normal and tumor epithelium in carcinoma in situ of the breast. Cell Biochem Biophys. 2009;53:145–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  202. Kamel K, O’Brien CJ, Zhdanov AV, Papkovsky DB, Clark AF, Stamer D, Irnaten M. Reduced oxidative phosphorylation and increased glycolysis in human glaucoma lamina cribrosa cells. Invest Ophthalmol Vis Sci. 2020;61:4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  203. Pires L, Nogueira MS, Pratavieira S, Moriyama LT, Kurachi C. Time-resolved fluorescence lifetime for cutaneous melanoma detection. Biomed Opt Express. 2014;5:3080–9.

    Article  PubMed  PubMed Central  Google Scholar 

  204. Kretschmer S, Pieper M, Hüttmann G, Bölke T, Wollenberg B, Marsh LM, Garn H, König P. Autofluorescence multiphoton microscopy for visualization of tissue morphology and cellular dynamics in murine and human airways. Lab Invest. 2016;96:918–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  205. Niesner R, Peker B, Schlüsche P, Gericke K-H. Noniterative biexponential fluorescence lifetime imaging in the investigation of cellular metabolism by means of NAD(P)H autofluorescence. ChemPhysChem. 2004;5:1141–9.

    Article  CAS  PubMed  Google Scholar 

  206. Rajoria S, Zhao L, Intes X, Barroso M. FLIM-FRET for cancer applications. Curr Mol Imaging. 2014;3:144–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  207. Ferri G, Tesi M, Massarelli F, Marselli L, Marchetti P, Cardarelli F. Metabolic response of Insulinoma 1E cells to glucose stimulation studied by fluorescence lifetime imaging. FASEB Bioadv. 2020;2:409–18.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  208. Ranjit S, Malacrida L, Jameson DM, Gratton E. Fit-free analysis of fluorescence lifetime imaging data using the phasor approach. Nat Protoc. 2018;13:1979–2004.

    Article  CAS  PubMed  Google Scholar 

  209. Ranjit S, Malacrida L, Gratton E. Differences between FLIM phasor analyses for data collected with the Becker and Hickl SPC830 card and with the FLIMbox card. Microsc Res Tech. 2018;81:980–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  210. Stuntz E, Gong Y, Sood D, Liaudanskaya V, Pouli D, Quinn KP, Alonzo C, Liu Z, Kaplan DL, Georgakoudi I. Endogenous two-photon excited fluorescence imaging characterizes neuron and astrocyte metabolic responses to manganese toxicity. Sci Rep. 2017;7:1041.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  211. Mothes R, Ulbricht C, Leben R, Günther R, Hauser AE, Radbruch H, Niesner R. Teriflunomide does not change dynamics of nadph oxidase activation and neuronal dysfunction during neuroinflammation. Front Mol Biosci. 2020;7:62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  212. Radbruch H, Bremer D, Guenther R, Cseresnyes Z, Lindquist R, Hauser AE, Niesner R. Ongoing oxidative stress causes subclinical neuronal dysfunction in the recovery phase of EAE. Front Immunol. 2016;7:92.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  213. Mossakowski AA, Pohlan J, Bremer D, Lindquist R, Millward JM, Bock M, Pollok K, Mothes R, Viohl L, Radbruch M, Gerhard J, Bellmann-Strobl J, Behrens J, Infante-Duarte C, Mähler A, Boschmann M, Rinnenthal JL, Füchtemeier M, Herz J, Pache FC, Bardua M, Priller J, Hauser AE, Paul F, Niesner R, Radbruch H. Tracking CNS and systemic sources of oxidative stress during the course of chronic neuroinflammation. Acta Neuropathol. 2015;130:799–814.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  214. Simpson DSA, Oliver PL. ROS generation in microglia: understanding oxidative stress and inflammation in neurodegenerative disease. Antioxidants (Basel). 2020;9:743.

    Article  CAS  Google Scholar 

  215. Hernandes MS, Britto LR. NADPH oxidase and neurodegeneration. Curr Neuropharmacol. 2012;10:321–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  216. Bedard K, Krause KH. The NOX family of ROS-generating NADPH oxidases: physiology and pathophysiology. Physiol Rev. 2007;87:245–313.

    Article  CAS  PubMed  Google Scholar 

  217. Bayerl SH, Niesner R, Cseresnyes Z, Radbruch H, Pohlan J, Brandenburg S, Czabanka MA, Vajkoczy P. Time lapse in vivo microscopy reveals distinct dynamics of microglia-tumor environment interactions—a new role for the tumor perivascular space as highway for trafficking microglia. Glia. 2016;64:1210–26.

    Article  PubMed  Google Scholar 

  218. Radbruch H, Mothes R, Bremer D, Seifert S, Kohler R, Pohlan J, Ostendorf L, Gunther R, Leben R, Stenzel W, Niesner RA, Hauser AE. Analyzing nicotinamide adenine dinucleotide phosphate oxidase activation in aging and vascular amyloid pathology. Front Immunol. 2017;8:844.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  219. Leben R, Köhler M, Radbruch H, Hauser AE, Niesner RA. Systematic enzyme mapping of cellular metabolism by phasor-analyzed label-free NAD(P)H fluorescence lifetime imaging. Int J Mol Sci. 2019;20:5565.

    Article  CAS  PubMed Central  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

This work was funded by an NIH Grant (1 R03 NS118230-01) awarded to KDB.

Author information

Authors and Affiliations

Authors

Contributions

NC, KDB, ED planned the manuscript. NC, KDB, SIA and ED wrote the manuscript and NC, KDB and ED edited the manuscript. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Kimberley D. Bruce.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This 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.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Cleland, N.R.W., Al-Juboori, S.I., Dobrinskikh, E. et al. Altered substrate metabolism in neurodegenerative disease: new insights from metabolic imaging. J Neuroinflammation 18, 248 (2021). https://doi.org/10.1186/s12974-021-02305-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12974-021-02305-w

Keywords