Next Article in Journal
Alginate Oligosaccharides Enhance Antioxidant Status and Intestinal Health by Modulating the Gut Microbiota in Weaned Piglets
Previous Article in Journal
Analysis of the Aging-Related AP2/ERF Transcription Factor Gene Family in Osmanthus fragrans
Previous Article in Special Issue
The Significance of Genetically Determined Methylation and Folate Metabolism Disorders in the Pathogenesis of Coronary Artery Disease: A Target for New Therapies?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Vitamin B12 Metabolism: A Network of Multi-Protein Mediated Processes

by
Patryk Mucha
1,
Filip Kus
1,2,
Dominik Cysewski
3,
Ryszard T. Smolenski
1 and
Marta Tomczyk
1,*
1
Department of Biochemistry, Medical University of Gdansk, 80-210 Gdansk, Poland
2
Laboratory of Protein Biochemistry, Intercollegiate Faculty of Biotechnology of University of Gdansk and Medical University of Gdansk, 80-307 Gdansk, Poland
3
Clinical Research Centre, Medical University of Bialystok, 15-276 Bialystok, Poland
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(15), 8021; https://doi.org/10.3390/ijms25158021
Submission received: 3 June 2024 / Revised: 12 July 2024 / Accepted: 19 July 2024 / Published: 23 July 2024

Abstract

:
The water-soluble vitamin, vitamin B12, also known as cobalamin, plays a crucial role in cellular metabolism, particularly in DNA synthesis, methylation, and mitochondrial functionality. Its deficiency can lead to hematological and neurological disorders; however, the manifestation of these clinical outcomes is relatively late. It leads to difficulties in the early diagnosis of vitamin B12 deficiency. A prolonged lack of vitamin B12 may have severe consequences including increased morbidity to neurological and cardiovascular diseases. Beyond inadequate dietary intake, vitamin B12 deficiency might be caused by insufficient bioavailability, blood transport disruptions, or impaired cellular uptake and metabolism. Despite nearly 70 years of knowledge since the isolation and characterization of this vitamin, there are still gaps in understanding its metabolic pathways. Thus, this review aims to compile current knowledge about the crucial proteins necessary to efficiently accumulate and process vitamin B12 in humans, presenting these systems as a multi-protein network. The epidemiological consequences, diagnosis, and treatment of vitamin B12 deficiency are also highlighted. We also discuss clinical warnings of vitamin B12 deficiency based on the ongoing test of specific moonlighting proteins engaged in vitamin B12 metabolic pathways.

1. Introduction

Vitamin B12 (cobalamin) was initially isolated in 1948 by two independent scientific teams—Folker’s in the USA and Smith’s in Great Britain—and identified: it was previously described as the antipernicious anemia factor [1]. Structurally, it possesses an organometallic nature, with the cyanide ion coordinated to the central cobalt atom. The primary Co ligand is surrounded by a corrin macrocycle containing four reduced pyrrole rings and seven amide side chains [2]. The precise structure varies among different corrinoids, where vitamin B12 is the only form active in animals and humans. The pyrrols constitute the equatorial plane of the ring, which is supplemented by two axial Co-ligands: a nucleotide base firmly attached to the ring and an exchangeable external ligand, e.g., water, cyanide, and methyl- or 5′-deoxyadenosyl group (as well as a number of other nucleophiles), see Figure 1. Hydroxycobalamin and cyanocobalamin are metabolically inactive; however, upon conversion to methylcobalamin and 5-deoxyadenosylcobalamin, the vitamins attain full activity [3]. Among these exchangeable groups, water represents the weakest coordination ligand, while CN-group exhibits the strongest coordination [4].
Humans cannot endogenously synthesize cobalamin and, consequently, must obtain it through dietary intake, predominantly sourced from meat, fish, dairy products, eggs, or some edible algae symbiotic with vitamin B12-producing microorganisms. Diminished intake or impaired absorption of vitamin B12 is usually caused by a decreased synthesis of a specific intestinal vitamin B12-carrier Intrinsic Factor (IF), see Figure 2, and the disease manifests itself in elderly individuals or in vegans adhering to dairy and egg-free diets [5]. In herbivores, the intestinal uptake of vitamin B12 is supported by the gut microbiota; whereas, in all carnivores and omnivores (including humans) the internal production of vitamin B12 is of very limited importance compared to dietary sources. Not all corrinoids produced in nature find their biological function in human metabolism and some of them might even disrupt the normal metabolism of this vitamin [6]. One of the examples is Coα-[α-(7-adenyl)]-Coβ-cyanocobamide, a corrinoid isolated from Lactobacillus reuteri and commonly named pseudovitamin B12. This analog shares a structure resembling cobalamin, except for the lower cobalt ligand. In this modification, nucleotide base adenine substitutes the 5,6-dimethylbenzimidazole [7]. Even though pseudovitamin B12 does not inhibit the uptake of vitamin B12 in rats [8], some studies suggest that it could potentially disrupt the absorption of a physiological trace concentration of vitamin B12 in the medium through the occupation of the vitamin B12-transporting blood protein transcobalamin (TC, the product of gene TCN2) [9].
The overall transportation chain of vitamin B12 is roughly outlined in Figure 2. The absorption of dietary vitamin B12 is preceded by the release of cobalamin bound to food proteins. First, gastric acid and proteases take action in the stomach. Afterward, a salivary protein called haptocorrin (HC, the product of gene TCN1) binds to liberated vitamin B12. The protein moiety of the HC-vitamin B12 complex is subsequently degraded by the proteases of the upper small intestine. Next, in the proximal ileum, cobalamin creates a complex with Cobalamin Binding Intrinsic Factor (IF), which is further recognized by certain receptors (first of all, the receptor cubam). Finally, the absorbed vitamin B12 is released from IF and excreted into the blood, where it is captured by TC [10]. The “active” transporter TC has a fast turnover rate and, under a steady state, caries about 20% of the vitamin B12 in the circulation [5]. The subsequent cellular uptake of vitamin B12 involves the CD320 receptor, present in all cell types [11]. Next, the vitamin B12-TC complex is transferred into lysosomes where the protein carrier breaks down. ATP Binding Cassette Subfamily D Member 4 protein (ABCD4, encoded by the gene of the same name) and lysosomal cobalamin transport-escort protein (LMBD1, LMBRD1 gene) are responsible for its further transport from the lysosomal lumen to the cytosol. When cobalamin is released into the cytosol, the cytoplasmic chaperons methylmalonic aciduria and homocystinuria type C protein (CblC, gene MMACHC) and methylmalonic aciduria and homocystinuria type D protein (CblD, gene MMADHC) initiate processing of the incoming vitamin, which starts with its reduction to Co2+ accompanied by removal of the original exchangeable group [12]. In the cytoplasm, vitamin B12 is converted to methylcobalamin (CH3Cbl, MeCbl), while adenosylcobalamin (AdoCbl) is synthesized in the mitochondria. MeCbl acts as a carrier of methyl groups. It is a co-factor of the enzyme methionine synthase (MS, gene MTR), which is involved in the transformation of homocysteine to methionine in the cytosol. These processes and subsequent metabolic reactions are involved in the synthesis of neurotransmitters, phospholipids, DNA, and RNA. On the other hand, AdoCbl acts as a cofactor of a mitochondrial enzyme, methylmalonyl-coenzyme A mutase (MCM, gene MMUT), which converts methylmalonyl-CoA to succinyl-CoA. This reaction is involved in the catabolism of cholesterol, fatty acids, and several amino acids [11]. The impairment of those enzymatic reactions leads to the elevation of homocysteine and methylmalonic acid, non-metabolizable end products, which is important in vitamin B12 deficiency diagnosis.
Due to its involvement in various cellular pathways (including isolation of nerve axons and DNA synthesis), alterations in vitamin B12 metabolism strongly correlate with cognitive impairment, hematological disorders, and cardiovascular diseases [13,14]. Thus, more in-depth research is necessary to fully elucidate the steps and compounds involved in vitamin B12 metabolic pathways.
This review aims to comprehensively summarize all of the identified proteins essential for the proper metabolism of vitamin B12 in humans and diseases related to its dysfunction. Furthermore, based on the collected data, we propose new diagnostic options as well as new warnings within the field of vitamin B12 deficiency research.

2. Proteins Involved in Vitamin B12 Metabolism

2.1. Gastrointestinal Vitamin B12 Transport

2.1.1. Haptocorrin (HC) and Intrinsic Factor (IF)

As previously elucidated, the absorption and utilization of vitamin B12 within the body require its initial binding to carrier proteins in the gastrointestinal tract. In humans, the two proteins facilitating this binding process are haptocorrin (HC) and Intrinsic Factor (IF). Despite their immunological distinctiveness, they exhibit structural similarities due to their primary and secondary protein structures [15,16]. In humans, parietal cells within the gastric mucosa of the ileum produce Intrinsic Factor (IF). The gene responsible for encoding IF is known as Gastric Intrinsic Factor (GIF) or Cobalamin Binding Intrinsic Factor (CBLIF), which is situated on chromosome 11 (11q12.1) in humans and on chromosome 19 in mice [17,18]. IF is made of 399 amino acid residues [16] and has a two-domain organization, with each domain having a mass of approx. 30 kDa and 20 kDa (including the conjugated sugar residues). Vitamin B12 is bound to the two domains via multiple bonds [16].
Initially named the R-binder or transcobalamin I (TCN1), haptocorrin (HC) is the first carrier protein for ingested vitamin B12. The investigation of the TCN1 gene sequence and its localization suggests that HC is an evolutionary product of CBLIF duplication [19]. This glycoprotein of 60 to 70 kDa [16,19] is present in various body fluids including saliva, breast milk, bile, tears, and blood plasma [17,20]. Upon ingestion, vitamin B12 is bound by HC secreted in saliva, which is responsible for transferring the free vitamin to the stomach. HC also binds dietary vitamin B12, released from food proteins in the stomach [17]. In the acidic pH of the stomach, HC is thought to protect the vitamin from partial hydrolysis [21]. Vitamin B12 remains bound to HC at the entrance to the duodenum, where pancreatic proteases, namely trypsin, chymotrypsin, and elastase, degrade HC, allowing for the release of vitamin B12 and its subsequent binding by IF, which was produced in the stomach but remained incapable of binding vitamin B12 until the acidic medium became neutralized [11]. Interestingly, deficiencies of pancreatic proteases can lead to defects in vitamin B12 absorption due to the inability to degrade HC and release the vitamin from the vitamin B12–HC complex. Patients with exocrine pancreatic insufficiency or chronic pancreatitis, conditions marked by inadequate delivery of pancreatic digestive enzymes to the intestine, often develop vitamin B12 malabsorption. However, these conditions rarely lead to severe vitamin B12 deficiency [22,23,24].
Studies on the exact role of HC in humans are limited due to the lack of homologous genes encoding HC in both mice and rats [25]. In these rodents, salivary glands produce another vitamin B12-binding protein transcobalamin (TC), which in humans is responsible for binding the vitamin in blood plasma [11]. This points out that details of the binding of ingested vitamin B12 vary between different species, where only some of them express HC [25].
After the release of vitamin B12 from HC in the duodenum, vitamin B12 binds to Intrinsic Factor (IF). IF is mostly found in gastric juice and ileal fluid in humans, being produced and secreted by parietal cells in the stomach. It is responsible for transferring vitamin B12 along the small intestine to the intestinal epithelial cells, which are mainly composed of enterocytes [16,21]. When vitamin B12 is bound to IF, its domains can be assembled. This mechanism ensures that only the IF–vitamin B12 complex, not the apoprotein present in a high excess, can be recognized and absorbed with the help of the specific receptor cubam [15,26,27].
IF deficiency can be caused by the presence of anti-IF or anti-parietal cell autoantibodies as well as the age-dependent or autoimmune degeneration of parietal cells, e.g., autoimmune gastritis (AIG). The presence of autoantibodies is often linked to Helicobacter pylori (H. pylori) infection. Patients with H. pylori infection have been found to have decreased vitamin B12 levels in serum [28]. The exact cause of AIG is still unknown and there is no treatment available. The wide range of unspecific clinical symptoms, which are often subtle (many patients are asymptomatic for a long period after the disease onset), result in difficulties in diagnosing the condition promptly [29]. The lack of IF in this condition leads to long-lasting vitamin B12 deficiency due to the inability to deliver vitamin B12 to the cells within the ileum and subsequent failure to distribute it throughout the body. The consequence of vitamin B12 deficiency in AIG is pernicious anemia (PA) [11,30]. Around 10–15% of patients with AIG develop PA, characterized by hematological, neurological impairments, or both. The hematological manifestation includes megaloblastic anemia, while the clinical neurologic manifestations of PA include weakness, depression, impaired nerve function, cognitive impairment, and sensory deficits [11,31].

2.1.2. Cubam Receptor

When the vitamin B12–IF complex reaches the intestine, it is internalized into enterocytes by receptor-mediated endocytosis. Cubam is a multiligand apical membrane receptor expressed in absorptive epithelia in several tissues. In the kidney, cubam is expressed in the proximal tubular epithelium where it mediates the reabsorption of proteins from the renal filtrate, such as albumin, apolipoprotein A-1, hemoglobin, transferrin, and vitamin-D binding protein, leading to reduced proteinuria [32,33]. Cubam is also present in the visceral yolk sac, where it is essential to coordinate maternal nutrient uptake and proper embryo development [34,35]. In the epithelium of the distal ileum, the only known function of cubam is to facilitate the uptake of the IF–vitamin B12 complex [33].
Cubam is composed of two proteins: cubilin (CUBN) and type-1 transmembrane protein amnionless (coded by gene AMN). Cubilin is a 460 kDa protein that contains three distinct regions: an N-terminal domain, responsible for membrane association; an extracellular fragment formed by epidermal growth factor-like repeats; and 27 domains of complement components C1r/C1s, epidermal growth factor-related protein 1 (Uegf), and bone morphogenetic protein 1 (Bmp1), CUB [36]. Four CUB domains, CUB5–8, are responsible for the interaction with the IF–cobalamin complex, as revealed by crystal structures [37]. The remaining domains are involved in the recognition and binding of other ligands, for example, during ultrafiltration and reabsorption in the kidney [38]. To form a receptor, three CUBN-subunits combine in a pin-shaped molecule that docks to transmembrane protein AMN. Cubilin lacks a transmembrane domain and, therefore, the interaction with amnionless is necessary for the formation of a functional cubam complex [32].
AMN-protein is a 45-kDa transmembrane protein also composed of three parts: an extracellular domain, a transmembrane helix, and a cytoplasmic domain responsible for clathrin-mediated internalization. In addition to anchoring CUBN-protein to the apical membrane of epithelia, amnionless is also required for biosynthetic processing and trafficking of cubilin to the plasma membrane. It also mediates cubam receptor recycling and endocytic signaling [39]. The cytoplasmic domain of amnionless features two putative FXNPXF endocytic signals that closely resemble the FXNPXY signal found in the low-density lipoprotein receptor (LDLR) superfamily. The related signaling involves clathrin-associated sorting of proteins (CLASPs) and promotes the clathrin-dependent internalization of various cubam ligands in an analogous way to the internalization of LDLRs [33]. The LDLR family also includes the protein Megalin, which strongly colocalizes with cubam, especially in the kidney where they cooperate in ligand recognition. However, in the intestine, cubam appears to be functioning independently of Megalin, suggesting that only AMN and CUBN protein moieties are required to form a receptor specific for IF–vitamin B12, internalization [40].
Mutations in either of the CUBN or AMN genes disrupt cubam receptor functioning and result in malabsorption of vitamin B12 and the development of Imerslund–Gräsbeck syndrome (IGS). IGS is a rare autosomal recessive disorder that appears in childhood. It is characterized by megaloblastic anemia as a consequence of vitamin B12 deficiency [41]. Most IGS patients also develop mild proteinuria and non-specific symptoms attributed to vitamin B12 deficiency [42], even though, if left untreated, the disease may result in neurological impairments and can be fatal. Treatment with parenteral vitamin B12 therapy, typically via regular intramuscular injections, is effective in managing the condition [43,44]. Recent data also demonstrate altered expression of amnionless in the kidney and the ileum with aging. This causes vitamin B12 deficiency irrespectively of its normal nutritional uptake as well as the development of age-related chronic diseases [45].

2.1.3. Multidrug Resistant Protein 1 (MRP1)

Upon the uptake to the enterocyte, the next step in vitamin B12 transport is its distribution to other cells and tissues in the body. To achieve this, vitamin B12 needs to be exported into the blood circulation and this step is mediated by multidrug resistant protein 1 (MRP1, gene ABCC1).
MRP1 has been described for the first time, in 1992 by Cole et al. This 190 kDa glycoprotein that has been shown to export a wide range of substrates including bioactive compounds, signaling molecules, metabolites, anti-cancer drugs, and xenobiotics, thus conferring multidrug resistance in cancer. For example, MRP1 is overexpressed in many hematological and solid tumors [46,47]. Under physiologic conditions, MRP1 is ubiquitously expressed in the body, mostly in the lung, testis, kidney, heart, placenta, small intestine, colon, brain, and peripheral blood mononuclear cells [47,48,49]. In the small intestine, MRP1 is mainly present in the basolateral membranes of the crypt cells [50]. In addition to the plasma membrane localization, MRP1 has also been shown to accumulate in subcellular organelles such as endocytic vesicles, trans-Golgi vesicles, and lysosomes [48,51], although its involvement in the intracellular transport of vitamin B12 has never been demonstrated. MRP1, similarly to lysosomal vitamin B12 exporter ABCD4, also belongs to the ABC transporters family, hence its alternative name ABCC1 [46].
MRP1 has been shown to efflux vitamin B12 from the intestinal epithelium to the blood circulation [52]. Mice deficient in MRP1 (MRP1(−/−)) have decreased levels of vitamin B12 in their plasma, simultaneously accumulating the vitamin in the ileum and colon. These studies also established that MRP1 transports the vitamin in its protein-free form, contrary to the previously hypothesized export of the vitamin in complex with transcobalamin [52,53]. In some patients with recessive hereditary vitamin B12 malabsorption, mutations in the MRP1-encoding gene ABCC1 were detected; however, they were ruled out as potential causes of vitamin B12 deficiency as they either appeared in the intronic gene fragments, which did not affect the open reading frame, or did not fulfill Mendelian rules for inheritance [54]. These studies were conducted on a small group (approx. 30) of patients, suggesting a need for large-scale epidemiological investigation to clarify the involvement of ABCC1 mutations in the development of vitamin B12 deficiency. Such investigations could potentially validate the redundancy in vitamin B12 export given the existence of alternative pathways aside from MRP1 export in humans. Among other closely related transporters of the ABC family with wide tissue distribution (ABCB6, ABCG2, MRP3, and MRP5), only MRP1 was shown to be involved in vitamin B12 efflux [52]. Moreover, as presented by Beedholm-Ebsen et al., MRP1(−/−) mice have not presented morphologic and metabolic abnormalities, suggesting that other transporters should compensate for the lack of MRP1 (also considering a passive transport of vitamin B12 through membranes herein).

2.2. Proteins Involved in Vitamin B12 Blood Transport

Transcobalamin (TC)

When vitamin B12 is exported into the blood circulation, it is bound by a 43 kDa protein transcobalamin, a former transcobalamin II (gene TCN2) [55]. In humans, TC is synthesized by a variety of cell types and tissues and secreted to blood plasma [56,57]. TC is structurally related to IF and HC, with some conserved regions sharing 60–80% homology; all three vitamin B12 transporters are derived from a common ancestral gene [53,58]. TC is responsible for the distribution of vitamin B12 into all of the cells in the body but only about 25% of the total vitamin B12 in plasma is bound by TC, while the rest is carried by circulating HC [59,60]. The complex of vitamin B12 and transcobalamin is referred to as holoTC and is the “active” form of vitamin B12 with a fast turnover rate, making vitamin B12 available to cells. In contrast, the biological function of the complex with HC (holoHC) is not clear [21].
HoloTC levels in serum have been suggested as a specific marker of early vitamin B12 deficiency diagnosis with the use of commercially available immunoassays for holoTC [61]. However, some studies indicate that this is a controversial determinant of vitamin B12 status. Some patients with severe vitamin B12 deficiency, resulting in PA, displayed falsely normal holoTC levels [62]. It was also hinted that the rationale behind the diagnostic holoTC immunoassay is based on a false assumption that holoTC is the first factor to respond to vitamin B12 deficiency, where the level of methylmalonic acid (MMA), a form of substrate for methylmalonyl-CoA mutase that requires vitamin B12 as a cofactor, and is more sensitive to vitamin B12 deficits in the early stage [63,64].
Mutations in the TCN2 gene can hamper or even prevent intracellular accumulation of cobalamin, causing a rare multisystemic disorder characterized by autosomal recessive inheritance with symptoms such as pancytopenia, megaloblastic anemia, developmental delay, diarrhea, psychomotor impairment, and immune system deficiency [65]. Only a few cases have been investigated so far, showing either a large deletion of the entire exon 8 leading to a premature stop codon (p.E371fsX372) [66] or a change in intron 4 leading to the loss of the start codon as a cause of gene alternation [67].

2.3. Proteins Involved in Intracellular Vitamin B12 Transport

2.3.1. CD320 Receptor

CD320, also known as the receptor for cobalamin-saturated transcobalamin (TCblR), is a small 282 amino acid protein with 29 kDa calculated molecular weight (observed 58 kDa, probably due to a high glycosylation) [68]. It belongs to the LDLR protein family. CD320 is localized on the cell surface and contains two domains connected by the transmembrane helix. These are the C-terminal cytoplasmic region and the extracellular N-terminal fragment with an epidermal growth factor (EGF) separating the two LDLR type A sites [69,70]. A soluble form (sCD320) was also found in the human serum fraction [71]. CD320 exhibits high specificity and affinity toward transcobalamin and the TC–cobalamin complex is bound to CD320 and absorbed by the cells in the presence of Ca2+ ions [72,73]. Initially, holo-TC binds to the N-terminal region of CD320 whereupon it is internalized into an endocytic vesicle. Next, endosomes with holo-320 complexes are fused with a lysosome, where TC and CD320 undergo dissociation and degradation. Subsequently, cobalamin is freed into the cytoplasm [74].
During the process in which holo-TC uptake receptors are broken down, the occurrence of CD320 on the cell surface is the direct result of its synthesis [75]. The level of the CD320 receptor is highly associated with the cell-division cycle and its presence on the cell surface is high in fast-growing cells, including cancer cells [74]. The impact of CD320 deficiency was tested in animal models with CD320 knock-out (KO) mice fed with a vitamin B12 deficient diet. Those animals showed increased MMA and homocysteine levels in the serum, as well as macrocytic anemic phenotype [76]. Neutrophil hypersegmentation is usually also a clinical hallmark of cobalamin uptake failure in patients [77]. Moreover, female mice with CD320 depletion and limited cobalamin intake exhibited infertility. Even though supplementation with cyanocobalamin resulted in a successful pregnancy, all litters died just after birth [76]. Other studies indicate that the knockdown of CD320 caused cognitive and learning disorders along with anxiety behavior corresponding to symptoms observed in humans, pointing out direct changes in the central nervous system [78]. It is suggested that observed alternation of DNA methylation in CD320 KO mice brains might cause neuron demyelination [79].
A single nucleotide polymorphism of CD320, rs2336573, leading to a glycine-to-arginine change has been identified in patients with lowered serum vitamin B12 levels [80]. Analogous outcomes occurred at a single codon deletion of c.262_264GAG (p.E88del). These observations have been linked with an increased MMA concentration in newborns [81]. However, in 2022, Pangilinan et al., using population data, estimated that approximately 85% of p.E88del homozygotes born in the same period did not have elevated propionylcarnitine (C3), suggesting that cobalamin metabolism in the majority of these infants with this genotype is unaffected. Thus, more studies are required to clarify the natural history of various CD320 variants across patient populations [82].

2.3.2. ATP Binding Cassette Subfamily D Member 4 (ABCD4) and the Lysosomal Cobalamin Transport Escort Protein (LMBD1)

Free vitamin B12 is released into the cytoplasm by an ATP (adenosine-5′- triphosphate) Binding Cassette Subfamily D Member protein (ABCD4, encoded by the gene of the same name), a 68.6 kDa exporter located on the lysosomal membrane [83]. ABCD4 consists of a transmembrane domain and a nucleotide-binding domain, which is involved in binding and hydrolysis of ATP [84]. It belongs to the family of ATP-binding cassette (ABC) transporters that mediate the transport of various substrates across the cell membrane against the concentration gradient. Most of the human ABC exporters are localized in a way that allows the substrate to access from the binding site, in a cis-acting manner. As revealed by cryo-electron microscopy (EM) studies, ABCD4 is a unique exporter, facilitating transfer in a trans-acting manner, thus allowing the transfer of vitamin B12 from the lysosome to the cytoplasm [85].
ABCD4 is a membrane endoplasmic reticulum (ER)-resident protein and requires translocation from the ER to the lysosome to facilitate vitamin B12 release. ABCD4 was found to colocalize with a 61.4 kDa membrane lysosomal cobalamin transport escort protein (LMBD1, encoded by gene LMBRD1), which localizes primarily to lysosomes [83,86]. Initially, LMBD1 was suggested to act as a lysosomal exporter of vitamin B12 [87]. However, later findings demonstrated that LMBD1 acts as an adaptor protein and mediates the translocation of ABCD4 from ER to the lysosomes [88]. Both proteins function in a coordinated manner and form a complex, as demonstrated by Surface Plasmon Resonance (SPR) [86].
Mutations in ABCD4 and LMBDR1 genes cause disease group cobalamine J and F diseases (cblJ, cblF), respectively [83,87]. At the cellular level, both errors are marked with a decreased function of methylmalonyl-CoA mutase and methionine synthase and an accumulation of free vitamin B12 in the lysosomes [89]. However, up to date, there is no clear biochemical or clinical distinction between cblJ and cblF. Only molecular genetic testing, either gene-targeted (RNA-sequencing and single-gene testing) or comprehensive (e.g., exome sequencing), was successfully applied to confirm each of the defects [90,91,92]. There is a wide spectrum of clinical presentations in both diseases, which may include hampered growth, developmental delay, developmental abnormalities progressive hyperpigmentation, cardiac defects, and hematologic impairments (such as macrocytic anemia, neutropenia thrombocytopenia, and pancytopenia). Neurological presentations may include general weakness, dizziness, delayed motor development, and mild intellectual disability [83,87,90,93].

2.3.3. Methylmalonic Aciduria and the Homocystinuria Type C Protein (MMACHC)

Cobalamin that is released from lysosomes into the cytosol is available for two cytoplasmic proteins—methylmalonic aciduria and homocystinuria type C protein (MMACHC or CblC, encoded by MMACHC gene) and type d protein (MMADHC or CblD, encoded by MMADHC gene). The analysis of individuals with inborn errors of cobalamin metabolism revealed that the MMACHC gene is localized to the 1p34.1 region of chromosome 1 and consists of five exons. A study by Wang et al. suggested that c.609G > A, c.658_660delAAG, c.80A > G, and c.482 G > A mutations were the most common mutations in the Chinese population, while the most frequent mutations in the Caucasian population were c.271dupA, c.394C > T, and c.331C > T [94,95]. The MMACHC gene product consists of 282 amino acids and has a mass of 31.7 kDa [96]. MMACHC manifests a substantial degree of evolutionary conservation across mammalian species and exhibits high expression levels in the fetal liver, with its presence also identified in a broad spectrum of tissues, including the spleen, thymus, and bone marrow [96].
The MMACHC gene encodes a protein possessing both chaperone and enzyme functionalities. Its inactivation disrupts the reduction in incoming cobalamin molecules, which is necessary for the following synthesis of two essential cobalamin forms: methylcobalamin (MeCbl) and adenosylcobalamin (AdoCbl), which are cofactors for enzymes dependent on vitamin B12 [97]. The defects in MMACHC impact a stage occurring after the cellular absorption of cobalamin but preceding the synthesis of AdoCbl and MeCbl [96]. MMACHC catalyzes the reductive decyanation of cyanocobalamin (CNCbl or vitamin B12 in conjunction with a flavoprotein oxidoreductase) [98]. Interestingly, CblC also exhibits the remarkable capability to catalyze the nucleophilic displacement of the alkyl group by glutathione when presented with an alkylcobalamin. However, different forms of cobalamin are processed with a different efficacy by the intracellular system of enzymes and chaperons (e.g., better for GSCbl in comparison to other Cbl-forms [99]). This leads to considerable variation in the rates of the intracellular accumulation of those Cbls and, consequently, to different rates of the coenzyme appearance as established for HOCbl and CNCbl [100]. The intricate protein–protein interaction between MMACHC and MS is widely acknowledged for its pivotal role in regulating intracellular cobalamin metabolism. Notably, the interaction of MS inactive isoforms impedes MMACHC activity [101]. MMCHC is also essential in the conversion of MMA to succinyl-CoA [98].
Mutations in the MMACHC gene can result in a rare genetic disorder referred to as methylmalonic aciduria and homocystinuria type C (cblC). They demonstrate deficiencies in both vitamin B12-dependent enzymatic functions crucial in mammals—MS and MUT [98]. Thus, this disorder is characterized by the accumulation of both MMA and homocysteine (Hcy) in serum [98]. Individuals affected by the cblC disorder manifest severe neurological and systemic metabolic abnormalities [97], such as atypical hemolytic uremic syndromes (aHUS), decline in renal function, idiopathic neuropathies, spinal cord degenerations, ataxias, recurrent thrombosis, visual field defects, maculopathy, and optic disc atrophy [102]. Interestingly, recent studies suggest that the involvement of cobalamin and MMACHC might have a potential role in the oncogenic transformation of specific tumor subtypes. Melanomas and cell lines derived from melanoma often exhibit a higher occurrence of elevated methylation in the MMACHC promoter compared to other cancer types. This fact implicates the inactivation of MMACHC as the causative factor [97,103]. However, Bauer et al. highlighted that the defect in metabolism observed in MeWo-LC1 was unique and that decreased MMACHC expression was not a cause of methionine dependence in the other melanoma lines [104].

2.3.4. Methylmalonic Aciduria and the Homocystinuria Type D Protein (MMADHC)

The MMADHC gene is associated with methylmalonic aciduria type D and homocystinuria and it is classified within the cblD complementation group [105]. The MMADHC gene encodes a protein consisting of 296 amino acids, with a molecular weight of 32.9 kDa. The MMADHC protein (also called CblD protein) exhibits a high degree of conservation across various mammalian species. The MMADHC gene encodes a robust N-terminal mitochondrial leader sequence along with a conserved vitamin B12-binding motif located at residues 81–86 [105]; however, a recent study has also shown cobalamin binding to a different region of the MMADHC protein [106]. It is present in both the mitochondria and cytoplasm, although the precise mechanisms that lead to this dual localization are not fully understood. The analysis of MMACHC and MMADHC gene expression patterns during mouse development unveils overlapping expression, particularly evident in the developing cardiac, respiratory, and nervous systems. It is suggested that these two proteins are co-expressed in all cells, as there is no discernible specificity for any particular organ, tissue, or cell type exclusively expressing MMADHC. Notably, MMADHC demonstrates expression in all cells of embryos, including those expressing MMACHC [107].
The physiological function of MMADHC is yet to be fully understood; however, it seems to play a role in the delivery of the cobalamin precursor to either cytoplasmic or mitochondrial sites of the production of the two active cobalamin forms, MeCbl and AdoCbl [108]. Plesa et. al., in 2011, employed a phage display to predict regions on MMADHC that bind to MMACHC. The study identified five distinct regions on MMADHC, with all aligning with residues C-terminal to Met116 [109] but not with the N-terminus (residues 1–61) [110]. MMADHC lacks any recognized enzymatic activity and it is suggested to be the first protein identified to repurpose the nitroreductase fold exclusively for protein–protein interactions [111].
The cblD complementation group can be linked to isolated methylmalonic aciduria (cblD-MMA), isolated homocystinuria (cblD-HC), or a combination of both methylmalonic aciduria and homocystinuria (cblD-MMA/HC) [112]. Individuals characterized by cblD-MMA have at least one altered allele that induces a premature stop codon in the N-terminal region of the protein. In contrast, those with cblD-HC possess at least one missense mutation, resulting in amino acid substitutions located in the C-terminal part of the protein [108]. The clinical presentation indicates that patients with cblD-MMA/HC and cblD-HC exhibit neurological and hematological symptoms. In contrast, cblD-MMA patients experience respiratory distress, hyperammonemia, and neurological symptoms [105,112,113].

2.4. Vitamin B12-Dependent Intracellular Enzymes

2.4.1. Methionine Synthase (MS)

Methionine synthase (MS, encoded by gene MTR) is a big ~140 kDa monomeric protein found in the cytosol [114] and it consists of four domains. In bacteria, the N-terminal site, dependent on the zinc ions, binds and activates Hcy while the following domain has a high affinity to methyltetrahydrofolate and might affect its activity [115,116]. The third module is vitamin B12-dependent and crucial in methyl group transfer from 5-methyltetrahydrofolate to Hcy with the production of tetrahydrofolate and methionine [116]. Finally, the C-terminal domain binds S-adenosylmethionine for reductive activation of MS [117], when cobalamin cofactor undergoes oxidation process [117]. The human enzyme displays a 58% identity with MS from Escherichia coli (E. coli) [118], which is one of the biggest proteins with 1227 residues and a molecular weight of ~136 kDa [119]. The remarkable conservation of amino acid residues across proteins indicates a high degree of similarity in properties and structures between enzymes from both humans and E. coli. Consequently, the information gleaned from the E. coli enzyme is presumed to apply to its human counterpart [118]. However, a crucial divergence is the lack of flavodoxin in mammals, a pivotal component in the reductive reactivation process observed in bacterial models [120].
In mice models, full depletion of the MTR gene caused pregnancy failure [114]. In proliferating cells of the skin and intestinal epitheliums, as well as in rat intestine crypts and proliferating Caco-2 cells (human epithelial colorectal adenocarcinoma cell line), there is a consistent increase in the transcription, protein expression, and activity of MS. Additionally, MS activity correlates with DNA methylation in rat intestine villi [121]. Moreover, in situ hybridization indicated that the human MTR gene is found on the 1q43 chromosome region [118] and its missense mutation or deletion of three base pairs was identified in the cblG group of patients with cobalamin metabolism disorder [122]. However, the second form of this disorder has also been recognized as “methylcobalamin deficiency cblE type” (cblE) and it is believed to arise from mutations occurring at distinct genetic loci [123]. This rare autosomal disease is associated with homocysteinemia, homocystinuria, and hypomethioninemia [122] and patients usually exhibit symptoms within the initial two years of their lives [124]. Reduced MS activity also leads to an elevation in the catalytic protein phosphatase 2A (PP2A), resulting in an imbalance in the phosphorylation/methylation of nucleocytoplasmic RNA binding proteins. This leads to compromised energy metabolism and neuroplasticity linked to inborn errors of Cbl metabolism (IECM) [125].

2.4.2. Methylmalonyl-CoA Mutase (MCM)

Methylmalonyl coenzyme A mutase (MCM, encoded by gene MMUT) is an enzymatic protein found in mitochondrial matrix with a ~78 kDa molecular weight [126]. The structure of MCM is highly conserved evolutionally; thus, this AdoCbl-dependent enzyme may be found in both eukaryotes and prokaryotes [126] but not plants [127]. MCM takes part in converting carbon skeletons of odd-chain fatty acids, cholesterol, and amino acids such as threonine, methionine, valine, and isoleucine into succinyl-CoA for further utilization in the tricarboxylic acid (TCA) cycle [128].
It has to be mentioned that the proper MCM activity depends on AdoCbl, the synthesis of which involves chaperones encoded by the cblA and cblB genes. The cblA gene encodes a GTPase methylmalonic aciduria type A (MMAA) and cblB encodes adenosyltransferase (ATR), with each playing unique roles in cofactor assimilation, loading, and repair of MCM [126]. ATR, also known as MMAB for its role in methylmalonic aciduria type B, is a multifunctional protein acting as both an enzyme and a chaperone in mitochondrial transport [129]. Its adenosyltransferase activity converts inactive cobalamin into AdoCbl and inorganic triphosphate in the presence of a reductant and ATP. Some studies show that inactivation of MCM occurs when AdoCbl is oxidized to HOCbl [130]. The inactivation of MCM occurs when the active adenosyl radical generated during the enzyme catalytic cycle reacts with a molecule other than the substrate, inactivating the enzyme. However, the formation of the MMUT-methylmalonic aciduria type A (MMAA) protein complex with MCM reduces the rate of oxidized cofactor formation, protecting enzyme activity. In detail, the enzyme is inactivated until the damaged cobalamin is removed and replaced by MMAA [131].
Human MCM is encoded by the MMUT gene, situated as a lone copy on chromosome 6 (6p21.3), and composed of 13 exons [132]. Its mutation leads to methylmalonic acidemia—a metabolic autosomal disease. Patients with acidemia suffer from vomiting followed by dehydration, muscle weakness, lethargy, mental incapacity, and, rarely, to death when not treated [133].
Two subgroups of MCM mutation have been recognized so far: the mut defect, characterized by residual activity in the presence of high concentrations of AdoCbl, and the mut0 defect, with complete loss of MCM activity [134]. Distinguishing between mut and mut0 subgroups can be challenging at times. Certain mut mutations may only impact the apparent Michaelis constant (Km), while others can influence both the apparent Km and the maximum velocity of an enzymatic reaction (Vmax) [133].

3. Vitamin B12 Deficiency

3.1. Epidemiology and Symptoms of Vitamin B12 Deficiency

The prevalence of vitamin B12 deficiency in the United States (US) ranges from 2% to 15%, depending on the biomarker (e.g., total/active vitamin B12 metabolites or combined index of vitamin B12 status) and its cut-off used, as well as on the age, gender, and ethnicity [135,136]. An analysis of National Health and Nutrition Examination Survey data from 2007 to 2018 revealed that about 3.6% of all adults aged 19 and older have a vitamin B12 deficiency (in the US, it is defined as serum vitamin B12 levels below 200 pg/mL (148 pmol/L)). The deficiency rate increases to 3.7% in individuals aged 60 and older. However, vitamin B12 insufficiency (defined as serum levels below 300 pg/mL (221 pmol/L)) is more prevalent, affecting approximately 12.5% of all adults aged 19 and older and 12.3% of those aged 60 and older [137]. During pregnancy, serum vitamin B12 levels often decline, sometimes dropping to subnormal levels, but typically return to normal after delivery [138]. Furthermore, vitamin B12 deficiency due to pernicious anemia is more common in people of northern European ancestry than people of African descent [139,140].
Food-bound cobalamin malabsorption (FBCM) is among the frequent causes of vitamin B12 deficiency [141]. FBCM is characterized by a hindered liberation and absorption of cobalamin from food sources [142]. Impaired release of cobalamin from its gastrointestinal transport proteins is commonly seen in conditions like gastritis or during medication use, reducing hydrochloric acid production. Milder symptoms may persist due to impaired IF-dependent transport functionality [143]. Proton pump inhibitors and histamine H2-receptor antagonists heighten the risk of future cobalamin deficiency, while cessation of medication reduces the risk. For example, the widespread use of acid-suppressing medications may lead to undiagnosed cobalamin deficiency [144].
The effects of vitamin B12 deficiency are mainly seen in the blood and nervous system. Thus, the classic manifestation was first identified as pernicious anemia [145]. Since then, the spectrum has shifted considerably, starting with the recognition that neurological manifestations (ataxia, cognitive decline leading to dementia, or psychiatric disorder) often predominate and can occur in the absence of hematological disorders [146]. Anemia is caused by disruption of DNA synthesis in a hematopoietic system, predominantly erythroid precursors. It is caused by a “folate trap”, i.e., when the low activity of MS (caused by a low MeCbl coenzyme level) is not sufficient to liberate free tetrahydrofolate from its inactive methyl-complex. Neurological complications are associated with impaired methylation of myelin lipids caused by insufficient production of S-adenosyl methionine (mediated by MeCbl). Such insufficiency damages the isolating myelin sheath of axons [31,147,148]. Moreover, lack of MeCbl or MTR inhibition leads to hyperhomocysteinemia linked with inflammation, oxidative stress, and microvascular disease that may exacerbate neurological symptoms [149].
In addition to the symptomatic presentation of vitamin B12 deficiency, subclinical cobalamin deficiency (SCCD) presents a condition characterized by biochemical aberrations in the absence of clinical signs [141,150]. SCCD is encountered more commonly than the overt deficiency and it often originates from FBCM [151,152]. Currently, SCCD is perceived as a transient state without overt cobalamin depletion, necessitating exploration into remediable etiologies. Patients exhibiting serum cobalamin levels between 110 and 148 pmol/L should be reexamined by other markers to the test (e.g., MMA) and undergo retesting within one to two months; those exhibiting normalization need not undergo further assessment [153]. Persistent low levels may warrant low-dose oral cobalamin supplementation alongside anti-IF antibody analysis. Neurological manifestations should prompt reassessment by a healthcare provider. Positive anti-IF antibody titers necessitate treatment for pernicious anemia, while negative results mandate reevaluation after three to four months [154]. Additional biochemical assessments are warranted if cobalamin levels remain consistently low.

3.2. Diagnosis and Treatment of Vitamin B12 Deficiency

The assessment of vitamin B12 status typically relies on measurements of its serum or plasma levels. Laboratories often define subnormal values as below 200 or 250 pg/mL (148 or 185 pmol/L). Typically, a value significantly below the lower limit of the reference range suggests probable vitamin B12 deficiency, while a value well above probably indicates its adequate status. However, exceptions arise in patients with pernicious anemia who have circulating antibodies against IF; their vitamin B12 levels may appear falsely normal despite deficiency, supposedly because of the immunological interference from anti-IF antibodies with the test setup [155]. Active vitamin B12 bound to transcobalamin (holotranscobalamin) theoretically offers increased sensitivity for detecting vitamin B12 deficiency compared to total serum vitamin B12 levels. Although the measurement of holotranscobalamin is gradually becoming available in clinical settings, it has demonstrated only marginal improvement over total vitamin B12 levels as a biomarker for deficiency [156,157].
Another valuable and highly sensitive marker of vitamin B12 status is the serum MMA level, with a threshold greater than 0.271 µmol/L indicative of deficiency [11]. However, MMA levels can be elevated due to renal insufficiency and tend to be higher in older adults. Genetic influences on MMA cutoffs have also been highlighted by genome-wide association studies. A common polymorphism (rs291466) in HIBCH, involved in valine catabolism, is prevalent (minor allele frequency 0.43) and homozygotes for the allele show 46% higher plasma MMA levels that potentially impact its efficacy of this metabolite as a vitamin B12 status indicator [158]. Besides MMA concentration, total plasma Hcy levels also rise rapidly to levels above 15 µmol/L, with declining vitamin B12 status. However, Hcy levels lack specificity, being influenced by factors such as folate levels or renal function decline [159]. To overcome these challenges, different combinations of the aforementioned markers are frequently used in analysis [160] and references.
Other biochemical parameters that might be useful in diagnosing vitamin B12-related disorders include methionine. Indeed, in remethylation disorders, methionine levels might be suppressed but, more frequently, they remain stable. Thus, measuring methionine may help differentiate remethylation disorders from classical homocystinuria in cases with elevated Hcy [161]. Remethylation defects are linked with impaired remethylation of homocysteine to methionine and thus lead to the accumulation of homocysteine and perturbation of numerous methylation reactions [162]. All genetic remethylation defects result in reduced methionine synthase activity due to various factors: decreased function of the enzyme itself or its associated enzyme, methionine synthase reductase; insufficient production of the cofactor methylcobalamin; or disrupted availability of the substrate methyl-tetrahydrofolate (MTHF) [163]
Optimal management of vitamin B12 deficiency requires identifying its cause, particularly differentiating between its low intake and malabsorption. The Schilling test, once the gold standard, is now rarely used due to the lack of radiolabeled vitamin B12 availability and cost constraints. The CobaSorb test, measuring serum holotranscobalamin levels of post-vitamin B12 dosing, is another opportunity [164,165,166]. Other tracers (e.g., low radioactive 14C- and nonradioactive 13C-labeled vitamin B12) are currently discussed as feasible alternatives to 57Co-labeled vitamin B12 of the Schilling text [167,168]. Alternative diagnostic approaches include detecting autoantibodies for pernicious anemia and assessing for atrophic gastritis during upper endoscopy [169]. While a negative result for anti-IF autoantibodies does not exclude pernicious anemia, a positive result is highly specific. However, a positive result for anti-H+/K+ ATPase autoantibodies lacks specificity, as it may occur in patients without pernicious anemia [170,171]. In pediatric cases, consideration of inborn errors of cobalamin metabolism via complementation phenotyping or genetic analysis is crucial for diagnosis. These tests help differentiate between different genetic causes, aiding in tailored treatment approaches for children with vitamin B12 deficiency [12,172].
Treatment for vitamin B12 deficiency varies depending on the underlying cause. In 2021, The British Society for Haematology (BSH) updated guidance on vitamin B12 replacement [173]. For non-diet-related deficiency, such as in pernicious anemia or other conditions, hydroxocobalamin 1 mg intramuscular every 2–3 months for life is recommended, though the optimal time intervals might significantly vary from patient to patient. For diet-related deficiency, individuals can either orally intake cyanocobalamin (50–150 micrograms daily) between meals or receive a twice-yearly hydroxocobalamin 1 mg injection. In cases of dietary deficiency, treatment duration may vary, with vegans potentially requiring lifelong treatment.

3.3. Long-Term Consequences of Vitamin B12 Deficiency

Vitamin B12 deficiency, owing to its pivotal role in cellular metabolism, can mimic a diverse array of clinical presentations as listed above. However, a manifestation of these clinical outcomes is relatively late. It results in the development of SCCD, which may have severe consequences, including increased risk of neurodegenerative and cardiovascular diseases or, possibly, enhanced cancer progression due to the accumulation of MMA according to [174].
In vegetarians, deficiency of vitamin B12 was mostly found to be associated with depression and adverse neurological function. Berkins et al. highlighted that the dietary intake of vitamin B12 might affect brain structure [175]. Interestingly, supplementation of vitamin B12 along with anti-depressant therapy greatly improved depressive symptoms [176]. Other studies conclude that cobalamin deficiency may have some similarities with multiple sclerosis. In both conditions, destroyed myelin sheaths release glutamate, which is known to possess exotoxicity [177]. It is well known that cognitive decline often coincides with a high prevalence of vitamin B12 deficiency, although causality in these correlations is not fully established [178]. Furthermore, while the actual percentage of reversible dementia in individuals with vitamin B12 insufficiency is deemed low, impaired vitamin B12 metabolism could potentially serve as one of several factors influencing the onset and progression of cognitive decline and dementia [179]. It could be caused by reduced methylation capacity, elevated Hcy levels, and the role of B vitamins in upholding the integrity of the blood–brain barrier [180]. Some researchers propose that it may act as a predisposing factor for Alzheimer’s disease [181]. Additionally, it is associated with various other neurological conditions including Wernicke’s encephalopathy, subacute combined degeneration of the spinal cord, and peripheral neuropathy [182].
It is well known that hyperhomocysteinemia, strongly linked to vitamin B12 deficiency, is a prominent risk factor for atherosclerotic vascular diseases affecting coronary, cerebral, and peripheral vessels, as well as arterial and venous thromboembolism [183]. The adverse effects of Hcy on the vascular endothelium and coagulation cascade, along with its procoagulant properties, encompass reduced binding of antithrombin III to endothelial heparan sulfate, heightened affinity between lipoprotein (a) and fibrin, increased tissue factor activity in endothelial cells, and inhibition of factor V inactivation by activated protein [184]. Supplementation with vitamin B12 led to the resolution of both clinical and biological abnormalities in all participants [185]. However, whether vitamin B12 deficiency directly influences thrombosis or acts through hyperhomocysteinemia or other factors remains to be fully understood [186]. Lifestyle factors like smoking, BMI, and physical activity may also contribute to the association between hyperhomocysteinemia and thromboembolism [187].
Interestingly, other compounds that are strongly linked to vitamin B12 deficiency (e.g., MMA) have been recently identified as oncometabolites. MMA is a metabolite of methylmalonyl coenzyme A; its level is increased because of impaired conversion to succinyl coenzyme A mediated by a coenzyme form of vitamin B12 (AdoCbl) [188]. A study published in 2020 by Gomes et al. revealed that MMA levels increase in the serum of older individuals and may have a function as a mediator of tumor progression. Indeed, treatment of adenocarcinoma human alveolar basal epithelial (A549) cells with MMA induced a complete pro-aggressive epithelial to mesenchymal transition-like phenotype with a decline in E-cadherin and a concurrent increase in fibronectin and vimentin [174]. Additionally, two years later, Gomes et al. demonstrated that the accumulation of MMA accelerates cancer cell aggressiveness in breast and lung cancers [174,189]. Intriguingly, in the context of breast cancer, the blockade of other metabolic pathways involving MMA or the activation of MMA production, coupled with vitamin B12 intracellular deficiency, resulted in enhanced tumor metastasis [174,189]. Recently, MMA was also found to promote metastasis in colon tumor models [190]. However, the direct effect of vitamin B12 deficiency on cancer onset, prognosis, and progression remains unresolved due to many inconsistent studies [191,192,193,194,195,196].

4. Discussion

In this review, we consolidate existing knowledge regarding the proteins involved in vitamin B12 metabolism and explore their role in the cellular processes as well as the impact of their dysfunction on overall metabolic integrity. Several proteins listed in Table 1, including transcobalamin, haptocorrin, Intrinsic Factor, cubilin, and amnionless, are already utilized in the diagnosis of vitamin B12 deficiency. However, other proteins are primarily investigated at the gene level due to their intracellular localization or low detection levels in serum, limiting their use in routine diagnostics. Nonetheless, the levels of some of these proteins may harbor novel information concerning the risk of complications associated with vitamin B12 deficiency.
In 2018, Tseng et al. established that mice models demonstrated that a loss of single allele Lmbd1 would result in ventricular hypertrophy, which was accompanied by enhanced myocardial glucose uptake [197]. Moreover, a microarray screening study revealed a 25% reduction in LMBD1 mRNA levels in the septa of 13 heart transplant patients with dilated cardiomyopathy compared to non-failing hearts [198]. This clinical observation, coupled with the findings from mice studies, suggests that diminished expression of LMBD1 protein may play a role in the development of cardiomyopathy. Thus, the single nucleotide polymorphism sites associated with decreased expression of the human LMBRD1 gene evaluation may not only indicate the cause of vitamin B12 deficiency but also prognose a risk of onset of cardiovascular diseases.
Interestingly, the diminished expression of other proteins involved in vitamin B12- has been noted in the brains of multiple sclerosis patients, highlighting the role of this protein in proper central nervous system function [199]. It was accompanied by worsening the disease by both CD320 genetic deletion and vitamin B12 deficiency [199]. These findings highlight a plausible connection between vitamin B12 deficiency and the progression of pathology in patients with multiple sclerosis. One may conclude that vitamin B12 deficiency caused by CD320 disruption might be linked with an increased risk of neurodegenerative diseases; thus, its evaluation might be an interesting prognostic parameter.
Another protein linked with vitamin B12 metabolism that exhibits multifunctional properties is MRP1. MRP1 transports a wide range of therapeutic agents as well as various physiological substrates and may contribute to drug resistance development in several cancers, including lung, breast, and prostate cancers, as well as childhood neuroblastoma [200]. MRP1’s role as a glutathione transporter creates a unique vulnerability in cancer cells with high MRP1 expression, providing a potential therapeutic opportunity for targeting cancer. Direct pharmacological inhibition of MRP1 shows promise and is extensively tested in preclinical and clinical studies [201,202,203]. However, non-selective MRP1 inhibition could potentially disrupt vitamin B12 metabolism and lead to its deficiency. Therefore, there is an emerging need to investigate the effects of MRP-1 inhibition on vitamin B12 metabolism during therapy.
Interestingly, it has been noted that the utilization of the common drug metformin in diabetic patients is linked to potential malabsorption and deficiency of vitamin B12 [204,205]. Numerous observational studies and meta-analyses have underscored a significant correlation between metformin usage and vitamin B12 deficiency [206]. Patients undergoing metformin therapy display reduced absorption of B12, leading to decreased levels of serum total vitamin B12 and holoTC (“active vitamin B12”). This phenomenon is attributed to calcium-dependent antagonism in the ileal membrane, which may be mitigated through calcium supplementation [207]. Given these findings, it is strongly advised that individuals using metformin undergo regular monitoring of their vitamin B12 levels. This precaution is crucial as prolonged metformin usage may deplete hepatic vitamin B12 reserves [208].

5. Conclusions

Despite the initiation of studies on vitamin B12 as early as in the 1948s, our understanding of its biochemical alterations remains incomplete. Over the past decade, there has been a notable surge in research attention directed toward vitamin B12 biochemical, physiological, and clinical facets. Surprisingly, some recent studies suggested that vitamin B12 may affect gene expression, cellular plasticity, and tissue repair [178,209]. However, the specific molecules and cellular changes mediated by these alterations are still not fully elucidated.
Besides the fact that vitamin B12 deficiency leads to the presence of serious clinical manifestations, numerous studies have highlighted the correlation between low vitamin B12 levels and the increased risk of civilization diseases. Importantly, some of the proteins involved in vitamin B12 metabolism may act as moonlighting proteins and their pharmacological modification already used in the treatment of other clinical stages may implicate also proper vitamin B12 metabolism.
Hence, there is a pressing need for more basic science studies to unravel the vitamin B12 metabolism complexity that may lower the risk of vitamin B12 deficiency in clinical studies as well as potentially identify novel diagnostic parameters that may contribute to evaluating the risk of vitamin B12 deficiency consequences.

Author Contributions

Conceptualization, P.M., F.K. and M.T.; writing—original draft preparation, P.M., F.K., D.C. and M.T.; writing—review and editing, M.T. and R.T.S.; supervision, M.T. and R.T.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Science Centre in Poland (2021/43/D/NZ4/02501).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Folkers, K. History of Vitamin B12: Pernicious Anemia to Crystalline Cyanocobalamin. In Vitamin B12; John Wiley & Sons: New York, NY, USA, 1982; Volume I. [Google Scholar]
  2. Randaccio, L.; Geremia, S.; Demitri, N.; Wuerges, J. Vitamin B12: Unique Metalorganic Compounds and the Most Complex Vitamins. Molecules 2010, 15, 3228–3259. [Google Scholar] [CrossRef]
  3. Obeid, R.; Fedosov, S.N.; Nexo, E. Cobalamin Coenzyme Forms Are Not Likely to Be Superior to Cyano- and Hydroxyl-cobalamin in Prevention or Treatment of Cobalamin Deficiency. Mol. Nutr. Food Res. 2015, 59, 1364–1372. [Google Scholar] [CrossRef]
  4. Kurmaev, E.Z.; Moewes, A.; Ouyang, L.; Randaccio, L.; Rulis, P.; Ching, W.Y.; Bach, M.; Neumann, M. The Electronic Structure and Chemical Bonding of Vitamin B12. Europhys. Lett. 2003, 62, 582–587. [Google Scholar] [CrossRef]
  5. Institute of Medicine (US) Standing Committee on the Scientific Evaluation of Dietary Reference Intakes and Its Panel on Folate, Other B Vitamins, and Choline. Dietary Reference Intakes for Thiamin, Riboflavin, Niacin, Vitamin B6, Folate, Vitamin B12, Pantothenic Acid, Biotin, and Choline; National Academies Press: Washington, DC, USA, 1998; ISBN 978-0-309-06554-2. [Google Scholar]
  6. Herrmann, W.; Obeid, R. Cobalamin deficiency. Subcell. Biochem. 2012, 56, 301–322. [Google Scholar] [CrossRef]
  7. Santos, F.; Vera, J.L.; Lamosa, P.; de Valdez, G.F.; de Vos, W.M.; Santos, H.; Sesma, F.; Hugenholtz, J. Pseudovitamin Is the Corrinoid Produced by Lactobacillus reuteri CRL1098 under Anaerobic Conditions. FEBS Lett. 2007, 581, 4865–4870. [Google Scholar] [CrossRef]
  8. Latner, A.L.; Raine, L. Effect of Analogues on the Uptake of Vitamin B12 by the Intact Rat. Nature 1957, 180, 1197–1198. [Google Scholar] [CrossRef]
  9. Bito, T.; Bito, M.; Hirooka, T.; Okamoto, N.; Harada, N.; Yamaji, R.; Nakano, Y.; Inui, H.; Watanabe, F. Biological Activity of Pseudovitamin B12 on Cobalamin-Dependent Methylmalonyl-CoA Mutase and Methionine Synthase in Mammalian Cultured COS-7 Cells. Molecules 2020, 25, 3268. [Google Scholar] [CrossRef]
  10. Watanabe, F. Vitamin B12 Sources and Bioavailability. Exp. Biol. Med. 2007, 232, 1266–1274. [Google Scholar] [CrossRef]
  11. Green, R.; Allen, L.H.; Bjørke-Monsen, A.-L.; Brito, A.; Guéant, J.-L.; Miller, J.W.; Molloy, A.M.; Nexo, E.; Stabler, S.; Toh, B.-H.; et al. Vitamin B12 Deficiency. Nat. Rev. Dis. Primers 2017, 3, 17040. [Google Scholar] [CrossRef]
  12. Gherasim, C.; Lofgren, M.; Banerjee, R. Navigating the B12 Road: Assimilation, Delivery, and Disorders of Cobalamin. J. Biol. Chem. 2013, 288, 13186–13193. [Google Scholar] [CrossRef]
  13. O’Leary, F.; Samman, S. Vitamin B12 in Health and Disease. Nutrients 2010, 2, 299–316. [Google Scholar] [CrossRef]
  14. Kozyraki, R.; Cases, O. Vitamin B12 Absorption: Mammalian Physiology and Acquired and Inherited Disorders. Biochimie 2013, 95, 1002–1007. [Google Scholar] [CrossRef]
  15. Wuerges, J.; Geremia, S.; Randaccio, L. Structural Study on Ligand Specificity of Human Vitamin B12 Transporters. Biochem. J. 2007, 403, 431–440. [Google Scholar] [CrossRef]
  16. Mathews, F.S.; Gordon, M.M.; Chen, Z.; Rajashankar, K.R.; Ealick, S.E.; Alpers, D.H.; Sukumar, N. Crystal Structure of Human Intrinsic Factor: Cobalamin Complex at 2.6-Å Resolution. Proc. Natl. Acad. Sci. USA 2007, 104, 17311–17316. [Google Scholar] [CrossRef]
  17. Alpers, D.H.; Russell-Jones, G. Gastric Intrinsic Factor: The Gastric and Small Intestinal Stages of Cobalamin Absorption. A Personal Journey. Biochimie 2013, 95, 989–994. [Google Scholar] [CrossRef]
  18. Hewitt, J.E.; Gordon, M.M.; Taggart, R.T.; Mohandas, T.K.; Alpers, D.H. Human Gastric Intrinsic Factor: Characterization of CDNA and Genomic Clones and Localization to Human Chromosome 11. Genomics 1991, 10, 432–440. [Google Scholar] [CrossRef]
  19. Yang, S.Y.; Coleman, P.S.; Dupont, B. The Biochemical and Genetic Basis for the Microheterogeneity of Human R-Type Vitamin B12 Binding Proteins. Blood 1982, 59, 747–755. [Google Scholar] [CrossRef]
  20. Furger, E.; Frei, D.C.; Schibli, R.; Fischer, E.; Prota, A.E. Structural Basis for Universal Corrinoid Recognition by the Cobalamin Transport Protein Haptocorrin. J. Biol. Chem. 2013, 288, 25466–25476. [Google Scholar] [CrossRef]
  21. Nielsen, M.J.; Rasmussen, M.R.; Andersen, C.B.F.; Nexø, E.; Moestrup, S.K. Vitamin B12 Transport from Food to the Body’s Cells—A Sophisticated, Multistep Pathway. Nat. Rev. Gastroenterol. Hepatol. 2012, 9, 345–354. [Google Scholar] [CrossRef]
  22. Henderson, J.T.; Warwick, R.R.G.; Simpson, J.D.; Shearman, D.J.C. Does Malabsorption of Vitamin B12 Occur in Chronic Pancreatitis? Lancet 1972, 300, 241–243. [Google Scholar] [CrossRef]
  23. Guéant, J.-L.; Champigneulle, B.; Gaucher, P.; Nicolas, J.-P. Malabsorption of Vitamin B12 in Pancreatic Insufficiency of the Adult and of the Child. Pancreas 1990, 5, 559–567. [Google Scholar] [CrossRef] [PubMed]
  24. Forsmark, C.E. Diagnosis and Management of Exocrine Pancreatic Insufficiency. Curr. Treat. Options Gastroenterol. 2018, 16, 306–315. [Google Scholar] [CrossRef] [PubMed]
  25. Hygum, K.; Lildballe, D.L.; Greibe, E.H.; Morkbak, A.L.; Poulsen, S.S.; Sorensen, B.S.; Petersen, T.E.; Nexo, E. Mouse Transcobalamin Has Features Resembling Both Human Transcobalamin and Haptocorrin. PLoS ONE 2011, 6, e20638. [Google Scholar] [CrossRef] [PubMed]
  26. Fedosov, S.N.; Fedosova, N.U.; Berglund, L.; Moestrup, S.K.; Nexø, E.; Petersen, T.E. Assembly of the Intrinsic Factor Domains and Oligomerization of the Protein in the Presence of Cobalamin. Biochemistry 2004, 43, 15095–15102. [Google Scholar] [CrossRef] [PubMed]
  27. Fedosov, S.N.; Fedosova, N.U.; Kräutler, B.; Nexø, E.; Petersen, T.E. Mechanisms of Discrimination between Cobalamins and Their Natural Analogues during Their Binding to the Specific B12-Transporting Proteins. Biochemistry 2007, 46, 6446–6458. [Google Scholar] [CrossRef] [PubMed]
  28. Ayesh, M.H.; Jadalah, K.; Al Awadi, E.; Alawneh, K.; Khassawneh, B. Association between Vitamin B12 Level and Anti-Parietal Cells and Anti-Intrinsic Factor Antibodies among Adult Jordanian Patients with Helicobacter Pylori Infection. Braz. J. Infect. Dis. 2013, 17, 629–632. [Google Scholar] [CrossRef]
  29. Lenti, M.V.; Rugge, M.; Lahner, E.; Miceli, E.; Toh, B.-H.; Genta, R.M.; De Block, C.; Hershko, C.; Di Sabatino, A. Autoimmune Gastritis. Nat. Rev. Dis. Primers 2020, 6, 56. [Google Scholar] [CrossRef] [PubMed]
  30. Htut, T.W.; Thein, K.Z.; Oo, T.H. Pernicious Anemia: Pathophysiology and Diagnostic Difficulties. J. Evid. Based Med. 2021, 14, 161–169. [Google Scholar] [CrossRef]
  31. Qudsiya, Z.; De Jesus, O. Subacute Combined Degeneration of the Spinal Cord; StatPearls Publishing: Treasure Island, FL, USA, 2022. [Google Scholar]
  32. Larsen, C.; Etzerodt, A.; Madsen, M.; Skjødt, K.; Moestrup, S.K.; Andersen, C.B.F. Structural Assembly of the Megadalton-Sized Receptor for Intestinal Vitamin B12 Uptake and Kidney Protein Reabsorption. Nat. Commun. 2018, 9, 5204. [Google Scholar] [CrossRef] [PubMed]
  33. Pedersen, G.A.; Chakraborty, S.; Steinhauser, A.L.; Traub, L.M.; Madsen, M. AMN Directs Endocytosis of the Intrinsic Factor-Vitamin B12 Receptor Cubam by Engaging ARH or Dab2. Traffic 2010, 11, 706–720. [Google Scholar] [CrossRef]
  34. Strope, S.; Rivi, R.; Metzger, T.; Manova, K.; Lacy, E. Mouse Amnionless, Which Is Required for Primitive Streak Assembly, Mediates Cell-Surface Localization and Endocytic Function of Cubilin on Visceral Endoderm and Kidney Proximal Tubules. Development 2004, 131, 4787–4795. [Google Scholar] [CrossRef]
  35. Perea-Gomez, A.; Cases, O.; Lelièvre, V.; Pulina, M.V.; Collignon, J.; Hadjantonakis, A.-K.; Kozyraki, R. Loss of Cubilin, the Intrinsic Factor-Vitamin B12 Receptor, Impairs Visceral Endoderm Endocytosis and Endodermal Patterning in the Mouse. Sci. Rep. 2019, 9, 10168. [Google Scholar] [CrossRef] [PubMed]
  36. Christensen, E.I.; Verroust, P.J.; Nielsen, R. Receptor-Mediated Endocytosis in Renal Proximal Tubule. Pflügers Arch. Eur. J. Physiol. 2009, 458, 1039–1048. [Google Scholar] [CrossRef] [PubMed]
  37. Andersen, C.B.F.; Madsen, M.; Storm, T.; Moestrup, S.K.; Andersen, G.R. Structural Basis for Receptor Recognition of Vitamin-B12–Intrinsic Factor Complexes. Nature 2010, 464, 445–448. [Google Scholar] [CrossRef] [PubMed]
  38. Amsellem, S.; Gburek, J.; Hamard, G.; Nielsen, R.; Willnow, T.E.; Devuyst, O.; Nexo, E.; Verroust, P.J.; Christensen, E.I.; Kozyraki, R. Cubilin Is Essential for Albumin Reabsorption in the Renal Proximal Tubule. J. Am. Soc. Nephrol. 2010, 21, 1859–1867. [Google Scholar] [CrossRef] [PubMed]
  39. Fyfe, J.C.; Madsen, M.; Højrup, P.; Christensen, E.I.; Tanner, S.M.; de la Chapelle, A.; He, Q.; Moestrup, S.K. The Functional Cobalamin (Vitamin B12)–Intrinsic Factor Receptor Is a Novel Complex of Cubilin and Amnionless. Blood 2004, 103, 1573–1579. [Google Scholar] [CrossRef]
  40. Jensen, L.L.; Andersen, R.K.; Hager, H.; Madsen, M. Lack of Megalin Expression in Adult Human Terminal Ileum Suggests Megalin-Independent Cubilin/Amnionless Activity during Vitamin B12 Absorption. Physiol. Rep. 2014, 2, e12086. [Google Scholar] [CrossRef] [PubMed]
  41. Gräsbeck, R. Imerslund-Gräsbeck Syndrome (Selective Vitamin B12 Malabsorption with Proteinuria). Orphanet J. Rare Dis. 2006, 1, 17. [Google Scholar] [CrossRef]
  42. Kingma, S.D.K.; Neven, J.; Bael, A.; Meuwissen, M.E.C.; van den Akker, M. Imerslund-Gräsbeck Syndrome: A Comprehensive Review of Reported Cases. Orphanet J. Rare Dis. 2023, 18, 291. [Google Scholar] [CrossRef]
  43. Storm, T.; Zeitz, C.; Cases, O.; Amsellem, S.; Verroust, P.J.; Madsen, M.; Benoist, J.-F.; Passemard, S.; Lebon, S.; Jønsson, I.M.; et al. Detailed Investigations of Proximal Tubular Function in Imerslund-Gräsbeck Syndrome. BMC Med. Genet. 2013, 14, 111. [Google Scholar] [CrossRef]
  44. Ercan, Z.; Demir, M.E.; Ulas, T.; Ingec, M.; Horoz, M. A Long-Term Follow-up of an Imerslund-Grasbeck Syndrome Patient with Proteinuria. Nefrologia 2013, 33, 147–148. [Google Scholar] [CrossRef] [PubMed]
  45. Pannérec, A.; Migliavacca, E.; De Castro, A.; Michaud, J.; Karaz, S.; Goulet, L.; Rezzi, S.; Ng, T.P.; Bosco, N.; Larbi, A.; et al. Vitamin B12 Deficiency and Impaired Expression of Amnionless during Aging. J. Cachexia Sarcopenia Muscle 2018, 9, 41–52. [Google Scholar] [CrossRef]
  46. Cole, S.P.C. Multidrug Resistance Protein 1 (MRP1, ABCC1), a “Multitasking” ATP-Binding Cassette (ABC) Transporter. J. Biol. Chem. 2014, 289, 30880–30888. [Google Scholar] [CrossRef] [PubMed]
  47. Cole, S.P.C.; Bhardwaj, G.; Gerlach, J.H.; Mackie, J.E.; Grant, C.E.; Almquist, K.C.; Stewart, A.J.; Kurz, E.U.; Duncan, A.M.V.; Deeley, R.G. Overexpression of a Transporter Gene in a Multidrug-Resistant Human Lung Cancer Cell Line. Science 1992, 258, 1650–1654. [Google Scholar] [CrossRef] [PubMed]
  48. Bakos, É.; Homolya, L. Portrait of Multifaceted Transporter, the Multidrug Resistance-Associated Protein 1 (MRP1/ABCC1). Pflugers Arch. Eur. J. Physiol. 2007, 453, 621–641. [Google Scholar] [CrossRef]
  49. Flens, M.J.; Zaman, G.J.; van der Valk, P.; Izquierdo, M.A.; Schroeijers, A.B.; Scheffer, G.L.; van der Groep, P.; de Haas, M.; Meijer, C.J.; Scheper, R.J. Tissue Distribution of the Multidrug Resistance Protein. Am. J. Pathol. 1996, 148, 1237–1247. [Google Scholar]
  50. Peng, K.-C.; Cluzeaud, F.; Bens, M.; Duong Van Huyen, J.-P.; Wioland, M.A.; Lacave, R.; Vandewalle, A. Tissue and Cell Distribution of the Multidrug Resistance-Associated Protein (MRP) in Mouse Intestine and Kidney. J. Histochem. Cytochem. 1999, 47, 757–767. [Google Scholar] [CrossRef]
  51. Rajagopal, A.; Simon, S.M. Subcellular Localization and Activity of Multidrug Resistance Proteins. Mol. Biol. Cell 2003, 14, 3389–3399. [Google Scholar] [CrossRef]
  52. Beedholm-Ebsen, R.; van de Wetering, K.; Hardlei, T.; Nexø, E.; Borst, P.; Moestrup, S.K. Identification of Multidrug Resistance Protein 1 (MRP1/ABCC1) as a Molecular Gate for Cellular Export of Cobalamin. Blood 2010, 115, 1632–1639. [Google Scholar] [CrossRef]
  53. Seetharam, B.; Yammani, R.R. Cobalamin Transport Proteins and Their Cell-Surface Receptors. Expert Rev. Mol. Med. 2003, 5, 1–18. [Google Scholar] [CrossRef]
  54. Shah, N.P.; Beech, C.M.; Sturm, A.C.; Tanner, S.M. Investigation of the ABC Transporter MRP1 in Selected Patients with Presumed Defects in Vitamin B12 Absorption. Blood 2011, 117, 4397–4398. [Google Scholar] [CrossRef] [PubMed]
  55. Quadros, E.V.; Rothenberg, S.P.; Pan, Y.C.; Stein, S. Purification and Molecular Characterization of Human Transcobalamin II. J. Biol. Chem. 1986, 261, 15455–15460. [Google Scholar] [CrossRef] [PubMed]
  56. Quadros, E.V.; Rothenberg, S.P.; Jaffe, E.A. Endothelial Cells from Human Umbilical Vein Secrete Functional Transcobalamin II. Am. J. Physiol. Cell Physiol. 1989, 256, C296–C303. [Google Scholar] [CrossRef] [PubMed]
  57. Quadros, E.V.; Regec, A.L.; Khan, K.M.F.; Quadros, E.; Rothenberg, S.P. Transcobalamin II Synthesized in the Intestinal Villi Facilitates Transfer of Cobalamin to the Portal Blood. Am. J. Physiol. Gastrointest. Liver Physiol. 1999, 277, G161–G166. [Google Scholar] [CrossRef] [PubMed]
  58. Li, N.; Seetharam, S.; Seetharam, B. Genomic Structure of Human Transcobalamin II: Comparison to Human Intrinsic Factor and Transcobalamin I. Biochem. Biophys. Res. Commun. 1995, 208, 756–764. [Google Scholar] [CrossRef] [PubMed]
  59. Wuerges, J.; Garau, G.; Geremia, S.; Fedosov, S.N.; Petersen, T.E.; Randaccio, L. Structural Basis for Mammalian Vitamin B12 Transport by Transcobalamin. Proc. Natl. Acad. Sci. USA 2006, 103, 4386–4391. [Google Scholar] [CrossRef] [PubMed]
  60. Obeid, R.; Morkbak, A.L.; Munz, W.; Nexo, E.; Herrmann, W. The Cobalamin-Binding Proteins Transcobalamin and Haptocorrin in Maternal and Cord Blood Sera at Birth. Clin. Chem. 2006, 52, 263–269. [Google Scholar] [CrossRef] [PubMed]
  61. Thorpe, S.J.; Rigsby, P.; Roberts, G.; Lee, A.; Hamilton, M.; Craig, D. An International Standard for Holotranscobalamin (HoloTC): International Collaborative Study to Assign a HoloTC Value to the International Standard for Vitamin B12 and Serum Folate. Clin. Chem. Lab. Med. CCLM 2016, 54, 1467–1472. [Google Scholar] [CrossRef] [PubMed]
  62. Knoepfel, C.; Michel Blanco, M.; Nydegger, U.; Risch, L.; Renz, H.; Risch, M. Failure of the Holotranscobalamin Assay in Vitamin B12-Deficient Patients. J. Lab. Med. 2018, 42, 141–147. [Google Scholar] [CrossRef]
  63. Golding, P.H. Experimental Vitamin B12 Deficiency in a Human Subject: A Longitudinal Investigation of the Performance of the Holotranscobalamin (HoloTC, Active-B12) Immunoassay. SpringerPlus 2016, 5, 184. [Google Scholar] [CrossRef]
  64. Golding, P.H. Holotranscobalamin (HoloTC, Active-B12) and Herbert’s Model for the Development of Vitamin B12 Deficiency: A Review and Alternative Hypothesis. SpringerPlus 2016, 5, 668. [Google Scholar] [CrossRef] [PubMed]
  65. Trakadis, Y.J.; Alfares, A.; Bodamer, O.A.; Buyukavci, M.; Christodoulou, J.; Connor, P.; Glamuzina, E.; Gonzalez-Fernandez, F.; Bibi, H.; Echenne, B.; et al. Update on Transcobalamin Deficiency: Clinical Presentation, Treatment and Outcome. J. Inherit. Metab. Dis. 2014, 37, 461–473. [Google Scholar] [CrossRef]
  66. Nissen, P.H.; Nordwall, M.; Hoffmann-Lücke, E.; Sorensen, B.S.; Nexo, E. Transcobalamin Deficiency Caused by Compound Heterozygosity for Two Novel Mutations in the TCN2 Gene: A Study of Two Affected Siblings, Their Brother, and Their Parents. J. Inherit. Metab. Dis. 2010, 33, 269–274. [Google Scholar] [CrossRef] [PubMed]
  67. Oshi, M.A.M.; Alfaifi, J.; Alqahtani, Y.A.M.; Aljabri, M.F.; Kamal, N.M.; Althopaity, J.; Althobaiti, K.A.; Almalki, A.M.; Abosabie, S.A.S.; Abosabie, S.A.; et al. Progressive Myoclonic Ataxia and Developmental/Epileptic Encephalopathy Associated with a Novel Homozygous Mutation in TCN2 Gene. Mol. Genet. Genom. Med. 2024, 12, e2282. [Google Scholar] [CrossRef]
  68. Quadros, E.V. Advances in the Understanding of Cobalamin Assimilation and Metabolism. Br. J. Haematol. 2010, 148, 195–204. [Google Scholar] [CrossRef] [PubMed]
  69. Cho, W.-J.; Choi, J.-S.; Park, C.-H.; Yoon, S.-O.; Jeoung, D.-I.; Kim, Y.-M.; Choe, J.-S. Expression of CD320 in Human B Cells in Addition to Follicular Dendritic Cells. BMB Rep. 2008, 41, 863–867. [Google Scholar] [CrossRef] [PubMed]
  70. Jiang, W.; Nakayama, Y.; Sequeira, J.M.; Quadros, E.V. Mapping the Functional Domains of TCblR/CD320, the Receptor for Cellular Uptake of Transcobalamin-bound Cobalamin. FASEB J. 2013, 27, 2988–2994. [Google Scholar] [CrossRef] [PubMed]
  71. Arendt, J.F.B.; Quadros, E.V.; Nexo, E. Soluble Transcobalamin Receptor, SCD320, Is Present in Human Serum and Relates to Serum Cobalamin—Establishment and Validation of an ELISA. Clin. Chem. Lab. Med. 2012, 50, 515–519. [Google Scholar] [CrossRef]
  72. Youngdahl-Turner, P.; Mellman, I.S.; Allen, R.H.; Rosenberg, L.E. Protein Mediated Vitamin Uptake. Exp. Cell. Res. 1979, 118, 127–134. [Google Scholar] [CrossRef]
  73. Alam, A.; Woo, J.-S.; Schmitz, J.; Prinz, B.; Root, K.; Chen, F.; Bloch, J.S.; Zenobi, R.; Locher, K.P. Structural Basis of Transcobalamin Recognition by Human CD320 Receptor. Nat. Commun. 2016, 7, 12100. [Google Scholar] [CrossRef]
  74. Gick, G.G.; Arora, K.; Sequeira, J.M.; Nakayama, Y.; Lai, S.-C.; Quadros, E.V. Cellular Uptake of Vitamin B12: Role and Fate of TCblR/CD320, the Transcobalamin Receptor. Exp. Cell Res. 2020, 396, 112256. [Google Scholar] [CrossRef]
  75. Quadros, E.V.; Jacobsen, D.W. The Dynamics of Cobalamin Utilization in L-1210 Mouse Leukemia Cells: A Model of Cellular Cobalamin Metabolism. Biochim. Biophys. Acta BBA Gen. Subj. 1995, 1244, 395–403. [Google Scholar] [CrossRef]
  76. Bernard, D.J.; Pangilinan, F.J.; Cheng, J.; Molloy, A.M.; Brody, L.C. Mice Lacking the Transcobalamin-Vitamin B12 Receptor, CD320, Suffer from Anemia and Reproductive Deficits When Fed Vitamin B12-Deficient Diet. Hum. Mol. Genet. 2018, 27, 3627–3640. [Google Scholar] [CrossRef]
  77. Westerman, D.; Evans, D.; Metz, J. Neutrophil Hypersegmentation in Iron Deficiency Anaemia: A Case-Control Study. Br. J. Haematol. 1999, 107, 512–515. [Google Scholar] [CrossRef]
  78. Arora, K.; Sequeira, J.M.; Hernández, A.I.; Alarcon, J.M.; Quadros, E.V. Behavioral Alterations Are Associated with Vitamin B12 Deficiency in the Transcobalamin Receptor/CD320 KO Mouse. PLoS ONE 2017, 12, e0177156. [Google Scholar] [CrossRef]
  79. Fernàndez-Roig, S.; Lai, S.-C.; Murphy, M.M.; Fernandez-Ballart, J.; Quadros, E.V. Vitamin B12 Deficiency in the Brain Leads to DNA Hypomethylation in the TCblR/CD320 Knockout Mouse. Nutr. Metab. 2012, 9, 41. [Google Scholar] [CrossRef]
  80. Kurnat-Thoma, E.L.; Pangilinan, F.; Matteini, A.M.; Wong, B.; Pepper, G.A.; Stabler, S.P.; Guralnik, J.M.; Brody, L.C. Association of Transcobalamin II (TCN2) and Transcobalamin II-Receptor (TCblR) Genetic Variations with Cobalamin Deficiency Parameters in Elderly Women. Biol. Res. Nurs. 2015, 17, 444–454. [Google Scholar] [CrossRef]
  81. Quadros, E.V.; Lai, S.-C.; Nakayama, Y.; Sequeira, J.M.; Hannibal, L.; Wang, S.; Jacobsen, D.W.; Fedosov, S.; Wright, E.; Gallagher, R.C.; et al. Positive Newborn Screen for Methylmalonic Aciduria Identifies the First Mutation in TCblR/CD320, the Gene for Cellular Uptake of Transcobalamin-Bound Vitamin B12. Hum. Mutat. 2010, 31, 924–929. [Google Scholar] [CrossRef]
  82. Pangilinan, F.; Watkins, D.; Bernard, D.; Chen, Y.; Dong, N.; Wu, Q.; Ozel-Abaan, H.; Kaur, M.; Caggana, M.; Morrissey, M.; et al. Probing the Functional Consequence and Clinical Relevance of CD320 p.E88del, a Variant in the Transcobalamin Receptor Gene. Am. J. Med. Genet. Part A 2022, 188, 1124–1141. [Google Scholar] [CrossRef]
  83. Coelho, D.; Kim, J.C.; Miousse, I.R.; Fung, S.; du Moulin, M.; Buers, I.; Suormala, T.; Burda, P.; Frapolli, M.; Stucki, M.; et al. Mutations in ABCD4 Cause a New Inborn Error of Vitamin B12 Metabolism. Nat. Genet. 2012, 44, 1152–1155. [Google Scholar] [CrossRef]
  84. Seeger, M.A.; van Veen, H.W. Molecular Basis of Multidrug Transport by ABC Transporters. Biochim. Biophys. Acta BBA Proteins Proteom. 2009, 1794, 725–737. [Google Scholar] [CrossRef]
  85. Xu, D.; Feng, Z.; Hou, W.-T.; Jiang, Y.-L.; Wang, L.; Sun, L.; Zhou, C.-Z.; Chen, Y. Cryo-EM Structure of Human Lysosomal Cobalamin Exporter ABCD4. Cell Res. 2019, 29, 1039–1041. [Google Scholar] [CrossRef]
  86. Deme, J.C.; Hancock, M.A.; Xia, X.; Shintre, C.A.; Plesa, M.; Kim, J.C.; Carpenter, E.P.; Rosenblatt, D.S.; Coulton, J.W. Purification and Interaction Analyses of Two Human Lysosomal Vitamin B12 Transporters: LMBD1 and ABCD4. Mol. Membr. Biol. 2014, 31, 250–261. [Google Scholar] [CrossRef]
  87. Rutsch, F.; Gailus, S.; Miousse, I.R.; Suormala, T.; Sagné, C.; Toliat, M.R.; Nürnberg, G.; Wittkampf, T.; Buers, I.; Sharifi, A.; et al. Identification of a Putative Lysosomal Cobalamin Exporter Altered in the CblF Defect of Vitamin B12 Metabolism. Nat. Genet. 2009, 41, 234–239. [Google Scholar] [CrossRef]
  88. Kawaguchi, K.; Okamoto, T.; Morita, M.; Imanaka, T. Translocation of the ABC Transporter ABCD4 from the Endoplasmic Reticulum to Lysosomes Requires the Escort Protein LMBD1. Sci. Rep. 2016, 6, 30183. [Google Scholar] [CrossRef]
  89. Watkins, D.; Rosenblatt, D.S. Lessons in Biology from Patients with Inherited Disorders of Vitamin B12 and Folate Metabolism. Biochimie 2016, 126, 3–5. [Google Scholar] [CrossRef]
  90. Sloan, J.L.; Carrillo, N.; Adams, D.; Vendiitti, C.P. Disorders of Intracellular Cobalamin Metabolism. In GeneReviews; Adam, M.P., Feldman, J., Mirzaa, G.M., Pagon, R.A., Wallace, S.E., Bean, L.J.H., Gripp, K.W., Amemiya, A., Eds.; University of Washington: Seattle, WA, USA, 2021. [Google Scholar]
  91. Kim, J.C.; Lee, N.-C.; Hwu, P.W.-L.; Chien, Y.-H.; Fahiminiya, S.; Majewski, J.; Watkins, D.; Rosenblatt, D.S. Late Onset of Symptoms in an Atypical Patient with the CblJ Inborn Error of Vitamin B12 Metabolism: Diagnosis and Novel Mutation Revealed by Exome Sequencing. Mol. Genet. Metab. 2012, 107, 664–668. [Google Scholar] [CrossRef]
  92. Miousse, I.R.; Watkins, D.; Rosenblatt, D.S. Novel Splice Site Mutations and a Large Deletion in Three Patients with the CblF Inborn Error of Vitamin B12 Metabolism. Mol. Genet. Metab. 2011, 102, 505–507. [Google Scholar] [CrossRef]
  93. Alfadhel, M.; Lillquist, Y.P.; Davis, C.; Junker, A.K.; Stockler-Ipsiroglu, S. Eighteen-year Follow-up of a Patient with Cobalamin F Disease (CblF): Report and Review. Am. J. Med. Genet. Part A 2011, 155, 2571–2577. [Google Scholar] [CrossRef]
  94. Lerner-Ellis, J.P.; Anastasio, N.; Liu, J.; Coelho, D.; Suormala, T.; Stucki, M.; Loewy, A.D.; Gurd, S.; Grundberg, E.; Morel, C.F.; et al. Spectrum of Mutations in MMACHC, Allelic Expression, and Evidence for Genotype-Phenotype Correlations. Hum. Mutat. 2009, 30, 1072–1081. [Google Scholar] [CrossRef]
  95. Wang, C.; Li, D.; Cai, F.; Zhang, X.; Xu, X.; Liu, X.; Zhang, C.; Wang, D.; Liu, X.; Lin, S.; et al. Mutation Spectrum of MMACHC in Chinese Pediatric Patients with Cobalamin C Disease: A Case Series and Literature Review. Eur. J. Med. Genet. 2019, 62, 103713. [Google Scholar] [CrossRef] [PubMed]
  96. Lerner-Ellis, J.P.; Tirone, J.C.; Pawelek, P.D.; Doré, C.; Atkinson, J.L.; Watkins, D.; Morel, C.F.; Fujiwara, T.M.; Moras, E.; Hosack, A.R.; et al. Identification of the Gene Responsible for Methylmalonic Aciduria and Homocystinuria, CblC Type. Nat. Genet. 2006, 38, 93–100. [Google Scholar] [CrossRef] [PubMed]
  97. Guéant, J.-L.; Chéry, C.; Oussalah, A.; Nadaf, J.; Coelho, D.; Josse, T.; Flayac, J.; Robert, A.; Koscinski, I.; Gastin, I.; et al. A PRDX1 Mutant Allele Causes a MMACHC Secondary Epimutation in CblC Patients. Nat. Commun. 2018, 9, 67. [Google Scholar] [CrossRef] [PubMed]
  98. Kim, J.; Gherasim, C.; Banerjee, R. Decyanation of Vitamin B12 by a Trafficking Chaperone. Proc. Natl. Acad. Sci. USA 2008, 105, 14551–14554. [Google Scholar] [CrossRef] [PubMed]
  99. Pezacka, E.; Green, R.; Jacobsen, D.W. Glutathionylcobalamin as an Intermediate in the Formation of Cobalamin Coenzymes. Biochem. Biophys. Res. Commun. 1990, 169, 443–450. [Google Scholar] [CrossRef]
  100. Mellman, I.; Willard, H.F.; Youngdahl-Turner, P.; Rosenberg, L.E. Cobalamin Coenzyme Synthesis in Normal and Mutant Human Fibroblasts. Evidence for a Processing Enzyme Activity Deficient in CblC Cells. J. Biol. Chem. 1979, 254, 11847–11853. [Google Scholar] [CrossRef] [PubMed]
  101. Bassila, C.; Ghemrawi, R.; Flayac, J.; Froese, D.S.; Baumgartner, M.R.; Guéant, J.-L.; Coelho, D. Methionine Synthase and Methionine Synthase Reductase Interact with MMACHC and with MMADHC. Biochim. Biophys. Acta BBA Mol. Basis Dis. 2017, 1863, 103–112. [Google Scholar] [CrossRef]
  102. Kalantari, S.; Brezzi, B.; Bracciamà, V.; Barreca, A.; Nozza, P.; Vaisitti, T.; Amoroso, A.; Deaglio, S.; Manganaro, M.; Porta, F.; et al. Adult-Onset CblC Deficiency: A Challenging Diagnosis Involving Different Adult Clinical Specialists. Orphanet J. Rare Dis. 2022, 17, 33. [Google Scholar] [CrossRef]
  103. Sorin, M.; Watkins, D.; Gilfix, B.M.; Rosenblatt, D.S. Methionine Dependence in Tumor Cells: The Potential Role of Cobalamin and MMACHC. Mol. Genet. Metab. 2021, 132, 155–161. [Google Scholar] [CrossRef]
  104. Bauer, W.G.; Watkins, D.; Zacharias, C.; Gilfix, B.M.; Rosenblatt, D.S. Growth Requirement for Methionine in Human Melanoma-Derived Cell Lines with Different Levels of MMACHC Expression and Methylation. Mol. Genet. Metab. 2024, 141, 108111. [Google Scholar] [CrossRef]
  105. Coelho, D.; Suormala, T.; Stucki, M.; Lerner-Ellis, J.P.; Rosenblatt, D.S.; Newbold, R.F.; Baumgartner, M.R.; Fowler, B. Gene Identification for the CblD Defect of Vitamin B12 Metabolism. N. Engl. J. Med. 2008, 358, 1454–1464. [Google Scholar] [CrossRef]
  106. Li, Z.; Mascarenhas, R.; Twahir, U.T.; Kallon, A.; Deb, A.; Yaw, M.; Penner-Hahn, J.; Koutmos, M.; Warncke, K.; Banerjee, R. An Interprotein Co–S Coordination Complex in the B12-Trafficking Pathway. J. Am. Chem. Soc. 2020, 142, 16334–16345. [Google Scholar] [CrossRef]
  107. Pupavac, M.; Garcia, M.A.M.; Rosenblatt, D.S.; Jerome-Majewska, L.A. Expression of Mmachc and Mmadhc during Mouse Organogenesis. Mol. Genet. Metab. 2011, 103, 401–405. [Google Scholar] [CrossRef] [PubMed]
  108. Stucki, M.; Coelho, D.; Suormala, T.; Burda, P.; Fowler, B.; Baumgartner, M.R. Molecular Mechanisms Leading to Three Different Phenotypes in the CblD Defect of Intracellular Cobalamin Metabolism. Hum. Mol. Genet. 2012, 21, 1410–1418. [Google Scholar] [CrossRef] [PubMed]
  109. Plesa, M.; Kim, J.; Paquette, S.G.; Gagnon, H.; Ng-Thow-Hing, C.; Gibbs, B.F.; Hancock, M.A.; Rosenblatt, D.S.; Coulton, J.W. Interaction between MMACHC and MMADHC, Two Human Proteins Participating in Intracellular Vitamin B12 Metabolism. Mol. Genet. Metab. 2010, 102, 139–148. [Google Scholar] [CrossRef] [PubMed]
  110. Deme, J.C.; Miousse, I.R.; Plesa, M.; Kim, J.C.; Hancock, M.A.; Mah, W.; Rosenblatt, D.S.; Coulton, J.W. Structural Features of Recombinant MMADHC Isoforms and Their Interactions with MMACHC, Proteins of Mammalian Vitamin B12 Metabolism. Mol. Genet. Metab. 2012, 107, 352–362. [Google Scholar] [CrossRef] [PubMed]
  111. Froese, D.S.; Kopec, J.; Fitzpatrick, F.; Schuller, M.; McCorvie, T.J.; Chalk, R.; Plessl, T.; Fettelschoss, V.; Fowler, B.; Baumgartner, M.R.; et al. Structural Insights into the MMACHC-MMADHC Protein Complex Involved in Vitamin B12 Trafficking. J. Biol. Chem. 2015, 290, 29167–29177. [Google Scholar] [CrossRef]
  112. Suormala, T.; Baumgartner, M.R.; Coelho, D.; Zavadakova, P.; Kožich, V.; Koch, H.G.; Berghaüser, M.; Wraith, J.E.; Burlina, A.; Sewell, A.; et al. The CblD Defect Causes Either Isolated or Combined Deficiency of Methylcobalamin and Adenosylcobalamin Synthesis. J. Biol. Chem. 2004, 279, 42742–42749. [Google Scholar] [CrossRef]
  113. Parini, R.; Furlan, F.; Brambilla, A.; Codazzi, D.; Vedovati, S.; Corbetta, C.; Fedeli, T.; Merinero, B.; Pérez, B.; Ugarte, M. Severe Neonatal Metabolic Decompensation in Methylmalonic Acidemia Caused by CblD Defect. JIMD Rep. 2013, 11, 133–137. [Google Scholar]
  114. Swanson, D.A.; Liu, M.-L.; Baker, P.J.; Garrett, L.; Stitzel, M.; Wu, J.; Harris, M.; Banerjee, R.; Shane, B.; Brody, L.C. Targeted Disruption of the Methionine Synthase Gene in Mice. Mol. Cell. Biol. 2001, 21, 1058–1065. [Google Scholar] [CrossRef]
  115. Christensen, E.I.; Birn, H. Megalin and Cubilin: Synergistic Endocytic Receptors in Renal Proximal Tubule. Am. J. Physiol. Ren. Physiol. 2001, 280, F562–F573. [Google Scholar] [CrossRef]
  116. Menezo, Y.; Elder, K.; Clement, A.; Clement, P. Folic Acid, Folinic Acid, 5 Methyl TetraHydroFolate Supplementation for Mutations That Affect Epigenesis through the Folate and One-Carbon Cycles. Biomolecules 2022, 12, 197. [Google Scholar] [CrossRef] [PubMed]
  117. Dixon, M.M.; Huang, S.; Matthews, R.G.; Ludwig, M. The Structure of the C-Terminal Domain of Methionine Synthase: Presenting S-Adenosylmethionine for Reductive Methylation of B12. Structure 1996, 4, 1263–1275. [Google Scholar] [CrossRef] [PubMed]
  118. Chen, L.H.; Liu, M.L.; Hwang, H.Y.; Chen, L.S.; Korenberg, J.; Shane, B. Human Methionine Synthase. CDNA Cloning, Gene Localization, and Expression. J. Biol. Chem. 1997, 272, 3628–3634. [Google Scholar] [CrossRef] [PubMed]
  119. Drennan, C.L.; Matthews, R.G.; Ludwig, M.L. Cobalamin-Dependent Methionine Synthase: The Structure of a Methylcobalamin-Binding Fragment and Implications for Other B12-Dependent Enzymes. Curr. Opin. Struct. Biol. 1994, 4, 919–929. [Google Scholar] [CrossRef]
  120. Matthews, R.G.; Sheppard, C.; Goulding, C. Methylenetetrahydrofolate Reductase and Methionine Synthase: Biochemistry and Molecular Biology. Eur. J. Pediatr. 1998, 157, S54–S59. [Google Scholar] [CrossRef] [PubMed]
  121. Guéant, J.-L.; Caillerez-Fofou, M.; Battaglia-Hsu, S.; Alberto, J.-M.; Freund, J.-N.; Dulluc, I.; Adjalla, C.; Maury, F.; Merle, C.; Nicolas, J.-P.; et al. Molecular and Cellular Effects of Vitamin B12 in Brain, Myocardium and Liver through Its Role as Co-Factor of Methionine Synthase. Biochimie 2013, 95, 1033–1040. [Google Scholar] [CrossRef] [PubMed]
  122. Gulati, S.; Baker, P.; Li, Y.N.; Fowler, B.; Kruger, W.; Brody, L.C.; Banerjee, R. Defects in Human Methionine Synthase in CblG Patients. Hum. Mol. Genet. 1996, 5, 1859–1865. [Google Scholar] [CrossRef] [PubMed]
  123. Watkin, D.; Rosenblatt, D.S. Functional Methionine Synthase Deficiency (CblE and CblG): Clinical and Biochemical Heterogeneity. Am. J. Med. Genet. 1989, 34, 427–434. [Google Scholar] [CrossRef]
  124. Watkins, D.; Ru, M.; Hwang, H.-Y.; Kim, C.D.; Murray, A.; Philip, N.S.; Kim, W.; Legakis, H.; Wai, T.; Hilton, J.F.; et al. Hyperhomocysteinemia Due to Methionine Synthase Deficiency, CblG: Structure of the MTR Gene, Genotype Diversity, and Recognition of a Common Mutation, P1173L. Am. J. Hum. Genet. 2002, 71, 143–153. [Google Scholar] [CrossRef]
  125. Guéant, J.-L.; Guéant-Rodriguez, R.-M.; Kosgei, V.J.; Coelho, D. Causes and Consequences of Impaired Methionine Synthase Activity in Acquired and Inherited Disorders of Vitamin B12 Metabolism. Crit. Rev. Biochem. Mol. Biol. 2022, 57, 133–155. [Google Scholar] [CrossRef] [PubMed]
  126. Froese, D.S.; Kochan, G.; Muniz, J.R.C.; Wu, X.; Gileadi, C.; Ugochukwu, E.; Krysztofinska, E.; Gravel, R.A.; Oppermann, U.; Yue, W.W. Structures of the Human GTPase MMAA and Vitamin B12-Dependent Methylmalonyl-CoA Mutase and Insight into Their Complex Formation. J. Biol. Chem. 2010, 285, 38204–38213. [Google Scholar] [CrossRef] [PubMed]
  127. Takahashi-Iñiguez, T.; García-Hernandez, E.; Arreguín-Espinosa, R.; Flores, M.E. Role of Vitamin B12 on Methylmalonyl-CoA Mutase Activity. J. Zhejiang Univ. Sci. B 2012, 13, 423–437. [Google Scholar] [CrossRef] [PubMed]
  128. Heuberger, K.; Bailey, H.J.; Burda, P.; Chaikuad, A.; Krysztofinska, E.; Suormala, T.; Bürer, C.; Lutz, S.; Fowler, B.; Froese, D.S.; et al. Genetic, Structural, and Functional Analysis of Pathogenic Variations Causing Methylmalonyl-CoA Epimerase Deficiency. Biochim. Biophys. Acta BBA Mol. Basis Dis. 2019, 1865, 1265–1272. [Google Scholar] [CrossRef] [PubMed]
  129. Gouda, H.; Mascarenhas, R.; Pillay, S.; Ruetz, M.; Koutmos, M.; Banerjee, R. Patient Mutations in Human ATP:Cob(I)Alamin Adenosyltransferase Differentially Affect Its Catalytic versus Chaperone Functions. J. Biol. Chem. 2021, 297, 101373. [Google Scholar] [CrossRef] [PubMed]
  130. Pratt, J.M. 988. The Chemistry of Vitamin B12. Part II. Photochemical Reactions. J. Chem. Soc. 1964, 5154–5160. [Google Scholar] [CrossRef]
  131. Takahashi-Iñiguez, T.; González-Noriega, A.; Michalak, C.; Flores, M.E. Human MMAA Induces the Release of Inactive Cofactor and Restores Methylmalonyl-CoA Mutase Activity through Their Complex Formation. Biochimie 2017, 142, 191–196. [Google Scholar] [CrossRef] [PubMed]
  132. Liu, M.-Y.; Liu, T.-T.; Yang, Y.-L.; Chang, Y.-C.; Fan, Y.-L.; Lee, S.-F.; Teng, Y.-T.; Chiang, S.-H.; Niu, D.-M.; Lin, S.-J.; et al. Mutation Profile of the MUT Gene in Chinese Methylmalonic Aciduria Patients. JIMD Rep. Case Res. Rep. 2012, 6, 55–64. [Google Scholar]
  133. Acquaviva, C.; Benoist, J.; Pereira, S.; Callebaut, I.; Koskas, T.; Porquet, D.; Elion, J. Molecular Basis of Methylmalonyl-CoA Mutase Apoenzyme Defect in 40 European Patients Affected by Mut° and Mut-Forms of Methylmalonic Acidemia: Identification of 29 Novel Mutations in the MUT Gene. Hum. Mutat. 2005, 25, 167–176. [Google Scholar] [CrossRef]
  134. Willard, H.F.; Rosenberg, L.E. Inherited Methylmalonyl CoA Mutase Apoenzyme Deficiency in Human Fibroblasts. J. Clin. Investig. 1980, 65, 690–698. [Google Scholar] [CrossRef]
  135. Mineva, E.M.; Sternberg, M.R.; Bailey, R.L.; Storandt, R.J.; Pfeiffer, C.M. Fewer US Adults Had Low or Transitional Vitamin B12 Status Based on the Novel Combined Indicator of Vitamin B12 Status Compared with Individual, Conventional Markers, NHANES 1999–2004. Am. J. Clin. Nutr. 2021, 114, 1070–1079. [Google Scholar] [CrossRef] [PubMed]
  136. Bailey, R.L.; Carmel, R.; Green, R.; Pfeiffer, C.M.; Cogswell, M.E.; Osterloh, J.D.; Sempos, C.T.; Yetley, E.A. Monitoring of Vitamin B-12 Nutritional Status in the United States by Using Plasma Methylmalonic Acid and Serum Vitamin B-12. Am. J. Clin. Nutr. 2011, 94, 552–561. [Google Scholar] [CrossRef] [PubMed]
  137. Zhou, Y.; Wang, A.; Yeung, L.F.; Qi, Y.P.; Pfeiffer, C.M.; Crider, K.S. Folate and Vitamin B12 Usual Intake and Biomarker Status by Intake Source in United States Adults Aged ≥19 y: NHANES 2007–2018. Am. J. Clin. Nutr. 2023, 118, 241–254. [Google Scholar] [CrossRef] [PubMed]
  138. Obeid, R.; Murphy, M.; Solé-Navais, P.; Yajnik, C. Cobalamin Status from Pregnancy to Early Childhood: Lessons from Global Experience. Adv. Nutr. 2017, 8, 971–979. [Google Scholar] [CrossRef] [PubMed]
  139. Röhrig, G.; Gütgemann, I.; Kolb, G.; Leischker, A. Klinisch-Hämatologisches Bild Des Vitamin-B12-Mangels Im Alter. Z. Gerontol. Geriatr. 2018, 51, 446–452. [Google Scholar] [CrossRef] [PubMed]
  140. Devi, A.; Rush, E.; Harper, M.; Venn, B. Vitamin B12 Status of Various Ethnic Groups Living in New Zealand: An Analysis of the Adult Nutrition Survey 2008/2009. Nutrients 2018, 10, 181. [Google Scholar] [CrossRef]
  141. Jajoo, S.S.; Zamwar, U.M.; Nagrale, P. Etiology, Clinical Manifestations, Diagnosis, and Treatment of Cobalamin (Vitamin B12) Deficiency. Cureus 2024, 16, e52153. [Google Scholar] [CrossRef]
  142. Dali-Youcef, N.; Andres, E. An Update on Cobalamin Deficiency in Adults. QJM Int. J. Med. 2009, 102, 17–28. [Google Scholar] [CrossRef]
  143. Andrès, E.; Loukili, N.H.; Noel, E.; Kaltenbach, G.; Abdelgheni, M.B.; Perrin, A.E.; Noblet-Dick, M.; Maloisel, F.; Schlienger, J.-L.; Blicklé, J.-F. Vitamin B12 (Cobalamin) Deficiency in Elderly Patients. Can. Med. Assoc. J. 2004, 171, 251–259. [Google Scholar] [CrossRef]
  144. Lam, J.R.; Schneider, J.L.; Zhao, W.; Corley, D.A. Proton Pump Inhibitor and Histamine 2 Receptor Antagonist Use and Vitamin B12 Deficiency. JAMA 2013, 310, 2435–2442. [Google Scholar] [CrossRef]
  145. Green, R. Vitamin B12 Deficiency from the Perspective of a Practicing Hematologist. Blood 2017, 129, 2603–2611. [Google Scholar] [CrossRef] [PubMed]
  146. Lindenbaum, J.; Healton, E.B.; Savage, D.G.; Brust, J.C.M.; Garrett, T.J.; Podell, E.R.; Margell, P.D.; Stabler, S.P.; Allen, R.H. Neuropsychiatric Disorders Caused by Cobalamin Deficiency in the Absence of Anemia or Macrocytosis. N. Engl. J. Med. 1988, 318, 1720–1728. [Google Scholar] [CrossRef]
  147. Miller, A.; Korem, M.; Almog, R.; Galboiz, Y. Vitamin B12, Demyelination, Remyelination and Repair in Multiple Sclerosis. J. Neurol. Sci. 2005, 233, 93–97. [Google Scholar] [CrossRef] [PubMed]
  148. Reynolds, E.H.; Linnell, J.C. Vitamin B12 Deficiency, Demyelination, and Multiple Sclerosis. Lancet 1987, 330, 920. [Google Scholar] [CrossRef] [PubMed]
  149. Troen, A.M.; Shea-Budgell, M.; Shukitt-Hale, B.; Smith, D.E.; Selhub, J.; Rosenberg, I.H. B-Vitamin Deficiency Causes Hyperhomocysteinemia and Vascular Cognitive Impairment in Mice. Proc. Natl. Acad. Sci. USA 2008, 105, 12474–12479. [Google Scholar] [CrossRef]
  150. Carmel, R. Diagnosis and Management of Clinical and Subclinical Cobalamin Deficiencies: Why Controversies Persist in the Age of Sensitive Metabolic Testing. Biochimie 2013, 95, 1047–1055. [Google Scholar] [CrossRef]
  151. Kulnigg-Dabsch, S. Autoimmune Gastritis. Wien. Med. Wochenschr. 2016, 166, 424–430. [Google Scholar] [CrossRef]
  152. Lahner, E.; Annibale, B. Pernicious Anemia: New Insights from a Gastroenterological Point of View. World J. Gastroenterol. 2009, 15, 5121–5128. [Google Scholar] [CrossRef]
  153. Carmel, R. Subclinical Cobalamin Deficiency. Curr. Opin. Gastroenterol. 2012, 28, 151–158. [Google Scholar] [CrossRef]
  154. Devalia, V.; Hamilton, M.S.; Molloy, A.M.; on Behalf of the British Committee for Standards in Haematology. Guidelines for the Diagnosis and Treatment of Cobalamin and Folate Disorders. Br. J. Haematol. 2014, 166, 496–513. [Google Scholar] [CrossRef] [PubMed]
  155. Carmel, R.; Agrawal, Y.P. Failures of Cobalamin Assays in Pernicious Anemia. N. Engl. J. Med. 2012, 367, 385–386. [Google Scholar] [CrossRef]
  156. Miller, J.W.; Garrod, M.G.; Rockwood, A.L.; Kushnir, M.M.; Allen, L.H.; Haan, M.N.; Green, R. Measurement of Total Vitamin B12 and Holotranscobalamin, Singly and in Combination, in Screening for Metabolic Vitamin B12 Deficiency. Clin. Chem. 2006, 52, 278–285. [Google Scholar] [CrossRef] [PubMed]
  157. Nexo, E.; Hoffmann-Lücke, E. Holotranscobalamin, a Marker of Vitamin B-12 Status: Analytical Aspects and Clinical Utility. Am. J. Clin. Nutr. 2011, 94, 359S–365S. [Google Scholar] [CrossRef] [PubMed]
  158. Molloy, A.M.; Pangilinan, F.; Mills, J.L.; Shane, B.; O’Neill, M.B.; McGaughey, D.M.; Velkova, A.; Abaan, H.O.; Ueland, P.M.; McNulty, H.; et al. A Common Polymorphism in HIBCH Influences Methylmalonic Acid Concentrations in Blood Independently of Cobalamin. Am. J. Hum. Genet. 2016, 98, 869–882. [Google Scholar] [CrossRef] [PubMed]
  159. Allen, L.H.; Miller, J.W.; de Groot, L.; Rosenberg, I.H.; Smith, A.D.; Refsum, H.; Raiten, D.J. Biomarkers of Nutrition for Development (BOND): Vitamin B-12 Review. J. Nutr. 2018, 148, 1995S–2027S. [Google Scholar] [CrossRef] [PubMed]
  160. Fedosov, S.N.; Brito, A.; Miller, J.W.; Green, R.; Allen, L.H. Combined Indicator of Vitamin B12 Status: Modification for Missing Biomarkers and Folate Status and Recommendations for Revised Cut-Points. Clin. Chem. Lab. Med. CCLM 2015, 53, 1215–1225. [Google Scholar] [CrossRef] [PubMed]
  161. Huemer, M.; Diodato, D.; Schwahn, B.; Schiff, M.; Bandeira, A.; Benoist, J.; Burlina, A.; Cerone, R.; Couce, M.L.; Garcia-Cazorla, A.; et al. Guidelines for Diagnosis and Management of the Cobalamin-related Remethylation Disorders CblC, CblD, CblE, CblF, CblG, CblJ and MTHFR Deficiency. J. Inherit. Metab. Dis. 2017, 40, 21–48. [Google Scholar] [CrossRef] [PubMed]
  162. Shen, W.; Gao, C.; Cueto, R.; Liu, L.; Fu, H.; Shao, Y.; Yang, W.Y.; Fang, P.; Choi, E.T.; Wu, Q.; et al. Homocysteine-Methionine Cycle Is a Metabolic Sensor System Controlling Methylation-Regulated Pathological Signaling. Redox Biol. 2020, 28, 101322. [Google Scholar] [CrossRef] [PubMed]
  163. Fowler, B. Homocysteine: Overview of Biochemistry, Molecular Biology, and Role in Disease Processes. Semin. Vasc. Med. 2005, 5, 77–86. [Google Scholar] [CrossRef]
  164. Hvas, A.-M.; Morkbak, A.L.; Hardlei, T.F.; Nexo, E. The Vitamin B12 Absorption Test, CobaSorb, Identifies Patients Not Requiring Vitamin B12 Injection Therapy. Scand. J. Clin. Lab. Investig. 2011, 71, 432–438. [Google Scholar] [CrossRef]
  165. Bor, M.V.; Çetin, M.; Aytaç, S.; Altay, C.; Nexo, E. Nonradioactive Vitamin B12 Absorption Test Evaluated in Controls and in Patients with Inherited Malabsorption of Vitamin B12. Clin. Chem. 2005, 51, 2151–2155. [Google Scholar] [CrossRef] [PubMed]
  166. Hvas, A.-M.; Morkbak, A.L.; Nexo, E. Plasma Holotranscobalamin Compared with Plasma Cobalamins for Assessment of Vitamin B12 Absorption; Optimisation of a Non-Radioactive Vitamin B12 Absorption Test (CobaSorb). Clin. Chim. Acta 2007, 376, 150–154. [Google Scholar] [CrossRef] [PubMed]
  167. Kurpad, A.V.; Pasanna, R.M.; Hegde, S.G.; Patil, M.; Mukhopadhyay, A.; Sachdev, H.S.; Bhat, K.G.; Sivadas, A.; Devi, S. Bioavailability and Daily Requirement of Vitamin B12 in Adult Humans: An Observational Study of Its Colonic Absorption and Daily Excretion as Measured by [13C]-Cyanocobalamin Kinetics. Am. J. Clin. Nutr. 2023, 118, 1214–1223. [Google Scholar] [CrossRef] [PubMed]
  168. Garrod, M.G.; Rossow, H.A.; Calvert, C.C.; Miller, J.W.; Green, R.; Buchholz, B.A.; Allen, L.H. 14C-Cobalamin Absorption from Endogenously Labeled Chicken Eggs Assessed in Humans Using Accelerator Mass Spectrometry. Nutrients 2019, 11, 2148. [Google Scholar] [CrossRef] [PubMed]
  169. Stabler, S.P. Vitamin B12 Deficiency. N. Engl. J. Med. 2013, 368, 2040–2042. [Google Scholar] [CrossRef] [PubMed]
  170. Toh, B.-H. Diagnosis and Classification of Autoimmune Gastritis. Autoimmun. Rev. 2014, 13, 459–462. [Google Scholar] [CrossRef] [PubMed]
  171. Lahner, E.; Norman, G.L.; Severi, C.; Encabo, S.; Shums, Z.; Vannella, L.; Fave, G.D.; Annibale, B. Reassessment of Intrinsic Factor and Parietal Cell Autoantibodies in Atrophic Gastritis With Respect to Cobalamin Deficiency. Am. J. Gastroenterol. 2009, 104, 2071–2079. [Google Scholar] [CrossRef] [PubMed]
  172. Froese, D.S.; Gravel, R.A. Genetic Disorders of Vitamin B 12 Metabolism: Eight Complementation Groups—Eight Genes. Expert Rev. Mol. Med. 2010, 12, e37. [Google Scholar] [CrossRef] [PubMed]
  173. British Society for Haematology BSH Guidance on B12 Replacement during the COVID-19 Pandemic. 2020. Available online: https://b-s-h.org.uk/media/18275/bsh-guidance-b12-replacement-covid-1901052020finalv.pdf (accessed on 1 June 2024).
  174. Gomes, A.P.; Ilter, D.; Low, V.; Endress, J.E.; Fernández-García, J.; Rosenzweig, A.; Schild, T.; Broekaert, D.; Ahmed, A.; Planque, M.; et al. Age-Induced Accumulation of Methylmalonic Acid Promotes Tumour Progression. Nature 2020, 585, 283–287. [Google Scholar] [CrossRef]
  175. Berkins, S.; Schiöth, H.B.; Rukh, G. Depression and Vegetarians: Association between Dietary Vitamin B6, B12 and Folate Intake and Global and Subcortical Brain Volumes. Nutrients 2021, 13, 1790. [Google Scholar] [CrossRef]
  176. Sangle, P.; Sandhu, O.; Aftab, Z.; Anthony, A.T.; Khan, S. Vitamin B12 Supplementation: Preventing Onset and Improving Prognosis of Depression. Cureus 2020, 12, e11169. [Google Scholar] [CrossRef] [PubMed]
  177. Frigo, M.; Cogo, M.G.; Fusco, M.L.; Gardinetti, M.; Frigeni, B. Glutamate and Multiple Sclerosis. Curr. Med. Chem. 2012, 19, 1295–1299. [Google Scholar] [CrossRef] [PubMed]
  178. Mathew, A.R.; Di Matteo, G.; La Rosa, P.; Barbati, S.A.; Mannina, L.; Moreno, S.; Tata, A.M.; Cavallucci, V.; Fidaleo, M. Vitamin B12 Deficiency and the Nervous System: Beyond Metabolic Decompensation—Comparing Biological Models and Gaining New Insights into Molecular and Cellular Mechanisms. Int. J. Mol. Sci. 2024, 25, 590. [Google Scholar] [CrossRef]
  179. Alruwaili, M.; Basri, R.; AlRuwaili, R.; Albarrak, A.M.; Ali, N.H. Neurological Implications of Vitamin B12 Deficiency in Diet: A Systematic Review and Meta-Analysis. Healthcare 2023, 11, 958. [Google Scholar] [CrossRef] [PubMed]
  180. Song, Y.; Quan, M.; Li, T.; Jia, J. Serum Homocysteine, Vitamin B12, Folate, and Their Association with Mild Cognitive Impairment and Subtypes of Dementia. J. Alzheimers Dis. 2022, 90, 681–691. [Google Scholar] [CrossRef] [PubMed]
  181. Moore, E.; Mander, A.; Ames, D.; Carne, R.; Sanders, K.; Watters, D. Cognitive Impairment and Vitamin B12: A Review. Int. Psychogeriatr. 2012, 24, 541–556. [Google Scholar] [CrossRef]
  182. Calderón-Ospina, C.A.; Nava-Mesa, M.O. B Vitamins in the Nervous System: Current Knowledge of the Biochemical Modes of Action and Synergies of Thiamine, Pyridoxine, and Cobalamin. CNS Neurosci. Ther. 2020, 26, 5–13. [Google Scholar] [CrossRef] [PubMed]
  183. Ganguly, P.; Alam, S.F. Role of Homocysteine in the Development of Cardiovascular Disease. Nutr. J. 2015, 14, 6. [Google Scholar] [CrossRef]
  184. Coppola, A.; Davi, G.; De Stefano, V.; Mancini, F.P.; Cerbone, A.M.; Di Minno, G. Homocysteine, Coagulation, Platelet Function, And Thrombosis. Semin. Thromb. Hemost. 2000, 26, 243–254. [Google Scholar] [CrossRef]
  185. Ammouri, W.; Tazi, Z.M.; Harmouche, H.; Maamar, M.; Adnaoui, M. Venous Thromboembolism and Hyperhomocysteinemia as First Manifestation of Pernicious Anemia: A Case Series. J. Med. Case Rep. 2017, 11, 250. [Google Scholar] [CrossRef]
  186. Azzini, E.; Raguzzini, A.; Polito, A. A Brief Review on Vitamin B12 Deficiency Looking at Some Case Study Reports in Adults. Int. J. Mol. Sci. 2021, 22, 9694. [Google Scholar] [CrossRef] [PubMed]
  187. Ospina-Romero, M.; Cannegieter, S.C.; den Heijer, M.; Doggen, C.J.M.; Rosendaal, F.R.; Lijfering, W.M. Hyperhomocysteinemia and Risk of First Venous Thrombosis: The Influence of (Unmeasured) Confounding Factors. Am. J. Epidemiol. 2018, 187, 1392–1400. [Google Scholar] [CrossRef]
  188. Çoker, M.; de Klerk, J.B.C.; Poll-The, B.T.; Huijmans, J.G.M.; Duran, M. Plasma Total Odd-chain Fatty Acids in the Monitoring of Disorders of Propionate, Methylmalonate and Biotin Metabolism. J. Inherit. Metab. Dis. 1996, 19, 743–751. [Google Scholar] [CrossRef] [PubMed]
  189. Gomes, A.P.; Ilter, D.; Low, V.; Drapela, S.; Schild, T.; Mullarky, E.; Han, J.; Elia, I.; Broekaert, D.; Rosenzweig, A.; et al. Altered Propionate Metabolism Contributes to Tumour Progression and Aggressiveness. Nat. Metab. 2022, 4, 435–443. [Google Scholar] [CrossRef] [PubMed]
  190. Hu, C.; Ye, M.; Bai, J.; Liu, P.; Lu, F.; Chen, J.; Yu, P.; Chen, T.; Shi, X.; Tang, Q. Methylmalonic Acid Promotes Colorectal Cancer Progression via Activation of Wnt/β-Catenin Pathway Mediated Epithelial–Mesenchymal Transition. Cancer Cell Int. 2023, 23, 131. [Google Scholar] [CrossRef]
  191. Loedin, A.K.; Speijer, D. Is There a Carcinogenic Risk Attached to Vitamin B12 Deficient Diets and What Should We Do About It? Reviewing the Facts. Mol. Nutr. Food Res. 2021, 65, 2000945. [Google Scholar] [CrossRef]
  192. Obeid, R. High Plasma Vitamin B12 and Cancer in Human Studies: A Scoping Review to Judge Causality and Alternative Explanations. Nutrients 2022, 14, 4476. [Google Scholar] [CrossRef]
  193. Haghighat, G.; Khajeh-Mehrizi, A.; Ranjbar, H. Evaluation of Serum Vitamin B12 Levels in Patients with Colon and Breast Cancer: A Case-Control Study. Int. J. Hematol. Oncol. Stem Cell Res. 2023, 17, 240–244. [Google Scholar] [CrossRef] [PubMed]
  194. Zhang, S.M.; Willett, W.C.; Selhub, J.; Hunter, D.J.; Giovannucci, E.L.; Holmes, M.D.; Colditz, G.A.; Hankinson, S.E. Plasma Folate, Vitamin B6, Vitamin B12, Homocysteine, and Risk of Breast Cancer. JNCI J. Natl. Cancer Inst. 2003, 95, 373–380. [Google Scholar] [CrossRef]
  195. Wu, W.; Kang, S.; Zhang, D. Association of Vitamin B6, Vitamin B12 and Methionine with Risk of Breast Cancer: A Dose–Response Meta-Analysis. Br. J. Cancer 2013, 109, 1926–1944. [Google Scholar] [CrossRef]
  196. Matejcic, M.; de Batlle, J.; Ricci, C.; Biessy, C.; Perrier, F.; Huybrechts, I.; Weiderpass, E.; Boutron-Ruault, M.C.; Cadeau, C.; His, M.; et al. Biomarkers of Folate and Vitamin B12 and Breast Cancer Risk: Report from the EPIC Cohort. Int. J. Cancer 2017, 140, 1246–1259. [Google Scholar] [CrossRef] [PubMed]
  197. Tseng, L.T.-L.; Lin, C.-L.; Pan, K.-H.; Tzen, K.-Y.; Su, M.-J.; Tsai, C.-T.; Li, Y.-H.; Li, P.-C.; Chiang, F.-T.; Chang, S.C.; et al. Single Allele Lmbrd1 Knockout Results in Cardiac Hypertrophy. J. Formos. Med. Assoc. 2018, 117, 471–479. [Google Scholar] [CrossRef]
  198. Barth, A.S.; Kuner, R.; Buness, A.; Ruschhaupt, M.; Merk, S.; Zwermann, L.; Kääb, S.; Kreuzer, E.; Steinbeck, G.; Mansmann, U.; et al. Identification of a Common Gene Expression Signature in Dilated Cardiomyopathy Across Independent Microarray Studies. J. Am. Coll. Cardiol. 2006, 48, 1610–1617. [Google Scholar] [CrossRef]
  199. Jonnalagadda, D.; Kihara, Y.; Groves, A.; Ray, M.; Saha, A.; Ellington, C.; Lee-Okada, H.-C.; Furihata, T.; Yokomizo, T.; Quadros, E.V.; et al. FTY720 Requires Vitamin B12-TCN2-CD320 Signaling in Astrocytes to Reduce Disease in an Animal Model of Multiple Sclerosis. Cell Rep. 2023, 42, 113545. [Google Scholar] [CrossRef]
  200. Munoz, M.; Henderson, M.; Haber, M.; Norris, M. Role of the MRP1/ABCC1 Multidrug Transporter Protein in Cancer. IUBMB Life 2007, 59, 752–757. [Google Scholar] [CrossRef] [PubMed]
  201. Li, F.; Mao, C.; Yeh, S.; Xin, J.; Wang, P.; Shi, Q.; Ming, X. Combinatory Therapy of MRP1-Targeted Photoimmunotherapy and Liposomal Doxorubicin Promotes the Antitumor Effect for Chemoresistant Small Cell Lung Cancer. Int. J. Pharm. 2022, 625, 122076. [Google Scholar] [CrossRef] [PubMed]
  202. Pietz, H.L.; Abbas, A.; Johnson, Z.L.; Oldham, M.L.; Suga, H.; Chen, J. A Macrocyclic Peptide Inhibitor Traps MRP1 in a Catalytically Incompetent Conformation. Proc. Natl. Acad. Sci. USA 2023, 120, e2220012120. [Google Scholar] [CrossRef] [PubMed]
  203. Hanssen, K.M.; Haber, M.; Fletcher, J.I. Targeting Multidrug Resistance-Associated Protein 1 (MRP1)-Expressing Cancers: Beyond Pharmacological Inhibition. Drug Resist. Updates 2021, 59, 100795. [Google Scholar] [CrossRef]
  204. Alhaji, J.H. Vitamin B12 Deficiency in Patients with Diabetes on Metformin: Arab Countries. Nutrients 2022, 14, 2046. [Google Scholar] [CrossRef]
  205. Bell, D.S.H. Metformin-induced Vitamin B12 Deficiency Can Cause or Worsen Distal Symmetrical, Autonomic and Cardiac Neuropathy in the Patient with Diabetes. Diabetes Obes. Metab. 2022, 24, 1423–1428. [Google Scholar] [CrossRef]
  206. Infante, M.; Leoni, M.; Caprio, M.; Fabbri, A. Long-Term Metformin Therapy and Vitamin B12 Deficiency: An Association to Bear in Mind. World J. Diabetes 2021, 12, 916–931. [Google Scholar] [CrossRef] [PubMed]
  207. Bauman, W.A.; Shaw, S.; Jayatilleke, E.; Spungen, A.M.; Herbert, V. Increased Intake of Calcium Reverses Vitamin B12 Malabsorption Induced by Metformin. Diabetes Care 2000, 23, 1227–1231. [Google Scholar] [CrossRef]
  208. Out, M.; Kooy, A.; Lehert, P.; Schalkwijk, C.A.; Stehouwer, C.D.A. Long-Term Treatment with Metformin in Type 2 Diabetes and Methylmalonic Acid: Post Hoc Analysis of a Randomized Controlled 4.3 Year Trial. J. Diabetes Complicat. 2018, 32, 171–178. [Google Scholar] [CrossRef] [PubMed]
  209. Kovatcheva, M.; Melendez, E.; Chondronasiou, D.; Pietrocola, F.; Bernad, R.; Caballe, A.; Junza, A.; Capellades, J.; Holguín-Horcajo, A.; Prats, N.; et al. Vitamin B12 Is a Limiting Factor for Induced Cellular Plasticity and Tissue Repair. Nat. Metab. 2023, 5, 1911–1930. [Google Scholar] [CrossRef]
Figure 1. Skeletal formula of cobalamin. R (green) = CN, HO, methyl, or 5-deoxyadenosyl, N (blue)—nitrogen, Co+2 (pink)—cobaltous cation, O (red)—oxygen, H (grey)—hydrogen, P (orange)—phosphorus, arrows—possible bindings.
Figure 1. Skeletal formula of cobalamin. R (green) = CN, HO, methyl, or 5-deoxyadenosyl, N (blue)—nitrogen, Co+2 (pink)—cobaltous cation, O (red)—oxygen, H (grey)—hydrogen, P (orange)—phosphorus, arrows—possible bindings.
Ijms 25 08021 g001
Figure 2. Adsorption, blood transport, and intracellular metabolism of vitamin B12. Cbl—cobalamin, HC—haptocorrin, IF—Intrinsic Factor, TC—transcobalamin II, MRP1—multidrug resistant protein 1, ABCD4—ATP binding cassette subfamily D member protein, LMBD1—lysosomal cobalamin transport escort protein, MMACHC—methylmalonic aciduria and homocystinuria type C protein, MMADHC—methylmalonic aciduria and homocystinuria type D protein, AdoCbl—adenosylcobalamin, CH3Cbl—methylcobalamin, MCM—methylmalonyl CoA mutase, MS—methionine synthase. Created with BioRender.com.
Figure 2. Adsorption, blood transport, and intracellular metabolism of vitamin B12. Cbl—cobalamin, HC—haptocorrin, IF—Intrinsic Factor, TC—transcobalamin II, MRP1—multidrug resistant protein 1, ABCD4—ATP binding cassette subfamily D member protein, LMBD1—lysosomal cobalamin transport escort protein, MMACHC—methylmalonic aciduria and homocystinuria type C protein, MMADHC—methylmalonic aciduria and homocystinuria type D protein, AdoCbl—adenosylcobalamin, CH3Cbl—methylcobalamin, MCM—methylmalonyl CoA mutase, MS—methionine synthase. Created with BioRender.com.
Ijms 25 08021 g002
Table 1. Proteins are involved in vitamin B12 metabolism.
Table 1. Proteins are involved in vitamin B12 metabolism.
ProteinShortnameGeneTypeFunctionPathologyRef.
Haptocorrin, Transcobalamin IHC, TC-ITCN1GlycoproteinBinding vitamin B12 in saliva and blood, transferring the vitamin to the stomach; Protection of vitamin B12 from hydrolysis in the acidic pH of the stomach.The inability of HC degradation and release of the vitamin from the vitamin B12–HC complex (caused by deficiency of pancreatic proteases) may lead to vitamin B12 malabsorption.[15,16,17,18,19,20]
Intrinsic factorIFGIF or CBLIFGlycoproteinBinding vitamin B12 in the duodenum.The lack of IF in the condition of
H. pylori infection or AIG leads to long-lasting vitamin B12 deficiency due to the inability to deliver vitamin B12 to the cells within the ileum.
[15,16,21]
Cubam
receptor
CubilinCUBNCUBNMultiligand apical membrane receptorInteraction with IF; Docking to transmembrane protein AMN; Forming with AMN a receptor specific for IF–vitamin B12 internalization.Damage of this cubam results in malabsorption of vitamin B12 and the development of Imerslund–Gräsbeck syndrome (IGS), a rare autosomal recessive disorder that appears in childhood, characterized by megaloblastic anemia.[32,33,36,37]
Type-1 transmembrane protein amnionlessAMNAMNClathrin-mediated internalization; Anchoring Cubilin; Forming with CUBN a receptor specific for IF–vitamin B12 internalization.
Multidrug resistant protein 1MRP1, ABCC1ABCC1ATP-binding cassette (ABC) transporters familyEffluxing vitamin B12 from the intestinal epithelium to the blood circulation.Mutations in the gene ABCC1 were detected in some patients with vitamin B12 malabsorption; however, it is an unlikely cause of Cbl deficiency in humans.[46,47,52,54]
TranscobalaminTC, TC-IITCN2β-globulin proteinBinding vitamin B12 in the bloodstream; Facilitated distribution of vitamin B12 into all cells in the body.Mutations in the gene can lead to intracellular cobalamin depletion, causing a rare multisystemic disorder characterized by autosomal
recessive inheritance.
[55,59,60,65,66,67]
Receptor for cluster of differentiation 320, Receptor for cobalamin-saturated transcobalaminCD320, TCblRCD320Low-density Lipoprotein receptor protein familyEndocytosis of the TC–cobalamin complex.Inconsistent studies: in 2010, Quadros et al. associated elevated MMA in five newborns with a mutation in CD320, while in 2022 Pangilinan et al. estimated that approximately 85% of infants with this mutation have unaffected cobalamin metabolism.[68,69,70,72,73,80,81,82]
ATP-binding
cassette sub-family D member 4
ABCD4, CblJ proteinABCD4Exporter
protein
Transport of free vitamin B12 into the cytoplasm.Mutations are related to disease groups cblJ and cblF. At the cellular level, both errors are marked with a decreased function of MMUT and MTR and an accumulation of free vitamin B12 in the lysosomes.[83,86,87,88,90,93]
Lysosomal
cobalamin transport escort protein
LMBD1, CblF proteinLMBRD1Adaptor
protein,
chaperone
Mediates the translocation of ABCD4 from ER to the lysosomes.
Methylmalonic
aciduria and
homocystinuria
type C protein
MMACH, CblC
protein
MMACHCChaperone,
enzyme
Potentially receiving the Cbl cargo upon its release from the lysosomal compartment; Catalyzing the reductive decyanation and delkylation of cobalamin.Mutations result in a rare genetic disorder referred to as methylmalonic aciduria and homocystinuria type C (cblC); deficiencies in both vitamin B12-dependent enzymatic functions MTR and MMUT—accumulation of both MMA and Hcy in serum.[94,96,97,98,102]
Methylmalonic
aciduria and
homocystinuria
type D protein
MMADHC, CblD
protein
MMADHCProbably
enzyme
Delivers Cbl-precursor to the production sites of MeCbl and AdoCb.The cblD complementation group can be linked to isolated methylmalonic aciduria (cblD-MMA), isolated homocystinuria (cblD-HC), or a combination of both methylmalonic aciduria and homocystinuria (cblD-MMA/HC).[105,107,108,111,112,113]
Methionine synthaseMTR, MSMTRCytoplasmic enzymeVitamin B12-dependent transfer of methyl group from
5-methyltetrahydrofolate to homocysteine with the production of tetrahydrofolate and methionine.
Missense mutation or deletion of three base pairs was identified in the cblG group of patients with cobalamin metabolism disorder and methylcobalamin deficiency cblE type, rare autosomal diseases associated with homocysteinemia, homocystinuria, and hypomethioninemia.[81,82,83,84,86,87,115,123,124]
Methylmalonyl-CoA mutaseMMUT, MCM, MUTMMUTMitochondrial enzymeVitamin B12-dependent
conversion methylmalonyl-CoA to succinyl-CoA.
Mutation leads to methylmalonic acidemia.[88,89,90,92,128,131]
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

Mucha, P.; Kus, F.; Cysewski, D.; Smolenski, R.T.; Tomczyk, M. Vitamin B12 Metabolism: A Network of Multi-Protein Mediated Processes. Int. J. Mol. Sci. 2024, 25, 8021. https://doi.org/10.3390/ijms25158021

AMA Style

Mucha P, Kus F, Cysewski D, Smolenski RT, Tomczyk M. Vitamin B12 Metabolism: A Network of Multi-Protein Mediated Processes. International Journal of Molecular Sciences. 2024; 25(15):8021. https://doi.org/10.3390/ijms25158021

Chicago/Turabian Style

Mucha, Patryk, Filip Kus, Dominik Cysewski, Ryszard T. Smolenski, and Marta Tomczyk. 2024. "Vitamin B12 Metabolism: A Network of Multi-Protein Mediated Processes" International Journal of Molecular Sciences 25, no. 15: 8021. https://doi.org/10.3390/ijms25158021

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