Next Article in Journal
Matrix Metalloproteinase-9 Expression Is Associated with the Absence of Response to Neoadjuvant Chemotherapy in Triple-Negative Breast Cancer Patients
Next Article in Special Issue
Pentylenetetrazole-Induced Seizures Are Increased after Kindling, Exhibiting Vitamin-Responsive Correlations to the Post-Seizures Behavior, Amino Acids Metabolism and Key Metabolic Regulators in the Rat Brain
Previous Article in Journal
Endothelin-1 in Health and Disease
Previous Article in Special Issue
Chronic Administration of Non-Constitutive Proteasome Inhibitor Modulates Long-Term Potentiation and Glutamate Signaling-Related Gene Expression in Murine Hippocampus
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Synthetic Thioesters of Thiamine: Promising Tools for Slowing Progression of Neurodegenerative Diseases †

by
Lucien Bettendorff
Laboratory of Neurophysiology, GIGA Neurosciences, University of Liège, 4000 Liège, Belgium
This article is the record of a presentation made by the author at the Burke Neurological Institute (Weill Cornell Medicine, White Plaines NY) on 1 November 2022.
Int. J. Mol. Sci. 2023, 24(14), 11296; https://doi.org/10.3390/ijms241411296
Submission received: 30 May 2023 / Revised: 5 July 2023 / Accepted: 6 July 2023 / Published: 10 July 2023

Abstract

:
Thiamine (vitamin B1) is essential for the brain. This is attributed to the coenzyme role of thiamine diphosphate (ThDP) in glucose and energy metabolism. The synthetic thiamine prodrug, the thioester benfotiamine (BFT), has been extensively studied and has beneficial effects both in rodent models of neurodegeneration and in human clinical studies. BFT has no known adverse effects and improves cognitive outcomes in patients with mild Alzheimer’s disease. In cell culture and animal models, BFT has antioxidant and anti-inflammatory properties that seem to be mediated by a mechanism independent of the coenzyme function of ThDP. Recent in vitro studies show that another thiamine thioester, O,S-dibenzoylthiamine (DBT), is even more efficient than BFT, especially with respect to its anti-inflammatory potency, and is effective at lower concentrations. Thiamine thioesters have pleiotropic properties linked to an increase in circulating thiamine concentrations and possibly in hitherto unidentified open thiazole ring derivatives. The identification of the active neuroprotective metabolites and the clarification of their mechanism of action open extremely promising perspectives in the field of neurodegenerative, neurodevelopmental, and psychiatric conditions. The present review aims to summarize existing data on the neuroprotective effects of thiamine thioesters and give a comprehensive account.

1. Introduction

Thiamine (vitamin B1, Figure 1), probably the most pleiotropic of the B vitamins, is involved in many cellular processes (energy metabolism, cell survival, antioxidant activity …). Thiamine diphosphate (ThDP), a coenzyme for over 20 enzymes (among which are 6 enzymes and enzyme complexes in mammalian cells), is involved in cell energy metabolism as well as other metabolic functions (Figure 2) [1]. Other derivatives (Figure 1), among which thiamine triphosphate (ThTP, [2,3,4,5]) and adenosine thiamine triphosphate (AThTP, [6,7]), are thought to be involved in coenzyme-independent functions.
ThDP can be hydrolyzed to thiamine monophosphate (ThMP) by various phosphatases [8,9]. Whether ThMP has a specific physiological function remains unknown, but it seems to have some antinociceptive effects mediated by prostatic acid phosphatase [10].
We recommend using the abbreviations ThMP, ThDP, ThTP rather than TMP, TDP, or TTP in order to avoid confusion with phosphorylated thymidine derivatives (officially TMP, TDP, and TTP) as happened in the past [11].
Figure 1. Structural formulas of thiamine and thiamine precursors (prodrugs) with high bioavailability. The sulfur atom of the thiamine thiazolium ring is shown in yellow. The alcohol group of thiamine is shown in green. Organic substituents are in blue, and the phosphate group of BFT is shown in red. SuBT (a symmetric dimer) is a disulfide, while BFT, DBT, and S-BT are thioesters. O-Benzoylthiamine (O-BT) results from the hydrolysis of the thioester in DBT or through an intramolecular rearrangement of S-BT followed by spontaneous thiazole ring closure [10]. Note that for all the open thiazole ring derivatives, the -CH3 group and the sulfur group must be in the “trans” position to allow thiamine to be formed [12]. This is the case for the (Z)-isomer of BFT shown here and which is the only one that should be called “benfotiamine”.
Figure 1. Structural formulas of thiamine and thiamine precursors (prodrugs) with high bioavailability. The sulfur atom of the thiamine thiazolium ring is shown in yellow. The alcohol group of thiamine is shown in green. Organic substituents are in blue, and the phosphate group of BFT is shown in red. SuBT (a symmetric dimer) is a disulfide, while BFT, DBT, and S-BT are thioesters. O-Benzoylthiamine (O-BT) results from the hydrolysis of the thioester in DBT or through an intramolecular rearrangement of S-BT followed by spontaneous thiazole ring closure [10]. Note that for all the open thiazole ring derivatives, the -CH3 group and the sulfur group must be in the “trans” position to allow thiamine to be formed [12]. This is the case for the (Z)-isomer of BFT shown here and which is the only one that should be called “benfotiamine”.
Ijms 24 11296 g001
Two recent clinical studies suggested that a thiamine prodrug, benfotiamine (BFT), might be helpful in slowing down the evolution of Alzheimer’s disease [13,14]. While many studies showed beneficial pharmacological effects of BFT in mouse models of neurodegenerative diseases [15,16,17], the mechanism of action is rarely discussed in terms of active metabolites and their molecular targets. In this review, after a brief introduction to the biochemical properties of thiamine and homeostasis, we will summarize the information available on the possible molecular targets of BFT and other thiamine prodrugs, as well as the possible metabolites involved. We clearly want to distinguish between physiological coenzyme effects and pharmacological, coenzyme-independent effects of thiamine prodrugs. With this respect, we will also examine the limitations and pitfalls of animal and cell culture models and the information that can be obtained.
Figure 2. Thiamine metabolism in a neuron. Thiamine is transported into the cells and then pyrophosphorylated by thiamine pyrophosphokinase (TPK). This free ThDP of high turnover can either bind to transketolase or be transported into mitochondria, where it binds to 2-oxoacid dehydrogenase apoenzymes (see Figure 3). The sum of bound ThDP represents the low turnover pool [9]. Excess cytosolic ThDP is hydrolyzed to ThMP by thiamine pyrophosphatases (ThDPase), and ThMP is hydrolyzed to thiamine by thiamine monophosphatases (ThMPase). A small fraction of mitochondrial ThDP can be phosphorylated to ThTP by FoF1-ATP synthase. ThTP can exit mitochondria and be hydrolyzed to ThDP by a very specific cytosolic 25-kDa thiamine triphosphatase (THTPA) [18,19]. Cytosolic ThDP can also be converted to AThTP by a ThDP adenylyl transferase [20].
Figure 2. Thiamine metabolism in a neuron. Thiamine is transported into the cells and then pyrophosphorylated by thiamine pyrophosphokinase (TPK). This free ThDP of high turnover can either bind to transketolase or be transported into mitochondria, where it binds to 2-oxoacid dehydrogenase apoenzymes (see Figure 3). The sum of bound ThDP represents the low turnover pool [9]. Excess cytosolic ThDP is hydrolyzed to ThMP by thiamine pyrophosphatases (ThDPase), and ThMP is hydrolyzed to thiamine by thiamine monophosphatases (ThMPase). A small fraction of mitochondrial ThDP can be phosphorylated to ThTP by FoF1-ATP synthase. ThTP can exit mitochondria and be hydrolyzed to ThDP by a very specific cytosolic 25-kDa thiamine triphosphatase (THTPA) [18,19]. Cytosolic ThDP can also be converted to AThTP by a ThDP adenylyl transferase [20].
Ijms 24 11296 g002
Figure 3. Role of ThDP-dependent enzymes in cell metabolism. TK, transketolase; PDHC, pyruvate dehydrogenase complex; OGDHC, 2-oxoglutarate dehydrogenase complex. The mitochondrial PDHC-catalyzed reaction is an important source of acetyl-CoA, the precursor of the neurotransmitter acetylcholine, via citrate as an intermediate [21]. Broken lines represent steps involving several reactions.
Figure 3. Role of ThDP-dependent enzymes in cell metabolism. TK, transketolase; PDHC, pyruvate dehydrogenase complex; OGDHC, 2-oxoglutarate dehydrogenase complex. The mitochondrial PDHC-catalyzed reaction is an important source of acetyl-CoA, the precursor of the neurotransmitter acetylcholine, via citrate as an intermediate [21]. Broken lines represent steps involving several reactions.
Ijms 24 11296 g003

2. Transport and Metabolism of Thiamine Derivatives

Biosynthesis of thiamine, and especially the thiazole moiety, is complex and slightly differs among bacteria, yeast, and plants [22,23,24]. As there are no thiamine synthesizing pathways in animals, thiamine biosynthesis will not be discussed here.
In humans, the vast amount of thiamine is of alimentary origin. After transport across the intestinal epithelium, thiamine enters the brain via two transporters of the folate/thiamine transporter family: SLC19A2 (thiamine transporter 1, ThTR1) and SLC19A3 (thiamine transporter 2, ThTR2). These high-affinity transporters function as thiamine/H+ antiporters [25,26,27]. The Km for thiamine is 2.5 ± 0.6 µM for ThTR1 [28] and 27 ± 8 nM for ThTR2 [29]. While these transporters are well characterized from a molecular point of view, their distribution in the human brain remains poorly understood, and to our knowledge, only one study partially addressed this issue [30]. According to this study, ThTR1 and ThTR2 are expressed in the blood vessels throughout the brain, with ThTR1 immunoreactivity present only at the luminal side of the blood vessels, while ThTR2 is expressed at the basement membrane and in perivascular pericytes. Hence, both types of transporters may be involved in thiamine transport across the blood–brain barrier. In the human brain, ThTR2 seems to have a higher expression pattern in neurons than ThTR1 [30].
Inside cells, thiamine is pyrophosphorylated to ThDP by thiamine pyrophosphokinase (TPK) [31,32]. The main fates of ThDP are binding to apotransketolase or transport into mitochondria [8,9]. ThDP is transported into mitochondria by a specific carrier (the mitochondrial thiamine pyrophosphate transporter, MTPPT), probably in exchange for nucleoside di- or tri-phosphates [33] (Figure 2). ThDP can be hydrolyzed to ThMP, which in turn can be hydrolyzed to thiamine by various phosphatases, including prostatic acid phosphatase [1,10]. Extracellular ThMP probably originates from cellular efflux through the reduced folate transporter 1 [34].
Triphosphorylated thiamine derivatives exist in most cells (for a discussion, see [4], Figure 2). ThTP is synthesized in brain mitochondria by oxidative phosphorylation, probably involving FOF1-ATP synthase [35]. AThTP, probably synthesized by a cytosolic thiamine diphosphate adenylyl transferase [7,20], exists in small amounts in mammalian tissues [36]. However, both compounds are probably not involved in the phenomena described here and will not be discussed further.

3. Coenzyme Function of Thiamine Diphosphate

3.1. ThDP-Dependent Enzymes in Mammals

ThDP is the coenzyme for many enzymes and enzyme complexes involved in carbohydrate and amino acid catabolism. In mammals, the most important are pyruvate (PDH, EC 1.2.4.1) and 2-oxoglutarate (OGDH, EC 1.2.4.2) dehydrogenases that catalyze two essential steps in citric acid metabolism ([26], Figure 3) and transketolase (TK, EC 2.2.1.1) in the pentose phosphate pathway (PPP). Hence, ThDP-dependent enzymes PDH and OGDH are responsible for nearly two-thirds of the CO2 released during the complete oxidation of glucose. Indeed, only about 5% of glucose is metabolized through the PPP in the brain [37].
As these enzymes catalyze essential steps in glucose oxidation, it is obvious that ThDP is an indispensable coenzyme for energy metabolism, and it is not surprising that thiamine deficiency will have deleterious effects on organs that are particularly dependent on oxidative metabolism, such as the nervous and cardiovascular systems. In addition to catalyzing essential reactions in cell energy metabolism [38], PDH and OGDH are also important for the synthesis of the neurotransmitters glutamate, GABA, and acetylcholine [39].
TK is a key rate-limiting enzyme of the PPP [40]. This pathway is essential for the synthesis of (1) NADPH, which is important for reductive biosynthesis and protection against oxidative stress, and (2) ribose 5-phosphate.
In addition, ThDP is a coenzyme for at least three other mammalian enzymes [41]:
  • Mitochondrial oxoacid dehydrogenase, involved in lysine-derived 2-oxoadipate catabolism [42,43];
  • Mitochondrial branched-chain 2-oxoacid dehydrogenase (EC 1.2.4.4) [44] is required for the catabolism of branched-chain amino acids;
  • Peroxisomal 2-hydroxyacyl-CoA lyase (EC 4.1.2.63) in α-oxidation of phytanoic acids [38].
However, the implication of the latter enzymes in the symptoms of thiamine deficiency disorders is less documented [26].

3.2. ThDP Mediated Catalysis

The ThDP molecule is composed of three parts: the catalytically active thiazole heterocycle, the pyrimidine ring, and the phosphorylated hydroxyethyl chain (Figure 4). Ronald Breslow was the first to propose that the thiazolium ring is the core element for catalysis by ThDP [45,46]. Though the pyrimidine ring of ThDP is not essential for catalysis, a conserved Glu residue in ThDP-dependent enzymes facilitates the adoption of the so-called V-conformation, forcing an angle between the thiazolium and pyrimidine rings in such a way as to bring closer the 4′-amino group (as an imino tautomer) and the thiazolium ring in order to favor deprotonation of the C-2 proton. The diphosphate group plays no role in catalysis and is only required for tight binding of ThDP to the apoenzymes.
In mammalian enzymes, ThDP catalyzes the cleavage of C-C bonds immediately adjacent to a carbonyl group (α cleavage). Indeed, cleavage of a C-C bond requires stabilization of the product, which is immediate for β cleavage but not for α cleavage, thus requiring an additional tool brought by the thiamine thiazole ring [47].

4. Thiamine Status in Humans and Inter-Organ Homeostasis

Thiamine status in humans is an important issue in view of the high number of populations at risk of thiamine deficiency around the world and, in particular, in children [48,49]. The clinical symptoms of thiamine deficiency syndromes are highly variable and affect the cardiovascular and nervous systems. Typical thiamine deficiency syndromes are beriberi and Wernicke’s encephalopathy. Both are highly reversible after administration of thiamine [50]. In addition to the severe forms of thiamine deficiency, subclinical thiamine deficiency can be relatively common in humans.

4.1. Thiamine Status in Humans

One of the central questions is to understand the distribution of thiamine derivatives between different organs. In rats, only 1.5% of total thiamine is found in the brain, while 22% is found in the liver [51]. Total thiamine levels are three to four times lower in the brain than in the liver [36]. More significantly, during thiamine deficiency, thiamine levels decrease less steeply in the brain than in the liver [51,52]. The picture that is emerging from rodent models suggests that when thiamine (or thiamine prodrugs) is administered, it is captured and phosphorylated in the liver and, to a lesser extent, in erythrocytes both compartments acting as transient buffers of blood thiamine levels. When thiamine levels normalize again, mainly through renal excretion, hepatocytes, and erythrocytes release excess thiamine, which is then eliminated. In contrast, brain thiamine levels are highly regulated such as to minimize variations in thiamine and thiamine phosphate levels in either direction.
In humans, we have only access to circulating thiamine levels, and it is not clear to what extent they reflect brain thiamine (and ThDP) levels. Indeed, we do not know how brain thiamine levels are regulated in humans, but there is no reason to assume that the mechanisms are different from those in rodents.
Therefore, an important issue is to know how much brain ThDP levels must decrease before overt symptoms of thiamine deficiency appear. In rats, brain thiamine levels must drop to about 10–20% of normal before symptoms of thiamine deficiency appear [36,52,53]. Expressed in pmol/mg, this is not much lower than normal levels in the human brain (Table 1). If we assume that ThDP requirements are not different between humans and rodents, then both circulating and brain ThDP levels in healthy humans are close to the threshold of deficiency (Table 1). This would explain that a subclinical thiamine status in humans is relatively frequent [49].
Hence, it seems that erythrocyte and liver thiamine content is more or less proportional to the plasma thiamine concentrations, while brain thiamine content is much more tightly controlled. The latter is only affected when plasma thiamine concentrations are decreased over an extended period during a chronic thiamine deficiency period.
Two features may play an important role in brain thiamine homeostasis: transport across the BBB and regulation of ThDP synthesis.

4.2. Thiamine Transport across the Blood–Brain Barrier Is Limited by Trans-Stimulation

When loading cultured mouse neuroblastoma cells with 14C-thiamine, we observed that after eliminating the extracellular tracer, the cells released 14C-thiamine into the extracellular medium. This release was stimulated by extracellular thiamine in the 0.1–1 µM range [60].
Years later, using an in situ rat brain perfusion method, it was demonstrated that thiamine efflux through the BBB was also trans-stimulated by thiamine in the sub-micromolar range [61]. Plasma thiamine concentrations are approximately 10 nM in humans (Table 1) and 100 nM in rodents. A trans-stimulation would occur when the plasma thiamine concentration is increased above 100 nM, which is easily attained in humans after administration of thiamine prodrugs.
The above data suggest that a mechanism protecting the brain from a thiamine overload is operating through a thiamine/thiamine exchange. It is not clear whether the known transporters can switch from a thiamine+/H+ to a thiamine+/thiamine+ exchange mode (Figure 5) or whether a specific protein, different from ThTR1 or ThTR2, mediates thiamine efflux.

4.3. TPK Is Regulated by Retroinhibition of ThDP

We observed numerous times that in animals treated with thiamine or BFT, blood thiamine levels may increase by several orders of magnitude, but ThDP levels are only slightly increased [56]. A similar observation was made in cultured cells exposed to BFT, suggesting a very tight control of ThDP synthesis [9,56].
Indeed, when studying the kinetics of recombinant mouse TPK in the forward (physiological) direction (thiamine + ATP ⇆ ThDP + AMP) (Figure 2 and Figure 6), we observed that the reaction stopped even when only a small proportion of thiamine had been converted to ThDP [32]. This could be explained by the observation that ThDP is a potent non-competitive inhibitor of TPK with a Ki of approximately 0.2–0.4 µM, of the same order of magnitude as free intracellular ThDP concentrations. This observation may explain why, even with a large excess of thiamine, ThDP does not accumulate within cells (Figure 6).
A possible justification for such behavior would be the protection of the brain from thiamine depletion. Indeed, if thiamine could be phosphorylated to ThDP without limit (the equilibrium of the reaction is strongly towards ThDP synthesis [32]), this could lead to a capture of plasma thiamine by red blood cells and hepatocytes (the latter having a high capacity transporter for thiamine, see Section 6.1) and depletion of the plasma thiamine pool, decreasing its availability for the brain [32].

5. Development of Synthetic Thiamine Derivatives with High Bioavailability

Thiamine transport across cell membranes is a relatively slow process [27]; therefore, several synthetic thiamine prodrugs with high bioavailability were developed, mainly with the aim of rapidly restoring brain thiamine levels after deficiency. These derivatives have in common that they are relatively hydrophobic, or they generate hydrophobic metabolites allowing a rapid diffusion across the intestinal permeability barrier [26].
We must distinguish between two chemically distinct types of prodrugs (Figure 7):
  • Disulfides, among which the most prominent representative is SuBT and its product of hydrolysis thiamine disulfide [63];
  • Thioesters, represented by BFT [56,64].
Both types of molecules have an open thiazole ring, but the distinction between the two groups is important as their metabolization requires different reactions to form thiamine: disulfides require reduction of the disulfide bond, while thioesters require hydrolysis.

6. BFT in Animal Studies

6.1. Administration of BFT in Mice Does Not Significantly Increase Brain ThDP Levels

There are only a few systematic pharmacokinetic studies with BFT, and those existing limit themselves to the determination of thiamine, ThMP, and ThDP in blood for the estimation of its bioavailability [65,66]. To our knowledge, no attempt has been made to detect BFT or potential byproducts (S-BT or O-BT, for instance, Figure 1). Among the many studies that tested the potential therapeutic properties of BFT, rare are those that tested their effect on levels of thiamine derivatives in the target organ (see [67], for instance). As late as 2007, no data on the effects of BFT on nervous tissue were available, and so it became overdue to check the effect of this molecule on brain thiamine levels.
Surprisingly, the administration of high levels of BFT increased total thiamine in the blood and liver, but not in the brain (Figure 8, [56]). The absence of the effect of BFT on brain thiamine levels, though it strongly increases circulating thiamine levels, can be explained by the above-mentioned trans-stimulation of thiamine uptake across the BBB.
In addition, while thiamine concentrations increased by two or three orders of magnitude above control levels in blood and liver, ThDP concentrations were only marginally increased, in agreement with a feedback inhibition of TPK by ThDP [32] (Figure 6).
A third factor might be a slower thiamine transport across the BBB than into hepatocytes. Indeed, in rats, radioactive thiamine entered the liver much more rapidly than the brain [68], which would suggest the involvement, in the liver, of a high-capacity thiamine transporter different from ThTR1 or ThTR2 (which are dependent on the H+-gradient). It has been suggested that the organic cation transporter 1 might be involved in thiamine uptake by hepatocytes [69,70,71]. It was also claimed that thiamine uptake in isolated rat hepatocytes is Na+-dependent [72,73], but this is probably rather a dependence on the Na+ gradient rather than Na+ itself [27]. Organic cation transporters are Na+-independent, but their electrogenic nature makes them indirectly dependent on the Na+ gradient via the membrane potential [69].

6.2. BFT Has Beneficial Effects in Animal Models of Neurodegenerative Diseases, Stress, and Anxiety

Paradoxically, in addition to the intriguing observation that BFT has apparently no significant effect on brain thiamine content, many studies document neurological effects of BFT while, at the same time, reporting no increase in brain ThDP levels, suggesting that the coenzyme function of ThDP is not involved.
For instance, Pan et al. (2010) [25], in a landmark study, investigated the effects of chronic treatment by BFT in amyloid precursor protein/presenilin (APP/PS1) double transgenic mice, a classical model of Alzheimer’s disease. Oral treatment (100–200 mg/kg/day) for 8 weeks enhanced spatial memory and reduced amyloid plaque numbers and phosphorylated tau levels in cortical areas. Remarkably, no such beneficial effects were seen when the mice were treated with the thiamine prodrug fursultiamine suggesting a certain specificity for BFT.
More recently, a similar treatment with high doses of BFT (200 mg/kg per day for 8 weeks) was applied in a mouse model of tauopathy, P301S mice [26]. In this model, BFT increased lifespan, prevented the death of spinal neurons, and improved behavioral deficits. It also decreased oxidative stress, inflammation, accumulation of advanced glycation end products (AGEs), tau phosphorylation, and formation of neurofibrillary tangles in the cerebral cortex and hippocampus.
In addition to these therapeutic actions in mouse models of neurodegeneration, BFT treatment counteracted anxiety and depression-like behavior as well as aggression linked to emotional stress. Stress was induced either by a modified swim test, the presence of a predator for five consecutive nights [48,49], or chronic ultrasound exposure for 20 days [50].
BFT also normalized plasticity markers, stimulated neurogenesis, and improved memory:
  • The administration of BFT decreased aggression, reversed ultrasound-induced changes in GluA1 and GluA2 subunit expression, and reversed the decreased expression of plasticity markers [74];
  • BFT (more efficiently than thiamine) counteracted the predator stress-induced decrease in the proliferation and survival of newborn immature neurons in the subgranular zone of the dentate gyrus [75].
These data suggest that BFT may have a therapeutic potential not only in neurodegenerative diseases but also in other brain pathologies, such as major depression linked to stressful events. At least some of these effects may be mediated by GSK-3β and possibly GSK-3α downregulation [16,76]. Changes in expression and phosphorylation of GSK3β were reported in mouse models of stress-induced anxiety and depression [76]. In the prefrontal cortex, mRNA levels of GSK3β were increased in the modified swim test and when the mice were exposed to predator stress. This increase was fully reversed when the animals were treated with BFT.

6.3. BFT Has Coenzyme-Independent Antioxidant and Anti-Inflammatory Properties

It is essential to know the mechanism by which BFT exerts these various effects.
In all the cited experiments, the mice received a normal chow, rich in thiamine, and we can suppose that all ThDP-dependent enzymes were saturated by the coenzyme ThDP. Moreover, treatment with BFT did not increase brain ThDP levels [15,16,56,75]. Hence, the observed effects are probably coenzyme independent.
According to Tapias et al., BFT activates the Nrf2/antioxidant responsive element (ARE) [15]. However, the activation of this pathway seems to require relatively high BFT concentrations (>100 µM) in cultured cells [77]. As such high concentrations of either BFT or one of its metabolites are unlikely to be reached in the brain parenchyma, and Nrf2 is unlikely to be a direct target of thiamine prodrugs in vivo. However, it is possible that the NrF2 antioxidant pathway is indirectly activated.
Gorlova et al. report a normalization of brain oxidative stress markers by BFT after ultrasound-induced aggression [74]. They suggest that this effect may be due to decreased GSK-3β activity, ultimately impacting Nrf2 signaling. Indeed GSK-3β phosphorylates Nrf2, resulting in its nuclear exclusion [78]. In this hypothesis, the effect of BFT would be on GSK-3β, or on a factor upstream of GSK-3β rather than Nrf2. Such a hypothesis would explain the activation of Nrf2 target genes by BFT in vivo [15].
BFT also has strong antioxidant effects in neuroblastoma cells [77] and BV2 microglial cells, probably targeting glutathione metabolism [79].
Microglial cells are resident macrophage-like immune cells of the nervous system. The overactivation of microglia induces the production of neurotoxic reactive oxygen and nitrogen species. This process is likely to play an important role in neurodegenerative processes [80].
In addition to antioxidant effects, BFT also significantly reduced anti-inflammatory markers in vivo, among which reduced iNOS expression and NF-κB p65 immunoreactivity [15,75,81]. BFT counteracts the morphological changes corresponding to the LPS-induced activation of the microglial cells. In addition, it decreases the production of proinflammatory mediators such as iNOS, COX-2, TNF-α, and IL6. These effects of BFT are likely mediated by the suppression of NF-κB translocation to the nucleus [82].

7. BFT in Human Clinical Studies

7.1. Thiamine Homeostasis Is Disrupted in Patients with Alzheimer’s Disease

There are only a handful of studies on thiamine status in Alzheimer’s patients, but they consistently show a lower status compared to age-matched controls. Indeed, several independent studies have reported abnormalities of thiamine homeostasis in patients suffering from neurodegenerative diseases. An early study reported a 20% decrease in ThDP levels in postmortem brains of patients with Alzheimer’s disease compared with adequately matched controls [83]. Moreover, a 40–50% decrease in postmortem ThDP levels was observed in patients with frontal lobe degeneration of the non-Alzheimer type [84], but in this study, the number of patients was small (n = 6). Much more recently, it was shown that patients with Alzheimer’s disease had 20–30% decreased blood ThDP concentrations compared to controls, while ThMP and thiamine levels were not impaired [85,86]. The same authors reported increased ThDPase and ThMPase activities which might contribute to the reduction of ThDP concentrations [87]. They also showed that ThDP reduction strongly correlates with brain glucose hypometabolism in Alzheimer’s disease [88], suggesting that a “subclinical” thiamine deficiency (i.e., not sufficient to produce overt symptoms of thiamine deficiency such as Wernicke’s encephalopathy, but nevertheless significant) might contribute to the clinical manifestations of Alzheimer’s disease.
It is to be expected that such a subclinical deficiency would be immediately corrected by the administration of thiamine or thiamine precursors. Indeed, it is well-known that thiamine deficiency disorders are readily reversible [49].
It is not clear why thiamine homeostasis is impaired in Alzheimer’s disease patients, but an explanation may come from a recent study showing a significantly lower expression of both THTR-1 and THTR-2 in the brains of Alzheimer’s disease patients leading to lower brain thiamine content [89]. The authors suggest that proinflammatory cytokines impair thiamine transport in these patients.

7.2. Therapeutic Effects of BFT on Patients with Alzheimer’s Disease

The first study testing BFT (300 mg daily over 18 months) in Alzheimer’s disease patients was carried out by the group of Chunjiu Zhong at Fudan University in Shanghai. The study was conducted on a relatively small sample of mild to moderate Alzheimer’s disease patients (n = 5) [13]. BFT significantly improved the cognitive abilities of the patients independently of brain amyloid accumulation.
A second study on a larger sample of patients (n = 34) was conducted by the group of Gary Gibson at the Burke Neurological Institute in New York [14]. BFT (2 × 300 mg for 12 months) slowed down cognitive decline compared to the placebo group, as viewed by the clinical dementia rating, and this effect was stronger in the APOE 4 non-carriers. Furthermore, BFT also significantly reduced increases in blood AGE charge. Serum metabolic and lipid profiling on a small subset of these patients showed that BFT normalized several amino acids (Tyr, Trp, His) and lipids (phosphatidylcholine) to levels observed in the control placebo group [90].
The results obtained in these clinical studies, though not conclusive, are indeed very promising. An important point concerns the doses used in both studies. In mice, the effects of BFT were much more spectacular than in the two human studies, but the doses used were much higher (200 mg/kg per day). Extrapolated to humans, such a dose would correspond to approximately 14 g of BFT per day. This could, of course, be one of the reasons for the differences observed between animal and human studies.
A further point requiring attention is that Alzheimer’s disease patients have lower brain and blood thiamine levels (see Section 7.1), a situation not yet reported in animal models [15,16]. The reason is probably that experimental animals are kept on a thiamine-rich chow, compensating for any effects of the pathology on thiamine transport. Therefore, we must expect that BFT administration in humans, in addition to its pharmacological effects, will correct thiamine levels.

8. O,S-Dibenzoylthiamine, a Promising New Thiamine Prodrug

In view of the above-mentioned considerations, it should be of interest to have a derivative that would cross membranes more rapidly than BFT. Indeed, the latter, though often and incorrectly referred to as a hydrophobic molecule, is insoluble in lipids and requires dephosphorylation to the lipid-soluble S-BT (Figure 1) by ectophosphatases. Hence, S-BT would be an obvious candidate, but it is relatively unstable and tends to spontaneously lose the benzoyl group and reform thiamine. Esterification of the thiamine hydroxyethyl chain increases the stability of open thiol forms. In the case of BFT, the presence of a phosphoryl group strongly alters the solubility of the molecule: it is soluble in aqueous solvents at slightly alkaline pH [91].
An alternative would be to replace the phosphoryl group with a more hydrophobic group. An obvious replacement possibility would be the substitution by a second benzoyl group.
This molecule, O,S-dibenzoylthiamine (DBT, sometimes called bentiamine, Figure 1), is commercially available, but strangely, at the beginning of our studies, only one scientific publication specifically pertaining to the use of this compound in mammals was available [92]. DBT has been known for a long time and is allowed as a food additive in Japan. DBT is composed of a thiamine molecule with an open thiazolium ring linked to two molecules of benzoate, one via a thioester and the other via an O-ester bond. Hence, its conversion to thiamine requires the action of two different enzymes: a thioesterase and an O-esterase [93]. No toxic or tumorigenic side effects have been reported [92], yet there are practically no data about its biological effects. In salmon yearlings, it was better tolerated than thiamine or BFT, and it also led to higher retention of thiamine over time [94].
Hence, this hitherto unexplored derivative could meet the requirements of safety and could be a superior alternative to BFT [93]. Therefore, we decided to test DBT first in cellular models and second in a transgenic mouse model of amyotrophic lateral sclerosis (ALS).

8.1. DBT Has Strong Antioxidant and Anti-Inflammatory Properties in Cultured Cells

As already observed for BFT, DBT has strong antioxidant and anti-inflammatory properties, the latter being even more powerful than those of BFT [93]. As shown for BFT, these effects are independent of the coenzyme role of ThDP.

8.1.1. DBT Has Powerful Antioxidant Effects Mediated by Glutathione

At low concentrations (5–10 µM), DBT protects neuroblastoma cells from paraquat toxicity by counteracting oxidative stress [93]. To obtain the same protection with BFT, 5–10-fold higher concentrations are required. DBT increases the synthesis of reduced glutathione and NADPH in an Nrf2-independent manner [93]. Indeed inhibition or knock-down of Nrf2 did not prevent the antioxidant effects of BFT or DBT [93]. A direct stoichiometric effect was also excluded.

8.1.2. DBT Has Powerful Anti-Inflammatory Effects Mediated by NF-κB

DBT has more potent anti-inflammatory properties than BFT. At 10–50 µM, it protects BV2 cells from LPS-induced inflammatory processes (increased expression of iNOS and TNF-α and production of nitric oxide) by translocation of NF-κB into the nucleus [93].

8.2. DBT Protects from Motor Neuron Degeneration in a Transgenic Mouse Model of ALS

Chronic administration of DBT (25 mg/kg per day) relieved depressive-like behavior in mice submitted to chronic ultrasound stress [93].
Moreover, DBT arrested motor dysfunction in Fused in Sarcoma (FUS) transgenic mice, a model of amyotrophic lateral sclerosis (ALS). Mutations in the gene encoding the RNA/DNA-binding FUS protein have been detected in familial ALS patients. FUS is a critical component of the oxidative damage repair complex which might explain its role in neurodegeneration. In the FUS [1–359]-transgenic mouse model of ALS, DBT-treated mice displayed improvements in physiological outcomes, motor function, and muscle atrophy compared to vehicle, and the treatment normalized brain levels of GSK-3β, GSK-3β mRNA, and IL-1β mRNA in the spinal cord [17].
No clinical data are available on DBT, but a very recent case report showed that BFT positively impacted a patient with ALS [95]. Though a placebo effect cannot be excluded, this study shows that therapy with these derivatives could be useful in patients suffering from ALS.

9. Hypotheses on the Mechanisms of Action of Synthetic Thiamine Derivatives

Thiamine prodrugs, and in particular BFT, have been studied at three different levels: (1) cultured cells, (2) rodent models of stress and neurodegenerative diseases, and (3) human patients with mild cognitive impairment or early symptoms of Alzheimer’s disease. These derivatives have strong antioxidant and anti-inflammatory properties, the latter more pronounced for DBT compared to BFT.
Among the various questions remaining, two are essential for understanding the issues: (1) which is the active metabolite of thiamine prodrugs and (2) which are the molecular targets. Our hypotheses concerning these points are summarized in Figure 9 and Figure 10.

9.1. The Pharmacological Effects of BFT and DBT Are Most Probably ThDP-Independent

In thiamine-deficient animals, BFT and DBT, even at low doses, will rapidly restore normal thiamine levels. Pharmacological (mainly antioxidant and anti-inflammatory) effects are observed at high doses (generally above 100 mg/kg). The puzzling, though many times reproduced, observation is that these pharmacological effects in animals are observed while thiamine concentration is only minimally increased in the brain and ThDP concentration not at all.
Unfortunately, this point is ignored in most studies, possibly because the non-coenzyme functions of thiamine are still frowned upon, despite many arguments in their favor [4,96,97,98]. Therefore, we carefully examined this possibility, and our conclusion is that the pharmacological action of DBT and SBT are most probably coenzyme-independent:
  • Most of the studies were conducted in non-deficient cells or animals. Normal chow is rich in thiamine, and it must be understood that thiamine prodrugs are administered under conditions of thiamine saturation, where thiamine-dependent enzymes are saturated with ThDP. Indeed, when tested, ThDP-dependent enzymes are not or, at best, minimally increased [15,93];
  • In animals, administration of thiamine prodrugs does not increase brain ThDP content [15,16,56,75]. This may be linked to feedback inhibition of TPK by ThDP (Figure 6) and, as already mentioned above, is not in favor of an implication of coenzyme function;
  • In cell culture models, the antioxidant and anti-inflammatory effects of thiamine prodrugs persist even when ThDP synthesis is blocked by pyrithiamine [77,93];
  • Many studies claim that BFT activates TK [67,99,100]. It is not clear how this could happen in the absence of a thiamine deficiency as, under normal conditions, TK is close to saturation with ThDP (see point 1). One possibility is that TK expression is upregulated at the mRNA level. Such mechanisms cannot be excluded as TK expression is controlled by Nrf2 [101]. However, using cultured neuroblastoma cells, we observed antioxidant effects of BFT and DBT, though neither TK activity nor expression was increased [93].
These results strongly suggest that ThDP-dependent enzymes are not involved in the pharmacological effects of thiamine prodrugs. Hence, the therapeutic effects of BFT and DBT cannot be reduced to only normalizing coenzyme levels.
If the actions of thiamine precursors are coenzyme-independent (we prefer coenzyme-independent rather than ThDP-independent as ThDP might have other functions in addition to being a coenzyme), this question is, of course, strongly related to their cellular metabolism responsible for the production of the active metabolite.

9.2. If Not ThDP, What Is the Active Metabolite of BFT and DBT?

This question is clearly essential if we aim to understand the action of these compounds and develop better-targeted and more efficient drugs. From the experiments carried out in our laboratory, we can deduce the scheme depicted in Figure 9.
  • BFT requires extracellular dephosphorylation to S-BT (occurring in the intestine in vivo). S-BT, as well as DBT and SuBT, cross plasma membranes by simple diffusion. Intracellular S-BT is subject to the action of thioesterases to yield the open thiol form of thiamine [77]. In an aqueous medium, S-BT spontaneously undergoes an intramolecular rearrangement to O-BT, probably as a result of acyl-migration from sulfur to oxygen [10]. O-BT is then spontaneously converted to thiamine. The conversion of S-BT to O-BT is favored by alkaline pH. Though O-BT can be enzymatically converted to thiamine in most tissues, it does not seem to be more efficient than thiamine in raising tissue thiamine content [102]. O-BT has not been pharmacologically tested, but its spontaneous synthesis from BFT has been reported [10], and docking studies suggested that O-BT might be a good ligand for Keap1, competing for Nrf2 and favoring the translocation of Nrf2 to the nucleus [15]. Therefore, it is not excluded that O-BT might be pharmacologically active. Thiochrome, an oxidized form of thiamine present in small amounts in cells, was not increased by treatment of Neuro2a cells with BFT [77];
  • DBT can either be hydrolyzed to S-BT or yield O-BT through the action of thioesters. O-BT might be transformed to S-BT by acyl migration from an O to an S atom [10]. S-BT can then proceed further to the open thiol form;
  • SuBT, which was not discussed here as it has not been tested in neurodegenerative diseases [63], requires hydrolysis to thiamine disulfide, which must be reduced to the open thiol form. The reductant could be reduced glutathione (GSH) which is the most abundant intracellular thiol [103], or NADPH in a reaction of the type catalyzed by glutathione reductase [104]. SuBT has antioxidant [77] and weak anti-inflammatory [93] properties at a level similar to BFT but, for the latter, did not match DBT;
  • All three thiamine prodrugs considered here will ultimately yield an open thiol form before being processed into thiamine [66]. The formation of the open thiol form from thiamine is unlikely under physiological conditions, but it is a thermodynamically stable form at alkaline pH (pH > 10) [105]. Nevertheless, a challenging hypothesis would be that there exist enzymes able to catalyze the formation of the open form from thiamine [105,106].
The chemistry of the open thiol ring is relatively unexplored. This form is short-lived, but it is possible that in the presence of oxidants (such as G-S-S-G), the open intermediate may react to form -S-S- disulfides with free cysteine thiols of proteins [107]. This might also be an interesting pathway of post-translational protein regulation [107]. Hence, it cannot be excluded that the open thiol form may be pharmacologically active.
Another possibility is that the active compound is simply thiamine formed from the open thiol form, and administration of prodrugs is pharmacologically more efficient than administration of thiamine, because it has a higher impact on intracellular thiamine concentrations. In cultured cells, intracellular thiamine (but not ThDP) concentrations were significantly correlated with cell survival after paraquat poisoning [77].
A further possibility would be that a presently unknown compound is formed from thiamine prodrugs that are stable enough to resist transport from the blood into the brain. So far, neither BFT nor S-BT has been demonstrated to appear in the blood. The absorption of DBT has not been studied, and it is not clear if any reaches the bloodstream. Even if some DBT reaches the bloodstream, it is highly improbable that any would reach the brain parenchyma as it would enter red blood cells and ultimately brain endothelial cells, where it would be transformed into thiamine. Hence, all experimental data suggest that most of the thiamine prodrugs end up as thiamine in the blood.
Though treatment with prodrugs does not lead to increased ThDP levels in the brain, unphosphorylated thiamine is consistently, but not always significantly, increased. Most studies use mouse brain samples of approximately 50–100 mg, which contain all structures, neurons, astrocytes, axons, and synapses. It is possible that increased thiamine levels are limited to very local structures within the sample, but with significant effect. Alternatively, a pharmacologically active compound may be formed from locally increased thiamine.
It is not clear which cell types (endothelial cells, astrocytes of neurons) are involved in this phenomenon. However, brain endothelial cells are in direct contact with blood, the compartment where the highest thiamine increase is observed, and these cells might act as thiamine sensors sending signals to neurons.

9.3. Which Are the Targets of BFT and DBT?

As described above, both BFT and DBT have antioxidant and anti-inflammatory effects. In addition, they seem to impact brain plasticity (Figure 10).
  • BFT and DBT (and, to a lesser extent, SuBT and thiamine) have strong antioxidant properties both in vitro [75,77,93] and in vivo [15,17,79]. In cultured neuroblastoma cells, thiamine prodrugs normalized paraquat reduced NADPH and GSH levels in cultured neuroblastoma cells but were independent of TK activity and expression or Nrf2/ARE signaling. These prodrugs clearly seem to be involved in maintaining a high [GSH]/[GSSG] ratio and high NADPH concentrations. These results might be explained by an effect on glucose 6-phosphate dehydrogenase (G6PD). Indeed, this rate-limiting enzyme of the PPP is highly regulated, both at the level of transcription and post-translation [108,109]. However, it must be kept in mind that in addition to the oxidative part of the PPP, other sources of NADPH, such as cytosolic dehydrogenases and mitochondrial trans-hydrogenase, exist [103,109];
  • Decreases in GSK-3β expression or activity regularly turn out after BFT or DBT treatment, in particular in in vivo studies [74,75,76,110] (but [15]). GSK-3β is a pleiotropic serine-threonine protein kinase able to phosphorylate over 100 substrates and play a role in numerous cellular functions affected in Alzheimer’s disease [111]. Decreased activity of GSK-3β activity by phosphorylation on Ser 9 may be responsible for changes in brain plasticity observed [17,74,77]. In addition, GSK-3β might be involved in the antioxidant response by activating Nrf2, an upstream regulator of G6PD and TK [112,113];
  • BFT and, in particular DBT, have strong anti-inflammatory properties that seem to be mediated by NF-κB. Indeed, these compounds block the translocation of its p65 subunit to the nucleus and suppress LPS-induced iNOS expression and cytokine release [79,93]. This is of particular interest in view of the recent results showing that cytokines decrease thiamine transporter expression and activity and that the expression of these transporters is reduced in the brains of Alzheimer’s disease patients [89]. According to our data, DBT is a much more potent anti-inflammatory agent than BFT (and SuBT) in cultured BV-2 cells. We have no explanation for this difference, especially as it does not seem to be more efficient in raising intracellular thiamine concentrations. The differences may be related to different metabolization schemes with different proportions of intermediates O-BT, open thiol form) generated (Figure 9). This might also suggest that antioxidant and anti-inflammatory properties are mediated via different molecular targets;
  • A significant amount of prodrug is already transformed into thiamine in the intestine, thereby affecting the gut microbiome. This might affect the enteric nervous system and, indirectly, the central nervous system. Indeed, thiamine affects the relative abundance of bacterial populations favoring those producing short-chain fatty acids [114]. A very appealing hypothesis would be that BFT corrects gut dysbiosis in Alzheimer’s patients, reducing inflammation and oxidative stress in the brain through the microbiota-gut-brain axis [115,116]. Interestingly NF-κB and Nrf2, both potential targets of BFT, seem to be involved in these processes. However, the gut microbiome hypothesis alone is not sufficient to explain the antioxidant and anti-inflammatory properties of BFT and DBT, as such effects were also observed in cell cultures, suggesting that other mechanisms coexist, as illustrated in Figure 10.

10. Conclusions

It is now well documented that human patients with Alzheimer’s disease have impaired thiamine homeostasis, and BFT administrations correct these abnormalities linked to the suboptimal working of ThDP-dependent enzymes [117]. However, we think that in addition to the coenzyme effects, thiamine prodrugs also exert coenzyme-independent effects.
As summarized in Figure 10, these effects are numerous. Some data discussed here were obtained in rodents, while others were obtained in cultured cells. The latter can constitute a simpler model for studying the metabolization of thiamine prodrugs than living animals. However, it is important to be aware of the limitations of cellular models. Indeed, cultured cells are directly exposed to the prodrugs, while brain cells are separated from these molecules by several compartments and barriers (intestinal barrier, blood, and the BBB).
Hence, the situation is different in animal models, where the prodrugs are administered orally. In the intestine, BFT is dephosphorylated and crosses the intestinal epithelium as S-BT or thiamine. There is so far no experimental evidence that these prodrugs reach the brain as such. Nevertheless, the results obtained with BFT and DBT in cultured cells are consistent with the data obtained in in vivo models, as they clearly show that they have antioxidant and anti-inflammatory properties. Some other studies also showed that BFT prevents the formation of AGEs [15,67]. The exact mechanism and the active derivative involved remain unknown, but all experiments are consistent with effects mediated by GSK-3β and NF-κB.
Considering that there exists no effective treatment for Alzheimer’s disease, we think that thiamine prodrugs have the potential to slow the evolution of the disease. Though only preliminary, present clinical studies show effects of BFT that are at least equal if not superior to monoclonal antibodies directed against ß-amyloid such as Lecanemab or Aduhelm, which are not yet convincing and have important side effects. Thiamine prodrugs have no reported side effects. Anti-tau antibodies do not seem to be effective in patients with early Alzheimer’s disease [118,119].
Taken together, those data suggest that both BFT and DBT have a therapeutic potential in brain pathologies associated with oxidative stress and inflammation, not only in neurodegenerative diseases, but also in major depression and aggression linked to stressful events. These effects seem to be independent of the coenzyme role of ThDP.

Funding

L.B. is Research Director at the “Fonds de la Recherche Scientifique—FNRS” (Belgium).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The author wishes to thank Adeline Deward (Illumine, http://www.illuminesciences.be/, accessed on 4 July 2023) for the design of Figure 2, Figure 3, Figure 5, Figure 6, Figure 9, and Figure 10.

Conflicts of Interest

The author declares no conflict of interest.

Abbreviations

AGEAdvanced Glycation end products
ALSAmyotrophic lateral sclerosis
BBBBlood–brain barrier
BFTBenfotiamine
DBTO,S-Dibenzoylthiamine
GSHReduced glutathione
G6PDGlucose 6-phosphate dehydrogenase
O-BTO-Benzoylthiamine
OGDHC2-Oxoglutarate dehydrogenase complex
PDHCPyruvate dehydrogenase complex
PPPPentose phosphate pathway
S-BTS-Benzoylthiamine
SuBTSulbutiamine
ThDPThiamine diphosphate
ThMPThiamine monophosphate
TKTransketolase
TPKThiamine pyrophosphokinase

References

  1. Bettendorff, L. The Chemistry, Biochemistry and Metabolism of Thiamin (Vitamin B1). In B Vitamins and Folate; Preedy, V.R., Ed.; Food and Nutritional Components in Focus; RSC Publishing: Cambridge, UK, 2012; pp. 71–92. [Google Scholar]
  2. Lakaye, B.; Wirtzfeld, B.; Wins, P.; Grisar, T.; Bettendorff, L. Thiamine Triphosphate, a New Signal Required for Optimal Growth of Escherichia coli during Amino Acid Starvation. J. Biol. Chem. 2004, 279, 17142–17147. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Gigliobianco, T.; Gangolf, M.; Lakaye, B.; Pirson, B.; von Ballmoos, C.; Wins, P.; Bettendorff, L. An Alternative Role of FoF1-ATP Synthase in Escherichia coli: Synthesis of Thiamine Triphosphate. Sci. Rep. 2013, 3, 1071. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Bettendorff, L. Update on Thiamine Triphosphorylated Derivatives and Metabolizing Enzymatic Complexes. Biomolecules 2021, 11, 1645. [Google Scholar] [CrossRef]
  5. Bettendorff, L.; Lakaye, B.; Kohn, G.; Wins, P. Thiamine Triphosphate: A Ubiquitous Molecule in Search of a Physiological Role. Metab. Brain Dis. 2014, 29, 1069–1082. [Google Scholar] [CrossRef]
  6. Bettendorff, L.; Wins, P. Thiamin Diphosphate in Biological Chemistry: New Aspects of Thiamin Metabolism, Especially Triphosphate Derivatives Acting Other than as Cofactors. FEBS J. 2009, 276, 2917–2925. [Google Scholar] [CrossRef] [PubMed]
  7. Bettendorff, L.; Wirtzfeld, B.; Makarchikov, A.F.; Mazzucchelli, G.; Frédérich, M.; Gigliobianco, T.; Gangolf, M.; De Pauw, E.; Angenot, L.; Wins, P. Discovery of a Natural Thiamine Adenine Nucleotide. Nat. Chem. Biol. 2007, 3, 211–212. [Google Scholar] [CrossRef]
  8. Bettendorff, L.; Wins, P.; Lesourd, M. Subcellular Localization and Compartmentation of Thiamine Derivatives in Rat Brain. Biochim. Biophys. Acta 1994, 1222, 1–6. [Google Scholar] [CrossRef]
  9. Bettendorff, L. The Compartmentation of Phosphorylated Thiamine Derivatives in Cultured Neuroblastoma Cells. Biochim. Biophys. Acta 1994, 1222, 7–14. [Google Scholar] [CrossRef] [PubMed]
  10. Hurt, J.K.; Coleman, J.L.; Fitzpatrick, B.J.; Taylor-Blake, B.; Bridges, A.S.; Vihko, P.; Zylka, M.J. Prostatic Acid Phosphatase Is Required for the Antinociceptive Effects of Thiamine and Benfotiamine. PLoS ONE 2012, 7, e48562. [Google Scholar] [CrossRef] [Green Version]
  11. Wu, C.-C.; Ekanem, T.I.; Phan, N.N.; Loan, D.T.T.; Hou, S.-Y.; Lee, K.-H.; Wang, C.-Y. Gene Signatures and Prognostic Analyses of the Tob/BTG Pituitary Tumor-Transforming Gene (PTTG) Family in Clinical Breast Cancer Patients. Int. J. Med. Sci. 2020, 17, 3112–3124. [Google Scholar] [CrossRef]
  12. Sambon, M.; Wins, P.; Bettendorff, L. Neuroprotective Effects of Thiamine and Precursors with Higher Bioavailability: Focus on Benfotiamine and Dibenzoylthiamine. Int. J. Mol. Sci. 2021, 22, 5418. [Google Scholar] [CrossRef] [PubMed]
  13. Pan, X.; Chen, Z.; Fei, G.; Pan, S.; Bao, W.; Ren, S.; Guan, Y.; Zhong, C. Long-Term Cognitive Improvement after Benfotiamine Administration in Patients with Alzheimer’s Disease. Neurosci. Bull. 2016, 32, 591–596. [Google Scholar] [CrossRef] [PubMed]
  14. Gibson, G.E.; Luchsinger, J.A.; Cirio, R.; Chen, H.; Franchino-Elder, J.; Hirsch, J.A.; Bettendorff, L.; Chen, Z.; Flowers, S.; Gerber, L.; et al. Benfotiamine and Cognitive Decline in Alzheimer’s Disease: Results of a Randomized Placebo-Controlled Phase IIa Clinical Trial. J. Alzheimers Dis. 2020, 78, 989–1010. [Google Scholar] [CrossRef]
  15. Tapias, V.; Jainuddin, S.; Ahuja, M.; Stack, C.; Elipenahli, C.; Vignisse, J.; Gerges, M.; Starkova, N.; Xu, H.; Starkov, A.A.; et al. Benfotiamine Treatment Activates the Nrf2/ARE Pathway and Is Neuroprotective in a Transgenic Mouse Model of Tauopathy. Hum. Mol. Genet. 2018, 27, 2874–2892. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Pan, X.; Gong, N.; Zhao, J.; Yu, Z.; Gu, F.; Chen, J.; Sun, X.; Zhao, L.; Yu, M.; Xu, Z.; et al. Powerful Beneficial Effects of Benfotiamine on Cognitive Impairment and Beta-Amyloid Deposition in Amyloid Precursor Protein/Presenilin-1 Transgenic Mice. Brain 2010, 133, 1342–1351. [Google Scholar] [CrossRef] [Green Version]
  17. Probert, F.; Gorlova, A.; Deikin, A.; Bettendorff, L.; Veniaminova, E.; Nedorubov, A.; Chaprov, K.D.; Ivanova, T.A.; Anthony, D.C.; Strekalova, T. In FUS [1−359]-tg Mice O,S-Dibenzoyl Thiamine Reduces Muscle Atrophy, Decreases Glycogen Synthase Kinase 3 Beta, and Normalizes the Metabolome. Biomed. Pharmacother. 2022, 156, 113986. [Google Scholar] [CrossRef] [PubMed]
  18. Lakaye, B.; Makarchikov, A.F.; Antunes, A.F.; Zorzi, W.; Coumans, B.; De Pauw, E.; Wins, P.; Grisar, T.; Bettendorff, L. Molecular Characterization of a Specific Thiamine Triphosphatase Widely Expressed in Mammalian Tissues. J. Biol. Chem. 2002, 277, 13771–13777. [Google Scholar] [CrossRef] [Green Version]
  19. Delvaux, D.; Kerff, F.; Murty, M.R.; Lakaye, B.; Czerniecki, J.; Kohn, G.; Wins, P.; Herman, R.; Gabelica, V.; Heuze, F.; et al. Structural Determinants of Specificity and Catalytic Mechanism in Mammalian 25-kDa Thiamine Triphosphatase. Biochim. Biophys. Acta 2013, 1830, 4513–4523. [Google Scholar] [CrossRef] [PubMed]
  20. Makarchikov, A.F.; Brans, A.; Bettendorff, L. Thiamine Diphosphate Adenylyl Transferase from E. Coli: Functional Characterization of the Enzyme Synthesizing Adenosine Thiamine Triphosphate. BMC Biochem. 2007, 8, 17. [Google Scholar] [CrossRef] [Green Version]
  21. Jankowska-Kulawy, A.; Klimaszewska-Łata, J.; Gul-Hinc, S.; Ronowska, A.; Szutowicz, A. Metabolic and Cellular Compartments of Acetyl-CoA in the Healthy and Diseased Brain. Int. J. Mol. Sci. 2022, 23, 10073. [Google Scholar] [CrossRef]
  22. Begley, T.P.; Ealick, S.E.; McLafferty, F.W. Thiamin Biosynthesis: Still Yielding Fascinating Biological Chemistry. Biochem. Soc. Trans. 2012, 40, 555–560. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Goyer, A. Thiamine in Plants: Aspects of Its Metabolism and Functions. Phytochemistry 2010, 71, 1615–1624. [Google Scholar] [CrossRef] [PubMed]
  24. Liu, Z.; Farkas, P.; Wang, K.; Kohli, M.-O.; Fitzpatrick, T.B. B Vitamin Supply in Plants and Humans: The Importance of Vitamer Homeostasis. Plant J. Cell Mol. Biol. 2022, 111, 662–682. [Google Scholar] [CrossRef]
  25. Said, H.M. Intestinal Absorption of Water-Soluble Vitamins in Health and Disease. Biochem. J. 2011, 437, 357–372. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Bettendorff, L. 10. Thiamine. In Present Knowledge in Nutrition, 11th ed.; Marriott, B., Birt, D.F., Stalling, V., Yates, A., Eds.; Basic Nutrition and Metabolism; Academic Press: Cambridge, MA, USA, 2020; Volume 1, p. 676. ISBN 978-0-323-66162-1. [Google Scholar]
  27. Bettendorff, L.; Wins, P. Mechanism of Thiamine Transport in Neuroblastoma Cells. Inhibition of a High Affinity Carrier by Sodium Channel Activators and Dependence of Thiamine Uptake on Membrane Potential and Intracellular ATP. J. Biol. Chem. 1994, 269, 14379–14385. [Google Scholar] [CrossRef]
  28. Dutta, B.; Huang, W.; Molero, M.; Kekuda, R.; Leibach, F.H.; Devoe, L.D.; Ganapathy, V.; Prasad, P.D. Cloning of the Human Thiamine Transporter, a Member of the Folate Transporter Family. J. Biol. Chem. 1999, 274, 31925–31929. [Google Scholar] [CrossRef] [Green Version]
  29. Said, H.M.; Balamurugan, K.; Subramanian, V.S.; Marchant, J.S. Expression and Functional Contribution of hTHTR-2 in Thiamin Absorption in Human Intestine. Am. J. Physiol. Gastrointest. Liver Physiol. 2004, 286, G491–G498. [Google Scholar] [CrossRef] [Green Version]
  30. Kevelam, S.H.; Bugiani, M.; Salomons, G.S.; Feigenbaum, A.; Blaser, S.; Prasad, C.; Haberle, J.; Baric, I.; Bakker, I.M.; Postma, N.L.; et al. Exome Sequencing Reveals Mutated SLC19A3 in Patients with an Early-Infantile, Lethal Encephalopathy. Brain 2013, 136, 1534–1543. [Google Scholar] [CrossRef] [Green Version]
  31. Nosaka, K.; Onozuka, M.; Kakazu, N.; Hibi, S.; Nishimura, H.; Nishino, H.; Abe, T. Isolation and Characterization of a Human Thiamine Pyrophosphokinase cDNA. Biochim. Biophys. Acta 2001, 1517, 293–297. [Google Scholar] [CrossRef]
  32. Sambon, M.; Pavlova, O.; Alhama-Riba, J.; Wins, P.; Brans, A.; Bettendorff, L. Product Inhibition of Mammalian Thiamine Pyrophosphokinase Is an Important Mechanism for Maintaining Thiamine Diphosphate Homeostasis. Biochim. Biophys. Acta Gen. Subj. 2022, 1866, 130071. [Google Scholar] [CrossRef]
  33. Lindhurst, M.J.; Fiermonte, G.; Song, S.; Struys, E.; De Leonardis, F.; Schwartzberg, P.L.; Chen, A.; Castegna, A.; Verhoeven, N.; Mathews, C.K.; et al. Knockout of Slc25a19 Causes Mitochondrial Thiamine Pyrophosphate Depletion, Embryonic Lethality, CNS Malformations, and Anemia. Proc. Natl. Acad. Sci. USA 2006, 103, 15927–15932. [Google Scholar] [CrossRef] [PubMed]
  34. Zhao, R.; Gao, F.; Goldman, I.D. Reduced Folate Carrier Transports Thiamine Monophosphate: An Alternative Route for Thiamine Delivery into Mammalian Cells. Am. J. Physiol. Cell Physiol. 2002, 282, C1512–C1517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Gangolf, M.; Wins, P.; Thiry, M.; El Moualij, B.; Bettendorff, L. Thiamine Triphosphate Synthesis in Rat Brain Occurs in Mitochondria and Is Coupled to the Respiratory Chain. J. Biol. Chem. 2010, 285, 583–594. [Google Scholar] [CrossRef] [Green Version]
  36. Gangolf, M.; Czerniecki, J.; Radermecker, M.; Detry, O.; Nisolle, M.; Jouan, C.; Martin, D.; Chantraine, F.; Lakaye, B.; Wins, P.; et al. Thiamine Status in Humans and Content of Phosphorylated Thiamine Derivatives in Biopsies and Cultured Cells. PLoS ONE 2010, 5, e13616. [Google Scholar] [CrossRef] [Green Version]
  37. Ben-Yoseph, O.; Camp, D.M.; Robinson, T.E.; Ross, B.D. Dynamic Measurements of Cerebral Pentose Phosphate Pathway Activity in Vivo Using [1,6-13C2,6,6-2H2]Glucose and Microdialysis. J. Neurochem. 1995, 64, 1336–1342. [Google Scholar] [CrossRef] [PubMed]
  38. Foulon, V.; Antonenkov, V.D.; Croes, K.; Waelkens, E.; Mannaerts, G.P.; Van Veldhoven, P.P.; Casteels, M. Purification, Molecular Cloning, and Expression of 2-Hydroxyphytanoyl- CoA Lyase, a Peroxisomal Thiamine Pyrophosphate-Dependent Enzyme That Catalyzes the Carbon-Carbon Bond Cleavage during Alpha-Oxidation of 3- Methyl-Branched Fatty Acids. Proc. Natl. Acad. Sci. USA 1999, 96, 10039–10044. [Google Scholar] [CrossRef]
  39. Hamel, E.; Butterworth, R.F.; Barbeau, A. Effect of Thiamine Deficiency on Levels of Putative Amino Acid Transmitters in Affected Regions of the Rat Brain. J. Neurochem. 1979, 33, 575–577. [Google Scholar] [CrossRef] [PubMed]
  40. Kochetov, G.A.; Solovjeva, O.N. Structure and Functioning Mechanism of Transketolase. Biochim. Biophys. Acta 2014, 1844, 1608–1618. [Google Scholar] [CrossRef]
  41. Bettendorff, L.; Wins, P. Cofactors and Coenzymes|Biochemistry of Thiamine and Thiamine Phosphate Compounds. In Encyclopedia of Biological Chemistry; Elsevier: Oxford, UK, 2021; Volume 1, pp. 302–313. [Google Scholar]
  42. Jordan, F.; Nemeria, N.; Gerfen, G. Human 2-Oxoglutarate Dehydrogenase and 2-Oxoadipate Dehydrogenase Both Generate Superoxide/H2O2 in a Side Reaction and Each Could Contribute to Oxidative Stress in Mitochondria. Neurochem. Res. 2019, in press. [Google Scholar] [CrossRef]
  43. Bunik, V.I.; Degtyarev, D. Structure-Function Relationships in the 2-Oxo Acid Dehydrogenase Family: Substrate-Specific Signatures and Functional Predictions for the 2-Oxoglutarate Dehydrogenase-like Proteins. Proteins 2007, 71, 874–890. [Google Scholar] [CrossRef]
  44. Navarro, D.; Zwingmann, C.; Butterworth, R.F. Impaired Oxidation of Branched-Chain Amino Acids in the Medial Thalamus of Thiamine-Deficient Rats. Metab. Brain Dis. 2008, 23, 445–455. [Google Scholar] [CrossRef] [PubMed]
  45. Breslow, R. On the Mechanism of Thiamine Action. IV.1 Evidence from Studies on Model Systems. J. Am. Chem. Soc. 1958, 80, 3719–3726. [Google Scholar] [CrossRef]
  46. Prajapati, S.; Rabe von Pappenheim, F.; Tittmann, K. Frontiers in the Enzymology of Thiamin Diphosphate-Dependent Enzymes. Curr. Opin. Struct. Biol. 2022, 76, 102441. [Google Scholar] [CrossRef] [PubMed]
  47. Voet, D.; Voet, J.G.; Pratt, C.W. Principles of Biochemistry; John Wiley & Sons: Singapore, 2013. [Google Scholar]
  48. Whitfield, K.C.; Smith, G.; Chamnan, C.; Karakochuk, C.D.; Sophonneary, P.; Kuong, K.; Dijkhuizen, M.A.; Hong, R.; Berger, J.; Green, T.J.; et al. High Prevalence of Thiamine (Vitamin B1) Deficiency in Early Childhood among a Nationally Representative Sample of Cambodian Women of Childbearing Age and Their Children. PLoS Negl. Trop. Dis. 2017, 11, e0005814. [Google Scholar] [CrossRef] [Green Version]
  49. Whitfield, K.C.; Bourassa, M.W.; Adamolekun, B.; Bergeron, G.; Bettendorff, L.; Brown, K.H.; Cox, L.; Fattal-Valevski, A.; Fischer, P.R.; Frank, E.L.; et al. Thiamine Deficiency Disorders: Diagnosis, Prevalence, and a Roadmap for Global Control Programs. Ann. N. Y. Acad. Sci. 2018, 1430, 3–43. [Google Scholar] [CrossRef]
  50. Johnson, C.R.; Fischer, P.R.; Thacher, T.D.; Topazian, M.D.; Bourassa, M.W.; Combs, G.F. Thiamin Deficiency in Low- and Middle-Income Countries: Disorders, Prevalences, Previous Interventions and Current Recommendations. Nutr. Health 2019, 25, 127–151. [Google Scholar] [CrossRef]
  51. Balaghi, M.; Pearson, W.N. Tissue and Intracellular Distribution of Radioactive Thiamine in Normal and Thiamine-Deficient Rats. J. Nutr. 1966, 89, 127–132. [Google Scholar] [CrossRef]
  52. Harata, N.; Iwasaki, Y.; Ohara, Y. Reappraisal of Regional Thiamine Content in the Central Nervous System of the Normal and Thiamine-Deficient Mice. Metab. Brain Dis. 1993, 8, 45–59. [Google Scholar] [CrossRef]
  53. McCandless, D.W.; Schenker, S.; Cook, M. Encephalopathy of Thiamine Deficiency: Studies of Intracerebral Mechanisms. J. Clin. Investig. 1968, 47, 2268–2280. [Google Scholar] [CrossRef]
  54. Tallaksen, C.M.; Bohmer, T.; Karlsen, J.; Bell, H. Determination of Thiamin and Its Phosphate Esters in Human Blood, Plasma, and Urine. Methods Enzymol. 1997, 279, 67–74. [Google Scholar]
  55. Nunes, P.T.; da Silva Oliveira, P.; Ferraz, V.; Ribeiro, A.M. Validation of a HPLC Method for Quantification of Thiamine and Its Phosphate Esters in Rat Brain Tissue. J. Behav. Brain Sci. 2017, 7, 79–93. [Google Scholar] [CrossRef]
  56. Volvert, M.L.; Seyen, S.; Piette, M.; Evrard, B.; Gangolf, M.; Plumier, J.C.; Bettendorff, L. Benfotiamine, a Synthetic S-Acyl Thiamine Derivative, Has Different Mechanisms of Action and a Different Pharmacological Profile than Lipid-Soluble Thiamine Disulfide Derivatives. BMC Pharmacol. 2008, 8, 10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Bettendorff, L.; Schoffeniels, E.; Naquet, R.; Silva-Barrat, C.; Riche, D.; Menini, C. Phosphorylated Thiamine Derivatives and Cortical Activity in the Baboon Papio Papio: Effect of Intermittent Light Stimulation. J. Neurochem. 1989, 53, 80–87. [Google Scholar] [CrossRef] [PubMed]
  58. Bettendorff, L.; Peeters, M.; Jouan, C.; Wins, P.; Schoffeniels, E. Determination of Thiamin and Its Phosphate Esters in Cultured Neurons and Astrocytes Using an Ion-Pair Reversed-Phase High-Performance Liquid Chromatographic Method. Anal. Biochem. 1991, 198, 52–59. [Google Scholar] [CrossRef]
  59. Makarchikov, A.F.; Lakaye, B.; Gulyai, I.E.; Czerniecki, J.; Coumans, B.; Wins, P.; Grisar, T.; Bettendorff, L. Thiamine Triphosphate and Thiamine Triphosphatase Activities: From Bacteria to Mammals. Cell. Mol. Life Sci. 2003, 60, 1477–1488. [Google Scholar] [CrossRef]
  60. Bettendorff, L. Thiamine Homeostasis in Neuroblastoma Cells. Neurochem. Int. 1995, 26, 295–302. [Google Scholar] [CrossRef]
  61. Lockman, P.R.; Mumper, R.J.; Allen, D.D. Evaluation of Blood-Brain Barrier Thiamine Efflux Using the in Situ Rat Brain Perfusion Method. J. Neurochem. 2003, 86, 627–634. [Google Scholar] [CrossRef] [Green Version]
  62. Thornalley, P.J.; Babaei-Jadidi, R.; Al Ali, H.; Rabbani, N.; Antonysunil, A.; Larkin, J.; Ahmed, A.; Rayman, G.; Bodmer, C.W. High Prevalence of Low Plasma Thiamine Concentration in Diabetes Linked to a Marker of Vascular Disease. Diabetologia 2007, 50, 2164–2170. [Google Scholar] [CrossRef] [Green Version]
  63. Starling-Soares, B.; Carrera-Bastos, P.; Bettendorff, L. Role of the Synthetic B1 Vitamin Sulbutiamine on Health. J. Nutr. Metab. 2020, 2020, 9349063. [Google Scholar] [CrossRef]
  64. Raj, V.; Ojha, S.; Howarth, F.C.; Belur, P.D.; Subramanya, S.B. Therapeutic Potential of Benfotiamine and Its Molecular Targets. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 3261–3273. [Google Scholar]
  65. Sheng, L.; Cao, W.; Lin, P.; Chen, W.; Xu, H.; Zhong, C.; Yuan, F.; Chen, H.; Li, H.; Liu, C.; et al. Safety, Tolerability and Pharmacokinetics of Single and Multiple Ascending Doses of Benfotiamine in Healthy Subjects. Drug Des. Devel. Ther. 2021, 15, 1101–1110. [Google Scholar] [CrossRef] [PubMed]
  66. Loew, D. Pharmacokinetics of Thiamine Derivatives Especially of Benfotiamine. Int. J. Clin. Pharmacol. Ther. 1996, 34, 47–50. [Google Scholar] [PubMed]
  67. Hammes, H.P.; Du, X.; Edelstein, D.; Taguchi, T.; Matsumura, T.; Ju, Q.; Lin, J.; Bierhaus, A.; Nawroth, P.; Hannak, D.; et al. Benfotiamine Blocks Three Major Pathways of Hyperglycemic Damage and Prevents Experimental Diabetic Retinopathy. Nat. Med. 2003, 9, 294–299. [Google Scholar] [CrossRef] [PubMed]
  68. Rindi, G.; Patrini, C.; Comincioli, V.; Reggiani, C. Thiamine Content and Turnover Rates of Some Rat Nervous Regions, Using Labeled Thiamine as a Tracer. Brain Res. 1980, 181, 369–380. [Google Scholar] [CrossRef] [PubMed]
  69. Koepsell, H. Organic Cation Transporters in Health and Disease. Pharmacol. Rev. 2020, 72, 253–319. [Google Scholar] [CrossRef]
  70. Kato, K.; Moriyama, C.; Ito, N.; Zhang, X.; Hachiuma, K.; Hagima, N.; Iwata, K.; Yamaguchi, J.; Maeda, K.; Ito, K.; et al. Involvement of Organic Cation Transporters in the Clearance and Milk Secretion of Thiamine in Mice. Pharm. Res. 2015, 32, 2192–2204. [Google Scholar] [CrossRef]
  71. Meyer, M.J.; Tuerkova, A.; Römer, S.; Wenzel, C.; Seitz, T.; Gaedcke, J.; Oswald, S.; Brockmöller, J.; Zdrazil, B.; Tzvetkov, M.V. Differences in Metformin and Thiamine Uptake between Human and Mouse Organic Cation Transporter OCT1:Structural Determinants and Potential Consequences for Intrahepatic Concentrations. Drug Metab. Dispos. Biol. Fate Chem. 2020, 48, 1380–1392. [Google Scholar] [CrossRef]
  72. Yoshioka, K. Some Properties of the Thiamine Uptake System in Isolated Rat Hepatocytes. Biochim. Biophys. Acta 1984, 778, 201–209. [Google Scholar] [CrossRef]
  73. Lumeng, L.; Edmondson, J.W.; Schenker, S.; Li, T.K. Transport and Metabolism of Thiamin in Isolated Rat Hepatocytes. J. Biol. Chem. 1979, 254, 7265–7268. [Google Scholar] [CrossRef]
  74. Gorlova, A.; Pavlov, D.; Anthony, D.C.; Ponomarev, E.; Sambon, M.; Proshin, A.; Shafarevich, I.; Babaevskaya, D.; Lesch, K.-P.; Bettendorff, L.; et al. Thiamine and Benfotiamine Counteract Ultrasound-Induced Aggression, Normalize AMPA Receptor Expression and Plasticity Markers, and Reduce Oxidative Stress in Mice. Neuropharmacology 2019, 156, 107543. [Google Scholar] [CrossRef]
  75. Vignisse, J.; Sambon, M.; Gorlova, A.; Pavlov, D.; Caron, N.; Malgrange, B.; Shevtsova, E.; Svistunov, A.; Anthony, D.C.; Markova, N.; et al. Thiamine and Benfotiamine Prevent Stress-Induced Suppression of Hippocampal Neurogenesis in Mice Exposed to Predation without Affecting Brain Thiamine Diphosphate Levels. Mol. Cell. Neurosci. 2017, 82, 126–136. [Google Scholar] [CrossRef] [PubMed]
  76. Markova, N.; Bazhenova, N.; Anthony, D.C.; Vignisse, J.; Svistunov, A.; Lesch, K.-P.; Bettendorff, L.; Strekalova, T. Thiamine and Benfotiamine Improve Cognition and Ameliorate GSK-3β-Associated Stress-Induced Behaviours in Mice. Prog. Neuropsychopharmacol. Biol. Psychiatry 2017, 75, 148–156. [Google Scholar] [CrossRef] [PubMed]
  77. Sambon, M.; Napp, A.; Demelenne, A.; Vignisse, J.; Wins, P.; Fillet, M.; Bettendorff, L. Thiamine and Benfotiamine Protect Neuroblastoma Cells against Paraquat and β-Amyloid Toxicity by a Coenzyme-Independent Mechanism. Heliyon 2019, 5, e01710. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Culbreth, M.; Aschner, M. GSK-3β, a Double-Edged Sword in Nrf2 Regulation: Implications for Neurological Dysfunction and Disease. F1000Research 2018, 7, 1043. [Google Scholar] [CrossRef] [PubMed]
  79. Bozic, I.; Savic, D.; Stevanovic, I.; Pekovic, S.; Nedeljkovic, N.; Lavrnja, I. Benfotiamine Upregulates Antioxidative System in Activated BV-2 Microglia Cells. Front. Cell. Neurosci. 2015, 9, 351. [Google Scholar] [CrossRef] [Green Version]
  80. Hansen, D.V.; Hanson, J.E.; Sheng, M. Microglia in Alzheimer’s Disease. J. Cell Biol. 2018, 217, 459–472. [Google Scholar] [CrossRef]
  81. Allowitz, K.V.; Yoo, J.J.; Taylor, J.R.; Baloch, O.A.; Harames, K.; Ramana, K.V. Therapeutic Potential of Vitamin B1 Derivative Benfotiamine from Diabetes to COVID-19. Future Med. Chem. 2022, 14, 809–826. [Google Scholar] [CrossRef]
  82. Bozic, I.; Savic, D.; Laketa, D.; Bjelobaba, I.; Milenkovic, I.; Pekovic, S.; Nedeljkovic, N.; Lavrnja, I. Benfotiamine Attenuates Inflammatory Response in LPS Stimulated BV-2 Microglia. PLoS ONE 2015, 10, e0118372. [Google Scholar] [CrossRef] [Green Version]
  83. Mastrogiacomo, F.; Bettendorff, L.; Grisar, T.; Kish, S.J. Brain Thiamine, Its Phosphate Esters, and Its Metabolizing Enzymes in Alzheimer’s Disease. Ann. Neurol. 1996, 39, 585–591. [Google Scholar] [CrossRef]
  84. Bettendorff, L.; Mastrogiacomo, F.; Wins, P.; Kish, S.J.; Grisar, T.; Ball, M.J. Low Thiamine Diphosphate Levels in Brains of Patients with Frontal Lobe Degeneration of the Non-Alzheimer’s Type. J. Neurochem. 1997, 69, 2005–2010. [Google Scholar] [CrossRef]
  85. Pan, X.; Fei, G.; Lu, J.; Jin, L.; Pan, S.; Chen, Z.; Wang, C.; Sang, S.; Liu, H.; Hu, W.; et al. Measurement of Blood Thiamine Metabolites for Alzheimer’s Disease Diagnosis. EBioMedicine 2016, 3, 155–162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Chen, Z.; Pan, X.; Fei, G.; Pan, S.; Bao, W.; Ren, S.; Guan, Y.; Zhong, C. Diagnostic Value of Blood Thiamine Metabolites in Alzheimer’s Disease Examined by 11C-PiB PET Scanning. Future Sci. OA 2017, 3, FSO172. [Google Scholar] [CrossRef] [PubMed]
  87. Pan, X.; Sang, S.; Fei, G.; Jin, L.; Liu, H.; Wang, Z.; Wang, H.; Zhong, C. Enhanced Activities of Blood Thiamine Diphosphatase and Monophosphatase in Alzheimer’s Disease. PLoS ONE 2017, 12, e0167273. [Google Scholar] [CrossRef] [PubMed]
  88. Sang, S.; Pan, X.; Chen, Z.; Zeng, F.; Pan, S.; Liu, H.; Jin, L.; Fei, G.; Wang, C.; Ren, S.; et al. Thiamine Diphosphate Reduction Strongly Correlates with Brain Glucose Hypometabolism in Alzheimer’s Disease, Whereas Amyloid Deposition Does Not. Alzheimer’s Res. Ther. 2018, 10, 26. [Google Scholar] [CrossRef] [Green Version]
  89. Ramamoorthy, K.; Yoshimura, R.; Al-Juburi, S.; Anandam, K.Y.; Kapadia, R.; Alachkar, A.; Abbott, G.W.; Said, H.M. Alzheimer’s Disease Is Associated with Disruption in Thiamin Transport Physiology: A Potential Role for Neuroinflammation. Neurobiol. Dis. 2022, 171, 105799. [Google Scholar] [CrossRef]
  90. Bhawal, R.; Fu, Q.; Anderson, E.T.; Gibson, G.E.; Zhang, S. Serum Metabolomic and Lipidomic Profiling Reveals Novel Biomarkers of Efficacy for Benfotiamine in Alzheimer’s Disease. Int. J. Mol. Sci. 2021, 22, 13188. [Google Scholar] [CrossRef]
  91. Wada, T.; Takagi, H.; Minakami, H.; Hamanaka, W.; Okamoto, K.; Ito, A.; Sahashi, Y. A New Thiamine Derivative, S-Benzoylthiamine O-Monophosphate. Science 1961, 134, 195–196. [Google Scholar] [CrossRef]
  92. Heywood, R.; Wood, J.D.; Majeed, S.K. Tumorigenic and Toxic Effect of O,S-Dibenzoyl Thiamine Hydrochloride in Prolonged Dietary Administration to Rats. Toxicol. Lett. 1985, 26, 53–58. [Google Scholar] [CrossRef]
  93. Sambon, M.; Gorlova, A.; Demelenne, A.; Alhama-Riba, J.; Coumans, B.; Lakaye, B.; Wins, P.; Fillet, M.; Anthony, D.C.; Strekalova, T.; et al. Dibenzoylthiamine Has Powerful Antioxidant and Anti-Inflammatory Properties in Cultured Cells and in Mouse Models of Stress and Neurodegeneration. Biomedicines 2020, 8, 361. [Google Scholar] [CrossRef]
  94. Ketola, H.G.; Isaacs, G.R.; Robins, J.S.; Lloyd, R.C. Effectiveness and Retention of Thiamine and Its Analogs Administered to Steelhead and Landlocked Atlantic Salmon. J. Aquat. Anim. Health 2008, 20, 29–38. [Google Scholar] [CrossRef]
  95. Mann, R.H. Impaired Thiamine Metabolism in Amyotrophic Lateral Sclerosis and Its Potential Treatment With Benfotiamine: A Case Report and a Review of the Literature. Cureus 2023, 15, e40511. [Google Scholar] [CrossRef] [PubMed]
  96. Bettendorff, L. Thiamine in Excitable Tissues: Reflections on a Non-Cofactor Role. Metab. Brain Dis. 1994, 9, 183–209. [Google Scholar] [CrossRef]
  97. Mkrtchyan, G.; Aleshin, V.; Parkhomenko, Y.; Kaehne, T.; Luigi Di Salvo, M.; Parroni, A.; Contestabile, R.; Vovk, A.; Bettendorff, L.; Bunik, V. Molecular Mechanisms of the Non-Coenzyme Action of Thiamin in Brain: Biochemical, Structural and Pathway Analysis. Sci. Rep. 2015, 5, 12583. [Google Scholar] [CrossRef] [Green Version]
  98. Aleshin, V.A.; Mkrtchyan, G.V.; Bunik, V.I. Mechanisms of Non-Coenzyme Action of Thiamine: Protein Targets and Medical Significance. Biochemistry 2019, 84, 829–850. [Google Scholar] [CrossRef] [PubMed]
  99. Katare, R.; Caporali, A.; Emanueli, C.; Madeddu, P. Benfotiamine Improves Functional Recovery of the Infarcted Heart via Activation of Pro-Survival G6PD/Akt Signaling Pathway and Modulation of Neurohormonal Response. J. Mol. Cell. Cardiol. 2010, 49, 625–638. [Google Scholar] [CrossRef] [Green Version]
  100. Balakumar, P.; Rohilla, A.; Krishan, P.; Solairaj, P.; Thangathirupathi, A. The Multifaceted Therapeutic Potential of Benfotiamine. Pharmacol. Res. 2010, 61, 482–488. [Google Scholar] [CrossRef] [PubMed]
  101. Xu, I.M.-J.; Lai, R.K.-H.; Lin, S.-H.; Tse, A.P.-W.; Chiu, D.K.-C.; Koh, H.-Y.; Law, C.-T.; Wong, C.-M.; Cai, Z.; Wong, C.C.-L.; et al. Transketolase Counteracts Oxidative Stress to Drive Cancer Development. Proc. Natl. Acad. Sci. USA 2016, 113, E725–E734. [Google Scholar] [CrossRef]
  102. Inoue, K.; Itokawa, Y. Metabolism of O-Benzoylthiamine in Animal Body. Biochem. Med. 1973, 8, 450–463. [Google Scholar] [CrossRef]
  103. Bettendorff, L. Reduced Nucleotides, Thiols and O2 in Cellular Redox Balance: A Biochemist’s View. Antioxidants 2022, 11, 1877. [Google Scholar] [CrossRef]
  104. Couto, N.; Wood, J.; Barber, J. The Role of Glutathione Reductase and Related Enzymes on Cellular Redox Homoeostasis Network. Free. Radic. Biol. Med. 2016, 95, 27–42. [Google Scholar] [CrossRef]
  105. Hopmann, R.F.W. The Alkali-Induced Transformations of Thiamin. Ann. N. Y. Acad. Sci. 1982, 378, 32–50. [Google Scholar] [CrossRef]
  106. Brown, R.D. Thiamin as a Catalyst of Chemiosmotoc Energy Transductions. Ann. N. Y. Acad. Sci. 1982, 378, 442–448. [Google Scholar] [CrossRef]
  107. Pavlova, O.; Stepanenko, S.; Chehivska, L.; Sambon, M.; Bettendorff, L.; Yu, P. Thiamine Deficiency in Rats Affects Thiamine Metabolism Possibly through the Formation of Oxidized Thiamine Pyrophosphate. Biochim. Biophys. Acta Gen. Subj. 2021, 1865, 129980. [Google Scholar] [CrossRef] [PubMed]
  108. Meng, Q.; Zhang, Y.; Hao, S.; Sun, H.; Liu, B.; Zhou, H.; Wang, Y.; Xu, Z.-X. Recent Findings in the Regulation of G6PD and Its Role in Diseases. Front. Pharmacol. 2022, 13, 932154. [Google Scholar] [CrossRef] [PubMed]
  109. Stincone, A.; Prigione, A.; Cramer, T.; Wamelink, M.M.C.; Campbell, K.; Cheung, E.; Olin-Sandoval, V.; Grüning, N.-M.; Krüger, A.; Tauqeer Alam, M.; et al. The Return of Metabolism: Biochemistry and Physiology of the Pentose Phosphate Pathway. Biol. Rev. Camb. Philos. Soc. 2015, 90, 927–963. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Pavlov, D.; Markova, N.; Bettendorff, L.; Chekhonin, V.; Pomytkin, I.; Lioudyno, V.; Svistunov, A.; Ponomarev, E.; Lesch, K.-P.; Strekalova, T. Elucidating the Functions of Brain GSK3α: Possible Synergy with GSK3β Upregulation and Reversal by Antidepressant Treatment in a Mouse Model of Depressive-like Behaviour. Behav. Brain Res. 2017, 335, 122–127. [Google Scholar] [CrossRef]
  111. Beurel, E.; Grieco, S.F.; Jope, R.S. Glycogen Synthase Kinase-3 (GSK3): Regulation, Actions, and Diseases. Pharmacol. Ther. 2015, 148, 114–131. [Google Scholar] [CrossRef] [Green Version]
  112. Soni, D.; Kumar, P. GSK-3β-Mediated Regulation of Nrf2/HO-1 Signaling as a New Therapeutic Approach in the Treatment of Movement Disorders. Pharmacol. Rep. 2022, 74, 557–569. [Google Scholar] [CrossRef] [PubMed]
  113. Tang, Y.-C.; Hsiao, J.-R.; Jiang, S.-S.; Chang, J.-Y.; Chu, P.-Y.; Liu, K.-J.; Fang, H.-L.; Lin, L.-M.; Chen, H.-H.; Huang, Y.-W.; et al. C-MYC-Directed NRF2 Drives Malignant Progression of Head and Neck Cancer via Glucose-6-Phosphate Dehydrogenase and Transketolase Activation. Theranostics 2021, 11, 5232–5247. [Google Scholar] [CrossRef]
  114. Park, J.; Hosomi, K.; Kawashima, H.; Chen, Y.-A.; Mohsen, A.; Ohno, H.; Konishi, K.; Tanisawa, K.; Kifushi, M.; Kogawa, M.; et al. Dietary Vitamin B1 Intake Influences Gut Microbial Community and the Consequent Production of Short-Chain Fatty Acids. Nutrients 2022, 14, 2078. [Google Scholar] [CrossRef]
  115. Radford-Smith, D.E.; Anthony, D.C. Prebiotic and Probiotic Modulation of the Microbiota-Gut-Brain Axis in Depression. Nutrients 2023, 15, 1880. [Google Scholar] [CrossRef]
  116. Das, T.K.; Ganesh, B.P. Interlink between the Gut Microbiota and Inflammation in the Context of Oxidative Stress in Alzheimer’s Disease Progression. Gut Microbes 2023, 15, 2206504. [Google Scholar] [CrossRef] [PubMed]
  117. Gibson, G.E.; Feldman, H.H.; Zhang, S.; Flowers, S.A.; Luchsinger, J.A. Pharmacological Thiamine Levels as a Therapeutic Approach in Alzheimer’s Disease. Front. Med. 2022, 9, 1033272. [Google Scholar] [CrossRef] [PubMed]
  118. Zheng, X.; Tang, Y.; Yang, Q.; Wang, S.; Chen, R.; Tao, C.; Zhang, P.; Fan, B.; Zhan, J.; Tang, C.; et al. Effectiveness and Safety of Anti-Tau Drugs for Alzheimer’s Disease: Systematic Review and Meta-Analysis. J. Am. Geriatr. Soc. 2022, 70, 3281–3292. [Google Scholar] [CrossRef] [PubMed]
  119. Florian, H.; Wang, D.; Arnold, S.E.; Boada, M.; Guo, Q.; Jin, Z.; Zheng, H.; Fisseha, N.; Kalluri, H.V.; Rendenbach-Mueller, B.; et al. Tilavonemab in Early Alzheimer’s Disease: Results from a Phase 2, Randomized, Double-Blind Study. Brain 2023, 145, 2275–2284. [Google Scholar] [CrossRef] [PubMed]
Figure 4. Mechanism of thiamine diphosphate catalysis. The apoenzymes force the thiamine molecule in the so-called V-conformation bringing the 4′-amino group close to the thiazolium C-2 proton, favoring deprotonation. The catalytic action of ThDP is linked to the ability of its thiazolium ring to form an ylide (carbanion) intermediate by deprotonation of the thiazolium C-2 and attack of the resulting carbanion on a substrate (here shown for pyruvate) carbonyl.
Figure 4. Mechanism of thiamine diphosphate catalysis. The apoenzymes force the thiamine molecule in the so-called V-conformation bringing the 4′-amino group close to the thiazolium C-2 proton, favoring deprotonation. The catalytic action of ThDP is linked to the ability of its thiazolium ring to form an ylide (carbanion) intermediate by deprotonation of the thiazolium C-2 and attack of the resulting carbanion on a substrate (here shown for pyruvate) carbonyl.
Ijms 24 11296 g004
Figure 5. High extracellular thiamine concentrations favor a thiamine+/thiamine+ over a thiamine+/H+ exchange mode across the BBB [61]. Such a trans-stimulation was also observed in cultured neuroblastoma cells [60]. It is not clear whether the same transport protein is involved in the two modes and which compartment of the BBB would be involved in such behavior.
Figure 5. High extracellular thiamine concentrations favor a thiamine+/thiamine+ over a thiamine+/H+ exchange mode across the BBB [61]. Such a trans-stimulation was also observed in cultured neuroblastoma cells [60]. It is not clear whether the same transport protein is involved in the two modes and which compartment of the BBB would be involved in such behavior.
Ijms 24 11296 g005
Figure 6. Regulation of ThDP synthesis by TPK in a red blood cell. Thiamine is transported into the cell by a thiamine+/H+ exchange, probably involving ThTR1 [62]. Inside the cells, thiamine is pyrophosphorylated by TPK to ThDP, which acts as a feedback inhibitor of TPK when intracellular thiamine concentration exceeds 0.5 µM. ThDP cannot exist across the plasma membrane by simple diffusion. ThDP binds to apotransketolase, and excess coenzyme is hydrolyzed to ThMP and thiamine. Note that in physiological solutions, the positive charge of thiamine is neutralized by a chloride anion.
Figure 6. Regulation of ThDP synthesis by TPK in a red blood cell. Thiamine is transported into the cell by a thiamine+/H+ exchange, probably involving ThTR1 [62]. Inside the cells, thiamine is pyrophosphorylated by TPK to ThDP, which acts as a feedback inhibitor of TPK when intracellular thiamine concentration exceeds 0.5 µM. ThDP cannot exist across the plasma membrane by simple diffusion. ThDP binds to apotransketolase, and excess coenzyme is hydrolyzed to ThMP and thiamine. Note that in physiological solutions, the positive charge of thiamine is neutralized by a chloride anion.
Ijms 24 11296 g006
Figure 7. Sequences of reactions leading from SuBT (hydrolysis + reduction) or BFT (2 × hydrolysis) to thiamine. Please note that with the symmetric dimer SuBT, two molecules of thiamine are formed. Here, we suppose that the cleavage of the disulfide link in SuBT occurs via reduction by another thiol, the most obvious being reduced glutathione.
Figure 7. Sequences of reactions leading from SuBT (hydrolysis + reduction) or BFT (2 × hydrolysis) to thiamine. Please note that with the symmetric dimer SuBT, two molecules of thiamine are formed. Here, we suppose that the cleavage of the disulfide link in SuBT occurs via reduction by another thiol, the most obvious being reduced glutathione.
Ijms 24 11296 g007
Figure 8. Thiamine, ThMP, and ThDP levels one hour after a single oral administration of benfotiamine (100 mg/kg) or control solution. The data were analyzed by MANOVA (Wilk’s Lambda, p = 0.0002, 0.0145, and 0.27 for blood, liver, and brain, respectively) followed by ANOVA for each thiamine compound (*, p < 0.05; **, p < 0.01). The results are expressed as mean ± SEM for 4 animals in each group. (Modified according to [56]).
Figure 8. Thiamine, ThMP, and ThDP levels one hour after a single oral administration of benfotiamine (100 mg/kg) or control solution. The data were analyzed by MANOVA (Wilk’s Lambda, p = 0.0002, 0.0145, and 0.27 for blood, liver, and brain, respectively) followed by ANOVA for each thiamine compound (*, p < 0.05; **, p < 0.01). The results are expressed as mean ± SEM for 4 animals in each group. (Modified according to [56]).
Ijms 24 11296 g008
Figure 9. Hypothesis concerning the metabolization of DBT, BFT, and sulbutiamine, based on the data published in the following papers [10,15,32,77,93]. (Figure modified from [12]).
Figure 9. Hypothesis concerning the metabolization of DBT, BFT, and sulbutiamine, based on the data published in the following papers [10,15,32,77,93]. (Figure modified from [12]).
Ijms 24 11296 g009
Figure 10. Summary diagram of the pharmacological effects of BFT and DBT. X is a hypothetical intermediate, possibly an open thiol form (Figure 9). (For explanations, see text).
Figure 10. Summary diagram of the pharmacological effects of BFT and DBT. X is a hypothetical intermediate, possibly an open thiol form (Figure 9). (For explanations, see text).
Ijms 24 11296 g010
Table 1. Levels of thiamine and ThDP levels in the blood and cerebral cortex of several species.
Table 1. Levels of thiamine and ThDP levels in the blood and cerebral cortex of several species.
Species (Tissue)ThiamineThMPThDP
Plasma (nmol/L)
Human 17.1 ± 1.65.8 ± 0.6-
Blood (nmol/L)
Human 24 ± 310 ± 4138 ± 33
Rat 3188 ± 72718 ± 901127 ± 55
Mouse 473 ± 10103 ± 37737 ± 97
Cerebral cortex (pmol/mg protein)
Human 20.2 ± 0.33.5 ± 2.621 ± 5
Baboon (P. papio) 58 ± 22.6 ± 0.443 ± 7
Rat 63.1 ± 1.04.0 ± 0.479 ± 5
Mouse 46 ± 0.88.1 ± 4.690 ± 8
1 [54], 2 [36], 3 [55], 4 [56], 5 modified according to [57], 6 [58]. For a more detailed list, see [36,59].
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Bettendorff, L. Synthetic Thioesters of Thiamine: Promising Tools for Slowing Progression of Neurodegenerative Diseases. Int. J. Mol. Sci. 2023, 24, 11296. https://doi.org/10.3390/ijms241411296

AMA Style

Bettendorff L. Synthetic Thioesters of Thiamine: Promising Tools for Slowing Progression of Neurodegenerative Diseases. International Journal of Molecular Sciences. 2023; 24(14):11296. https://doi.org/10.3390/ijms241411296

Chicago/Turabian Style

Bettendorff, Lucien. 2023. "Synthetic Thioesters of Thiamine: Promising Tools for Slowing Progression of Neurodegenerative Diseases" International Journal of Molecular Sciences 24, no. 14: 11296. https://doi.org/10.3390/ijms241411296

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

Article Metrics

Back to TopTop