Skip to main content

The role and mechanism of TXNDC5 in diseases

Abstract

Thioredoxin domain-containing protein 5 (TXNDC5) is a member of the protein disulfide isomerase (PDI) family. It can promote the formation and rearrangement of disulfide bonds, ensuring proper protein folding. TXNDC5 has three Trx-like domains, which can act independently to introduce disulfide bonds rapidly and disorderly. TXNDC5 is abnormally expressed in various diseases, such as cancer, rheumatoid arthritis (RA), etc. It can protect cells from oxidative stress, promote cell proliferation, inhibit apoptosis and promote the progression of disease. Aberrant expression of TXNDC5 in different diseases suggests its role in disease diagnosis. In addition, targeting TXNDC5 in the treatment of diseases has shown promising application prospects. This article reviews the structure and function of TXNDC5 as well as its role and mechanism in cancer, RA and other diseases.

Background

Thioredoxin domain-containing protein 5 (TXNDC5) is a member of the protein disulfide isomerase (PDI) family, and it is also known as ERp46, PDIA15, HCC-2 or Endo-PDI. TXNDC5 is a protein with a molecular weight of 48 kDa [1]. It was first found to be highly expressed in endothelial cells, liver cells and hypoxic tissues, such as cancer endothelial cells and atherosclerotic plaques [1, 2]. The TXNDC5 protein is mainly expressed in the endoplasmic reticulum (ER), and it is also expressed in the cytoplasm and envelope [3]. The TXNDC5 gene is located on chromosome 6p24.3 and encodes six indirect variants, two of which (TXNDC5-001 and TXNDC5-003) can be translated into proteins. TXNDC5-001 is a full-length protein containing 432 amino acids, and TXNDC5-003 isoform consists of 389 amino acids [4]. The current research on TXNDC5 always refers to the full-length one. In addition, compared with archetypal PDI family members, TXNDC5 has unique structure and function. Since the discovery of TXNDC5, many studies have found its abnormal expression in various diseases. Single nucleotide polymorphisms (SNPs) in the TXNDC5 gene are associated with the risk of various diseases (Table 1). In addition, TXNDC5 plays an important role in cell proliferation, apoptosis, migration and antioxidative stress. Besides, it has shown a promising role in the diagnosis and treatment of diseases. The role of TXNDC5 in diseases has been extensively studied in recent years. Among them, cancers and rheumatoid arthritis (RA) are most widely studied. In this paper, we summarize the structural and functional characteristics of TXNDC5 and its role in cancer, RA and other diseases.

Table 1 TXNDC5 SNP-related diseases

TXNDC5 structure

TXNDC5 is a member of the PDI family. PDI consists of four Trx-like domains (a, b, bʹ, aʹ) and a C-terminal domain (c) (Fig. 1a). The a and aʹ domains have redox activity due to the CxxC motif, and it can also catalyse the formation of natural disulfide bonds [5]. The b and b' domains have no CxxC motif and, therefore, no oxidoreductase activity. However, the bʹ domain is mainly responsible for providing substrate binding sites [6, 7]. In addition, it binds to unfolded or partially folded protein substrates [8]. Unlike archetypal PDI, TXNDC5 contains three Trx-like domains (a0, a and aʹ) [1, 9] (Fig. 1b) that all have redox active sites. In addition, the three different domains are connected by approximately 20 amino acid residues [10]. Three Trx-like domains of TXNDC5 can act independently to accelerate protein folding, while the Trx-like domains of PDI require synergy [10]. In the absence of the b domain, the three catalytic domains of TXNDC5 were found to bind peptides containing aromatic or alkaline residues [11]. Besides, cysteine sites of Trx-like domains Cys88, 217 and 350 are considered with supposed protein binding function [10] (Fig. 1c). The interaction observed in TXNDC5 crystals suggests that the third catalytic domain of TXNDC5 may bind to the substrate through small hydrophobic pockets on the surface of the Trx-like domain or by exposing the Trp349 residues [12]. A specific structural feature of TXNDC5 is the presence of a lysine residue (Lys344) near the second cysteine in the CxxC motif [12]. The lysine plays a role in the catalytic reaction by forming a salt bridge with glutamate 378 (Glu378) on the adjacent β chain. However, in other archetypal members of PDIs, glutamate plays a catalytic role by accepting the proton of the second cysteine [12]. Despoina et al. found that the reactivity of Trx-fold proteins is not determined solely by their CxxC motif. The cysteine pKa values, which determine reactivity, are fine-tuned by interactions involving several residues distant in sequence but spatially close to the CxxC motif [13]. In PDIA1, arginine residue Arg415 can participate in the catalytic reaction by regulating the pKa value of the second cysteine in the CxxC motif [14]. In the third Trx-like domain of TXNDC5, it was observed that Arg415 did not insert into the hydrophobic nucleus but bonded to the carbonyl oxygen of Pro397 by hydrogen bonds, so it couldn’t regulate the cysteine PKa value [11].

Fig. 1
figure 1

Scheme showing archetypal PDI and TXNDC5 domains and the role of TXNDC5 in diseases. Compared with the archetypal PDI domain (a), TXNDC5 has three a domains and no b domains (b); c: three TrX-like domains of TXNDC5 and their substrate binding sites; TXNDC5can be used as disease diagnostic biomarker (d) and treatment target (e). The surrounding text indicates the acting substrate or signal pathway. CC cervical cancer, ccRCC clear cell renal cell carcinoma, CRC colorectal cancer; CRPC castration-resistant prostate cancer, CTLC cutaneous T-cell lymphoma, GC gastric cancer, HCC hepatocellular carcinoma, HL hepatic Lipidosis, hMF human myocardial fibrosis, LC lung cancer, PF pulmonary fibrosis, LSCC laryngeal squamous cell carcinoma, PC pancreatic cancer, RA rheumatoid arthritis, RCC renal cell carcinoma, RF renal fibrosis, RRMM refractory/relapsed multiple myeloma;

TXNDC5 function

Similar to most members of the PDI family, TXNDC5 can catalyse the formation and rearrangement of disulfide bonds, thereby promoting the correct folding of proteins. In addition, the CxxC motif plays a key role in this process. Properly formed disulfide bonds can enhance the stability of proteins and protect them from damage [15, 16]. Both PDI and TXNDC5 can catalyse the formation of regulatory disulfide in endoplasmic reticulum oxidoreductase 1α (Ero1α), thereby preventing excessive oxidative stress caused by excessive H2O2 production. Ero1 mainly interacts with archetypal PDI family members, whereas TXNDC5 mainly interacts with peroxiredoxin-4 (Prx4) [10, 17]. Hydrogen peroxide generated during Ero1α–PDI interaction can act on the catalytic oxidation reaction between Prx4 and TXNDC5 to accelerate protein folding [10]. In addition, TXNDC5 can interact with peroxidized Prx2 and participate in redox signal transduction [18].

The three Trx-like domains of TXNDC5 can act independently to introduce disulfide bonds rapidly but disorderly without efficient selectivity. The two Trx-like domains of PDI act synergistically, so it is not as fast as TXNDC5. However, PDI can selectively introduce disulfide bonds and correct for nonnatural disulfide bonds. Therefore, TXNDC5 and PDI may act at different stages of oxidative protein folding [10, 19]. In addition, TXNDC5 has a protective effect on endothelial cells in a hypoxic environment. Inhibiting the expression of TXNDC5 can lead to a decrease in the secretion of molecules that protect cells from hypoxia, such as adrenomedullin, endothelin-1, and CD105 [1]. Moreover, TXNDC5 is involved in TNFα-induced angiogenesis [20]. Knockout of TXNDC5 can prevent TNFα-induced phosphorylation of ERK1/2. Phosphorylated ERK is an important substance that promotes the expression of two important angiogenesis-related proteases, MMP-9 and cathepsin B [20]. TXNDC5 also has a molecular chaperone function that can promote the proper protein folding [21, 22]. ER degradation enhancing α -Mannosidase-like protein 3 (EDEM3) is a key enzyme that can remove mannose from N-polysaccharide and repair the abnormal folding protein. TXNDC5 can bind to EDEM3 and trigger its mannose pruning activity to correct misfolded proteins. [23].

TXNDC5 and cancers

Currently, TXNDC5 is considered as a cancer-enhancing gene [24, 25]. It can protect cancer cells from oxidative stress and promote cancer growth, proliferation, migration and angiogenesis. Xu et al. found that TXNDC5 was a susceptibility gene for cervical cancer [26]. SNP analysis of the TXNDC5 gene found that two SNPs (rs1225944 and rs1225943) of TXNDC5 are related to the development of hepatocellular carcinoma (HCC) [27]. In addition, our research group found that the TXNDC5 gene increases the risk of cervical cancer, oesophageal cancer and liver cancer [21]. TXNDC5 is highly expressed in various cancer tissues. Therefore, many researchers have conducted extensive studies on its role in cancers.

The mechanism of TXNDC5 promoting cancers

TXNDC5 can promote cancer development in many ways, and they are summarized in Table 2. Like other PDI family members, TXNDC5 can participate in the formation of disulfide bonds, promoting the correct folding of proteins in the ER and preventing endoplasmic reticulum stress. However, PDIs can protect cells under both normal and hypoxic conditions, whereas TXNDC5 has this function only under hypoxic conditions [1].

Table 2 Cancer-promoting mechanism of TXNDC5

The expression of TXNDC5 is upregulated by hypoxia [21]. In vivo studies showed that hypoxia induces TXNDC5 expression by upregulating hypoxia inducible factor-1α (HIF-1α), thereby inhibiting hypoxia-induced ROS/ER stress signalling and promoting the reproduction and survival of colorectal cancer (CRC) cells [28]. However, TXNDC5 has no obvious effect on CRC cells under normoxia [28]. TXNDC5 can promote the progression of castration-resistant prostate cancer (CRPC) through androgen receptor (AR) signal transduction [29]. The specific mechanism is that androgen deprivation therapy (ADT) induces hypoxia in prostate cancer patients. In addition, hypoxia leads to increased expression of AR and TXNDC5. TXNDC5 can directly interact with AR and enhance its stability and transcriptional activity. Besides, TXNDC5 can make the Akt, ERK1/2, MET, and HER2 signalling pathways sensitive to hypoxia.

TXNDC5 can regulate the cell cycle. TXNDC5 can promote cell proliferation by regulating the cell cycle. Upregulating the expression of TXNDC5 in gastric cancer cells can significantly increase the percentage of cells in the G2/M phase, whereas downregulating can increase the percentage of cells in the 0/G1 phase [25]. In CRPC, TXNDC5 promotes cancer cell proliferation by causing cell cycle disorders between G2/M and S phases in vivo and in vitro [29].

TXNDC5 promotes cancer through distinct pathways. In vitro experiments have found that inhibiting the expression of TXNDC5 in cervical cancer cells leads to increased expression of SERPINF1 and TRAF1 [26]. SERPINF1 and TRAF1 can inhibit angiogenesis and metastasis and induce apoptosis. In lung cancer cells, TXNDC5 exerts a cancer-promoting effect by combining with sulfiredoxin (Srx) [30]. Srx can promote the proliferation, colony formation, and metastasis of lung cancer cells by activating mitogen-activated protein kinase cascades. TXNDC5 interacts with Srx to promote the retention of Srx in the ER, thereby protecting ER homeostasis of lung cancer cells and promoting the growth, proliferation, and invasion of cells [31]. In addition, TXNDC5 can interact with adiponectin receptor 1 (AdipoR1) in HeLa cells. Knocking out TXNDC5 can change the expression and distribution of AdipoR1 and adiponectin signals [32]. However, Duivenvoorden et al. found that TXNDC5 cannot establish a stable interaction with AdipoR1 in renal cell carcinoma (RCC) cells [33]. TXNDC5 may act through transient binding with AdipoR1. Alternatively, the difference in the reaction environment may affects the observed results of the experiment. In short, the relationship between AdipoR1 and TXNDC5 needs further research and exploration.

Expression regulation of TXNDC5 in cancers

As mentioned above, the expression of TXNDC5 is up-regulated by hypoxia. In particular, the expression of TXNDC5 in non-small-cell lung cancer tissues is upregulated, but this regulation is not controlled by hypoxia [34]. Indeed, TXNDC5 expression is also regulated by a variety of molecules (Table 3). CIRC_0000517 is upregulated in HCC tissues and cells. It can promote TXNDC5 expression by downregulating miR-1296-5p, thereby promoting cell viability, proliferation, colony formation, and inhibiting apoptosis [35]. The combination of circRNA-104718 and miR-218-5p in HCC cells could promote the expression of TXNDC5 [36]. Human ether a-go-go-related gene 1 (HERG1) in esophageal squamous cell carcinoma (ESCC) can promote the expression of TXNDC5 and promote cancer progression by activating the PI3K/AKT pathway [37]. The nuclear receptor 4A1 (NR4A1) in pancreatic cancer cells can promote the expression of the TXNDC5 gene to resist ROS/ER stress and inhibit cell apoptosis [38]. NR4A1 has also been found to regulate the TXNDC5 gene in renal cell adenocarcinoma [39]. Inhibiting NR4A1 in rhabdomyosarcoma (RMS) cells can reduce the expression of TXNDC5, thereby reducing the production of ROS and IL-24 and inhibiting the proliferation, survival and migration of RMS cells [40].

Table 3 Expression regulation mechanism of TXNDC5 in cancer cells

The role of TXNDC5 in cancer diagnosis and treatment

TXNDC5 may be used as a diagnostic marker for a variety of cancers (Fig. 1d). Evidence obtained by immunohistochemical staining and Western blot showed that TXNDC5 expression was altered in a variety of cancers. Tan et al. found that the expression of TXNDC5 was positively correlated with the TNM stage of colorectal cancer [28]. TXNDC5 was upregulated in poorly differentiated HCC but did not change in highly differentiated HCC [41]. In addition, the expression of TXNDC5 was negatively correlated with the overall survival rate in RCC [42] and gastric cancer [24]. However, lung cancer patients with high TXNDC5 levels have a longer survival time [31]. These studies indicated that TXNDC5 may be used as a prognostic marker for cancers, and the same TXNDC5 expression status in different diseases may indicate different prognostic results.

Variant gastritis (VG) is a precancerous lesion of gastric cancer. Detection of differentially expressed proteins in VG and normal tissues revealed significant differences in TXNDC5 expression [25]. In addition, the content of TXNDC5 in relatively early stage lung cancer is higher than that in normal tissue [34]. The above results indicate that the change in TXNDC5 expression may have occurred in the cancers’ early stages. However, the sample size of the above studies was small, and the application prospects of TXNDC5 as a cancer marker need to be verified by large-scale experiments.

TXNDC5 can be used as a therapeutic target for cancers (Fig. 1e). Inhibiting TXNDC5 expression in cervical cancer [21], castration-resistant prostate cancer [29], RCC [34], gastric cancer [25], laryngeal squamous cell carcinoma (LSCC) [43], pancreatic cancer [38], and liver cancer [35] can promote cell apoptosis and inhibit cell proliferation and migration. The above results suggest that TXNDC5 can be used as a therapeutic target for cancers. Knockout of TXNDC5 can make clear cell renal cell carcinoma (ccRCC) cells allergic to chemotherapy drugs (such as camptothecin and 5-fluorouracil) and inhibit the growth, migration, and invasion of ccRCC cells [42]. Biochip research showed that after TXNDC5 knockdown, the expression of JAG1, STAT1, fibronectin, GLS, EFEMP1, PLK2, and EDN2 genes was downregulated, whereas caspase-7, CCDC80, CDK14, and dbicd2 genes were upregulated. These results indicate that TXNDC5 may play an important role in the carcinogenesis and development of ccRCC through complex signal crosstalk [42]. In addition, TXNDC5 may be a potential therapeutic target for LSCC. Cetuximab and cisplatin are commonly used chemotherapy drugs for LSCC. Peng et al. found that cetuximab could reduce TXNDC5 expression, thereby increasing the production of ROS and enhancing the endoplasmic reticulum stress-related apoptosis of LSCC cells [43].

TXNDC5 and rheumatoid arthritis (RA)

In 2009, our research group found that TXNDC5 was specifically expressed in the tissues of RA patients compared with osteoarthritis and ankylosing spondylitis [44]. High expression of TXNDC5 in the blood and synovial fluid of RA patients can be detected in the early and continuous course of the disease [44]. The expression of TXNDC5 in tissues is positively correlated with the histological inflammatory score of osteoarthritis, chronic pyrophosphate arthropathy and RA [45]. In vivo experiments showed that TXNDC5-Tg mice are susceptible to collagen-induced arthritis (CIA); In vitro experiments showed that increased TXNDC5 expression in Rheumatoid arthritis synovial fibroblast-like cells (RASFs) could promote cell invasion, migration and secretion of TNF-α, IL-1α, IL-1β and IL-17. After treatment with anti-TXNDC5 siRNA, the above effects were weakened [46]. In addition, SNP analysis showed that 9 SNPs (rs9505298, rs41302895, rs1225936, rs1225938, rs372578, rs443861, rs408014, rs9392189, and rs2743992) were associated with RA [47].

In recent years, it has been found that TXNDC5 promotes the process of RA. TXNDC5 can protect synovial fibroblasts from the harmful effects of ER stress, thereby promoting the survival of cells [48]. In addition, TXNDC5 can promote angiogenesis and inflammation by inducing the secretion of vascular endothelial growth factor (VEGF), IL-6, and IL-8 [48]. TXNDC5 can inhibit the expression of insulin-like growth factor-binding protein-1 (IGFBP1). Insulin-like growth factor 1 (IGF-1) is a hormone with a structure similar to insulin, and it can bind to the IGF receptor. After binding, IGF exhibits tyrosine kinase activity. It can activate signalling pathways, such as phosphoinositide 3-kinase/protein kinase B (PI3K/AKT) and RAS/MAPK, thereby affecting cell proliferation, differentiation, survival and apoptosis [49]. TXNDC5 can also regulate NF-κB signalling by co-acting with the heat-shock-protein-70 (HSC70) [50]. In the absence of lipopolysaccharide (LPS), HSC70 can accelerate the degradation of IκBβ; In the presence of LPS, HSC70 can transfer more IκBβ into the nucleus [50]. TXNDC5 can regulate NF-κB signalling by retaining HSC70 in the cytoplasmic compartment. In addition, the expression of TXNDC5 can be regulated by miR573. Inhibiting miR-573 in RASFs can upregulate the expression of TXNDC5. TXNDC5 can enhance the migration of RASFs and cartilage destruction in RA by regulating the expression of matrix metalloproteinase-3 (MMP-3) [51]. The role and mechanism of TXNDC5 in RA mentioned above are shown in Table 4 and Fig. 2.

Table 4 Role and mechanism of TXNDC5 in RA
Fig. 2
figure 2

Role of TXNDC5 in rheumatoid arthritis (RA). TXNDC5 regulates RA progression, such as fibroblast proliferation and migration, inflammation, cartilage destruction and angiogenesis

TXNDC5 and other diseases

In addition to cancers and RA, the SNPs of TXNDC5 are also related to many other diseases. A study in Korean population found that three exon SNPs (rs1043784, rs7764128, and rs8643) in TXNDC5 are related to the pathogenesis/susceptibility of nonpigmented vitiligo [52]. The rs13873 SNP of the TXNDC5 gene and the rs1225934–rs13873 haplotype of the BMP6–TXNDC5 gene are related to sustained attention impairment in patients with schizophrenia [53]. In addition, a SNP related to TXNDC5 (rs13196892: TXNDC5 | MUTED) is one of the newly discovered menopausal/age-related SNPs in longevity populations [54]. These SNPs may cause disease by affecting the function of TXNDC5. In addition, TXNDC5 was found to be significantly increased in probands with potential new diseases, indicating that the gene may not easily tolerate copy damage [55]. Thus, the relationship between the SNPs of TXNDC5 and disease are worth further exploration.

An Arg345Trp (R345W) mutation in Fibrin-3 (F3) causes the rare, autosomal dominant macular dystrophy, Malattia Leventinese. John et al. found that TXNDC5 can interact with wild-type (WT) and R345W mutant F3 [56]. In vivo and in vitro experiments found that when interferon-stimulated gene 15 (Isg 15) was knocked out, the expression of TXNDC5 was reduced. It was suggested that Isg15 mediated its proviral role in hepatitis C virus replication through TXNDC5 [57]. Therefore, TXNDC5 is expected to serve as a target for gene or drug therapy in an effort to alter the fate of the above diseases.

TXNDC5 was increased in human and mouse atherosclerotic lesions. In addition, it could increase proteasome-mediated degradation of heat shock factor 1, leading to reduced heat shock protein 90 and accelerated eNOS (endothelial nitric oxide synthase) protein degradation. TXNDC5 deletion markedly increased eNOS protein and reduced atherosclerosis in ApoE −/− mice [58]. In addition, inhibiting the expression of TXNDC5 can reduce damage from isoproterenol to the heart, relieve fibrosis, and improve myocardial function [59]. In further research, it was found that in human cardiac fibroblasts (hCF), TGFβ1 can upregulate the expression of TXNDC5. TXNDC5 can promote ECM protein folding, thereby enhancing the JNK signaling pathway and promoting the activation and proliferation of hCF [59]. Similarly, the role of TXNDC5 in renal and pulmonary fibrosis via the TGF-β pathway has been revealed by related studies [60, 61]. In addition, it was found that inhibiting activity of IRE1α endoribonuclease can reduce the expression of TXNDC5 in activated fibroblasts and delay lung fibrosis induced by crystalline silica. [62].

Histone deacetylase inhibitor (HDACi) is an effective drug for cutaneous T-cell lymphoma (CTLC), but drug resistance often occurs in patients. The increased level of histone acetylation in drug-resistant CTLC patients is associated with increased expression of many genes, including TXNDC5 [63]. Therefore, the detection of TXNDC5 can be used to determine whether CTLC patients are resistant to HDACi. Analogously, the high expression of TXNDC5 in plasma cells of bortezomib-resistant refractory/relapsed multiple myeloma (RRMM) patients suggests that TXNDC5 may be a bortezomib resistance marker or a potential RRMM treatment target [64]. TXNDC5 expression was increased in the renal medulla samples of autosomal dominant tubulointerstitial kidney disease–UMOD model mice compared with normal mice [65]. In addition, the expression of TXNDC5 in peripheral blood mononuclear cells of the elderly is lower than that of young people, suggesting that expression changes in TXNDC5 may be related to the ageing of endothelial cells [66]. Kim et al. found that TXNDC5 expression decreased with age and may serve as a biomarker for skin ageing [67]. In addition, the TXNDC5 content in peripheral blood B cells can be used as a marker of the identification of seroconversion after COVID-19 vaccination [68]. Compared with wild-type mice, the expression of TXNDC5 were increased in the liver tissues of apoE-ko mice supplemented with squalene. The increase in TXNDC5 expression was related to liver fat content. This evidence suggests that TXNDC5 can be used as a target of squalene to combat hepatic steatosis [69]. The expression of TXNDC5 in lipid rafts of human umbilical vein endothelial cells treated with atorvastatin was upregulated, and the combination of TXNDC5 and Nox2 also increased. Atorvastatin may inhibit the activity of Nox2 by promoting the membrane translocation of TXNDC5 and the combination of TXNDC5 and Nox2, thereby exerting an antioxidant effect [70]. In addition, simvastatin enhances docetaxel-induced toxicity of human prostate cancer cells. In addition, TXNDC5 was found to be involved in this effect as the targets of simvastatin [71]. In summary, TXNDC5 plays an important role in various diseases, and it shows a potential in the diagnosis and treatment of diseases. The complex mechanisms and role of TXNDC5 in various diseases deserve further exploration.

Conclusions

TXNDC5 is a member of the PDI family. It has three Trx-like domains that can promote correct protein folding. In addition, it can protect cells from oxidative stress, promote proliferation and inhibit apoptosis. High expression of TXNDC5 can promote the progression of cancer, RA, fibrosis, atherosclerosis and other diseases. Current studies indicated that TXNDC5 has a promising future in the diagnosis and treatment of diseases. Targeting TXNDC5 has also shown a significant role in the treatment of diseases. However, the limitations of detection methods and inclusion in larger databases make its role in disease difficult to fully understand. In the future, it is of great significance to further clarify the role and mechanism of TXNDC5 in different diseases. From laboratory to practical application, the study on TXNDC5 needs to be validated by large-scale clinical trials.

Availability of data and materials

Not applicable as this manuscript is a review article.

Abbreviations

AdipoR1:

Adiponectin receptor

ADT:

Androgen deprivation therapy

AR:

Androgen receptor

CIA:

Collagen-induced arthritis

CRC:

Colorectal cancer

CRPC:

Castration-resistant prostate cancer

CTLC:

Cutaneous T-cell lymphoma

EDEM3:

ER degradation enhancing α-mannosidase-like protein 3

Ero1α:

Endoplasmic reticulum oxidoreductase 1α

ESCC:

Esophageal squamous cell carcinoma

HCC:

Hepatocellular carcinoma

hCF:

Human cardiac fibroblasts

HDACi:

Histone deacetylase inhibitor

HERG1:

Human ether a-go-go-related gene 1

HIF-1α:

Hypoxia inducible factor-1α

HSC70:

Heat-shock-protein-70

IGF-1:

Insulin-like growth factor 1

IGFBP1:

Insulin-like growth factor-binding protein-1

Isg 15:

Interferon-stimulated gene 15

LPS:

Lipopolysaccharide

LSCC:

Laryngeal squamous cell carcinoma

MMP-3:

Matrix metalloproteinase-3

NR4A1:

Nuclear receptor 4A1

PDI:

Protein disulfide isomerase

Prx4:

Peroxiredoxin-4

RA:

Rheumatoid arthritis

RASFs:

Rheumatoid arthritis synovial fibroblast-like cells

RCC:

Renal cell carcinoma

RMS:

Rhabdomyosarcoma

RRMM:

Refractory/relapsed multiple myeloma

SNP:

Single nucleotide polymorphism

Srx:

Sulfiredoxin

TXNDC5:

Thioredoxin domain-containing protein 5

VEGF:

Vascular endothelial growth factor

VG:

Variant gastritis

References

  1. Sullivan DC, Huminiecki L, Moore JW, Boyle JJ, Poulsom R, Creamer D, et al. EndoPDI, a novel protein-disulfide isomerase-like protein that is preferentially expressed in endothelial cells acts as a stress survival factor. J Biol Chem. 2003;278(47):47079–88. https://doi.org/10.1074/jbc.M308124200.

    Article  CAS  PubMed  Google Scholar 

  2. Knoblach B, Keller BO, Groenendyk J, Aldred S, Zheng J, Lemire BD, et al. ERp19 and ERp46, new members of the thioredoxin family of endoplasmic reticulum proteins. Mol Cell Proteom. 2003;2(10):1104–19. https://doi.org/10.1074/mcp.M300053-MCP200.

    Article  CAS  Google Scholar 

  3. Chawsheen HA, Ying Q, Jiang H, Wei Q. A critical role of the thioredoxin domain containing protein 5 (TXNDC5) in redox homeostasis and cancer development. Genes Dis. 2018;5(4):312–22. https://doi.org/10.1016/j.gendis.2018.09.003.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Horna-Terron E, Pradilla-Dieste A, Sanchez-de-Diego C, Osada J. TXNDC5, a newly discovered disulfide isomerase with a key role in cell physiology and pathology. Int J Mol Sci. 2014;15(12):23501–18. https://doi.org/10.3390/ijms151223501.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Tian G, Xiang S, Noiva R, Lennarz WJ, Schindelin H. The crystal structure of yeast protein disulfide isomerase suggests cooperativity between its active sites. Cell. 2006;124(1):61–73. https://doi.org/10.1016/j.cell.2005.10.044.

    Article  CAS  PubMed  Google Scholar 

  6. Klappa P, Ruddock LW, Darby NJ, Freedman RB. The b’ domain provides the principal peptide-binding site of protein disulfide isomerase but all domains contribute to binding of misfolded proteins. EMBO J. 1998;17(4):927–35. https://doi.org/10.1093/emboj/17.4.927.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Pirneskoski A, Klappa P, Lobell M, Williamson RA, Byrne L, Alanen HI, et al. Molecular characterization of the principal substrate binding site of the ubiquitous folding catalyst protein disulfide isomerase. J Biol Chem. 2004;279(11):10374–81. https://doi.org/10.1074/jbc.M312193200.

    Article  CAS  PubMed  Google Scholar 

  8. Irvine AG, Wallis AK, Sanghera N, Rowe ML, Ruddock LW, Howard MJ, et al. Protein disulfide-isomerase interacts with a substrate protein at all stages along its folding pathway. PLoS ONE. 2014;9(1):e82511. https://doi.org/10.1371/journal.pone.0082511.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Kozlov G, Määttänen P, Thomas DY, Gehring K. A structural overview of the PDI family of proteins. FEBS J. 2010;277(19):3924–36. https://doi.org/10.1111/j.1742-4658.2010.07793.x.

    Article  CAS  PubMed  Google Scholar 

  10. Kojima R, Okumura M, Masui S, Kanemura S, Inoue M, Saiki M, et al. Radically different thioredoxin domain arrangement of ERp46, an efficient disulfide bond introducer of the mammalian PDI family. Structure. 2014;22(3):431–43. https://doi.org/10.1016/j.str.2013.12.013.

    Article  CAS  PubMed  Google Scholar 

  11. Funkner A, Parthier C, Schutkowski M, Zerweck J, Lilie H, Gyrych N, et al. Peptide binding by catalytic domains of the protein disulfide isomerase-related protein ERp46. J Mol Biol. 2013;425(8):1340–62. https://doi.org/10.1016/j.jmb.2013.01.029.

    Article  CAS  PubMed  Google Scholar 

  12. Gulerez IE, Kozlov G, Rosenauer A, Gehring K. Structure of the third catalytic domain of the protein disulfide isomerase ERp46. Acta Crystallogr Sect F Struct Biol Cryst Commun. 2012;68(Pt 4):378–81. https://doi.org/10.1107/S1744309112005866.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Mavridou DAI, Saridakis E, Kritsiligkou P, Mozley EC, Ferguson SJ, Redfield C. An extended active-site motif controls the reactivity of the thioredoxin fold. J Biol Chem. 2014;289(12):8681–96. https://doi.org/10.1074/jbc.M113.513457.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Karala AR, Lappi AK, Ruddock LW. Modulation of an active-site cysteine pKa allows PDI to act as a catalyst of both disulfide bond formation and isomerization. J Mol Biol. 2010;396(4):883–92. https://doi.org/10.1016/j.jmb.2009.12.014.

    Article  CAS  PubMed  Google Scholar 

  15. Okada S, Matsusaki M, Arai K, Hidaka Y, Inaba K, Okumura M, et al. Coupling effects of thiol and urea-type groups for promotion of oxidative protein folding. Fold Chem Commun (Camb). 2019;55(6):759–62. https://doi.org/10.1039/c8cc08657e.

    Article  CAS  Google Scholar 

  16. Okumura M, Shimamoto S, Hidaka Y. A chemical method for investigating disulfide-coupled peptide and protein folding. FEBS J. 2012;279(13):2283–95. https://doi.org/10.1111/j.1742-4658.2012.08596.x.

    Article  CAS  PubMed  Google Scholar 

  17. Jessop CE, Watkins RH, Simmons JJ, Tasab M, Bulleid NJ. Protein disulphide isomerase family members show distinct substrate specificity: P5 is targeted to BiP client proteins. J Cell Sci. 2009;122(Pt 23):4287–95. https://doi.org/10.1242/jcs.059154.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Pace Paul E, Peskin Alexander V, Han MH, Hampton Mark B, Winterbourn CC. Hyperoxidized peroxiredoxin 2 interacts with the protein disulfide- isomerase ERp46. Biochem J. 2013;453(3):475–85. https://doi.org/10.1042/bj20130030.

    Article  CAS  PubMed  Google Scholar 

  19. Sato Y, Kojima R, Okumura M, Hagiwara M, Masui S, Maegawa K, et al. Synergistic cooperation of PDI family members in peroxiredoxin 4-driven oxidative protein folding. Sci Rep. 2013;3:2456. https://doi.org/10.1038/srep02456.

    Article  PubMed  PubMed Central  Google Scholar 

  20. Camargo LL, Babelova A, Mieth A, Weigert A, Mooz J, Rajalingam K, et al. Endo-PDI is required for TNFalpha-induced angiogenesis. Free Radic Biol Med. 2013;65:1398–407. https://doi.org/10.1016/j.freeradbiomed.2013.09.028.

    Article  CAS  PubMed  Google Scholar 

  21. Chang X, Xu B, Wang L, Wang Y, Wang Y, Yan S. Investigating a pathogenic role for TXNDC5 in tumors. Int J Oncol. 2013;43(6):1871–84. https://doi.org/10.3892/ijo.2013.2123.

    Article  CAS  PubMed  Google Scholar 

  22. Freedman RB, Hirst TR, Tuite MF. Protein disulphide isomerase: building bridges in protein folding. Trends Biochem Sci. 1994;19(8):331–6. https://doi.org/10.1016/0968-0004(94)90072-8.

    Article  CAS  PubMed  Google Scholar 

  23. Yu S, Ito S, Wada I, Hosokawa N. ER-resident protein 46 (ERp46) triggers the mannose-trimming activity of ER degradation-enhancing alpha-mannosidase-like protein 3 (EDEM3). J Biol Chem. 2018;293(27):10663–74. https://doi.org/10.1074/jbc.RA118.003129.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Wu Z, Zhang L, Li N, Sha L, Zhang K. An immunohistochemical study of thioredoxin domain-containing 5 expression in gastric adenocarcinoma. Oncol Lett. 2015;9(3):1154–8. https://doi.org/10.3892/ol.2014.2832.

    Article  PubMed  Google Scholar 

  25. Zhang L, Hou Y, Li N, Wu K, Zhai J. The influence of TXNDC5 gene on gastric cancer cell. J Cancer Res Clin Oncol. 2010;136(10):1497–505. https://doi.org/10.1007/s00432-010-0807-x.

    Article  CAS  PubMed  Google Scholar 

  26. Xu B, Li J, Liu X, Li C, Chang X. TXNDC5 is a cervical tumor susceptibility gene that stimulates cell migration, vasculogenic mimicry and angiogenesis by down-regulating SERPINF1 and TRAF1 expression. Oncotarget. 2017;8(53):91009–24. https://doi.org/10.18632/oncotarget.18857.

    Article  PubMed  PubMed Central  Google Scholar 

  27. Park MS, Kim SK, Shin HP, Lee SM, Chung JH. TXNDC5 gene polymorphism contributes to increased risk of hepatocellular carcinoma in the Korean male population. Anticancer Res. 2013;33(9):3983–7.

    CAS  PubMed  Google Scholar 

  28. Tan F, Zhu H, He X, Yu N, Zhang X, Xu H, et al. Role of TXNDC5 in tumorigenesis of colorectal cancer cells: In vivo and in vitro evidence. Int J Mol Med. 2018;42(2):935–45. https://doi.org/10.3892/ijmm.2018.3664.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Wang L, Song G, Chang X, Tan W, Pan J, Zhu X, et al. The role of TXNDC5 in castration-resistant prostate cancer—involvement of androgen receptor signaling pathway. Oncogene. 2014;34(36):4735–45. https://doi.org/10.1038/onc.2014.401.

    Article  CAS  PubMed  Google Scholar 

  30. Wei Q, Jiang H, Xiao Z, Baker A, Young MR, Veenstra TD, et al. Sulfiredoxin Peroxiredoxin IV axis promotes human lung cancer progression through modulation of specific phosphokinase signaling. Proc Natl Acad Sci USA. 2011;108(17):7004–9. https://doi.org/10.1073/pnas.1013012108.

    Article  PubMed  PubMed Central  Google Scholar 

  31. Chawsheen HA, Jiang H, Ying Q, Ding N, Thapa P, Wei Q. The redox regulator sulfiredoxin forms a complex with thioredoxin domain-containing 5 protein in response to ER stress in lung cancer cells. J Biol Chem. 2019;294(22):8991–9006. https://doi.org/10.1074/jbc.RA118.005804.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Charlton HK, Webster J, Kruger S, Simpson F, Richards AA, Whitehead JP. ERp46 binds to AdipoR1, but not AdipoR2, and modulates adiponectin signalling. Biochem Biophys Res Commun. 2010;392(2):234–9. https://doi.org/10.1016/j.bbrc.2010.01.029.

    Article  CAS  PubMed  Google Scholar 

  33. Duivenvoorden WC, Paschos A, Hopmans SN, Austin RC, Pinthus JH. Endoplasmic reticulum protein ERp46 in renal cell carcinoma. PLoS ONE. 2014;9(3):e90389. https://doi.org/10.1371/journal.pone.0090389.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Vincent EE, Elder DJ, Phillips L, Heesom KJ, Pawade J, Luckett M, et al. Overexpression of the TXNDC5 protein in non-small cell lung carcinoma. Anticancer res. 2011;31(5):1577–82.

    CAS  PubMed  Google Scholar 

  35. Zang H, Li Y, Zhang X, Huang G. Circ_0000517 contributes to hepatocellular carcinoma progression by upregulating TXNDC5 via sponging miR-1296–5p. Cancer Manag Res. 2020;12:3457–68. https://doi.org/10.2147/cmar.s244024.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Yu J, Yang M, Zhou B, Luo J, Zhang Z, Zhang W, et al. CircRNA-104718 acts as competing endogenous RNA and promotes hepatocellular carcinoma progression through microRNA-218–5p/TXNDC5 signaling pathway. Clin Sci. 2019;133(13):1487–503. https://doi.org/10.1042/CS20190394.

    Article  CAS  Google Scholar 

  37. Wang H, Yang X, Guo Y, Shui L, Li S, Bai Y, et al. HERG1 promotes esophageal squamous cell carcinoma growth and metastasis through TXNDC5 by activating the PI3K/AKT pathway. J Exp Clin Cancer Res. 2019;38(1):324. https://doi.org/10.1186/s13046-019-1284-y.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Lee SO, Jin UH, Kang JH, Kim SB, Guthrie AS, Sreevalsan S, et al. The orphan nuclear receptor NR4A1 (Nur77) regulates oxidative and endoplasmic reticulum stress in pancreatic cancer cells. Mol Cancer Res. 2014;12(4):527–38. https://doi.org/10.1158/1541-7786.MCR-13-0567.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Hedrick E, Lee SO, Kim G, Abdelrahim M, Jin UH, Safe S, et al. Nuclear receptor 4A1 (NR4A1) as a drug target for renal cell adenocarcinoma. PLoS ONE. 2015;10(6):e0128308. https://doi.org/10.1371/journal.pone.0128308.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Hedrick E, Mohankumar K, Lacey A, Safe S. Inhibition of NR4A1 promotes ROS accumulation and IL24-dependent growth arrest in rhabdomyosarcoma. Mol Cancer Res. 2019;17(11):2221–32. https://doi.org/10.1158/1541-7786.MCR-19-0408.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Nissom PM, Lo SL, Lo JC, Ong PF, Lim JW, Ou K, et al. Hcc-2, a novel mammalian ER thioredoxin that is differentially expressed in hepatocellulr carcinoma. FEBS Lett. 2006;580(9):2216–26. https://doi.org/10.1016/j.febslet.2006.03.029.

    Article  CAS  PubMed  Google Scholar 

  42. Mo R, Peng J, Xiao J, Ma J, Li W, Wang J, et al. High TXNDC5 expression predicts poor prognosis in renal cell carcinoma. Tumor Biol. 2016;37(7):9797–806. https://doi.org/10.1007/s13277-016-4891-7.

    Article  CAS  Google Scholar 

  43. Peng F, Zhang H, Du Y, Tan P. Cetuximab enhances cisplatin-induced endoplasmic reticulum stress-associated apoptosis in laryngeal squamous cell carcinoma cells by inhibiting expression of TXNDC5. Mol Med Rep. 2018;17(3):4767–76. https://doi.org/10.3892/mmr.2018.8376.

    Article  CAS  PubMed  Google Scholar 

  44. Chang X, Cui Y, Zong M, Zhao Y, Yan X, Chen Y, et al. Identification of proteins with increased expression in rheumatoid arthritis synovial tissues. J Rheumatol. 2009;36(5):872–80. https://doi.org/10.3899/jrheum.080939.

    Article  CAS  PubMed  Google Scholar 

  45. de Seny D, Bianchi E, Baiwir D, Cobraiville G, Collin C, Deliège M, et al. Proteins involved in the endoplasmic reticulum stress are modulated in synovitis of osteoarthritis, chronic pyrophosphate arthropathy and rheumatoid arthritis, and correlate with the histological inflammatory score. Sci Rep. 2020;10(1):14159. https://doi.org/10.1038/s41598-020-70803-7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Wang L, Zheng Y, Xu H, Yan X, Chang X. Investigate pathogenic mechanism of TXNDC5 in rheumatoid arthritis. PLoS ONE. 2013;8(1):e53301. https://doi.org/10.1371/journal.pone.0053301.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Chang X, Zhao Y, Yan X, Pan J, Fang K, Wang L. Investigating a pathogenic role for TXNDC5 in rheumatoid arthritis. Arthritis Res Ther. 2011;13(4):R124. https://doi.org/10.1186/ar3429.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Lu Q, Wang J, Zhang X, Tian R, Qiao L, Ge L, et al. TXNDC5 protects synovial fibroblasts of rheumatoid arthritis from the detrimental effects of endoplasmic reticulum stress. Intractable Rare Dis Res. 2020;9(1):23–9. https://doi.org/10.5582/irdr.2019.01139.

    Article  PubMed  PubMed Central  Google Scholar 

  49. Li J, Xu B, Wu C, Yan X, Zhang L, Chang X. TXNDC5 contributes to rheumatoid arthritis by down-regulating IGFBP1 expression. Clin Exp Immunol. 2018;192(1):82–94. https://doi.org/10.1111/cei.13080.

    Article  CAS  PubMed  Google Scholar 

  50. Wang L, Dong H, Song G, Zhang R, Pan J, Han J. TXNDC5 synergizes with HSC70 to exacerbate the inflammatory phenotype of synovial fibroblasts in rheumatoid arthritis through NF-κB signaling. Cell Mol Immunol. 2018;15(7):685–96. https://doi.org/10.1038/cmi.2017.20.

    Article  CAS  PubMed  Google Scholar 

  51. Wang L, Song G, Zheng Y, Wang D, Dong H, Pan J, et al. miR-573 is a negative regulator in the pathogenesis of rheumatoid arthritis. Cell Mol Immunol. 2015;13(6):839–49. https://doi.org/10.1038/cmi.2015.63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Jeong KH, Shin MK, Uhm YK, Kim HJ, Chung JH, Lee MH. Association of TXNDC5 gene polymorphisms and susceptibility to nonsegmental vitiligo in the Korean population. Br J Dermatol. 2010;162(4):759–64. https://doi.org/10.1111/j.1365-2133.2009.09574.x.

    Article  CAS  PubMed  Google Scholar 

  53. Lin SH, Liu CM, Liu YL, Shen-Jang Fann C, Hsiao PC, Wu JY, et al. Clustering by neurocognition for fine mapping of the schizophrenia susceptibility loci on chromosome 6p. Genes Brain Behav. 2009;8(8):785–94. https://doi.org/10.1111/j.1601-183X.2009.00523.x.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Bae H, Lunetta KL, Murabito JM, Andersen SL, Schupf N, Perls T, et al. Genetic associations with age of menopause in familial longevity. Menopause. 2019;26(10):1204–12. https://doi.org/10.1097/GME.0000000000001367.

    Article  PubMed  PubMed Central  Google Scholar 

  55. Murcia Pienkowski V, Kucharczyk M, Młynek M, Szczałuba K, Rydzanicz M, Poszewiecka B, et al. Mapping of breakpoints in balanced chromosomal translocations by shallow whole-genome sequencing points to EFNA5, BAHD1 and PPP2R5E as novel candidates for genes causing human mendelian disorders. J Med genet. 2019;56(2):104–12. https://doi.org/10.1136/jmedgenet-2018-105527.

    Article  CAS  PubMed  Google Scholar 

  56. Hulleman JD, Genereux JC, Nguyen A. Mapping wild-type and R345W fibulin-3 intracellular interactomes. Exp Eye Res. 2016;153:165–9. https://doi.org/10.1016/j.exer.2016.10.017.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Real CI, Megger DA, Sitek B, Jahn-Hofmann K, Ickenstein LM, John MJ, et al. Identification of proteins that mediate the pro-viral functions of the interferon stimulated gene 15 in hepatitis C virus replication. Antiviral Res. 2013;100(3):654–61. https://doi.org/10.1016/j.antiviral.2013.10.009.

    Article  CAS  PubMed  Google Scholar 

  58. Yeh CF, Cheng SH, Lin YS, Shentu TP, Huang RT, Zhu J, et al. Targeting mechanosensitive endothelial TXNDC5 to stabilize eNOS and reduce atherosclerosis in vivo. Sci Adv. 2022;8(3):eabl8096. https://doi.org/10.1126/sciadv.abl8096.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Shih YC, Chen CL, Zhang Y, Mellor RL, Kanter EM, Fang Y, et al. Endoplasmic reticulum protein TXNDC5 augments myocardial fibrosis by facilitating extracellular matrix protein folding and redox-sensitive cardiac fibroblast activation. Circ Res. 2018;122(8):1052–68. https://doi.org/10.1161/CIRCRESAHA.117.312130.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Chen YT, Jhao PY, Hung CT, Wu YF, Lin SJ, Chiang WC, et al. Endoplasmic reticulum protein TXNDC5 promotes renal fibrosis by enforcing TGF-β signaling in kidney fibroblasts. J Clin Investig. 2021;131(5):e143645. https://doi.org/10.1172/jci143645.

    Article  CAS  PubMed Central  Google Scholar 

  61. Lee TH, Yeh CF, Lee YT, Shih YC, Chen YT, Hung CT, et al. Fibroblast-enriched endoplasmic reticulum protein TXNDC5 promotes pulmonary fibrosis by augmenting TGFβ signaling through TGFBR1 stabilization. Nature commun. 2020;11(1):4254. https://doi.org/10.1038/s41467-020-18047-x.

    Article  CAS  Google Scholar 

  62. Chen X, Li C, Liu J, He Y, Wei Y, Chen J. Inhibition of ER stress by targeting the IRE1α-TXNDC5 pathway alleviates crystalline silica-induced pulmonary fibrosis. Int Immunopharmacol. 2021;95:107519. https://doi.org/10.1016/j.intimp.2021.107519.

    Article  CAS  PubMed  Google Scholar 

  63. Andrews JM, Schmidt JA, Carson KR, Musiek AC, Mehta-Shah N, Payton JE. Novel cell adhesion/migration pathways are predictive markers of HDAC inhibitor resistance in cutaneous T cell lymphoma. EBioMedicine. 2019;46:170–83. https://doi.org/10.1016/j.ebiom.2019.07.053.

    Article  PubMed  PubMed Central  Google Scholar 

  64. Dytfeld D, Luczak M, Wrobel T, Usnarska-Zubkiewicz L, Brzezniakiewicz K, Jamroziak K, et al. Comparative proteomic profiling of refractory/relapsed multiple myeloma reveals biomarkers involved in resistance to bortezomib-based therapy. Oncotarget. 2016;7(35):56726–36. https://doi.org/10.18632/oncotarget.11059.

    Article  PubMed  PubMed Central  Google Scholar 

  65. Kemter E, Frohlich T, Arnold GJ, Wolf E, Wanke R. Mitochondrial dysregulation secondary to endoplasmic reticulum stress in autosomal dominant tubulointerstitial kidney disease—UMOD (ADTKD-UMOD). Sci Rep. 2017;7:42970. https://doi.org/10.1038/srep42970.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Noh JY, Oh SH, Lee JH, Kwon YS, Ryu DJ, Lee KH. Can blood components with age-related changes influence the ageing of endothelial cells? Exp Dermatol. 2010;19(4):339–46. https://doi.org/10.1111/j.1600-0625.2009.01010.x.

    Article  CAS  PubMed  Google Scholar 

  67. Kim SH, Kim JH, Suk JM, Lee YI, Kim J, Lee JH, et al. Identification of skin aging biomarkers correlated with the biomechanical properties. Skin Res Technol. 2021;27(5):940–7. https://doi.org/10.1111/srt.13046.

    Article  PubMed  Google Scholar 

  68. Huang D, Liu AYN, Leung KS, Tang NLS. Direct measurement of b lymphocyte gene expression biomarkers in peripheral blood transcriptomics enables early prediction of vaccine seroconversion. Genes. 2021;12(7):971. https://doi.org/10.3390/genes12070971.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Ramirez-Torres A, Barcelo-Batllori S, Martinez-Beamonte R, Navarro MA, Surra JC, Arnal C, et al. Proteomics and gene expression analyses of squalene-supplemented mice identify microsomal thioredoxin domain-containing protein 5 changes associated with hepatic steatosis. J Proteom. 2012;77:27–39. https://doi.org/10.1016/j.jprot.2012.07.001.

    Article  CAS  Google Scholar 

  70. Gu MX, Fu Y, Sun XL, Ding YZ, Li CH, Pang W, et al. Proteomic analysis of endothelial lipid rafts reveals a novel role of statins in antioxidation. J Proteom Res. 2012;11(4):2365–73. https://doi.org/10.1021/pr300098f.

    Article  CAS  Google Scholar 

  71. Jung EJ, Chung KH, Kim CW. Identification of simvastatin-regulated targets associated with JNK activation in DU145 human prostate cancer cell death signaling. BMB Rep. 2017;50(9):466–71. https://doi.org/10.5483/bmbrep.2017.50.9.087.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

No funding was received for conducting this study.

Author information

Authors and Affiliations

Authors

Contributions

XW and XC had the idea for the article, XW and HL performed the literature search and drafted the work, and XC critically revised the work. All authors reviewed the manuscript. All author read and approved the final manuscript.

Corresponding author

Correspondence to Xiaotian Chang.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher's Note

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

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wang, X., Li, H. & Chang, X. The role and mechanism of TXNDC5 in diseases. Eur J Med Res 27, 145 (2022). https://doi.org/10.1186/s40001-022-00770-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40001-022-00770-4

Keywords