Skip to content
Publicly Available Published by De Gruyter December 12, 2012

Proteomics of vitamin B12 processing

  • Luciana Hannibal

    Luciana Hannibal is a Research Associate in the Department of Pathobiology, Lerner Research Institute, Cleveland Clinic, Cleveland, Ohio, USA. Her research focuses on the molecular mechanisms of vitamin B12 processing and chaperone-mediated intracellular B12 transport to client enzymes. She also studies the structure-function relationships that govern catalysis by nitric oxide synthases and related hemoproteins.

    EMAIL logo
    , Patricia M. DiBello

    Patricia M. DiBello is a Research Associate in the Department of Cellular and Molecular Medicine, Lerner Research Institute, Cleveland Clinic, Cleveland, Ohio. Her research focuses on the molecular mechanisms of alcoholic liver disease (ALD) and the relationship between dysregulated vitamin B12, folate and homocysteine metabolism in the pathogenesis of ALD.

    and Donald W. Jacobsen

    Donald W. Jacobsen is Professor of Molecular Medicine, Department of Molecular Medicine, Cleveland Clinic Lerner College of Medicine, Case Western Reserve University and a member of the Professional Staff, Department of Cellular and Molecular Medicine, Lerner Research Institute, Cleveland Clinic, Cleveland, Ohio. His laboratory focuses on the basic and clinical research of vitamin B12, folate, homocysteine and one-carbon metabolism.

Abstract

The causes of cobalamin (B12, Cbl) deficiency are multifactorial. Whether nutritional due to poor dietary intake, or functional due to impairments in absorption or intracellular processing and trafficking events, the major symptoms of Cbl deficiency include megaloblastic anemia, neurological deterioration and in extreme cases, failure to thrive and death. The common biomarkers of Cbl deficiency (hyperhomocysteinemia and methylmalonic acidemia) are extremely valuable diagnostic indicators of the condition, but little is known about the changes that occur at the protein level. A mechanistic explanation bridging the physiological changes associated with functional B12 deficiency with its intracellular processers and carriers is lacking. In this article, we will cover the effects of B12 deficiency in a cblC-disrupted background (also referred to as MMACHC) as a model of functional Cbl deficiency. As will be shown, major protein changes involve the cytoskeleton, the neurological system as well as signaling and detoxification pathways. Supplementation of cultured MMACHC-mutant cells with hydroxocobalamin (HOCbl) failed to restore these variants to the normal phenotype, suggesting that a defective Cbl processing pathway produces irreversible changes at the protein level.

Introduction

Vitamin B12 (cobalamin, Cbl), an ancient vitamin and ‘Nature’s most beautiful cofactor’ [1, 2] is required by all cells in the body. Humans rely on dietary supplies of the vitamin since a biosynthesis pathway is lacking in higher organisms. Only a relatively few archaea and bacteria express the 30 or so enzymes required to synthesize the complex cobalt-containing macrocycle and its attached dimethylbenzimidazole moiety (Figure 1) [4]. In mammals, Cbl serves as a cofactor for methionine synthase (MS) and methylmalonyl-CoA mutase (MUT). Besides these canonical functions, new roles have been described for Cbl including intracellular signaling [5], apoptosis [6–8], oxidative stress [9] and cytokine and growth factor-mediated regulation [10, 11]. The exact mechanisms underlying these non-canonical actions remain largely unexplored. Dietary vitamin B12 enters the gastrointestinal tract first by complexation with the B12-binder haptocorrin (HC), which is present in saliva. Once it reaches the stomach, the vitamin is relayed to a second B12-binder, intrinsic factor (IF). Absorption of B12 occurs in ileal enterocytes of the lower intestine, where the vitamin dissociates from IF and binds to apo-transcobalamin (TC), the cellular transporter of vitamin B12[12]. The transcobalamin receptor (TCblR) captures holo-TC (TC•XCbl) from circulation and internalizes the complex by absorptive endocytosis [12]. In the acidic milieu of the endosomal compartment, holo-TC dissociates from its receptor and TCblR recycles back to the cell surface [13]. Fusion of late endosomes carrying the holo-TC with lysosomes results in the proteolytic degradation of transcobalamin. Cobalamin is thereby released within the lysosome and exported to the cytosol via the cblF gene product LMBD1 [14–21]. It was recently reported that the cblJ gene product, ABCD4, may work in conjunction with LMBD1 to mediate transport out of the lysosome to the cytosol [22]. Cobalamin then undergoes processing and trafficking to cytosolic methionine synthase, or to mitochondrial methylmalonyl-CoA mutase.

Figure 1 Structure of cobalamin.The base-off conformation (free α-axial ligand) of methylcobalamin (MeCbl) is shown. The upper, β-axial position can be occupied by different ligands, including cyanide (CNCbl, vitamin B12), 5′-adenosyl (AdoCbl) and hydroxo (HOCbl). The structure of MeCbl was taken from the MMACHC•MeCbl complex (PDB 3SC0 [3]). The Figure was generated with PyMol software (DeLano Scientific LLC).
Figure 1

Structure of cobalamin.

The base-off conformation (free α-axial ligand) of methylcobalamin (MeCbl) is shown. The upper, β-axial position can be occupied by different ligands, including cyanide (CNCbl, vitamin B12), 5′-adenosyl (AdoCbl) and hydroxo (HOCbl). The structure of MeCbl was taken from the MMACHC•MeCbl complex (PDB 3SC0 [3]). The Figure was generated with PyMol software (DeLano Scientific LLC).

Cellular processing of B12: MMACHC (CblC)

Much of our knowledge of Cbl processing (defined as removal of the upper axial ligand with either concerted or subsequent reduction of the cobalt center) arose from ex vivo studies with fibroblasts from patients carrying inborn errors of Cbl metabolism. Mutations in the genes that encode the enzymes or proteins involved in Cbl processing, trafficking and biosynthesis are defined by Cbl complementation groups (cblA-cblG and mut) [23, 24]. A study by Chu et al. suggested that dietary methylcobalamin (MeCbl) and adenosylcobalamin (AdoCbl) must undergo processing of their upper axial ligand prior to their incorporation into MS and MUT, respectively [25]. The first case report of functional Cbl deficiency caused by an inborn error of metabolism was provided by Harvey Mudd et al., more than 40 years ago [26]. The patient under study belonged to the cblC complementation group and presented with combined homocystinuria and methylmalonic aciduria [26]. Cultured patient fibroblasts displayed slightly reduced uptake of Cbl with respect to normal skin fibroblasts, efflux of Cbl at long incubation times, and impaired biosynthesis of both MeCbl and AdoCbl [26]. This was the first evidence that the gene responsible for the cblC complementation group was required for a step prior to both cofactors biosynthesis. A number of patients presenting with both early and late onset of the cblC disease were reported thereafter, which amounts to more than 360 cases to date [27].

In vitro assays with cell extracts showed that cblC fibroblasts possessed reduced Cbl β-transferase activity and/or Cbl reductase activity [28–32]. These studies suggested that the methylmalonic aciduria combined with homocystinuria type C (MMACHC) protein was involved in processing of the upper axial ligand of Cbls and/or reduction of the cobalt center. The work by Pezacka and Jacobsen also revealed a requirement for the most abundant intracellular thiol, glutathione, in a step preceding cofactor biosynthesis [28–32]. It has been postulated that MMACHC protein, the product of the cblC gene, is the immediate downstream acceptor of the Cbl cargo exiting the lysosome and the protein responsible for processing of the upper axial ligand of incoming dietary Cbls [33]. It was not until 2006 that the gene responsible for the cblC phenotype was identified and characterized [34]. According to the primary structure, the MMACHC protein is not a member of any previously identified gene family [34]. Although it is well-conserved among mammals, its C-terminal end does not seem to be conserved in eukaryotes outside mammalia, and no homologous proteins are found in prokaryotes [34]. The MMACHC gene is expressed in most tissues. High mRNA levels were detected in fetal liver with lower levels being detected in spleen, lymph node, thymus and bone marrow, and no message was detected in peripheral blood leukocytes [34]. Work by Koutmos et al. showed that a truncated form of cblC lacking the last 38 amino acid residues is predominantly expressed in most tissues [3]. Importantly, a comprehensive examination of the mitochondrion proteome identified MMACHC as one of its resident proteins [35]. How MMACHC is transported into the mitochondrion and its role in this compartment remains to be elucidated.

Biophysical and structural characterization of the B12-processing enzyme MMACHC and its interactions with MMADHC (CblD)

The mystery of how decyanation of cyanocobalamin (CNCbl) occurs was recently solved by the in vitro studies of Kim et al. [36]. The reductive decyanation of CNCbl is catalyzed by the MMACHC protein in the presence of a flavoprotein reductase and NADPH [36]. The authors reported that cblC bound both MeCbl and AdoCbl inducing their base-off conformation [36]; however, it did not catalyze the dealkylation of MeCbl and AdoCbl, the two major dietary forms of Cbl. This intriguing finding was re-examined via ex vivo studies [37], and a new function was uncovered for the cblC protein: MMACHC is also a Cbl dealkylase [37]. Mechanistically, dealkylation of AdoCbl and MeCbl is distinct from the decyanation pathway. MMACHC catalyzes the dealkylation of alkylcobalamins by a reaction involving the nucleophilic attack of the Co–C bond by the thiolate anion of glutathione [38]. Demethylation of MeCbl was much faster than the removal of the 5′-adenosyl group from AdoCbl (11.7±0.2 and 0.150±0.006/h, respectively) [38]. In addition, MMACHC was capable of dealkylating a series of MeCbl analogues namely, ethylcobalamin, propylcobalamin, butylcobalamin, pentylcobalamin and hexylcobalamin [37]. MMACHC catalyzed the removal of the alkyl group at the upper axial position of all of the MeCbl analogs, however, the rate of dealkylation decreased with increasing alkyl chain length [38]. Whether the latter is a result of conformational alterations in the MMACHC protein induced by the more bulky alkyl moieties or due to an unfavorable incorporation of the longer alkyl carbocations into glutathione remains to be elucidated.

An interesting feature of MMACHC is its stability. Froese et al. demonstrated that MMACHC is naturally thermolabile (Tm=39°C) and that some of the most frequent mutations that occur in humans exacerbate this property [39] as well as its ability to bind Cbls [40]. Studies with the bovine isoform of MMACHC revealed that the reduced form of glutathione stabilizes MMACHC, suggesting that intracellular redox control could play a role in the regulation of the protein’s lifetime [41–44]. Koutmos [3] and Froese [45] have independently obtained high-resolution X-ray crystal structures of MMACHC. The work by Koutmos et al. revealed that MMACHC possesses an N-terminal flavodoxin nitroreductase domain, which can use FMN or FAD to catalyze the reductive decyanation of CNCbl [3]. MMACHC possesses a large cavity for binding B12 in its base-off configuration (Figure 2, panel A), a binding mode thought to facilitate the reductive removal of the cyanide group at the β-axial position. Unlike other Cbl-dependent enzymes, the base-off Cbl binding by MMACHC does not involve the coordination of a His residue from the protein backbone [3]. Binding of MeCbl to MMACHC induces measurable conformational changes in three different loop-structured domains around the B12 cavity (Figure 2, panel A, arrows). Froese et al. elucidated the first structure of MMACHC bound to AdoCbl [45]. The overall fold of MMACHC does not differ markedly from that reported of MMACHC complexed with MeCbl, but revealed a highly conserved dimerization cap for the β-axial 5′-adenosyl ligand, and an arginine-rich glutathione-binding pocket up above the β-axial ligand position [45]. Importantly, the arginine-rich pocket comprises residues Arg161, Arg206 and Arg230 (Figure 2, panel B, blue) all of which are sites for point mutations that occur in humans, leading to cblC disease. A citrate molecule from the solvent was identified in the region predicted to be occupied by glutathione during catalysis (Figure 2, panel B, red). Froese et al. further showed that recombinant mutant Arg206Gln was insoluble suggesting a structural role for this residue, and that mutants Arg161Gln and Arg230Gln abolished GSH binding and dealkylase activity, which is a strong indication that these amino acid residues are critical for GSH binding [45]. The authors noted that FMN, and to a lesser extent Cbl, induces the dimerization of MMACHC, a previously unrecognized feature of the protein [45].

Figure 2 Structures of MMACHC.(A) Apo-MMACHC (PDB 3SBZ, cyan) and the MMACHC•MeCbl complex (PDB 3SC0 [3], green). The arrows indicate the major regions that undergo conformational change upon binding of Cbl. (B) The MMACHC•AdoCbl complex (PDB 3SOM [45]) displays three conserved Arg residues within the Arg-rich pocket (Arg161, Arg206 and Arg230, blue) and a citrate molecule (red) at a site predicted to be occupied by GSH under physiological conditions. The Figures were generated with PyMol software (DeLano Scientific LLC).
Figure 2

Structures of MMACHC.

(A) Apo-MMACHC (PDB 3SBZ, cyan) and the MMACHC•MeCbl complex (PDB 3SC0 [3], green). The arrows indicate the major regions that undergo conformational change upon binding of Cbl. (B) The MMACHC•AdoCbl complex (PDB 3SOM [45]) displays three conserved Arg residues within the Arg-rich pocket (Arg161, Arg206 and Arg230, blue) and a citrate molecule (red) at a site predicted to be occupied by GSH under physiological conditions. The Figures were generated with PyMol software (DeLano Scientific LLC).

The exact sequence of events that occur upon processing of the upper axial ligand of Cbl by MMACHC prior to its delivery to the two Cbl-dependent enzymes is unknown. Methylmalonic aciduria combined with homocystinuria type D (MMADHC), the product of the cblD gene has been proposed to have an adapter function to target newly processed Cbl into MS or MUT [3, 24–45]. The cblD gene has been recently mapped to chromosome 2q23.2 and has been designated MMADHC [46]. The predicted gene product has sequence similarity with a bacterial ATP-binding cassette transporter, possesses a putative Cbl binding motif and a putative mitochondrial targeting sequence [46]. The remarkable heterogeneity [47] of the cblD disorder in humans can be partially rationalized in terms of mutation sites. Mutations affecting the putative Cbl binding site would lead to a disrupted delivery of newly processed Cbls to both MS and MUT, resulting in combined homocystinuria and methylmalonic aciduria (a phenotype common to the cblC disorder), whereas mutations in the mitochondrial targeting sequence would result in disrupted delivery of Cbl to MUT, and therefore, would lead to isolated methylmalonic aciduria. Work by Stucki et al. indicated that MMADHC is a single protein with two different domains that interact with either cytosolic or mitochondrial targets [48]. Plesa et al. were the first to demonstrate that MMACHC and MMADHC interact [49]. Surface plasmon resonance studies showed that the two proteins interact at a 1:1 stoichiometry with an affinity in the mid-nanomolar range [49]. Phage display assays identified five putative MMACHC-binding sites. Two of these belong within the ABC-transporter homology domain, and the other three potential binding sites comprise residues in the C-terminal portion of MMADHC [49]. Further, phage display studies performed by Deme et al. identified several putative MMACHC binding sites in MMADHC, as well as self-binding regions in MMACHC [50]. In agreement with previous findings by Plesa et al., the authors proposed that the MMADHC functionality involves its C-terminal residues, which interact with MMACHC to orchestrate the fate of newly processed Cbl. Interestingly, none of these studies found evidence for Cbl binding to MMADHC [49, 50], despite the presence of a partially conserved B12-binding motif in its primary sequence.

Metabolic profile of cells with a defective MMACHC protein

An assessment of the metabolic and morphological characteristics of three cblC cell lines [WG1801 (c.217C>T/c.217C>T), WG2176 (c.1–234A>G/c.609G>A) and WG3354 (c.435_436delAT/c.435_436delAT)] was first conducted to determine whether this genetic system would be a suitable model of cobalamin deficiency.

The Cbl processing activity of these cblC cell lines has been characterized ex vivo [37]. All cblC cell lines released increased levels of homocysteine and methylmalonic acid compared to the normal cell line [51]. Supplementation with hydroxocobalamin (HOCbl) reduced the levels of methylmalonic acid exported by the cblC cell lines, but failed to substantially diminish the levels of homocysteine [51]. All cblC cell lines displayed reduced uptake and impaired processing of HOCbl compared to the normal cell line [51]. This is in line with the finding that all cblC mutant cell lines take up or retain lower amounts of Cbl compared to the normal fibroblasts [26], a phenomenon well-mirrored in total levels of intracellular folates [51]. Therefore, it appears that the positive response of the cblC cell lines to HOCbl supplementation (lowering of methylmalonic acid) may not be directly associated with the activity of mitochondrial MUT. An altered cellular morphology was observed when cblC cells were seeded at very low densities, but this vanished once the cultures became confluent. Whether this is an unwanted effect of elevated homocysteine and methylmalonic acid, or the presence of a defective MMACHC protein is uncertain. Altogether, the selected cblC cell lines displayed properties consistent with a model of functional deficiency.

Proteomics of defective B12 processing in cblC cell lines

The recent years have witnessed much progress in the elucidation of the genes, the proteins and the diverse phenotypes associated with B12 processing and trafficking disorders [6, 9, 27, 34, 36, 38–40, 45, 49–71]. However, considerably less is known about the changes that accompany Cbl deficiency at the protein level. The proteome of fibroblasts of three genetically unrelated, severely ill cblC patients were examined in order to identify changes associated with functional Cbl deficiency [51]. The effect of supplementation with HOCbl, a form of Cbl used in the treatment of cblC patients, was also evaluated [51]. Proteomic studies performed with diseased individuals are often challenged by the availability of their cells or tissues, and the examination of the cblC fibroblast proteome by two-dimensional difference gel electrophoresis (2-D DIGE) was no exception. Skin fibroblasts were the only sample available from the patients, WG1801, WG2176 and WG3354, all of whom had deceased by the time the study was conducted. Likewise, fibroblasts from healthy family members without cblC disease were not available. The problem of small sample size (three cblC fibroblasts samples and three from normal individuals), was partially solved by identifying protein expression differences that were common to all three cblC fibroblasts sources and by excluding the changes that were particular to an individual sample. Although these hurdles prevent the direct extrapolation of protein changes identified in fibroblasts to other cell types unless appropriate validation is conducted, the study provides a global view of the changes associated with the cblC disease by utilizing cells that are genetically stable, easy to grow and for which genetically unrelated controls are readily available (fibroblasts from normal patients).

Table 1 shows the changes in protein expression in cblC vs. normal fibroblasts. Major changes occurred in proteins related to cytoskeleton, the nervous system, signaling and cellular detoxification [51]. Table 2 presents the changes in protein expression in cblC vs. normal fibroblasts supplemented with 723 nM (1 μg/mL) HOCbl. Supplementation of the cell cultures with HOCbl led to a global downregulation of the cblC fibroblast proteome. The target proteins, namely MMACHC, MS and MUT could not be detected in the study, possibly due to their naturally low levels of expression and the limit of detection of the 2-D DIGE technique (~0.25 ng protein) [72]. An independent study by Sourmala et al. demonstrated that MS activity of 5 cblC fibroblast patient cell lines was four to five-fold lower than normal fibroblasts, and that MUT activity was two to five times lower in cblC compared to normal fibroblasts [73]. No differences in the expression of HDL binding protein, collagen VI α1 precursor, eukaryotic elongation factor 2 (eEF2), glutathione-S-transferase omega 1 (GSTO1), GST M3 and glyceraldehyde-3-phosphate dehydrogenase (GAPDH) were observed between the normal and cblC fibroblasts upon supplementation with HOCbl, which suggests that supplying the vitamin corrected for the alterations caused by its deficiency, an effect that must be independent of the cblC processing route.

Table 1

Protein changes in cblC vs. normal fibroblasts grown in the absence of exogenous hydroxocobalamin (n=3, α=0.05). Adapted from Hannibal et al. [51].

ProteinGene IDFold changeMascot scoret-test (p-Value)
Proteins whose expression was upregulated in cblC fibroblasts
 HDL binding protein427162802.878070.029
 Eukaryotic translation elongation factor 245034832.512340.042
 Collagen type VI, α2, isoform 2C21155270622.12950.028
 Collagen type VI, α1 precursor871963392.16620.028
 Ribosomal protein S1450320512.452600.033
 H2B histone family member A45042572.45900.033
Proteins whose expression was downregulated in cblC fibroblasts
 Serine (or Cys) proteinase inhibitor (protease inhibitor 6 or serpin B6)41152086−2.0923220.032
 Caldesmon-1, isoform 2, Ct truncated4826657−2.099560.032
 β-actin4501885−2.1111160.024
 Tubulin α, ubiquitous57013276−2.179070.028
 Chloride intracellular channel 47330335−2.0415070.018
 Collagen type VI, α2, isoform 2C21155270622.12950.028
 Collagen type VI, α1 precursor871963392.16620.028
 Ubiquitin carboxyl esterase L121361091−2.1912610.021
 Plastin 37549809−2.914710.023
 Vimentin62414289−2.4436520.032
 Glutathione transferase omega 14758484−2.11720.0096
 Glutathione transferase4504183−2.848210.016
 Glutathione transferase M32306552−2.192810.021
 PDI associated 3 precursor21361657−2.95290.023
 SH3 domain binding Glu-rich protein like 313775198−2.233120.042
 S100 Ca-binding protein A67657532−2.231210.042
 GAPDH7669492−2.117610.024
Table 2

Protein changes in cblC vs. normal fibroblasts supplemented with 723 nM hydroxocobalamin (n=3, α=0.05). The proteins are listed in descending order of fold change in expression. Adapted from Hannibal et al. [51], with permission.

ProteinGene IDFold changeMascot scoret-test (p-Value)
Hsp 90 protein 1, β20149594−2.0317580.042
Hsp 90 α (cytosolic), class A, member 1, isoform 240254816−2.0313930.042
Transgelin 24507357−2.0313640.04
Annexin V4502107−2.0812850.031
Phosphoglycerate dehydrogenase23308577−2.1323160.033
Inosine monophosphate dehydrogenase 266933016−2.137780.042
Hsp70 protein 8, isoform 15729877−2.1522180.042
Annexin VI isoform 171773329−2.1516780.042
ATPase, H+ transporting, lysosomal19913424−2.157010.042
Voltage dependent anion channel 14507879−2.176570.034
Voltage dependent anion channel 242476281−2.174050.034
S100 Ca-binding protein A67657532−2.191410.031
Peroxirredoxin 2, isoform a32189392−2.238990.031
Septin 118922712−2.2512370.031
Cofilin 15031635−2.264730.031
Hsp70, protein 5,16507237−2.2833100.031
PDI associated 44758304−2.284350.031
Peroxiredoxin 14505591−2.292690.031
Lamin A/C, isoform 25031875−2.326720.049
Dihydropyrimidase -like 24503377−2.316220.049
Chaperonine containing TCP1, sub363162572−2.38290.049
Peroxiredoxin 64758638−2.346000.042
PDI associated 3 precursor21361657−2.3811400.042
Tubulin α614389309−2.3810140.042
Chloride intracellular Channel 47330335−2.386210.034
DJ-131543380−2.521520.031
Glutathione transferase4504183−2.549260.031
Tryptophanyl-tRNA synthetase isoform a47419914−2.6321290.033
Collagen type VI, isoform 2C217402875−2.729530.039
Vinculin isoform VCL4507877−2.726600.039
Ubiquitin and ribosomal protein S27a precursor4506713−2.793820.031
Annexin V, A2, isoform I50845388−2.93550.039
Triosephosphate isomerase 14507645−2.939540.039
SH3 domain binding Glu-rich protein like 313775198−3.15650.032
Ubiquitin carboxyl esterase L121361091−3.169490.037
Vimentin62414289−5.0339080.031

Cytoskeleton: assembly and remodeling

Substantial changes in protein expression levels were identified for cytoskeletal proteins with structural and regulatory roles. These included collagen VI, α1 and 2C2 isoforms, vimentin (VIM), tubulin-α (TUBA1B), β-actin (ACTB), vinculin isoform, plastin 3 (PLS3), lamin A/C isoform 2, chaperonin TCP1, caldesmon 1 (CALD1), cofilin 1 and transgelin 2. Changes in cytoskeletal proteins have been also reported for a patient cell line belonging to the cblD complementation group [74] and also, for human fibroblasts [75] and colonocytes [76] under conditions of folate deficiency. Collagen VI, isoform 2C2 was upregulated in cblC fibroblasts. Supplementation of the cblC cultures with HOCbl downregulated its expression. Mutations in the genes that code for collagen VI subunits result in Ullrich syndrome [77] and Bethlem myopathy [78], an autosomal dominant disorder. An upregulation of collagen VI α2 was also noted in a patient with cblD disease [74]. The patient described in the study presented with isolated methylmalonic aciduria. Another study performed in human smooth muscle cells demonstrated that high levels of homocysteine cause an upregulation in the production of collagen, which could be related to the pathogenesis of homocystinuria [79]. Patients with untreated homocystinuria have widespread premature atherosclerosis with intimal thickening and collagen-rich fibrous plaques [80]. An altered collagen expression may contribute to the pathology of cblC given the deregulation in the fibroblasts from patients (elevated Hcy and increased collagen expression). VIM was also downregulated in cblC fibroblasts. VIM is a cytoskeletal protein whose major role is stabilizing the architecture of the cytoplasm. An in vitro study conducted by Mor-Vaknin et al. revealed that monocyte-derived macrophages secrete VIM into the extracellular space [81]. Secretion of VIM was stimulated by tumor necrosis factor-α (TNF-α) and inhibited by IL10, suggesting that the protein is also involved in the immune response [81]. It was found that cblC fibroblasts expressed lower levels of VIM compared to normal cells, and, interestingly, downregulation was further exacerbated by supplementation with HOCbl [51]. The abnormal expression of VIM in cblC fibroblasts could be responsible for their altered cellular morphology. Less is known about the roles of other cytoskeleton-related proteins identified in the study. However, it appears that their functions are broadly related to the regulation of smooth muscle, non-muscle contraction (CALD1, a calmodulin and actin-binding protein) [82] and axonogenesis (PLS3) [83] and its related neuromuscular disease, spinal muscular athrophy [84]. Neurological and muscular disorders commonly seen in cblC patients are often alleviated by the administration of HOCbl. PLS3 levels corrected to the normal pattern of expression upon supplementation with HOCbl, which could partly account for the improvement observed in some cblC patients.

Nervous system and signaling

Ubiquitin carboxy-terminal hydrolase L1 (UCHL1) is another of the downregulated proteins in cblC fibroblasts. UCHL1, also known as PGP9.5, is an essential component of the ubiquitin-proteasome system (UPS), a major pathway for protein degradation [85]. UCHL1 is one of the most abundant proteins in the brain and is predominantly localized in neurons [86] and cells of the diffuse neuroendocrine system [87]. Downregulation and extensive oxidative modification of UCHL1 occurs in brain tissue of patients with Alzheimer’s as well as Parkinson’s diseases [88–90]. Downregulation of UCHL1 in cblC fibroblasts was not restored to normal levels upon supplementation with HOCbl. Therefore, downregulation of UCHL1 could contribute to the neurocognitive manifestations of the cblC disorder and the poor improvement observed after treatment with HOCbl [91]. Three other proteins were downregulated in cblC fibroblasts grown in the presence of exogenous HOCbl: DJ-1 (Parkinson‘s disease protein 7), dihydropyrimidase-like 2 (DPYLS2), and annexin V A2 isoform I. DJ-1 belongs to a family of peptidases that act as a positive regulator of androgen receptor-dependent transcription. DJ-1 may also function as a redox-sensitive chaperone and it is thought to protect neurons from oxidative damage [92]. Defects in this gene are the cause of early-onset Parkinson‘s disease 7 [92, 93]. DPYLS2 shares homology with dihydropyrimidase and it is expressed actively in the fetal and neonatal brains of mammals and chickens. Little is known about this family of proteins, however, they are thought to be intracellular transducers in the development of the nervous system [94].

Gene regulation and protein synthesis

Fibroblasts with a defective cblC gene grown in the absence of exogenous HOCbl displayed an upregulation in the expression levels of eEF2, ribosomal protein S14 (RPS14) and H2B member A, and a downregulation in the levels of the calcium binding protein S100A6 [51]. Supplementation with HOCbl restored the expression of eEF2, RPS14 and H2B to normal levels in cblC fibroblasts, however, the expression pattern of S100A6 was unaffected. Two other proteins, septin 11 (SEPT11) and ubiquitin and ribosomal protein S27a precursor (URPS27a) were dowregulated in cblC fibroblast grown in the presence of HOCbl. Septins are a novel family of GTP-binding proteins that appear to play an important role in cytokinesis, membrane dynamics, vesicle trafficking, apoptosis, and cell polarity [95]. SEPT11 was also downregulated in fibroblasts from patients with the cblD disorder [74].

Intracellular trafficking and protein folding

A number of proteins involved in protein folding and intracellular trafficking were downregulated in cblC fibroblasts compared to the normal cell line, a pattern that was not reversed by exogenous supplies of HOCbl. These include: protein disulfide isomerase (PDI) precursor 3 and PDI associated 4, heat shock protein 90 (Hsp90) 1 and Hsp90 α-class A member 1 isoform 2, heat shock protein 70 (Hsp70) proteins 5 and 8, annexin VI isoform 1, annexin V, annexin V A2, a lysosomal – H+ transporting ATPase, and voltage dependent anion channels (VDACs) 1 and 2. Importantly, PDI, Hsp70 and Hsp90 play important roles in folding of newly synthesized proteins or stabilizing and refolding of denatured proteins after stress [96, 97]. Annexins are a family of Ca2+-dependent and membrane-binding proteins, which are involved in membrane trafficking and various other processes including signaling, proliferation, differentiation, and inflammation [98–100]. Lysosomal H+-transporting ATPase is a vacuolar enzyme that mediates acidification of eukaryotic intracellular organelles, a critical step for processes such as protein sorting, zymogen activation, receptor-mediated endocytosis and synaptic vesicle proton gradient generation [101]. The VDACs are the major channels by which small hydrophilic molecules cross the mitochondrial outer membrane. Evidence suggests that VDAC isoforms in mammals may act in the cross-talk between mitochondria and the cytoplasm by direct interaction with enzymes involved in energy metabolism and proteins involved in mitochondrial-induced apoptosis [102]. VDACs also interact with anti-apoptotic proteins from the Bcl-2 family, and this interaction inhibits the release of apoptogenic proteins from the mitochondrion [103]. The expression of both VDAC1 and VDAC2 was diminished in cblC fibroblasts compared to normal cells, a pattern that also prevailed when cells were grown in the presence of HOCbl.

General metabolism and cellular detoxification

A number of proteins involved in general metabolism and cellular detoxification were identified as downregulated in the cblC fibroblast proteome. These include: high density lipoprotein binding protein (HDLBP), GAPDH and glutathione-S-transferase (GST) (various isoforms) among others. Three isoforms of GST were downregulated in cblC fibroblasts grown without HOCbl supplementation: GST Ω1, GST, and GST M3. Of these, only GST remained downregulated under conditions of HOCbl supplementation, whereas the expression levels GST Ω1 and GST M3 did not differ significantly from that of normal cell lines. Glutathione transferases utilize glutathione to detoxify drugs, xenobiotics, and oxidants. A recent report showed strong associations between the age of onset of Alzheimer’s and Parkinson’s diseases and polymorphisms of GST Ω1 and 2 [104]. The mu (M) class of GST functions in the detoxification of carcinogens, therapeutic drugs, environmental toxins, and products of oxidative stress by conjugation with GSH [105]. Immunoblotting analysis revealed that GST M3 is the predominant isoform in the brain [106]. Activity assays confirmed that cblC fibroblasts have reduced total GST activity [51]. Reduced GST expression and activity may compromise the detoxification of metabolites, which in turn could aggravate the manifestation of the cblC disease.

Conclusions and future perspective

The elucidation of the biophysical and structural properties of MMACHC led to a greater understanding of its function including its interactions with downstream MMADHC. The instability of MMACHC suggests it is likely short-lived in vivo, or that its existence may require an unknown stabilizing factor.

The cblC fibroblast proteome exhibited expression patterns that are significantly different from that of the normal cultured skin fibroblasts. A defective or absent MMACHC protein caused profound changes of cellular metabolism and regulation, including cytoskeleton assembly, nervous system proteins, signaling, and cellular detoxification [51]. A number of the proteins identified in the study have been linked to skeletal and muscular diseases as well as neurological diseases, which concurs with the clinical manifestations of the cblC disorder. The identification of proteins whose expression was altered by the cblC mutation could be useful targets for further research, and perhaps, for designing alternative therapies to alleviate the symptoms of the cblC disease. For instance, administration of UCHL1, a protein that is mutated or downregulated in patients with Alzheimer’s and Parkinson’s disease, and also downregulated in cblC fibroblasts, was shown to alleviate the β-amyloid-induced synaptic dysfunction and memory loss associated with a mouse model of Alzheimer’s disease [107]. Likewise, it is possible that therapies utilized to treat patients with skeletal and muscular diseases would be effective in ameliorating such affections in cblC patients; we now know that at least some of the proteins involved in the progression of these diseases have a common set of contributors, the cytoskeletal proteins. Although these notions remain largely speculative until further research is conducted, these results provide a platform for investigating the protein-phenotype relationships underlying the most common inborn error of vitamin B12 metabolism.


Corresponding author: Luciana Hannibal, PhD, Department of Pathobiology (NC2–104), Lerner Research Institute, Cleveland Clinic, 9500 Euclid Avenue, Cleveland, OH 44195, USA, Phone: +1 216 4459761, Fax: +1 216 6360104

About the authors

Luciana Hannibal

Luciana Hannibal is a Research Associate in the Department of Pathobiology, Lerner Research Institute, Cleveland Clinic, Cleveland, Ohio, USA. Her research focuses on the molecular mechanisms of vitamin B12 processing and chaperone-mediated intracellular B12 transport to client enzymes. She also studies the structure-function relationships that govern catalysis by nitric oxide synthases and related hemoproteins.

Patricia M. DiBello

Patricia M. DiBello is a Research Associate in the Department of Cellular and Molecular Medicine, Lerner Research Institute, Cleveland Clinic, Cleveland, Ohio. Her research focuses on the molecular mechanisms of alcoholic liver disease (ALD) and the relationship between dysregulated vitamin B12, folate and homocysteine metabolism in the pathogenesis of ALD.

Donald W. Jacobsen

Donald W. Jacobsen is Professor of Molecular Medicine, Department of Molecular Medicine, Cleveland Clinic Lerner College of Medicine, Case Western Reserve University and a member of the Professional Staff, Department of Cellular and Molecular Medicine, Lerner Research Institute, Cleveland Clinic, Cleveland, Ohio. His laboratory focuses on the basic and clinical research of vitamin B12, folate, homocysteine and one-carbon metabolism.

LH dedicates this article to Perry and Sasha.

Conflict of interest statement

Author’s conflict of interest disclosure: The authors state that there are no conflicts of interest regarding the publication of this article.

Research funding: This work was supported in part by grants HL71907 (DWJ) and HL52234 (DWJ) from the National Institutes of Health (USA) and 11POST650034 (LH) from the American Heart Association (USA).

Employment or leadership: None declared.

Honorarium: None declared.

References

1. Drennan CL, Huang S, Drummond JT, Matthews RG, Lidwig ML. How a protein binds B12: a 3.0 A X-ray structure of B12-binding domains of methionine synthase. Science 1994;266:1669–74.10.1126/science.7992050Search in Google Scholar PubMed

2. Putre L. B12: the beautiful molecule. Cleve Clin Mag 2010;7:20–6.Search in Google Scholar

3. Koutmos M, Gherasim C, Smith JL, Banerjee R. Structural basis of multifunctionality in a vitamin B12-processing enzyme. J Biol Chem 2011;286:29780–7.10.1074/jbc.M111.261370Search in Google Scholar PubMed PubMed Central

4. Martens JH, Barg H, Warren MJ, Jahn D. Microbial production of vitamin B12. Appl Microbiol Biotechnol 2002;58:275–85.10.1007/s00253-001-0902-7Search in Google Scholar PubMed

5. Okada K, Tanaka H, Temporin K, Okamoto M, Kuroda Y, Moritomo H, et al. Akt/mammalian target of rapamycin signaling pathway regulates neurite outgrowth in cerebellar granule neurons stimulated by methylcobalamin. Neurosci Lett 2011;495:201–4.10.1016/j.neulet.2011.03.065Search in Google Scholar PubMed

6. Jorge-Finnigan A, Gamez A, Perez B, Ugarte M, Richard E. Different altered pattern expression of genes related to apoptosis in isolated methylmalonic aciduria cblB type and combined with homocystinuria cblC type. Biochim Biophys Acta 2010;1802:959–67.10.1016/j.bbadis.2010.08.002Search in Google Scholar PubMed

7. Richard E, Alvarez-Barrientos A, Perez B, Desviat LR, Ugarte M. Methylmalonic acidaemia leads to increased production of reactive oxygen species and induction of apoptosis through the mitochondrial/caspase pathway. J Pathol 2007;213: 453–61.10.1002/path.2248Search in Google Scholar PubMed

8. Orozco-Barrios CE, Battaglia-Hsu SF, Arango-Rodriguez ML, Ayala-Davila J, Chery C, Alberto JM, et al. Vitamin B12-impaired metabolism produces apoptosis and Parkinson phenotype in rats expressing the transcobalamin-oleosin chimera in substantia nigra. PLoS One 2009;4:e8268.10.1371/journal.pone.0008268Search in Google Scholar PubMed PubMed Central

9. Richard E, Jorge-Finnigan A, Garcia-Villoria J, Merinero B, Desviat LR, Gort L, et al. Genetic and cellular studies of oxidative stress in methylmalonic aciduria (MMA) cobalamin deficiency type C (cblC) with homocystinuria (MMACHC). Hum Mutat 2009;30:1558–66.10.1002/humu.21107Search in Google Scholar PubMed

10. Scalabrino G. The multi-faceted basis of vitamin B12 (cobalamin) neurotrophism in adult central nervous system: lessons learned from its deficiency. Prog Neurobiol 2009;88:203–20.10.1016/j.pneurobio.2009.04.004Search in Google Scholar PubMed

11. Scalabrino G. Vitamin-regulated cytokines and growth factors in the CNS and elsewhere. J Neurochem 2009;111:1309–26.10.1111/j.1471-4159.2009.06417.xSearch in Google Scholar PubMed

12. Quadros EV, Nakayama Y, Sequeira JM. The protein and the gene encoding the receptor for the cellular uptake of transcobalamin-bound cobalamin. Blood 2009;113:186–92.10.1182/blood-2008-05-158949Search in Google Scholar

13. Amagasaki T, Green R, Jacobsen DW. Expression of transcobalamin II receptors by human leukemia K562 and HL-60 cells. Blood 1990;76:1380–6.10.1182/blood.V76.7.1380.1380Search in Google Scholar

14. Watkins D, Rosenblatt DS. Failure of lysosomal release of vitamin B12: a new complementation group causing methylmalonic aciduria (cblF). Am J Hum Genet 1986;39:404–8.Search in Google Scholar

15. Shih VE, Axel SM, Tewksbury JC, Watkins D, Cooper BA, Rosenblatt DS. Defective lysosomal release of vitamin B12 (cb1F): a hereditary cobalamin metabolic disorder associated with sudden death. Am J Med Genet 1989;33:555–63.10.1002/ajmg.1320330431Search in Google Scholar

16. Vassiliadis A, Rosenblatt DS, Cooper BA, Bergeron JJ. Lysosomal cobalamin accumulation in fibroblasts from a patient with an inborn error of cobalamin metabolism (cblF complementation group): visualization by electron microscope radioautography. Exp Cell Res 1991;195:295–302.10.1016/0014-4827(91)90376-6Search in Google Scholar

17. Laframboise R, Cooper BA, Rosenblatt DS. Malabsorption of vitamin B12from the intestine in a child with cblF disease: evidence for lysosomal-mediated absorption. Blood 1992;80:291–2.10.1182/blood.V80.1.291.bloodjournal801291Search in Google Scholar

18. Waggoner DJ, Ueda K, Mantia C, Dowton SB. Methylmalonic aciduria (cblF): case report and response to therapy. Am J Med Genet 1998;79:373–5.10.1002/(SICI)1096-8628(19981012)79:5<373::AID-AJMG8>3.0.CO;2-KSearch in Google Scholar

19. Rutsch F, Gailus S, Miousse IR, Suormala T, Sagne C, Toliat MR, et al. Identification of a putative lysosomal cobalamin exporter altered in the cblF defect of vitamin B12 metabolism. Nat Genet 2009;41:234–9.10.1038/ng.294Search in Google Scholar

20. Oladipo O, Rosenblatt DS, Watkins D, Miousse IR, Sprietsma L, Dietzen DJ, et al. Cobalamin F disease detected by newborn screening and follow-up on a 14-year-old patient. Pediatrics 2011;128:e1636–40.10.1542/peds.2010-3518Search in Google Scholar

21. Miousse IR, Watkins D, Rosenblatt DS. Novel splice site mutations and a large deletion in three patients with the cblF inborn error of vitamin B12 metabolism. Mol Genet Metab 2011;102:505–7.10.1016/j.ymgme.2011.01.002Search in Google Scholar

22. Coelho D, Kim JC, Miousse IR, Fung S, du Moulin M, Buers I, et al. Mutations in ABCD4 cause a new inborn error of vitamin B12 metabolism. Nat Genet 2012;44:1152–5.10.1038/ng.2386Search in Google Scholar

23. Watkins D, Rosenblatt DS. Inborn errors of cobalamin absorption and metabolism. Am J Med Genet C Semin Med Genet 2011;157:33–44.10.1002/ajmg.c.30288Search in Google Scholar

24. Froese DS, Gravel RA. Genetic disorders of vitamin B12 metabolism: eight complementation groups–eight genes. Expert Rev Mol Med 2010;12:e37.10.1017/S1462399410001651Search in Google Scholar

25. Chu RC, Begley JA, Colligan PD, Hall CA. The methylcobalamin metabolism of cultured human fibroblasts. Metabolism 1993;42:315–9.10.1016/0026-0495(93)90080-8Search in Google Scholar

26. Mudd SH, Levy HL, Abeles RH. A derangment in B12 metabolism leading to homocystinemia, cystathioninemia and methylmalonic aciduria. Biochem Biophys Res Commun 1969;35:121–6.10.1016/0006-291X(69)90491-4Search in Google Scholar

27. Lerner-Ellis JP, Anastasio N, Liu J, Coelho D, Suormala T, Stucki M, et al. Spectrum of mutations in MMACHC, allelic expression, and evidence for genotype-phenotype correlations. Hum Mutat 2009;30:1072–81.10.1002/humu.21001Search in Google Scholar

28. Pezacka EH, Denison CJ, Green R, Jacobsen DW. Biosynthesis of methylcobalamin: chemical model studies with thiol-cobalamin adducts and S-adenosylmethionine. J Cell Physiol 1988;107:860a.Search in Google Scholar

29. Pezacka E, Green R, Jacobsen DW. Glutathionylcobalamin as an intermediate in the formation of cobalamin coenzymes. Biochem Biophys Res Commun 1990;169:443–50.10.1016/0006-291X(90)90351-MSearch in Google Scholar

30. Pezacka EH, Green R, Jacobsen DW. Intracellular cobalamin metabolism: a thiol-cobalamin adduct as an intermediate in cobalamin coenzyme formation. FASEB J 1990;4:A2126.Search in Google Scholar

31. Pezacka EH. Identification and characterization of two enzymes involved in the intracellular metabolism of cobalamin. Cyanocobalamin beta-ligand transferase and microsomal cob(III)alamin reductase. Biochim Biophys Acta 1993;1157:167–77.Search in Google Scholar

32. Pezacka EH, Rosenblatt DS. Intracellular metabolism of cobalamin. Altered activities of B-axial-ligand transferase and microsomal cob(III)alamin reductase in cbl C and cbl D fibroblasts. In: Bhatt, James, Besser, Bottazzo, Keen, editors. Advances in Thomas Addison’s Diseases. Bristol: J. Endocrinol. Ltd., 1994:315–23.Search in Google Scholar

33. Banerjee R. B12 trafficking in mammals: a for coenzyme escort service. ACS Chem Biol 2006;1:149–59.10.1021/cb6001174Search in Google Scholar PubMed

34. Lerner-Ellis JP, Tirone JC, Pawelek PD, Dore C, Atkinson JL, Watkins D, et al. Identification of the gene responsible for methylmalonic aciduria and homocystinuria, cblC type. Nat Genet 2006;38:93–100.10.1038/ng1683Search in Google Scholar PubMed

35. Pagliarini DJ, Calvo SE, Chang B, Sheth SA, Vafai SB, Ong SE, et al. A mitochondrial protein compendium elucidates complex I disease biology. Cell 2008;134:112–23.10.1016/j.cell.2008.06.016Search in Google Scholar PubMed PubMed Central

36. Kim J, Gherasim C, Banerjee R. Decyanation of vitamin B12 by a trafficking chaperone. Proc Natl Acad Sci USA 2008;105:14551–4.10.1073/pnas.0805989105Search in Google Scholar PubMed PubMed Central

37. Hannibal L, Kim J, Brasch NE, Wang S, Rosenblatt DS, Banerjee R, et al. Processing of alkylcobalamins in mammalian cells: a role for the MMACHC (cblC) gene product. Mol Genet Metab 2009;97:260–6.10.1016/j.ymgme.2009.04.005Search in Google Scholar PubMed PubMed Central

38. Kim J, Hannibal L, Gherasim C, Jacobsen DW, Banerjee R. A human vitamin B12 trafficking protein uses glutathione transferase activity for processing alkylcobalamins. J Biol Chem 2009;284:33418–24.10.1074/jbc.M109.057877Search in Google Scholar PubMed PubMed Central

39. Froese DS, Healy S, McDonald M, Kochan G, Oppermann U, Niesen FH, et al. Thermolability of mutant MMACHC protein in the vitamin B12-responsive cblC disorder. Mol Genet Metab 2010;100:29–36.10.1016/j.ymgme.2010.02.005Search in Google Scholar PubMed PubMed Central

40. Froese DS, Zhang J, Healy S, Gravel RA. Mechanism of vitamin B12-responsiveness in cblC methylmalonic aciduria with homocystinuria. Mol Genet Metab 2009;98:338–43.10.1016/j.ymgme.2009.07.014Search in Google Scholar PubMed

41. Park J, Jeong J, Kim J. Destabilization of a bovine B12 trafficking chaperone protein by oxidized form of glutathione. Biochem Biophys Res Commun 2012;420:547–51.10.1016/j.bbrc.2012.03.031Search in Google Scholar PubMed

42. Jeong J, Kim J. Glutathione increases the binding affinity of a bovine B12 trafficking chaperone bCblC for vitamin B12. Biochem Biophys Res Commun 2011;412:360–5.10.1016/j.bbrc.2011.07.103Search in Google Scholar PubMed

43. Jeong J, Ha TS, Kim J. Protection of aquo/hydroxocobalamin from reduced glutathione by a B12 trafficking chaperone. BMB Rep 2011;44:170–5.10.5483/BMBRep.2011.44.3.170Search in Google Scholar PubMed

44. Park J, Kim J. Glutathione and vitamin B12 cooperate in stabilization of a B12 trafficking chaperone protein. Protein J 2012;31:158–65.10.1007/s10930-011-9385-2Search in Google Scholar PubMed

45. Froese DS, Krojer T, Wu X, Shrestha R, Kiyani W, von Delft F, et al. Structure of MMACHC reveals an arginine-rich pocket and a domain-swapped dimer for its B12 processing function. Biochemistry 2012;51:5083–90.10.1021/bi300150ySearch in Google Scholar PubMed

46. Coelho D, Suormala T, Stucki M, Lerner-Ellis JP, Rosenblatt DS, Newbold RF, et al. Gene identification for the cblD defect of vitamin B12 metabolism. N Engl J Med 2008;358:1454–64.10.1056/NEJMoa072200Search in Google Scholar PubMed

47. Miousse IR, Watkins D, Coelho D, Rupar T, Crombez EA, Vilain E, et al. Clinical and molecular heterogeneity in patients with the cblD inborn error of cobalamin metabolism. J Pediatr 2009;154:551–6.10.1016/j.jpeds.2008.10.043Search in Google Scholar PubMed

48. Stucki M, Coelho D, Suormala T, Burda P, Fowler B, Baumgartner MR. Molecular mechanisms leading to three different phenotypes in the cblD defect of intracellular cobalamin metabolism. Hum Mol Genet 2012;21:1410–8.10.1093/hmg/ddr579Search in Google Scholar PubMed

49. Plesa M, Kim J, Paquette SG, Gagnon H, Ng-Thow-Hing C, Gibbs BF, et al. Interaction between MMACHC and MMADHC, two human proteins participating in intracellular vitamin B12 metabolism. Mol Genet Metab 2011;102:139–48.10.1016/j.ymgme.2010.10.011Search in Google Scholar PubMed

50. Deme JC, Miousse IR, Plesa M, Kim JC, Hancock MA, Mah W, et al. Structural features of recombinant MMADHC isoforms and their interactions with MMACHC, proteins of mammalian vitamin B12 metabolism. Mol Genet Metab 2012;107:352–62.10.1016/j.ymgme.2012.07.001Search in Google Scholar PubMed

51. Hannibal L, DiBello PM, Yu M, Miller A, Wang S, Willard B, et al. The MMACHC proteome: hallmarks of functional cobalamin deficiency in humans. Mol Genet Metab 2011;103:226–39.10.1016/j.ymgme.2011.03.008Search in Google Scholar PubMed PubMed Central

52. Adams D, Venditti CP. Disorders of intracellular cobalamin metabolism. GeneReviews TM 2009. In: Pagon RA, Bird TD, Dolan CR, Stephens K, Adam MP, editors. Seattle (WA): University of Washington, Seattle, 2008.Search in Google Scholar

53. Morel CF, Lerner-Ellis JP, Rosenblatt DS. Combined methylmalonic aciduria and homocystinuria (cblC): phenotype-genotype correlations and ethnic-specific observations. Mol Genet Metab 2006;88:315–21.10.1016/j.ymgme.2006.04.001Search in Google Scholar PubMed

54. Ben-Omran TI, Wong H, Blaser S, Feigenbaum A. Late-onset cobalamin-C disorder: a challenging diagnosis. Am J Med Genet A 2007;143A:979–84.10.1002/ajmg.a.31671Search in Google Scholar PubMed

55. Heil SG, Hogeveen M, Kluijtmans LA, van Dijken PJ, van de Berg GB, Blom HJ, et al. Marfanoid features in a child with combined methylmalonic aciduria and homocystinuria (CblC type). J Inherit Metab Dis 2007;30:811.10.1007/s10545-007-0546-6Search in Google Scholar PubMed

56. Sharma AP, Greenberg CR, Prasad AN, Prasad C. Hemolytic uremic syndrome (HUS) secondary to cobalamin C (cblC) disorder. Pediatr Nephrol 2007;22:2097–103.10.1007/s00467-007-0604-1Search in Google Scholar PubMed

57. Gerth C, Morel CF, Feigenbaum A, Levin AV. Ocular phenotype in patients with methylmalonic aciduria and homocystinuria, cobalamin C type. J AAPOS 2008;12:591–6.10.1016/j.jaapos.2008.06.008Search in Google Scholar PubMed

58. Thauvin-Robinet C, Roze E, Couvreur G, Horellou MH, Sedel F, Grabli D, et al. The adolescent and adult form of cobalamin C disease: clinical and molecular spectrum. J Neurol Neurosurg Psychiatry 2008;79:725–8.10.1136/jnnp.2007.133025Search in Google Scholar PubMed

59. Loewy AD, Niles KM, Anastasio N, Watkins D, Lavoie J, Lerner-Ellis JP, et al. Epigenetic modification of the gene for the vitamin B12 chaperone MMACHC can result in increased tumorigenicity and methionine dependence. Mol Genet Metab 2009;96:261–7.10.1016/j.ymgme.2008.12.011Search in Google Scholar PubMed

60. Tang H, Hao H, Tang SH, Chen X, Liu F, Cha QB, et al. [Mutation analysis of the MMACHC gene in a pedigree with methylmalonic aciduria]. Zhonghua Yi Xue Yi Chuan Xue Za Zhi 2009;26:62–5.Search in Google Scholar

61. Wang F, Han LS, Hu YH, Yang YL, Ye J, Qiu WJ, et al. [Analysis of gene mutations in Chinese patients with methylmalonic acidemia and homocysteinemia]. Zhonghua Er Ke Za Zhi 2009;47:189–93.Search in Google Scholar

62. Brunel-Guitton C, Costa T, Mitchell GA, Lambert M. Treatment of cobalamin C (cblC) deficiency during pregnancy. J Inherit Metab Dis Epub 2010 Sep 10. DOI: 10.1007/s10545-010-9202-7.10.1007/s10545-010-9202-7Search in Google Scholar PubMed

63. Frattini D, Fusco C, Ucchino V, Tavazzi B, Della Giustina E. Early onset methylmalonic aciduria and homocystinuria cblC type with demyelinating neuropathy. Pediatr Neurol 2010;43:135–8.10.1016/j.pediatrneurol.2010.04.007Search in Google Scholar PubMed

64. Wang F, Han L, Yang Y, Gu X, Ye J, Qiu W, et al. Clinical, biochemical, and molecular analysis of combined methylmalonic acidemia and hyperhomocysteinemia (cblC type) in China. J Inherit Metab Dis Epub 2010 Oct 6. DOI: 10.1007/s10545-010-9217-0.10.1007/s10545-010-9217-0Search in Google Scholar PubMed

65. Weisfeld-Adams JD, Morrissey MA, Kirmse BM, Salveson BR, Wasserstein MP, McGuire PJ, et al. Newborn screening and early biochemical follow-up in combined methylmalonic aciduria and homocystinuria, cblC type, and utility of methionine as a secondary screening analyte. Mol Genet Metab 2010;99:116–23.10.1016/j.ymgme.2009.09.008Search in Google Scholar PubMed PubMed Central

66. Chang JT, Chen YY, Liu TT, Liu MY, Chiu PC. Combined methylmalonic aciduria and homocystinuria cblC type of a Taiwanese infant with c.609G>A and C.567dupT mutations in the MMACHC gene. Pediatr Neonatol 2011;52:223–6.10.1016/j.pedneo.2011.05.006Search in Google Scholar PubMed

67. Martinelli D, Deodato F, Dionisi-Vici C. Cobalamin C defect: natural history, pathophysiology, and treatment. J Inherit Metab Dis 2011;34:127–35.10.1007/s10545-010-9161-zSearch in Google Scholar PubMed

68. Pupavac M, Garcia MA, Rosenblatt DS, Jerome-Majewska LA. Expression of MMACHC and MMADHC during mouse organogenesis. Mol Genet Metab 2011;103:401–5.10.1016/j.ymgme.2011.04.004Search in Google Scholar PubMed

69. Carrillo-Carrasco N, Chandler RJ, Venditti CP. Combined methylmalonic acidemia and homocystinuria, cblC type. I. Clinical presentations, diagnosis and management. J Inherit Metab Dis 2012;35:91–102.10.1007/s10545-011-9364-ySearch in Google Scholar PubMed PubMed Central

70. Menni F, Testa S, Guez S, Chiarelli G, Alberti L, Esposito S. Neonatal atypical hemolytic uremic syndrome due to methylmalonic aciduria and homocystinuria. Pediatr Nephrol 2012;27:1401–5.10.1007/s00467-012-2152-6Search in Google Scholar

71. Wang X, Sun W, Yang Y, Jia J, Li C. A clinical and gene analysis of late-onset combined methylmalonic aciduria and homocystinuria, cblC type, in China. J Neurol Sci 2012;318:155–9.10.1016/j.jns.2012.04.012Search in Google Scholar

72. Experience the power of 2-D electrophoresis with 2-D DIGE. GE Healthcare Life Sciences, 2012. Available from: https://www.gelifesciences.com/gehcls_images/GELS/Related%20Content/Files/1347435454111/litdoc11001340AE_20120914132155.pdf. Accessed October, 2012.Search in Google Scholar

73. Suormala T, Baumgartner MR, Coelho D, Zavadakova P, Kozich V, Koch HG, et al. The cblD defect causes either isolated or combined deficiency of methylcobalamin and adenosylcobalamin synthesis. J Biol Chem 2004;279:42742–9.10.1074/jbc.M407733200Search in Google Scholar

74. Richard E, Monteoliva L, Juarez S, Perez B, Desviat LR, Ugarte M, et al. Quantitative analysis of mitochondrial protein expression in methylmalonic acidemia by two-dimensional difference gel electrophoresis. J Proteome Res 2006;5:1602–10.10.1021/pr050481rSearch in Google Scholar

75. Katula KS, Heinloth AN, Paules RS. Folate deficiency in normal human fibroblasts leads to altered expression of genes primarily linked to cell signaling, the cytoskeleton and extracellular matrix. J Nutr Biochem 2007;18:541–52.10.1016/j.jnutbio.2006.11.002Search in Google Scholar

76. Duthie SJ, Mavrommatis Y, Rucklidge G, Reid M, Duncan G, Moyer MP, et al. The response of human colonocytes to folate deficiency in vitro: functional and proteomic analyses. J Proteome Res 2008;7:3254–66.10.1021/pr700751ySearch in Google Scholar

77. Jimenez-Mallebrera C, Maioli MA, Kim J, Brown SC, Feng L, Lampe AK, et al. A comparative analysis of collagen VI production in muscle, skin and fibroblasts from 14 Ullrich congenital muscular dystrophy patients with dominant and recessive COL6A mutations. Neuromuscul Disord 2006;16: 571–82.10.1016/j.nmd.2006.07.015Search in Google Scholar

78. Kanagawa M, Toda T. The genetic and molecular basis of muscular dystrophy: roles of cell-matrix linkage in the pathogenesis. J Hum Genet 2006;51:915–26.10.1007/s10038-006-0056-7Search in Google Scholar

79. Majors AK, Sengupta S, Jacobsen DW, Pyeritz RE. Upregulation of smooth muscle cell collagen production by homocysteine-insight into the pathogenesis of homocystinuria. Mol Genet Metab 2002;76:92–9.10.1016/S1096-7192(02)00030-6Search in Google Scholar

80. Majors A, Ehrhart LA, Pezacka EH. Homocysteine as a risk factor for vascular disease – enhanced collagen production and accumulation by smooth muscle cells. Arterioscler Thromb Vasc Biol 1997;17:2074–81.10.1161/01.ATV.17.10.2074Search in Google Scholar PubMed

81. Mor-Vaknin N, Punturieri A, Sitwala K, Markovitz DM. Vimentin is secreted by activated macrophages. Nat Cell Biol 2003;5:59–63.10.1038/ncb898Search in Google Scholar

82. Yoshio T, Morita T, Kimura Y, Tsujii M, Hayashi N, Sobue K. Caldesmon suppresses cancer cell invasion by regulating podosome/invadopodium formation. FEBS Lett 2007;581: 3777–82.10.1016/j.febslet.2007.06.073Search in Google Scholar

83. Lin CS, Park T, Chen ZP, Leavitt J. Human plastin genes. Comparative gene structure, chromosome location, and differential expression in normal and neoplastic cells. J Biol Chem 1993;268:2781–92.10.1016/S0021-9258(18)53842-4Search in Google Scholar

84. Oprea GE, Krober S, McWhorter ML, Rossoll W, Muller S, Krawczak M, et al. Plastin 3 is a protective modifier of autosomal recessive spinal muscular atrophy. Science 2008;320:524–7.10.1126/science.1155085Search in Google Scholar PubMed PubMed Central

85. Gong B, Leznik E. The role of ubiquitin C-terminal hydrolase L1 in neurodegenerative disorders. Drug News Perspect 2007;20:365–70.10.1358/dnp.2007.20.6.1138160Search in Google Scholar PubMed

86. Wilson PO, Barber PC, Hamid QA, Power BF, Dhillon AP, Rode J, et al. The immunolocalization of protein gene product 9.5 using rabbit polyclonal and mouse monoclonal antibodies. Br J Exp Pathol 1988;69:91–104.Search in Google Scholar

87. Doran JF, Jackson P, Kynoch PA, Thompson RJ. Isolation of PGP 9.5, a new human neurone-specific protein detected by high-resolution two-dimensional electrophoresis. J Neurochem 1983;40:1542–7.10.1111/j.1471-4159.1983.tb08124.xSearch in Google Scholar PubMed

88. Castegna A, Aksenov M, Thongboonkerd V, Klein JB, Pierce WM, Booze R, et al. Proteomic identification of oxidatively modified proteins in Alzheimer’s disease brain. Part II: dihydropyrimidinase-related protein 2, alpha-enolase and heat shock cognate 71. J Neurochem 2002;82:1524–32.10.1046/j.1471-4159.2002.01103.xSearch in Google Scholar PubMed

89. Choi J, Levey AI, Weintraub ST, Rees HD, Gearing M, Chin LS, et al. Oxidative modifications and down-regulation of ubiquitin carboxyl-terminal hydrolase L1 associated with idiopathic Parkinson’s and Alzheimer’s diseases. J Biol Chem 2004;279:13256–64.10.1074/jbc.M314124200Search in Google Scholar PubMed

90. Butterfield DA, Gnjec A, Poon HF, Castegna A, Pierce WM, Klein JB, et al. Redox proteomics identification of oxidatively modified brain proteins in inherited Alzheimer’s disease: an initial assessment. J Alzheimers Dis 2006;10:391–7.10.3233/JAD-2006-10407Search in Google Scholar PubMed

91. Andersson HC, Marble M, Shapira E. Long-term outcome in treated combined methylmalonic acidemia and homocystinemia. Genet Med 1999;1:146–50.10.1097/00125817-199905000-00006Search in Google Scholar PubMed

92. Abou-Sleiman PM, Healy DG, Quinn N, Lees AJ, Wood NW. The role of pathogenic DJ-1 mutations in Parkinson’s disease. Ann Neurol 2003;54:283–6.10.1002/ana.10675Search in Google Scholar

93. Annesi G, Savettieri G, Pugliese P, D’Amelio M, Tarantino P, Ragonese P, et al. DJ-1 mutations and parkinsonism-dementia-amyotrophic lateral sclerosis complex. Ann Neurol 2005;58:803–7.10.1002/ana.20666Search in Google Scholar

94. Kitamura K, Takayama M, Hamajima N, Nakanishi M, Sasaki M, Endo Y, et al. Characterization of the human dihydropyrimidinase-related protein 2 (DRP-2) gene. DNA Res 1999;6:291–7.10.1093/dnares/6.5.291Search in Google Scholar

95. Ito H, Iwamoto I, Morishita R, Nozawa Y, Narumiya S, Asano T, et al. Possible role of Rho/Rhotekin signaling in mammalian septin organization. Oncogene 2005;24:7064–72.10.1038/sj.onc.1208862Search in Google Scholar

96. Appenzeller-Herzog C, Ellgaard L. The human PDI family: versatility packed into a single fold. Biochim Biophys Acta 2008;1783:535–48.10.1016/j.bbamcr.2007.11.010Search in Google Scholar

97. Gregersen N. Protein misfolding disorders: pathogenesis and intervention. J Inherit Metab Dis 2006;29:456–70.10.1007/s10545-006-0301-4Search in Google Scholar

98. Raynal P, Pollard HB. Annexins: the problem of assessing the biological role for a gene family of multifunctional calcium- and phospholipid-binding proteins. Biochim Biophys Acta 1994;1197:63–93.10.1016/0304-4157(94)90019-1Search in Google Scholar

99. Gerke V, Creutz CE, Moss SE. Annexins: linking Ca2+ signalling to membrane dynamics. Nat Rev Mol Cell Biol 2005;6:449–61.10.1038/nrm1661Search in Google Scholar PubMed

100. Grewal T, Enrich C. Molecular mechanisms involved in Ras inactivation: the annexin A6-p120GAP complex. Bioessays 2006;28:1211–20.10.1002/bies.20503Search in Google Scholar PubMed

101. Stevens TH, Forgac M. Structure, function and regulation of the vacuolar (H+)-ATPase. Annu Rev Cell Dev Biol 1997;13:779–808.10.1146/annurev.cellbio.13.1.779Search in Google Scholar PubMed

102. Meins T, Vonrhein C, Zeth K. Crystallization and preliminary X-ray crystallographic studies of human voltage-dependent anion channel isoform I (HVDAC1). Acta Crystallogr Sect F Struct Biol Cryst Commun 2008;64:651–5.10.1107/S174430910801676XSearch in Google Scholar

103. Hiller S, Garces RG, Malia TJ, Orekhov VY, Colombini M, Wagner G. Solution structure of the integral human membrane protein VDAC-1 in detergent micelles. Science 2008;321: 1206–10.10.1126/science.1161302Search in Google Scholar

104. Takeshita H, Fujihara J, Takastuka H, Agusa T, Yasuda T, Kunito T. Diversity of glutathione s-transferase omega 1 (a140d) and 2 (n142d) gene polymorphisms in worldwide populations. Clin Exp Pharmacol Physiol 2009;36:283–6.10.1111/j.1440-1681.2008.05058.xSearch in Google Scholar

105. Inskip A, Elexperu-Camiruaga J, Buxton N, Dias PS, MacIntosh J, Campbell D, et al. Identification of polymorphism at the glutathione S-transferase, GSTM3 locus: evidence for linkage with GSTM1*A. Biochem J 1995;312: 713–6.10.1042/bj3120713Search in Google Scholar

106. Campbell E, Takahashi Y, Abramovitz M, Peretz M, Listowsky I. A distinct human testis and brain mu-class glutathione S-transferase. Molecular cloning and characterization of a form present even in individuals lacking hepatic type mu isoenzymes. J Biol Chem 1990;265:9188–93.10.1016/S0021-9258(19)38830-1Search in Google Scholar

107. Gong B, Cao Z, Zheng P, Vitolo OV, Liu S, Staniszewski A, et al. Ubiquitin hydrolase Uch-L1 rescues beta-amyloid-induced decreases in synaptic function and contextual memory. Cell 2006;126:775–88.10.1016/j.cell.2006.06.046Search in Google Scholar PubMed

Received: 2012-8-31
Accepted: 2012-10-23
Published Online: 2012-12-12
Published in Print: 2013-03-01

©2013 by Walter de Gruyter Berlin Boston

Downloaded on 25.5.2024 from https://www.degruyter.com/document/doi/10.1515/cclm-2012-0568/html
Scroll to top button