Skip to main content

Severe pantothenic acid deficiency induces alterations in the intestinal mucosal proteome of starter Pekin ducks

Abstract

Background

Pantothenic acid deficiency (PAD) results in growth depression and intestinal hypofunction of animals. However, the underlying molecular mechanisms remain to be elucidated. Mucosal proteome might reflect dietary influences on physiological processes.

Results

A total of 128 white Pekin ducks of one-day-old were randomly assigned to two groups, fed either a PAD or a pantothenic acid adequate (control, CON) diet. After a 16-day feeding period, two ducks from each replicate were sampled to measure plasma parameters, intestinal morphology, and mucosal proteome. Compared to the CON group, high mortality, growth retardation, fasting hypoglycemia, reduced plasma insulin, and oxidative stress were observed in the PAD group. Furthermore, PAD induced morphological alterations of the small intestine indicated by reduced villus height and villus surface area of duodenum, jejunum, and ileum. The duodenum mucosal proteome of ducks showed that 198 proteins were up-regulated and 223 proteins were down-regulated (> 1.5-fold change) in the PAD group compared to those in the CON group. Selected proteins were confirmed by Western blotting. Pathway analysis of these proteins exhibited the suppression of glycolysis and gluconeogenesis, fatty acid beta oxidation, tricarboxylic acid cycle, oxidative phosphorylation, oxidative stress, and intestinal absorption in the PAD group, indicating impaired energy generation and abnormal intestinal absorption. We also show that nine out of eleven proteins involved in regulation of actin cytoskeleton were up-regulated by PAD, probably indicates reduced intestinal integrity.

Conclusion

PAD leads to growth depression and intestinal hypofunction of ducks, which are associated with impaired energy generation, abnormal intestinal absorption, and regulation of actin cytoskeleton processes. These findings provide insights into the mechanisms of intestinal hypofunction induced by PAD.

Background

Pantothenic acid is a precursor of two coenzymes, coenzyme A (CoA) and acyl-carrier-protein. The coenzymes of this vitamin participate in various metabolic reactions, such as glucose, fatty acids and amino acids entering into energy-yielding tricarboxylic acid (TCA) cycle, fatty acid oxidation and synthesis, cholesterol synthesis, acetylcholine synthesis, and heme synthesis etc. [1, 2]. Its importance is highlighted by the adverse effects of pantothenic acid deficiency (PAD) in mammals such as rats, cats, and pigs, including growth depression, skin lesions, diarrhea, loss of hair [2,3,4,5,6]. Also in poultry studies, PAD results in growth retardation, poor feathering, dermatosis, and high mortality in chicks, turkeys, geese, and ducks [7,8,9,10,11,12,13]. It has been demonstrated extensively that pantothenic acid can keep the structure of intestine integrity and maintain the intestinal function of animals [3, 14,15,16]. Intestinal hypofunction was shown to be a major consequence of PAD in rats, dogs, cats, chicks, and fish, such as intestinal ulceration, diarrhea, and colitis [4, 5, 17,18,19,20,21]. Rats deficient in pantothenic acid exhibit duodenitis and duodenal ulcers [18], as well as duodenal changes including eventual atrophy of crypts, diminution in size of villi [16]. In addition, previous studies in fish showed that PAD decreased intestinal digestive and absorptive capacities by reducing the activities of both intestinal brush border enzymes and digestive enzymes [14, 21].

In addition, pantothenic acid can protect cell membrane against peroxidative damage by increasing glutathione content [22,23,24]. Previous studies have shown that dietary PAD could lead to oxidative stress in geese [9] and fish [21]. It is proposed that oxidative stress induced by PAD leads to intestinal injury and hypofunction. So far, the detailed mechanisms of growth depression and intestinal hypofunction due to PAD are still unclear. In order to understand the underlying mechanisms, we established a PAD duck model, identified an overview of underlying processes and the extend of alterations of intestinal mucosa using a proteomic approach.

Results

Mortality and growth performance

The mortality of PAD ducks was greater than the control (CON) birds (P < 0.001; Table 1). In comparison with the CON group, average daily weight gain (ADG) and average daily feed intake (ADFI) were declined in the PAD group, while feed conversion ratio (FCR) was increased (P < 0.001; Table 1).

Table 1 Growth performance on day 16 of ducks in the pantothenic acid-deficient (PAD) and control (CON) group

Plasma parameters

Plasma pantothenic acid concentration decreased by 85%, while plasma alkaline phosphatase (ALP) activity decreased by approximately 70% as a result of PAD when compared to the CON group (P < 0.001; Table 2). Compared to the CON group, plasma glucose and insulin contents were declined in the PAD group (P < 0.01; Table 2). Plasma glucagon did not differ between the PAD and CON group (P > 0.05; Table 2). Compared to the CON group, plasma malondialdehyde (MDA) content was increased, while plasma total superoxide dismutase (T-SOD) activity was decreased in the PAD group (P < 0.001; Table 2).

Table 2 Plasma parameters of 16-day-old ducks in the pantothenic acid-deficient (PAD) and control (CON) group

Intestinal morphology analyses

Dietary PAD resulted in morphological alterations of the small intestine of Pekin ducks. Compared to the CON group, PAD reduced villus height and villus surface area of duodenum, jejunum, and ileum (P < 0.001; Table 3). Villus width and crypt depth of both duodenum and jejunum were not affected by PAD in ducks (P > 0.05; Table 3). Compared to the CON group, PAD decreased crypt depth of ileum in ducks (P < 0.001; Table 3) but not villus width of ileum (P > 0.05; Table 3).

Table 3 Intestinal mucosal histomorphology of 16-day-old ducks in the pantothenic acid-deficient (PAD) and control (CON) group

Changes in the intestinal mucosal proteomics of duck in response to PAD

A total of 22,973 peptide belonging to 3345 proteins were identified in the duodenum mucosa of two groups. Comparisons of the relative abundance of proteins from mucosa of PAD ducks with those of CON ducks showed that a total of 421 proteins showed a fold change (FC) > 1.5, of which 198 proteins were up-regulated and 223 proteins were down-regulated. The complete list of proteins altered by PAD is presented in Additional file 1.

All the differentially expressed proteins in mucosa in response to PAD were used to conduct GO categories of biological process, cellular component, and molecular function, and pathway analysis. As shown in Fig. 1, the top 15 enriched terms under biological process included oxidation-reduction process, small molecule metabolic process, generation of precursor metabolites and energy, purine nucleoside triphosphate metabolic process, monocarboxylic acid metabolic process, nucleoside triphosphate metabolic process, carboxylic acid metabolic process, ATP metabolic process, oxoacid metabolic process, ribonucleoside monophosphate metabolic process, organic acid metabolic process, nucleoside monophosphate metabolic process, ribonucleoside triphosphate metabolic process, coenzyme metabolic process, and cellular respiration. The top 15 enriched terms under cellular component included extracellular exosome, extracellular vesicle, extracellular organelle, vesicle, mitochondrion, cytoplasmic part, myelin sheath, cytoplasm, brush border, adherens junction, organelle inner membrane, focal adhesion, cell-substrate adherens junction, mitochondrial inner membrane, and cluster of actin-based cell projections. The top 15 enriched terms under molecular function included cell adhesion molecule binding, oxidoreductase activity, coenzyme binding, identical protein binding, cadherin binding, actin filament binding, actin binding, aminopeptidase activity, RNA binding, fatty-acyl-CoA binding, exopeptidase activity, cytoskeletal protein binding, hydro-lyase activity, glutathione transferase activity, and S100 protein binding.

Fig. 1
figure 1

Top 15 significantly terms enriched biological processes, cellular components, and molecular functions

Based on the pathway analysis by Kyoto Encyclopedia of Genes and Genomes (KEGG), the significantly affected pathways were glycolysis and gluconeogenesis, amino acid metabolism, cori cycle, fatty acid beta oxidation, striated muscle contraction, TCA cycle, trans-sulfuration pathway, PPAR signaling pathway, cytoplasmic ribosomal proteins, vitamin A and carotenoid metabolism, oxidative stress, regulation of actin cytoskeleton, glutathione metabolism, and oxidative phosphorylation (Fig. 2). Among them, energy metabolism (glycolysis and gluconeogenesis, fatty acid beta oxidation, TCA cycle, and oxidative phosphorylation), regulation of actin cytoskeleton, and oxidative stress were selected in the present study and the proteins involved in these processes are listed in Table 4.

Fig. 2
figure 2

The pathway analysis by KEGG on differentially expressed proteins

Table 4 Selected differentially expressed proteins in duodenum mucosa caused by pantothenic acid deficiency

Of the proteins enriched in glycolysis and gluconeogenesis, eleven proteins were down-regulated (fructose-bisphosphate aldolase (ALDOB), hexokinase domain containing 1 (HKDC1), alpha-enolase (ENO1), fructose-bisphosphatase 1 (FBP1), glyceraldehyde-3-phosphate dehydrogenase (GAPDH), pyruvate kinase (PKM), ATP-dependent 6-phosphofructokinase (PFKP), triosephosphate isomerase (TPI1), phosphoglycerate kinase (PGK1), dihydrolipoyl dehydrogenase (DLD), and pyruvate dehydrogenase E1 beta subunit (PDHB)), while two proteins were up-regulated (fructose-bisphosphate aldolase (ALDOA) and hexokinase 1 (HK1)).

Nine proteins involved in fatty acid beta oxidation were all down-regulated in the PAD group, including acyl-CoA synthetase long chain family member 5 (ACSL5), acyl-CoA dehydrogenase family member 11 (ACAD11), medium-chain specific acyl-CoA dehydrogenase (ACADM), carnitine O-acetyltransferase (CRAT), DLD, hydroxyacyl-CoA dehydrogenase trifunctional multienzyme complex subunit beta (HADHB), enoyl-CoA delta isomerase 2 (ECI2), acyl-coenzyme A oxidase (ACOX1), and acyl-coenzyme A oxidase (ACOX2).

Of the proteins enriched in TCA cycle, six proteins were down-regulated (DLD, aconitase 1 (Fragment) (ACO1), aconitate hydratase, mitochondrial (ACO2), malate dehydrogenase (fragment) (MDH1), malate dehydrogenase (MDH2), and pyruvate dehydrogenase E1 beta subunit (PDHB)), while one protein was up-regulated (isocitrate dehydrogenase [NADP] (IDH1)).

Five proteins were involved in the oxidative phosphorylation, which were all down-regulated in the PAD group (NADH:ubiquinone oxidoreductase subunit A5 (NDUFA5), NADH dehydrogenase [ubiquinone] 1 alpha subcomplex subunit 6 (Fragment) (NDUFA6), ATP synthase, H+ transporting, mitochondrial Fo complex subunit B1 (ATP5F1), ATP synthase subunit d, mitochondrial (Fragment) (ATP5H), and ATP synthase subunit O, mitochondrial (Fragment) (ATP5O)).

Of the proteins involved in regulation of actin cytoskeleton, two proteins were down-regulated (Villin 1 (VIL1) and Ezrin (EZR)), while nine proteins were up-regulated (vimentin (VIM), tropomyosin 3 (TPM3), tropomyosin alpha-1 chain (TPM1), myosin light chain 1 (MYL1), myosin light chain kinase (Fragment) (MYLK), actinin alpha 1 (ACTN1), fibronectin 1 (FN1), vinculin (VCL), and moesin (MSN)).

Of the proteins involved in oxidative stress, three proteins were all down-regulated (amine oxidase (MAOA), catalase (CAT), and microsomal glutathione S-transferase 1 (Fragment) (MGST1)), while one protein was up-regulated (glutamate-cysteine ligase catalytic subunit (GCLC)).

Besides, we found that four proteins were involved in intestinal absorption, which were all down-regulated by PAD, including solute carrier family 2 member 2 (SLC2A2), VIL1, EZR, and monoacylglycerol O-acyltransferase 2 (MOGAT2).

Western blot analyses

To validate the isobaric tags for relative and absolute quantification (iTRAQ) results, the abundance of ACADM and GAPDH were analyzed by Western blot. Compared to the CON group, the protein expressions of ACADM and GAPDH were decreased in the PAD group (Fig. 3), which were consistent with the iTRAQ results.

Fig. 3
figure 3

Western blot analysis of medium-chain-specific acyl-CoA dehydrogenase (ACADM; A) and glyceraldehyde-3-phosphate dehydrogenase (GAPDH; B) protein abundance of mucosal tissue of ducks in the pantothenic acid deficient (PAD) and Control (CON) groups. Loading control, histone H3, was used to normalize the levels of ACADM and GAPDH. Representative Western blots are shown. The images of complete Western blots are shown in Additional file 2. Values are means with their standard errors. a, b Mean values with unlike letters were significantly different (P < 0.05). Data were analyzed by the Student’s t test (n = 4)

Discussion

Previous studies reported that dietary PAD causes deficiency symptoms such as growth retardation, dermatosis, diarrhea, and even death in both mammals [2,3,4,5,6] and poultry [7,8,9,10,11,12,13, 19, 25]. In ducks, growth retardation and excessive exudate from the eyes were exhibited in response to PAD [13]. Consistent with previous studies, the PAD ducks showed growth depression, exudate on eyelids, poor feathering, and a high mortality rate in the present study. Furthermore, poor pantothenic acid status was observed in PAD ducks indicated by reduced plasma pantothenic acid concentration, which is a sensitive biomarker for pantothenic acid status [21, 26, 27]. These results indicate that a severe PAD animal model was successfully established.

It has been shown that pantothenic acid can maintain intestinal function by keeping the structure of intestine integrity of animals [3, 14,15,16]. Insufficient of pantothenic acid would cause intestinal hypofunction in rats, dogs, cats, chicks and fish [4, 5, 17,18,19,20,21]. And a feature common to experimental PAD in different animals is intestinal ulceration [4, 5, 17,18,19,20]. Duodenitis and duodenal ulcers [18], as well as duodenal changes including eventual atrophy of crypts, diminution in size of villi, were observed in PAD rats [16]. In agreement with the previous studies, we observed small intestine morphological changes indicated by the reduced villus height and villus surface area in PAD ducks in the present study, indicating damage to the intestinal epithelium.

Furthermore, PAD leads to abnormalities in carbohydrate metabolism. Low fasting blood glucose levels and increased sensitivity to insulin were found in PAD rats and dogs [28,29,30,31,32]. In agreement with previous studies, PAD caused fasting hypoglycemia and decreased plasma insulin level in ducks in the present study, indicating abnormal glucose metabolism.

Together, dietary PAD of ducks resulted in growth retardation, alterations of intestinal morphology and function, and abnormal glucose metabolism. However, limited data are available currently concerning the molecular mechanisms behind. Therefore, we used a proteomic approach, for the first time, to investigate the metabolic disorder of small intestine induced by PAD to explain intestinal hypofunction and growth depression. Proteomic analysis revealed 421 differentially expressed proteins in the mucosa of PAD ducks compared to CON birds, indicating an important impact of pantothenic acid on intestinal function. The differentially expressed proteins are mainly enriched in glycolysis and gluconeogenesis, fatty acid beta oxidation, oxidative phosphorylation, TCA cycle, intestinal absorption, regulation of actin cytoskeleton, and oxidative stress. These processes probably underlie the intestinal hypofunction and growth depression induced by PAD. Notably, the CoA-binding proteins were reduced in the intestinal mucosa of PAD ducks, such as ACSL5, ACAD11, ACADM, HADHB, ECI2, ACOX1, and ACOX2. This finding supports the hypothesis that CoA-binding proteins may be depressed as a result of dietary PAD.

Glycolysis and gluconeogenesis

A total of thirteen proteins participating in the glycolysis and gluconeogenesis pathway were altered by PAD. Of these, twelve proteins were involved in glycolysis, two enhanced (ALDOA and HK1) and ten diminished (ALDOB, HKDC1, ENO1, GAPDH, PKM, PFKP, TPI1, PGK1, DLD, and PDHB). HKDC1 and HK1 are two isozymes of hexokinases, which mediate the initial step of glycolysis by catalyzing phosphorylation of D-glucose to D-glucose 6-phosphate [33]. PFKP catalyzes the phosphorylation of fructose 6-phosphate to fructose 1,6-bisphosphate by ATP [34]. ALDOA and ALDOB are two isoforms of aldolase family which are located in skeletal muscle and liver tissue respectively, cleaves fructose-1,6-bisphosphate to triose phosphates [35]. TPI1 catalyzes the interconversion between dihydroxyacetone phosphate and D-glyceraldehyde-3-phosphate [36]. GAPDH catalyzes the conversion of glyceraldehyde 3-phosphate into 1,3-diphosphoglycerate [37]. PGK1 catalyzes one of the two ATP producing reactions in the glycolytic pathway via the reversible conversion of 1,3-diphosphoglycerate to 3-phosphoglycerate [38]. ENO1 catalyzes the formation of phosphoenolpyruvate from 2-phosphoglycerate [39]. PKM catalyzes the second ATP generation reaction in the glycolytic pathway via conversion of phosphoenolpyruvate to pyruvate [40]. PDHB is a subunit of pyruvate dehydrogenase (E1). As the E3 component of pyruvate dehydrogenase complex, DLD oxidizes dihydrolipoic acid to lipoic acid [41]. Pyruvate dehydrogenase complex converts pyruvate to acetyl-CoA [42, 43]. Therefore, ten out of twelve proteins (ALDOB, HKDC1, ENO1, GAPDH, PKM, PFKP, TPI1, PGK1, DLD, and PDHB) were down-regulated in the PAD group suggests that the rate of glycolysis in the intestinal mucosa may be impaired. Besides, one protein (FBP1) involved in gluconeogenesis process was down-regulated by PAD, indicating a depressed gluconeogenesis. FBP1, a rate-limiting enzyme in gluconeogenesis, catalyzes the hydrolysis of fructose 1,6-bisphosphate to fructose 6-phosphate.

Fatty acid beta oxidation

In the present study, PAD down-regulated 9 proteins involved in fatty acid beta oxidation, including ACSL5, ACAD11, ACADM, CRAT, DLD, HADHB, ECI2, ACOX1, and ACOX2. Notably, among these, ACSL5, ACAD11, ACADM, HADHB, ECI2, ACOX1, and ACOX2 are belong to CoA-binding proteins, which were all reduced due to PAD. ACSL5 belongs to the acyl-CoA synthetase family, catalyzing free fatty acids into fatty acyl-CoA esters, which play key roles in lipid biosynthesis and fatty acid degradation [44]. ACAD11 and ACADM are two members of fatty acyl-CoA dehydrogenases that catalyze the first step in each cycle of fatty acid beta oxidation [45]. CRAT catalyzes the reversible transfer of an acetyl group from acyl-CoA to carnitine [46]. ECI2 is involved in the beta oxidation of unsaturated fatty acids, catalyzing the formation of 2-trans-enoyl-CoA from 3-cis or trans-enoyl-CoA [47]. HADHB catalyzes the final step of beta-oxidation, in which 3-ketoacyl CoA is cleaved by the thiol group of another molecule of Coenzyme A [48]. ACOX1 and ACOX2 catalyze the first and rate-limiting step of peroxisomal fatty acid beta oxidation, the desaturation of acyl-CoAs to 2-trans-enoyl-CoAs. ACOX1 catalyzes medium to very long straight-chain fatty acids [49], while ACOX2 catalyze the CoA-esters of very long-chain fatty acids, branched-chain fatty acids and the C27-bile acid intermediates [50]. The down-regulation of all these proteins involved in the fatty acid beta oxidation process in PAD ducks may indicate that this process is impaired. This implication is supported by previous finding in rats that PAD reduced CoA and short-chain acyl-CoA contents [51], as well as hepatic peroxisomal fatty acid beta oxidation [6].

TCA cycle

PAD down-regulated six proteins involved in the TCA cycle, including DLD, ACO1, ACO2, MDH1, MDH2, and PDHB, and up-regulated one protein, IDH1. PDHB is a subunit of pyruvate dehydrogenase (E1). As the E3 component of pyruvate dehydrogenase complex and α-ketoglutarate dehydrogenase complex, DLD oxidizes dihydrolipoic acid to lipoic acid [41]. Pyruvate dehydrogenase complex converts pyruvate to acetyl-CoA [42, 43]. α-ketoglutarate dehydrogenase catalyzes the oxidation of α-ketoglutarate to form succinyl-CoA, which is a rate-limiting enzyme of the TCA cycle [52, 53]. ACO1 and ACO2 are two isozymes of aconitate hydratase located in cytoplasmic and mitochondrial respectively, which catalyzes the isomerization of citrate to isocitrate via cis-aconitate. MDH1 and MDH2 are two isozymes of malate dehydrogenase located in cytoplasmic and mitochondrial respectively, which catalyzes the reversible interconversion of malate and oxaloacetate [54]. IDH1 is a subunit of isocitrate dehydrogenase, catalyzing the oxidative decarboxylation of isocitrate into α-ketoglutarate [55]. Six out of seven proteins (DLD, ACO1, ACO2, MDH1, MDH2, and PDHB) were down-regulated due to PAD, indicating a decreased TCA cycle.

Oxidative phosphorylation

PAD down-regulated 5 proteins involved in oxidative phosphorylation, including NDUFA5, NDUFA6, ATP5F1, ATP5H, and ATP5PO. NDUFA5 and NDUFA6, the subunits of complex I, are involved in complex I assembly [56, 57]. ATP5F1, ATP5H, and ATP5PO, the subunits of complex V, are involved in complex V assembly [58,59,60]. The down-regulated expression of proteins involved in oxidative phosphorylation, probably indicates that this process is impaired by PAD.

Together, our duodenum mucosal proteomic analysis revealed glycolysis and gluconeogenesis, fatty acid beta oxidation, TCA cycle, and oxidative phosphorylation are probably impaired in response to PAD, consequently resulting in insufficient energy generation in the small intestine and subsequent growth retardation.

Regulation of actin cytoskeleton

Eleven proteins altered by PAD have actin-binding domains and play direct roles in the organization of structure of the cytoskeleton, including two down-regulated proteins (VIL1 and EZR) and nine up-regulated proteins (ACTN1, FN1, MSN, MYL1, MYLK, VCL, TPM3, TPM1, and VIM). VIL1 and EZR (also known as villin 2; VIL2) are microvillar proteins in intestinal epithelial cells [61]. VIL1 is an epithelial cell-specific actin-binding protein that regulates cell migration, cell death, cell morphology, and epithelial-to-mesenchy-mal transition [62, 63]. EZR is critical for the de novo lumen formation and expansion during villus morphogenesis, and EZR absence resulted in abnormal villus morphogenesis [64, 65].

Actinin is a component of stress fibers and links the cytoskeleton to adherens-type junctions. As one isoform of actinin, ACTN1 plays a major role in cell migration and adhesion [66]. Besides, ACTN1 can directly binds to VCL, and the two proteins cooperate to organize the cytoskeleton at adhesion junctions [67]. FN1, a glycoprotein component of the extracellular matrix, has a role in cell adhesion and migration [68]. VIM induces changes in cell shape, motility, and adhesion during the epithelial to mesenchymal transition [69]. MYLK induces contraction of the perijunctional actomyosin ring through myosin II regulatory light chain phosphorylation, and thereby increases intestinal epithelial permeability when activated [70]. TPM3 and TPM1 are two isoforms of tropomyosins which play important roles in the regulation of assembly, stability, and motility of the intestinal epithelial cells [71]. Collectively, the over expression of nine proteins (ACTN1, FN1, MSN, MYL1, MYLK, VCL, TPM3, TPM1, and VIM) and down-regulation of two proteins (VIL1 and EZR) in response to PAD probably indicates orchestrated regulation of actin cytoskeleton dynamics and the negative impact on intestinal integrity. Remodeling of the cytoskeleton is fundamental in proliferation, apoptosis, cell invasion and metastasis [72]. Therefore, these altered proteins involved in regulation of actin cytoskeleton in response to PAD probably resulted in morphological changes of small intestine, such as atrophy of their intestinal villus. Furthermore, it is reported that ATP depletion uncouples the gate and fence functions of the tight junction and induces actin network dissolution of epithelial cells [73]. The alterations of intestinal morphology and actin cytoskeleton due to PAD in the present study may be attributed to ATP depletion indicated by impaired glycolysis, fatty acid beta oxidation, TCA cycle, and oxidative phosphorylation processes.

Oxidative stress

PAD down-regulated three proteins involved in the oxidative stress, including MAOA, CAT, and MGST1, and up-regulated one protein, GCLC. MAOA metabolizes dopamine to dihydroxyphenylacetic acid and H2O2, a potential source of reactive oxygen species [74]. Both CAT and MGST1 take part in the oxidative stress defense as its scavenging of H2O2 [75, 76]. GCLC is a subunit of glutamate cysteine ligase which catalyzes the rate-limiting step in reduced glutathione (GSH) synthesis. And glutamate cysteine ligase is often activated to increase cellular GSH content in response to oxidative stress [77]. Collectively, the reduction of MAOA, CAT, and MGST1 in PAD ducks, as well as the enhanced GCLC, indicates small intestinal oxidative stress was induced. This is supported by the results of increased plasma MDA content and decreased T-SOD activity in the present study and is consistent with previous studies in geese [9] and fish [21]. It has been shown that pantothenic acid protects cells against oxidative stress by promoting glutathione levels and cellular repair mechanisms [22,23,24]. Furthermore, PAD induced oxidative stress in ducks, which may be associated with intestinal injury and morphological alterations.

Intestinal absorption

A novel and important finding of this study is that four proteins (SLC2A2, VIL1, EZR, and MOGAT2) involved in intestinal absorption were down-regulated in duodenum mucosa as a result of PAD. Intestinal SLC2A2 (GLUT2) is known as a means to transfer glucose and fructose from the lumen to the bloodstream and, thereby, to provide sugar to tissue. And intestinal SLC2A2 deletion in mice induced glucose malabsorption visualized by the delay in the distribution of oral sugar in tissues, as well as decreased microvillus length and body weight gain [78]. It is reported pantothenic acid appears to be part of a glucose carrier system [79], therefore, PAD may direct reduce SLC2A2 and resulted in abnormal glucose absorption and hypoglycemia in ducks. Villin (VIL1 and EZR) is also involved in the absorptive and secretory function of epithelial cells by modulating F-actin polymerization/depolymerization. Specifically, villin-depleted mice showed a reduction in intestinal glucose absorption [80]. Decreased protein expression of SLC2A2, VIL1, and EZR due to PAD probably impair glucose absorption system, which may provide a possible explanation for fasting hypoglycemia that is seen. MOGAT2 plays a central role in absorption of dietary fat in the small intestine by catalyzing the resynthesis of triacylglycerol in enterocytes [81]. MGAT2 deficient specifically in the small intestine showed a delay in fat absorption in mice [82]. Therefore, a reduction of SLC2A2, VIL1, EZR, and MOGAT2 suggests that glucose and fat malabsorption in the small intestine may be induced by PAD, which probably leads to growth depression observed. This is in line with the previous findings in fish that PAD decreased the digestive and absorptive capacities indicated by the reduced the activities of both intestinal brush border enzymes and digestive enzymes [14, 21].

Conclusion

PAD caused growth retardation, fasting hypoglycemia, morphological alterations of the small intestine, and oxidative stress in ducks. We have performed a proteomic profiling further to investigate the effects of PAD on duodenum mucosal proteome of ducks. The results indicated that PAD may suppress energy generation processes such as glycolysis and gluconeogenesis, fatty acid beta oxidation, TCA cycle, and oxidative phosphorylation, leading to impaired ATP generation. Furthermore, PAD may induce glucose and fat malabsorption in the small intestine indicated by four diminished proteins involved. Besides, PAD probably leads to negative impacts on intestinal integrity and morphology because most of proteins involved in regulation of actin cytoskeleton were up-regulated. To be sum, PAD causes intestinal hypofunction and growth depression probably by impairing glycolysis and gluconeogenesis, fatty acid beta oxidation, TCA cycle, oxidative phosphorylation, actin cytoskeleton, and intestinal absorption processes. These findings add to our understanding of the mechanisms of intestinal hypofunction due to PAD.

Methods

Animals ethics statement

All experimental procedures with ducks were performed according to the Guidelines for Experimental Animals established by the Ministry of Science and Technology (Beijing, China). Ethical approval on animal survival was given by the animal ethics committee of the Institute of Animal Sciences (IAS), Chinese Academy of Agricultural Sciences (CAAS, Beijing, China) with the following reference number: IASCAAS-2019–19.

Animals and housing

A total of 128 one-day-old male white Pekin ducks (Anas platyrhynchos) were assigned randomly to one of two dietary treatments of 8 replicate pens with 8 birds per pen. The ducks were either fed with a PAD diet or a pantothenic acid adequate (control, CON) diet. From hatch to 16 days of age, all ducks were reared in pens in a temperature-controlled room with feed and water ad libitum. And the light was continuously on.

Diet

The pantothenic acid-deficient basal diet was formulated in line with NRC (1994) [83], containing 4.65 mg pantothenic acid /kg of diet (Table 5). To produce the PAD diet and control diet, the basal diets were supplemented with 0 and 8 mg/kg of crystalline calcium pantothenate respectively (purity, 99%; Xinfu Technology Co. Ltd., Hangzhou, China). The pantothenic acid concentration for the control diet (12.65 mg/kg) met the recommendation provided by NRC (1994) [83].

Table 5 Composition of pantothenic acid-deficient basal diet from hatch to 16 days of age (% as-fed)

Sampling

At 16 days of age, after overnight fasting, duck weight and feed intake from each pen were recorded to calculate ADG, ADFI, and FCR. ADFI and FCR were all corrected for mortality. Two ducks from each pen were randomly selected for sampling as follows. Blood was taken via jugular vein to harvest plasma, stored at − 20 °C until further analysis. Thereafter, these selected ducks were sacrificed by CO2 inhalation and intestinal tissues were obtained immediately. The intestinal sections were divided into the duodenum, jejunum, and ileum. And they were rinsed with physiological saline and then cut into 1 cm length segments, fixed in 10% neutral formalin used for histological analysis. Duodenum mucosa from the remaining segment was obtained as described previously [84], snap frozen in liquid nitrogen, and stored at − 80 °C until further analysis.

Pantothenic acid content

Feed and plasma pantothenic acid concentrations were measured by HPLC (Agilent 1290) coupled with triple quadrupole mass spectrometry (Agilent 6470) according to the methods described previously [85]. Before LC/MS analysis, feed and plasma samples were pretreated according to the methods described previously [86, 87]. The peak was identified by the pure authentic standards (Sigma-Aldrich, St. Louis, MO, USA).

Plasma parameters

Plasma glucose and ALP were determined by commercial kits following the manufacturer’s instructions (BioSino Bio-technology and Science Inc., Beijing, China). Plasma MDA, T-SOD, insulin, and glucagon were measured by commercial kits following the manufacturer’s protocols (Nanjing Jiancheng Institute of Bioengineering, Nanjing, Jiangsu, China).

Intestinal morphology assessment

The duodenum, jejunum, and ileum sections were embedded in paraffin and transversely sectioned in (4 μm thick) and stained with hematoxylin and eosin following deparaffinization and dehydration. Intestinal tissues and structures were observed using a BH2 Olympus microscope (Olympus, Tokyo, Japan) and analyzed using an image analysis system (Olympus 6.0). Villus height, villus width, crypt depth, and villus surface area were assessed following the method as described previously [84].

Duodenum mucosal proteomics

Three individual duodenum mucosa samples were randomly chosen from each group to conduct the iTRAQ assays. Proteins were extracted and digested as described previously [88]. The digested samples were labelled following the manufacturer’s protocols with an 8-plex iTRAQ kit (AB SCIEX, Foster City, USA). The PAD samples were labelled with iTRAQ tags 113, 114, and 115, while the CON samples were labelled with tags 116, 117, and 118. The labelled samples were combined and fractionated into 20 fractions by HPLC (DINOEX Ultimate 3000 BioRS, Thermo Fisher, Waltham, MA, USA). We performed LC-electrospray ionization-MS/MS analysis on a Triple TOF 5600 plus system (AB SCIEX, Framingham, USA). The collected raw MS/MS data were searched according to the Uniprot-Swissport Anas platyrhynchos Database UniProt_Mallard_8839 using ProteinPilot Software (version 4.5, AB SCIEX). To minimize false-positive results, we counted only peptides at the 95% CI and a false discovery rate < 0.01, containing at least one unique peptide. For protein quantitation, a differentially expressed protein containing at least two unique spectra was used with a FC value > 1.5 or < − 1.5 with P < 0.05 between the PAD and CON groups. For functional annotation, the differentially expressed proteins induced by PAD were performed KEGG pathway and Gene Ontology (GO) enrichment analysis using ClueGo software as described previously [88].

Western blot analyses

Western blot analysis of two proteins, ACADM and GAPDH, were performed following the method as described [89]. Primary antibodies (1 μg/ml) against ACADM (ab92461; Abcam) and GAPDH (HX1828; Huaxingbio) were used. Histone H3 (A2348; ABclonal) served as a loading control.

Statistical analyses

For results of growth performance, plasma parameters, and intestinal mucosal histomorphology, data were analyzed using the Student’s t test procedures of SAS software (SAS Institute Inc., 2011). The replicate pen of 8 ducks for growth performance or two ducks for other indices served as the experimental unit. The variability in the data was expressed as the standard error of the means (SEM). Differences between means were considered statistically significant at P < 0.05.

Availability of data and materials

The data sets supporting the conclusions of this article are included within the article and its additional file. The mass spectrometry proteomics data have been deposited to the ProteomeXchange Consortium via the PRIDE (https://www.ebi.ac.uk/pride/archive/) partner repository with the data set identifier PXD026710.

Abbreviations

ADG:

average daily weight gain

ADFI:

average daily feed intake

ALP:

alkaline phosphatase

CoA:

coenzyme A

CON:

control

FC:

fold change

FCR:

feed conversion ratio

GO:

Gene Ontology

HPLC:

high performance liquid chromatography

iTRAQ:

isobaric tags for relative and absolute quantification

KEGG:

Kyoto Encyclopedia of Genes and Genomes

MDA:

malondialdehyde

PAD:

pantothenic acid deficiency

SEM:

standard error of the mean

TCA:

tricarboxylic acid

T-SOD:

total superoxide dismutase

References

  1. Miller JW, Rucker RB. Pantothenic acid. In: Erdman JW, Macdonald IA, Zeisel SH, editors. Present knowledge in nutrition. 10th ed. Washington: Wiley-Blackwell; 2012. p. 375–90. https://0-doi-org.brum.beds.ac.uk/10.1002/9781119946045.ch24.

    Chapter  Google Scholar 

  2. Smith CM, Song WO. Comparative nutrition of pantothenic acid. J Nutr Biochem. 1996;7(6):312–21. https://0-doi-org.brum.beds.ac.uk/10.1016/0955-2863(96)00034-4.

    Article  CAS  Google Scholar 

  3. Nelson RA. Intestinal transport, coenzyme a, and colitisin pantothenic acid deficiency. Am J Clin Nutr. 1968;21(5):495–501. https://0-doi-org.brum.beds.ac.uk/10.1093/ajcn/21.5.495.

    Article  CAS  PubMed  Google Scholar 

  4. Zucker TF. Pantothenic acid deficiency and its effect on the integrity and functions of the intestines. Am J Clin Nutr. 1958;6(1):65–74. https://0-doi-org.brum.beds.ac.uk/10.1093/ajcn/6.1.65.

    Article  CAS  PubMed  Google Scholar 

  5. Gershoff SN, Gottlieb LS. Pantothenic acid deficiency in cats. J Nutr. 1964;82(1):135–8. https://0-doi-org.brum.beds.ac.uk/10.1093/jn/82.1.135.

    Article  CAS  PubMed  Google Scholar 

  6. Youssef JA, Song WO, Badr MZ. Mitochondrial, but not peroxisomal, β-oxidation of fatty acids is conserved in coenzyme A-deficient rat liver. Mol Cell Biochem. 1997;175(1/2):37–42. https://0-doi-org.brum.beds.ac.uk/10.1023/A:1006877021617.

    Article  CAS  PubMed  Google Scholar 

  7. Bauerrnfeind JC, Norris LC, Heuser GF. The pantothenic acid requirement of chicks. Poult Sci. 1942;21(2):142–6. https://0-doi-org.brum.beds.ac.uk/10.3382/ps.0210142.

    Article  Google Scholar 

  8. Jukes TH. The pantothenic acid requirement of the chick. J Biol Chem. 1939;129(1):225–31. https://0-doi-org.brum.beds.ac.uk/10.1016/S0021-9258(18)73664-8.

    Article  CAS  Google Scholar 

  9. Wang B, Zhang X, Yue B, Ge W, Zhang M, Ma C, et al. Effects of pantothenic acid on growth performance, slaughter performance, lipid metabolism, and antioxidant function of Wulong geese aged one to four weeks. Anim Nutr. 2016;2(4):312–7. https://0-doi-org.brum.beds.ac.uk/10.1016/j.aninu.2016.07.005.

    Article  PubMed  PubMed Central  Google Scholar 

  10. Kratzer FH, Williams D. The pantothenic acid requirement of poults for early growth. Poult Sci. 1948;27(4):518–23. https://0-doi-org.brum.beds.ac.uk/10.3382/ps.0270518.

    Article  CAS  Google Scholar 

  11. Hegsted DM, Riggs TR. The pantothenic acid requirements of chicks receiving a purified diet. J Nutr. 1949;37(3):361–7. https://0-doi-org.brum.beds.ac.uk/10.1093/jn/37.3.361.

    Article  CAS  PubMed  Google Scholar 

  12. Lepkovsky S, Bird FH, Kratzer FH, Asmundson VS. The comparative requirements of chicks and Turkey poults for pantothenic acid. Poult Sci. 1945;24(4):335–9. https://0-doi-org.brum.beds.ac.uk/10.3382/ps.0240335.

    Article  CAS  Google Scholar 

  13. Hegsted DM, Perry RL. Nutritional studies with the duck V. riboflavin and pantothenic acid requirements. J Nutr. 1948;35(4):411–7. https://0-doi-org.brum.beds.ac.uk/10.1093/jn/35.4.411.

    Article  CAS  PubMed  Google Scholar 

  14. Wen ZP, Zhou XQ, Feng L, Jiang J, Liu Y. Effect of dietary pantothenic acid supplement on growth, body composition and intestinal enzyme activities of juvenile Jian carp (Cyprinus carpio var. Jian). Aquac Nutr. 2009;15(5):470–6. https://0-doi-org.brum.beds.ac.uk/10.1111/j.1365-2095.2008.00612.x.

    Article  CAS  Google Scholar 

  15. Sabui S, Kapadia R, Ghosal A, Schneider M, Lambrecht NWG, Said HM. Biotin and pantothenic acid oversupplementation to conditional SLC5A6 KO mice prevents the development of intestinal mucosal abnormalities and growth defects. Am J Physiol Cell Physiol. 2018;315(1):C73–9. https://0-doi-org.brum.beds.ac.uk/10.1152/ajpcell.00319.2017.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Seronde J. The pathogenesis of duodenal ulcer disease in the pantothenate-deficient rat. Yale J BioMed. 1963;36:141–56.

    Google Scholar 

  17. Ellestad-Sayed JJ, Nelson RA, Adson MA, Palmer WM, Soule EH. Pantothenic acid, coenzyme a, and human chronic ulcerative and granulomatous colitis. Am J Clin Nutr. 1976;29(12):1333–8. https://0-doi-org.brum.beds.ac.uk/10.1093/ajcn/29.12.1333.

    Article  CAS  PubMed  Google Scholar 

  18. Berg B. Duodenitis and duodenal ulcers produced in rats by pantothenic acid deficiency. Br J Exp Pathol. 1959;40:371–4.

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Gries CL, Scott ML. The pathology of thiamin, riboflavin, pantothenic acid and niacin deficiencies in the chick. J Nutr. 1972;102(10):1269–85. https://0-doi-org.brum.beds.ac.uk/10.1093/jn/102.10.1269.

    Article  CAS  PubMed  Google Scholar 

  20. McKibbin JM, Black S, Elvehjem CA. The essential nature of pantothenic acid and another alkali labile factor in the nutrition of the dog. Am J Phyisology. 1940;130(2):365–72. https://0-doi-org.brum.beds.ac.uk/10.1152/ajplegacy.1940.130.2.365.

    Article  CAS  Google Scholar 

  21. Qian Y, Li XF, Zhang DD, Cai DS, Tian HY, Liu WB. Effects of dietary pantothenic acid on growth, intestinal function, anti-oxidative status and fatty acids synthesis of juvenile blunt snout bream Megalobrama amblycephala. PLoS One. 2015;10(3):e0119518. https://0-doi-org.brum.beds.ac.uk/10.1371/journal.pone.0119518.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Slyshenkov VS, Piwocka K, Sikora E, Wojtczak L. Pantothenic acid protects Jurkat cells against ultraviolet light-induced apoptosis. Free Radical Bio Med. 2001;30(11):1303–10. https://0-doi-org.brum.beds.ac.uk/10.1016/S0891-5849(01)00531-7.

    Article  CAS  Google Scholar 

  23. Walczak-Jedrzejowska R, Wolski JK, Slowikowska-Hilczer J. The role of oxidative stress and antioxidants in male fertility. Cent Eur J Urol. 2013;66:60–7.

    Article  CAS  Google Scholar 

  24. Slyshenkov VS, Rakowska M, Moiseenok AG, Wojtczak L. Pantothenic acid and its derivatives protect Ehrlich ascites tumor cells against lipid peroxidation. Free Radical Bio Med. 1995;19(6):767–72. https://0-doi-org.brum.beds.ac.uk/10.1016/0891-5849(95)00084-B.

    Article  CAS  Google Scholar 

  25. Beer AE, Scott ML, Nesheim MC. The effects of graded levels of pantothenic acid on the breeding performance of white leghorn pullets. Br Poult Sci. 1963;4(3):243–53. https://0-doi-org.brum.beds.ac.uk/10.1080/00071666308415501.

    Article  Google Scholar 

  26. Lin YH, Lin HY, Shiau SY. Estimation of dietary pantothenic acid requirement of grouper, Epinephelus malabaricus according to physiological and biochemical parameters. Aquaculture. 2012;324:92–6.

    Article  Google Scholar 

  27. Shiau SY, Hsu CW. Dietary pantothenic acid requirement of juvenile grass shrimp, Penaeus monodon. J Nutr. 1999;129(3):718–21. https://0-doi-org.brum.beds.ac.uk/10.1093/jn/129.3.718.

    Article  CAS  PubMed  Google Scholar 

  28. Arnrich L, Hurley LS, Forker BR, Morgan AF. Response to stress by riboflavin-deficient and pantothenic acid-deficient dogs. Am J Phys. 1956;184(3):515–20. https://0-doi-org.brum.beds.ac.uk/10.1152/ajplegacy.1956.184.3.515.

    Article  CAS  Google Scholar 

  29. Arnrich L, Nelson MR, Gram MR, Morgan AF. Effect of adrenal hormones on carbohydrate metabolism in riboflavin and pantothenic acid-deficient dogs. Am J Phys. 1956;186(3):427–34. https://0-doi-org.brum.beds.ac.uk/10.1152/ajplegacy.1956.186.3.427.

    Article  CAS  Google Scholar 

  30. Hurley LS, Morgan AF. Carbohydrate metabolism and adrenal cortical function in the pantothenic acid-deficient rat. J Biol Chem. 1952;195(2):583–90. https://0-doi-org.brum.beds.ac.uk/10.1016/S0021-9258(18)55765-3.

    Article  CAS  PubMed  Google Scholar 

  31. Schultz RB, Winters RW, Krehl WA. The adrenal cortex of the pantothenic acid deficient rat: modification of the lesion by ACTH and cortisone treatment. Endocrinology. 1952;51(4):336–43. https://0-doi-org.brum.beds.ac.uk/10.1210/endo-51-4-336.

    Article  CAS  PubMed  Google Scholar 

  32. Winters RW, Schultz RB, Krehl WA. The adrenal cortex of the pantothenic acid-deficient rat: carbohydrate metabolism. Endocrinology. 1952;50(4):388–98. https://0-doi-org.brum.beds.ac.uk/10.1210/endo-50-4-388.

    Article  CAS  PubMed  Google Scholar 

  33. Ludvik AE, Pusec CM, Priyadarshini M, Angueira AR, Guo C, Lo A, et al. HKDC1 is a novel hexokinase involved in whole-body glucose use. Endocrinology. 2016;157(9):3452–61. https://0-doi-org.brum.beds.ac.uk/10.1210/en.2016-1288.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Lee JH, Liu R, Li J, Zhang C, Wang Y, Cai Q, et al. Stabilization of phosphofructokinase 1 platelet isoform by AKT promotes tumorigenesis. Nat Commun. 2017;8(1):949. https://0-doi-org.brum.beds.ac.uk/10.1038/s41467-017-00906-9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Dawson NJ, Biggar KK, Storey KB. Characterization of fructose-1, 6-bisphosphate aldolase during anoxia in the tolerant turtle, Trachemys scripta elegans: an assessment of enzyme activity, expression and structure. PLoS One. 2013;8(7):e68830. https://0-doi-org.brum.beds.ac.uk/10.1371/journal.pone.0068830.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Rodríguez-Almazán C, Arreola R, Rodríguez-Larrea D, Aguirre-López B, Gómez-Puyou MT, Pérez-Montfort R, et al. Structural basis of human triosephosphate isomerase deficiency mutation E104D is related to alterations of a conserved water network at the dimer interface. J Biol Chem. 2008;283(34):23254–63. https://0-doi-org.brum.beds.ac.uk/10.1074/jbc.M802145200.

    Article  CAS  PubMed  Google Scholar 

  37. Ercolani L, Florence B, Denaro M, Alexander M. Isolation and complete sequence of a functional human glyceraldehyde-3-phosphate dehydrogenase gene. J Biol Chem. 1988;263(30):15335–41. https://0-doi-org.brum.beds.ac.uk/10.1016/S0021-9258(19)37593-3.

    Article  CAS  PubMed  Google Scholar 

  38. Li X, Jiang Y, Meisenhelder J, Yang W, Hawke DH, Zheng Y, et al. Mitochondria-translocated PGK1 functions as a protein kinase to coordinate glycolysis and the TCA cycle in tumorigenesis. Mol Cell. 2016;61(5):705–19. https://0-doi-org.brum.beds.ac.uk/10.1016/j.molcel.2016.02.009.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Ejeskär K, Krona C, Carén H, Zaibak F, Li L, Martinsson T, et al. Introduction of in vitro transcribed ENO1 mRNA into neuroblastoma cells induces cell death. BMC Cancer. 2005;5(1). https://0-doi-org.brum.beds.ac.uk/10.1186/1471-2407-5-161.

  40. Christofk HR, Heiden MGV, Wu N, Asara JM, Cantley LC. Pyruvate kinase M2 is a phosphotyrosine-binding protein. Nature. 2008;452(7184):181–6. https://0-doi-org.brum.beds.ac.uk/10.1038/nature06667.

    Article  CAS  PubMed  Google Scholar 

  41. Carothers DJ, Pons G, Patel MS. Dihydrolipoamide dehydrogenase: functional similarities and divergent evolution of the pyridine nucleotide-disulfide oxidoreductases. Arch Biochem Biophys. 1989;268(2):409–25. https://0-doi-org.brum.beds.ac.uk/10.1016/0003-9861(89)90309-3.

    Article  CAS  PubMed  Google Scholar 

  42. Sugden MC, Holness MJ. Recent advances in mechanisms regulating glucose oxidation at the level of the pyruvate dehydrogenase complex by PDKs. Am J Physiol Endocrinol Metab. 2003;284(5):E855–62. https://0-doi-org.brum.beds.ac.uk/10.1152/ajpendo.00526.2002.

    Article  CAS  PubMed  Google Scholar 

  43. Holness M, Sugden M. Regulation of pyruvate dehydrogenase complex activity by reversible phosphorylation. Biochem Soc Trans. 2003;31(6):1143–51. https://0-doi-org.brum.beds.ac.uk/10.1042/bst0311143.

    Article  CAS  PubMed  Google Scholar 

  44. Watkins PA, Maiguel D, Jia Z, Pevsner J. Evidence for 26 distinct acyl-coenzyme a synthetase genes in the human genome. J Lipid Res. 2007;48(12):2736–50. https://0-doi-org.brum.beds.ac.uk/10.1194/jlr.M700378-JLR200.

    Article  CAS  PubMed  Google Scholar 

  45. Ghisla S, Thorpe C. Acyl-CoA dehydrogenases. Eur J Biochem. 2004;271(3):494–508. https://0-doi-org.brum.beds.ac.uk/10.1046/j.1432-1033.2003.03946.x.

    Article  CAS  PubMed  Google Scholar 

  46. Godárová A, Litzlbauer E, Brunner S, Agu A, Lohninger A. Hofbauer. L-carnitine regulates mRNA expression levels of the carnitine acyltransferases - CPT1A, CPT2, and CRAT. Monatsh Chem. 2005;136(8):1349–63. https://0-doi-org.brum.beds.ac.uk/10.1007/s00706-005-0336-5.

    Article  CAS  Google Scholar 

  47. Weeghel MV, Brinke HT, Lenthe HV, Kulik W, Minkler PE, Stoll MSK, et al. Functional redundancy of mitochondrial enoyl-CoA isomerases in the oxidation of unsaturated fatty acids. FASEB J. 2012;26(10):4316–26. https://0-doi-org.brum.beds.ac.uk/10.1096/fj.12-206326.

    Article  CAS  PubMed  Google Scholar 

  48. Das AM, Illsinger S, Lucke T, Hartmann H, Ruiter JPN, Steuerwald U, et al. Isolated mitochondrial long-chain ketoacyl-CoA thiolase deficiency resulting from mutations in the HADHB gene. Clin Chem. 2006;52(3):530–4. https://0-doi-org.brum.beds.ac.uk/10.1373/clinchem.2005.062000.

    Article  CAS  PubMed  Google Scholar 

  49. Jiao Y, Zan LS, Liu YF, Wang HB. Molecular characterization, polymorphism of the ACOX1 gene and association with ultrasound traits in Bos taurus. Genet Mol Res. 2011;10(3):1948–57. https://0-doi-org.brum.beds.ac.uk/10.4238/vol10-3gmr1036.

    Article  CAS  PubMed  Google Scholar 

  50. Ferdinandusse S, Denis S, Roermund CWTV, Preece MA, Koster J, Ebberink MS, et al. A novel case of ACOX2 deficiency leads to recognition of a third human peroxisomal acyl-CoA oxidase. BBA-Mol Basis Dis. 1864;2018:952–8.

    Google Scholar 

  51. Moiseenok AG, Sheibak VM, Gurinovich VA. Hepatic CoA, S-acyl-CoA, biosynthetic precursors of the coenzyme and pantothenate-protein complexes in dietary pantothenic acid deficiency. Int J Vitam Nutr Res. 1987;57(1):71–7.

    CAS  PubMed  Google Scholar 

  52. Gibson GE, Blass JP, Beal MF, Bunik V. The α-ketoglutarate-dehydrogenase complex. Mol Neurobiol. 2005;31(1-3):43–63. https://0-doi-org.brum.beds.ac.uk/10.1385/MN:31:1-3:043.

    Article  CAS  PubMed  Google Scholar 

  53. Zündorf G, Kahlert S, Bunik V, Reiser G. α-Ketoglutarate dehydrogenase contributes to production of reactive oxygen species in glutamate-stimulated hippocampal neurons in situ. Neuroscience. 2009;158(2):610–6. https://0-doi-org.brum.beds.ac.uk/10.1016/j.neuroscience.2008.10.015.

    Article  CAS  PubMed  Google Scholar 

  54. Lo AS, Liew CT, Ngai SM, Tsui SK, Fung KP, Lee CY, et al. Developmental regulation and cellular distribution of human cytosolic malate dehydrogenase (MDH1). J Cell Biochem. 2005;94(4):763–73. https://0-doi-org.brum.beds.ac.uk/10.1002/jcb.20343.

    Article  CAS  PubMed  Google Scholar 

  55. Grassian AR, Parker SJ, Davidson SM, Divakaruni AS, Green CR, Zhang X, et al. IDH1 mutations alter citric acid cycle metabolism and increase dependence on oxidative mitochondrial metabolism. Cancer Res. 2014;74(12):3317–31. https://0-doi-org.brum.beds.ac.uk/10.1158/0008-5472.CAN-14-0772-T.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Lazarou M, McKenzie M, Ohtake A, Thorburn DR, Ryan MT. Analysis of the assembly profiles for mitochondrial-and nuclear-DNA-encoded subunits into complex I. Mol Cell Biol. 2007;27(12):4228–37. https://0-doi-org.brum.beds.ac.uk/10.1128/MCB.00074-07.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Stroud DA, Surgenor EE, Formosa LE, Reljic B, Frazier AE, Dibley MG, et al. Accessory subunits are integral for assembly and function of human mitochondrial complex I. Nature. 2016;538(7623):123–6. https://0-doi-org.brum.beds.ac.uk/10.1038/nature19754.

    Article  CAS  PubMed  Google Scholar 

  58. Rönn T, Poulsen P, Tuomi T, Isomaa B, Groop L, Vaag A, et al. Genetic variation in ATP5O is associated with skeletal muscle ATP50 mRNA expression and glucose uptake in young twins. PLoS One. 2009;4(3):e4793. https://0-doi-org.brum.beds.ac.uk/10.1371/journal.pone.0004793.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Chinopoulos C. ATP synthase complex and the mitochondrial permeability transition pore: poles of attraction. EMBO Rep. 2007;18:1041–2.

    Article  Google Scholar 

  60. Brüggemann M, Gromes A, Poss M, Schmidt D, Klümper N, Tolkach Y, et al. Systematic analysis of the expression of the mitochondrial ATP synthase (complex V) subunits in clear cell renal cell carcinoma. Transl Oncol. 2017;10(4):661–8. https://0-doi-org.brum.beds.ac.uk/10.1016/j.tranon.2017.06.002.

    Article  PubMed  PubMed Central  Google Scholar 

  61. Saleh HS, Merkel U, Geißler KJ, Sperka T, Sechi A, Breithaupt C, et al. Properties of an ezrin mutant defective in F-actin binding. J Mol Biol. 2009;385(4):1015–31. https://0-doi-org.brum.beds.ac.uk/10.1016/j.jmb.2008.11.051.

    Article  CAS  PubMed  Google Scholar 

  62. Khurana S, George SP. Regulation of cell structure and function by actin-binding proteins: villin's perspective. FEBS Lett. 2008;582(14):2128–39. https://0-doi-org.brum.beds.ac.uk/10.1016/j.febslet.2008.02.040.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Friederich E, Vancompernolle K, Louvard D, Vandekerckhove J. Villin function in the Organization of the Actin Cytoskeleton correlation of in vivo effects to its biochemical activities in vitro. J Biol Chem. 1999;274(38):26751–60. https://0-doi-org.brum.beds.ac.uk/10.1074/jbc.274.38.26751.

    Article  CAS  PubMed  Google Scholar 

  64. Saotome I, Curto M, McClatchey AI. Ezrin is essential for epithelial organization and villus morphogenesis in the developing intestine. Dev Cell. 2004;6(6):855–64. https://0-doi-org.brum.beds.ac.uk/10.1016/j.devcel.2004.05.007.

    Article  CAS  PubMed  Google Scholar 

  65. Casaletto JB, Saotome I, Curto M, McClatchey AI. Ezrin-mediated apical integrity is required for intestinal homeostasis. PNAS. 2011;108(29):11924–9. https://0-doi-org.brum.beds.ac.uk/10.1073/pnas.1103418108.

    Article  PubMed  PubMed Central  Google Scholar 

  66. Otey CA, Carpen O. α-Actinin revisited: a fresh look at an old player. Cell Motil Cytoskel. 2004;58(2):104–11. https://0-doi-org.brum.beds.ac.uk/10.1002/cm.20007.

    Article  CAS  Google Scholar 

  67. Bois PRJ, Borgon RA, Vonrhein C, Izard T. Structural dynamics of α-actinin-vinculin interactions. Mol Cell Biol. 2005;25(14):6112–22. https://0-doi-org.brum.beds.ac.uk/10.1128/MCB.25.14.6112-6122.2005.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Goossens K, Soom AV, Zeveren AV, Favoreel H, Peelman LJ. Quantification of fibronectin 1 (FN1) splice variants, including two novel ones, and analysis of integrins as candidate FN1 receptors in bovine preimplantation embryos. BMC Dev Biol. 2009;9(1):1. https://0-doi-org.brum.beds.ac.uk/10.1186/1471-213X-9-1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Mendez MG, Kojima SI, Goldman RD. Vimentin induces changes in cell shape, motility, and adhesion during the epithelial to mesenchymal transition. FASEB J. 2010;24(6):1838–51. https://0-doi-org.brum.beds.ac.uk/10.1096/fj.09-151639.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Cunningham KE, Turner JR. Myosin light chain kinase: pulling the strings of epithelial tight junction function. Ann N Y Acad Sci. 2012;1258(1):34–42. https://0-doi-org.brum.beds.ac.uk/10.1111/j.1749-6632.2012.06526.x.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Broschat K, Burgess DR. Low Mr tropomyosin isoforms from chicken brain and intestinal epithelium have distinct actin-binding properties. J Biol Chem. 1986;261(28):13350–9. https://0-doi-org.brum.beds.ac.uk/10.1016/S0021-9258(18)69311-1.

    Article  CAS  PubMed  Google Scholar 

  72. Jaffe AB, Hall A. Rho GTPases in transformation and metastasis. Adv Cancer Res. 2002;84:57–80. https://0-doi-org.brum.beds.ac.uk/10.1016/S0065-230X(02)84003-9.

    Article  CAS  PubMed  Google Scholar 

  73. Bacallao R, Garfinkel A, Monke S, Zampighi G, Mandel LJ. ATP depletion: a novel method to study junctional properties in epithelial tissues. I. Rearrangement of the actin cytoskeleton. J Cell Sci. 1994;107(12):3301–13. https://0-doi-org.brum.beds.ac.uk/10.1242/jcs.107.12.3301.

    Article  CAS  PubMed  Google Scholar 

  74. Maragos WF, Young KL, Altman CS, Pocernich CB, Drake J, Butterfield DA, et al. Striatal damage and oxidative stress induced by the mitochondrial toxin malonate are reduced in clorgyline-treated rats and MAO-A deficient mice. Neurochem Res. 2004;29(4):741–6. https://0-doi-org.brum.beds.ac.uk/10.1023/B:NERE.0000018845.82808.45.

    Article  CAS  PubMed  Google Scholar 

  75. Salvi M, Battaglia V, Brunati AM, Rocca NL, Tibaldi E, Pietrangeli P, et al. Catalase takes part in rat liver mitochondria oxidative stress defense. J Biol Chem. 2007;282(33):24407–15. https://0-doi-org.brum.beds.ac.uk/10.1074/jbc.M701589200.

    Article  CAS  PubMed  Google Scholar 

  76. Siritantikorn A, Johansson K, Åhlen K, Rinaldi R, Suthiphongchai T, Wilairat P, et al. Protection of cells from oxidative stress by microsomal glutathione transferase 1. Biochem Bioph Res Co. 2007;355(2):592–6. https://0-doi-org.brum.beds.ac.uk/10.1016/j.bbrc.2007.02.018.

    Article  CAS  Google Scholar 

  77. Wild AC, Mulcahy RT. Regulation of γ-glutamylcysteine synthetase subunit gene expression: insights into transcriptional control of antioxidant defenses. Free Radic Res. 2000;32(4):281–301. https://0-doi-org.brum.beds.ac.uk/10.1080/10715760000300291.

    Article  CAS  PubMed  Google Scholar 

  78. Schmitt CC, Aranias T, Viel T, Chateau D, Gall ML, Waligora-Dupriet AJ, et al. Intestinal invalidation of the glucose transporter GLUT2 delays tissue distribution of glucose and reveals an unexpected role in gut homeostasis. Mol Metab. 2017;6(1):61–72. https://0-doi-org.brum.beds.ac.uk/10.1016/j.molmet.2016.10.008.

    Article  CAS  PubMed  Google Scholar 

  79. Huan PH, Hung LV. Use of the double Thiry-Vella loop in the study of the effects of pantothenic acid on intestinal absorption of glucose. Br J Nutr. 1972;28(3):405–8. https://0-doi-org.brum.beds.ac.uk/10.1079/BJN19720049.

    Article  CAS  PubMed  Google Scholar 

  80. Mziaut H, Mulligan B, Hoboth P, Otto O, Ivanova A, Herbig M, et al. The F-actin modifier villin regulates insulin granule dynamics and exocytosis downstream of islet cell autoantigen 512. Mol Metab. 2016;5(8):656–68. https://0-doi-org.brum.beds.ac.uk/10.1016/j.molmet.2016.05.015.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Gao Y, Nelson DW, Banh T, Yen MI, Yen CLE. Intestine-specific expression of MOGAT2 partially restores metabolic efficiency in Mogat2-deficient mice. J Lipid Res. 2013;54(6):1644–52. https://0-doi-org.brum.beds.ac.uk/10.1194/jlr.M035493.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Nelson DW, Gao Y, Yen MI, Yen CLE. Intestine-specific deletion of acyl-CoA: monoacylglycerol acyltransferase (MGAT) 2 protects mice from diet-induced obesity and glucose intolerance. J Biol Chem. 2014;289(25):17338–49. https://0-doi-org.brum.beds.ac.uk/10.1074/jbc.M114.555961.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. National Research Council. Nutrient Requirements of Poultry. 9th ed. Washington: National Academies Press; 1994.

    Google Scholar 

  84. Tan J, Applegate TJ, Liu S, Guo Y, Eicher S. Supplemental dietary L-arginine attenuates intestinal mucosal disruption during a coccidial vaccine challenge in broiler chickens. Br J Nutr. 2014;112(7):1098–109. https://0-doi-org.brum.beds.ac.uk/10.1017/S0007114514001846.

    Article  CAS  PubMed  Google Scholar 

  85. Lu B, Ren Y, Huang B, Liao W, Cai Z, Tie X. Simultaneous determination of four water-soluble vitamins in fortified infant foods by ultra-performance liquid chromatography coupled with triple quadrupole mass spectrometry. J Chromatogr Sci. 2008;46(3):225–32. https://0-doi-org.brum.beds.ac.uk/10.1093/chromsci/46.3.225.

    Article  CAS  PubMed  Google Scholar 

  86. Woollard DC, Indyk HE, Christiansen SK. The analysis of pantothenic acid in milk and infant formulas by HPLC. Food Chem. 2000;69(2):201–8. https://0-doi-org.brum.beds.ac.uk/10.1016/S0308-8146(99)00255-1.

    Article  CAS  Google Scholar 

  87. Petteys BJ, Frank EL. Rapid determination of vitamin B2 (riboflavin) in plasma by HPLC. Clin Chim Acta. 2011;412(1-2):38–43. https://0-doi-org.brum.beds.ac.uk/10.1016/j.cca.2010.08.037.

    Article  CAS  PubMed  Google Scholar 

  88. Tang J, Hu J, Xue M, Guo Z, Xie M, Zhang B, et al. Maternal diet deficient in riboflavin induces embryonic death associated with alterations in the hepatic proteome of duck embryos. Nutr Metab. 2019;16(1):19. https://0-doi-org.brum.beds.ac.uk/10.1186/s12986-019-0345-8.

    Article  CAS  Google Scholar 

  89. Zhu Y-W, Lu L, Li W-X, Zhang L-Y, Ji C, Lin X, et al. Effect of dietary manganese on antioxidant status and expression levels of heat-shock proteins and factors in tissues of laying broiler breeders under normal and high environmental temperatures. Br J Nutr. 2015;114(12):1965–74. https://0-doi-org.brum.beds.ac.uk/10.1017/S0007114515003803.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgments

We thank the members of our laboratory for their helpful comments on the manuscript. We thank Wuhan GeneCreate Biological Engineering Co., Ltd. for performing the mass spectrometric analysis.

Funding

This work was supported by the earmarked fund for the China Agriculture Research System (CARS-42), the science and technology innovation project of the Chinese Academy of Agricultural Sciences (CXGC-IAS-09). The funding bodys had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

JT performed the study, was involved in all aspects of analysis and drafted the manuscript. YF, MX, ZZ and SH were involved in experimental design. BZ, YW, ZG performed data analysis and sample analysis. JT, YF, SL and SH were involved in the data interpretation and writing. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Shuisheng Hou.

Ethics declarations

Ethics approval and consent to participate

All experimental procedures with ducks were performed according to the Guidelines for Experimental Animals established by the Ministry of Science and Technology (Beijing, China). Ethical approval on animal survival was given by the animal ethics committee of the Institute of Animal Sciences (IAS), Chinese Academy of Agricultural Sciences (CAAS, Beijing, China) with the following reference number: IASCAAS-2019–19.

Consent for publication

Not applicable.

Competing interests

All authors declare that they have no conflict of interests.

Additional information

Publisher’s Note

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

Supplementary Information

Additional file 1.

List of differentially expressed proteins in duodenum mucosa caused by pantothenic acid deficiency.

Additional file 2 Fig. S1.

Complete images of Western blots shown in Fig. 3. Western blot analysis of medium-chain-specific acyl-CoA dehydrogenase (ACADM) and glyceraldehyde-3-phosphate dehydrogenase (GAPDH) protein abundance of mucosal tissue of ducks in the pantothenic acid deficient (PAD) and Control (CON) groups. Histone H3 served as a loading control.

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

Tang, J., Feng, Y., Zhang, B. et al. Severe pantothenic acid deficiency induces alterations in the intestinal mucosal proteome of starter Pekin ducks. BMC Genomics 22, 491 (2021). https://0-doi-org.brum.beds.ac.uk/10.1186/s12864-021-07820-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://0-doi-org.brum.beds.ac.uk/10.1186/s12864-021-07820-x

Keywords