Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Influence of postruminal casein infusion and exogenous glucagon-like peptide 2 administration on the jejunal mucosal transcriptome in cattle

  • Ronald J. Trotta ,

    Roles Conceptualization, Data curation, Formal analysis, Funding acquisition, Investigation, Methodology, Project administration, Visualization, Writing – original draft, Writing – review & editing

    ronald.trotta@uky.edu

    Affiliation Department of Animal and Food Sciences, University of Kentucky, Lexington, Kentucky, United States of America

  • Kendall C. Swanson,

    Roles Conceptualization, Data curation, Funding acquisition, Investigation, Methodology, Project administration, Resources, Supervision, Writing – review & editing

    Affiliation Department of Animal Science, North Dakota State University, Fargo, North Dakota, United States of America

  • James L. Klotz,

    Roles Methodology, Resources, Writing – review & editing

    Affiliation Forage-Animal Production Research Unit, United States Department of Agriculture, Agricultural Research Service, Lexington, Kentucky, United States of America

  • David L. Harmon

    Roles Conceptualization, Data curation, Funding acquisition, Investigation, Methodology, Project administration, Resources, Supervision, Writing – review & editing

    Affiliation Department of Animal and Food Sciences, University of Kentucky, Lexington, Kentucky, United States of America

Abstract

We previously demonstrated that postruminal casein infusion and exogenous glucagon-like peptide 2 (GLP-2) administration independently stimulated growth and carbohydrase activity of the pancreas and jejunal mucosa in cattle. The objective of the current study was to profile the jejunal mucosal transcriptome of cattle using next-generation RNA sequencing in response to postruminal casein infusion and exogenous GLP-2. Twenty-four Holstein steers [250 ± 23.1 kg body weight (BW)] received a continuous abomasal infusion of 3.94 g raw corn starch/kg of BW combined with either 0 or 1.30 g casein/kg of BW for 7 d. Steers received subcutaneous injections at 0800 and 2000 h to provide either 0 or 100 μg GLP-2/kg of BW per day. At the end of the 7-d treatment period, steers were slaughtered for collection of the jejunal mucosa. Total RNA was extracted from jejunal mucosal tissue, strand-specific cDNA libraries were prepared, and RNA sequencing was conducted to generate 150-bp paired-end reads at a depth of 40 M reads per sample. Differentially expressed genes (DEG), KEGG pathway enrichment, and gene ontology enrichment were determined based on the FDR-corrected P-value (padj). Exogenous GLP-2 administration upregulated (padj < 0.05) 667 genes and downregulated 1,101 genes of the jejunal mucosa. Sphingolipid metabolism, bile secretion, adherens junction, and galactose metabolism were among the top KEGG pathways enriched with upregulated DEG (padj < 0.05) in response to exogenous GLP-2 administration. The top gene ontologies enriched with upregulated DEG (padj < 0.05) in response to exogenous GLP-2 administration included nutrient metabolic processes, brush border and bicellular tight junction assembly, and enzyme and transporter activities. Exogenous GLP-2 administration increased or tended to increase (padj < 0.10) brush border carbohydrase (MGAM, LCT, TREH), hexose transporter (SLC5A1, SLC2A2), and associated transcription factor (HNF1, GATA4, KAT2B) mRNA expression of the jejunal mucosa. Gene ontologies and KEGG pathways that were downregulated (padj < 0.05) in response to exogenous GLP-2 were related to genetic information processing. Postruminal casein infusion downregulated (padj < 0.05) 7 jejunal mucosal genes that collectively did not result in enriched KEGG pathways or gene ontologies. This study highlights some of the transcriptional mechanisms associated with increased growth, starch assimilation capacity, and barrier function of the jejunal mucosa in response to exogenous GLP-2 administration.

Introduction

In North American beef cattle and dairy cattle production systems, grain-based diets containing moderate to large proportions of starch are typically fed to increase the net energy concentration of the diet to allow for more efficient growth and improved product quality. Depending on grain source and processing methods, up to 40% of dietary starch intake can escape ruminal fermentation and flow to the small intestine for potential enzymatic digestion [1]. Enzymatic digestion of starch in the small intestine can be energetically more efficient than ruminal fermentation of starch because absorption and oxidation of glucose provides more energy to the host than the production and oxidation of short-chain fatty acids [2]. However, small intestinal starch digestibility in ruminants is potentially limited by inadequate production of pancreatic and/or small intestinal carbohydrases [3, 4].

In prior studies, postruminal casein infusion increased pancreatic mass [5], pancreatic α-amylase mRNA expression, activity, and secretion [510], small intestinal starch digestibility [1117], and net portal glucose absorption [18] in ruminants. Exogenous glucagon-like peptide 2 (GLP-2) has been shown to increase small intestinal mucosal growth by activating intracellular signaling pathways which promote cell proliferation and survival and inhibit cell proteolysis and apoptosis [1921]. In pre-term parenterally-fed neonatal pigs, exogenous GLP-2 increased jejunal sucrase-isomaltase and maltase-glucoamylase mRNA expression, as well as, jejunal maltase and sucrase activity [22]. These prior studies suggest that postruminal casein infusion and exogenous GLP-2 administration would have independent effects on pancreatic and small intestinal growth and carbohydrase activities necessary for small intestinal starch hydrolysis.

With the goal of independently stimulating pancreatic α-amylase activity and jejunal α-glucosidase activity, we recently evaluated the effects of postruminal casein infusion and exogenous GLP-2 administration on growth, cellularity, and enzyme activity of the pancreas and small intestine in cattle [23]. In accordance with our hypotheses, postruminal casein infusion and exogenous GLP-2 administration differentially stimulated the growth and enzyme activities of the pancreas and small intestinal mucosa in cattle abomasally infused with starch [23]. Postruminal casein infusion increased pancreatic mass, protein content, and α-amylase activity [23]. Exogenous GLP-2 administration increased jejunal mass, mucosal mass, DNA content, and α-glucosidase activity (sum of maltase, isomaltase, and glucoamylase activity) [23]. To better understand the hydrolytic limitations of small intestinal starch digestion in cattle, deeper investigations into the molecular mechanisms associated with increased jejunal growth and α-glucosidase activity are needed.

Understanding epithelial cell function during periods of compromised nutrient utilization by the gastrointestinal tract, such as limitations in small intestinal starch hydrolysis, is a key opportunity for future research [24]. Although there have been several investigations focused on adaptation of the ruminal epithelium to high-starch diets [2529], there is limited research examining functional measurements of the small intestinal epithelium combined with high-throughput molecular-based techniques to ascribe important phenotypes of ruminant animal production. It was hypothesized that RNA sequencing analyses would lead to the discovery of novel transcriptional mechanisms affected by postruminal casein infusion and exogenous GLP-2 administration. Therefore, the objective of this experiment was to evaluate the effects of postruminal casein infusion and exogenous GLP-2 administration on transcriptomic pathways and functions of the jejunal mucosa using next-generation RNA sequencing.

Materials and methods

All surgical, animal care, and experimental protocols were approved by the University of Kentucky Animal Care and Use Committee (2020–3479).

Animals, diet, and abomasal catheterization surgery

The experimental design was previously described by Trotta [23]. Briefly, 24 Holstein steers [initial body weight (BW) = 250 ± 23.1 kg; 10–12 months of age] were initially housed in individual pens (3 m × 3 m) in the Intensive Research Building of the University of Kentucky C. Oran Little Agricultural Research Center in Versailles, KY. Before surgery, steers were deprived of feed (36 h) and water (12 h), fitted with temporary jugular vein catheters (Mila International Inc., Florence, KY), and received a subcutaneous injection of ceftiofur (6.6 mg/kg of BW) and intravenous injection of flunixin meglumine (1 mg/kg of BW). Steers were sedated by intravenous injection of xylazine (0.088 mg/kg of BW) with ketamine hydrochloride (1.76 mg/kg of BW), intubated, and anesthesia was maintained with isoflurane in O2 using a ventilator in left lateral recumbency. The depth of anesthesia was monitored through observations of eye reflex and heart rate throughout the surgery. Steers were surgically fitted with infusion catheters in the abomasum that were constructed of Tygon tubing (6.35-mm internal diameter; Saint-Gobain North America, Malvern, PA) [30]. Flunixin meglumine was administered intravenously 24-h post-surgery and steers were monitored for abnormal feed intake, fecal consistency, behavior, and mobility for 7 d, and rectal temperatures were monitored for 3 d post-surgery.

Steers were limit-fed 5.65 ± 0.468 kg of an alfalfa hay cube-based diet (82.9% alfalfa hay cubes, 16.2% finely ground corn, 0.45% trace mineral premix, 0.45% vitamin premix; 40.7% neutral detergent fiber, 33.3% acid detergent fiber, 15.4% crude protein; 13.8% total starch, 1.32 Mcal/kg net energy for maintenance) on a dry matter (DM) basis. The diet was formulated to supply 1.33 times the net energy required for maintenance and exceed requirements for vitamins, minerals, ruminally degradable protein, and metabolizable protein for a steer gaining 0.46 kg/d [31]. Rations were provided twice daily at 0800 and 2000 h in equal portions. Steers were adapted to the basal diet for at least 7 d before the start of the treatment period.

Experimental design

Steers were transferred to metabolism stalls (1.2 m × 2.4 m) for the 7-d treatment period. Due to a limited number of stalls and infusion apparatus, steers were stratified by BW into 6 replicate blocks. Steers were then randomly assigned to one of four treatments within each replicate block. The experimental design was a randomized complete block design with a 2 × 2 factorial arrangement of treatments. All steers were abomasally infused with 3.94 ± 0.245 g raw corn starch · kg BW-1 · d-1. Along with raw corn starch, steers were abomasally infused with either water or 10% (wt:wt) sodium caseinate solution to provide 0 or 1.30 ± 0.299 g sodium caseinate · kg BW-1 · d-1. Additionally, steers received subcutaneous injection treatments in two equal portions daily at 0800 and 2000 h. Subcutaneous injection treatments were either vehicle (5 g/L bovine serum albumin (BSA) diluted in 9 g/L NaCl) or 100 μg GLP-2 in vehicle · kg BW-1 · d-1. This resulted in four treatments: 1) abomasal water + vehicle injection (control), 2) abomasal water + GLP-2 injection (GLP-2), 3) abomasal 10% sodium caseinate solution + vehicle injection (casein), and 4) abomasal 10% sodium caseinate solution + GLP-2 injection (casein + GLP-2). All steers received abomasal infusion and subcutaneous injection treatments for 7 d. The 7-d treatment period was chosen because postruminal casein infusion has been shown to increase small intestinal starch disappearance within 6 d [12] and acute (1 d) or chronic (10 d) exposure to GLP-2 at the same dosage level achieved pharmacological plasma concentrations of GLP-2 in cattle [32].

Glucagon-like peptide 2 preparation and administration

Glucagon-like peptide 2 was synthesized (Alan Scientific Inc., Gaithersburg, MD) based on the native bovine sequence [33, 34]. The purity (96%) was quantified by the manufacturer using high-performance liquid chromatography. The GLP-2 solution was prepared to a final concentration of 5 mg/mL by dissolving GLP-2 in 5 g/L BSA in 9 g/L NaCl solution. Aliquots were frozen in 10 mL screw-cap tubes at -20°C until use. Intestinotrophic effects of GLP-2 are greater when GLP-2 was administered subcutaneously compared with intramuscular or intraperitoneal injections [35]. Therefore, treatments were administered subcutaneously in the neck region in front of the shoulder. Left and right lateral administration sites were alternated for morning and evening injections, respectively.

Nutrient infusion preparation and techniques

Due to large inter- and intra-animal variability in postruminal starch flows when high-starch diets are fed to cattle [3, 36], postruminal starch flow was controlled via continuous abomasal infusion of starch. Raw corn starch (Clinton 185 Corn Starch; Archer Daniels Midland Company, Decatur, IL) was chosen as the starch source to facilitate the greatest limitations in small intestinal starch digestibility, as the action of pancreatic α-amylase and small intestinal α-glucosidases are required for complete hydrolysis to glucose. The chosen amount of infused raw corn starch was predicted to limit the extent of small intestinal starch disappearance [37] and mimic natural amounts of postruminal starch flow in cattle fed high-grain diets [38]. Additionally, infusion of 986 g of raw corn starch per day equates to 41.1 g/h, which does not cause diarrhea and appreciable amounts (10 to 15 g/h) of infused carbohydrate were expected to pass the ileocecal junction [39]. Sodium caseinate (AMCO Proteins, Burlington, NJ) was chosen as the protein source because of its relatively high solubility in water and because approximately 70–85% of luminal amino acids are transported across the apical membrane as small peptides [40]. The amount of casein was chosen relative to the amount of corn starch infused (1 part casein: 3 parts corn starch) and because previous experiments have demonstrated increased small intestinal starch disappearance [11, 15] and pancreatic α-amylase secretion [7] at similar levels.

Two containers of infusate suspensions per steer were prepared daily, immediately before infusion, for use over 12-h intervals. Infusate suspensions were maintained through continuous stirring via a stir bar with stir plate. Infusate suspensions (3.26 ± 0.109 kg) were prepared using the appropriate amount of raw corn starch, 5% (wt:wt) CrEDTA solution as an indigestible flow marker [41], and tap water every 12 h for each steer. A 10% (wt:wt) solution of sodium caseinate [42] was prepared with warm tap water in an insulated kettle (LEC-40; Legion Industries Inc., Waynesville, GA) using an industrial electric mixer (CDP3330; Baldor Electric Co., Ft. Smith, AR). The solution was mixed for at least 2 h and then was allowed to rest for 8 h to allow for air bubbles to escape [43]. The final solution was stored in 18.9-L buckets at -20°C until use. For casein treatments, the sodium caseinate solution replaced tap water in the infusate suspension. Corn starch and casein solute concentrations and abomasal infusion rates were similar to those reported by Kreikemeier [39] and Richards [15]. Abomasal infusion treatments were continuously infused (271 ± 9.08 g infusate/h) through Tygon tubing (3.18-mm internal diameter) using a multi-channel peristaltic pump (205U/CA; Watson-Marlow, Falmouth, United Kingdom). Infusion tubing was inserted into abomasal infusion catheters and secured with hose clamps. The amount infused was determined by the weight of the residual infusate after each 12-h period. The infusate containers were placed on top of a portable shelf (> 2 m height) to aid in the pumping of the suspension. Measures of fecal flow and postruminal starch disappearance from steers used in the current study were previously reported [23].

Tissue collection

Upon completion of the 7-d infusion period, the infusion tubing was disconnected from abomasal infusion catheters and steers were transported to the University of Kentucky Meats Laboratory within 1 h. Steers were humanely stunned via captive bolt and exsanguinated. After slaughter, the viscera were rapidly removed and separated for individual weights and subsample collection. The pyloric and ileocecal junctions were cut to separate the small intestine from the abomasum and cecum, respectively, and the mesentery was cut to separate the entire small intestine from the viscera. The small intestine was measured in length by looping the intestine between pegs at either end of a 2.43-m board [44]. The small intestine was separated into the duodenum (0.1 to 1.1 m caudal to the pyloric sphincter), jejunum (proximal half of non-duodenal small intestine), and ileum (distal half of non-duodenal small intestine) [45]. After the small intestinal length was measured, digesta was gently removed from the lumen of the small intestine. Each intestinal section (duodenum, jejunum, ileum) was weighed and sampled (1 m) from the midpoint. Each 1-m segment was cut into three equidistant subsamples, everted, and rinsed in ice-cold saline. The first subsample was cut laterally, scraped with a glass microscope slide, and 200 mg of mucosa was weighed and preserved with RNAlater stabilization solution (Thermo Scientific Inc., Waltham, MA) in RNase-free tubes. Tubes were initially stored at 4°C for 24 h to allow for tissue penetration and then were transferred to -80°C storage, as recommended by the manufacturer. The second subsample was scraped as described previously and the mucosa was flash-frozen in liquid nitrogen in an aluminum pouch and stored at -80°C for cellularity and enzymatic analyses. The third subsample was weighed, mucosa scraped from the entire section, and then the isolated mucosa was weighed. Measures of small intestinal length and mass, cellularity, and enzyme activity were previously published [23].

RNA isolation, library preparation, and sequencing

The jejunal mucosa was selected for RNA isolation and sequencing because small intestinal α-glucosidase activity (maltase, isomaltase, glucoamylase), Na+-dependent glucose uptake activity, and mRNA expression of intestinal glucose transporters (SLC5A1 and SLC2A2) is greatest in the jejunum of cattle [44, 46, 47]. Additionally, relative mRNA expression of the GLP-2 receptor (GLP2R) is greater in the jejunum than other gastrointestinal tissues [48]. Approximately 200 mg of jejunal mucosal tissue was shipped for RNA extraction, library preparation, and RNA sequencing by Novogene Corporation Inc. (Sacramento, CA). Total RNA was extracted from jejunal mucosal tissue using TRIzol reagent (Thermo Fisher Scientific Inc., Waltham, MA). RNA concentration (1,003 ± 863 ng/μL) was measured at 260 nm using a microvolume UV-VIS spectrophotometer (NanoDrop; Thermo Fisher Scientific Inc., Waltham, MA) and purity was assessed using the 260/280 nm ratio (range = 1.8 to 2.2) for protein contamination and 260/230 nm ratio (≥ 1.8) for nucleic acid contamination. Samples were assessed for RNA integrity (mean ± standard deviation = 6.59 ± 1.30; range = 4.2 to 9; median = 7) using the RNA 6000 Nano Kit (Agilent Technologies Inc., Santa Clara, CA) and automated electrophoresis (2100 Bioanalyzer Instrument; Agilent Technologies Inc., Santa Clara, CA). One sample in the control treatment was not prepared for RNA sequencing due to low RNA integrity (RIN < 4).

Strand-specific cDNA library preparation was conducted by Novogene Corporation Inc. (Sacramento, CA). Messenger RNA was isolated from total RNA using poly-T oligo-attached magnetic beads. Fragmentation was facilitated using divalent cations under elevated temperature in First Strand Synthesis Reaction Buffer (5X). First-strand cDNA was synthesized using random hexamer primer and Moloney murine leukemia virus reverse transcriptase and second-strand cDNA was synthesized using DNA polymerase I and RNAse H. Remaining overhangs were converted to blunt ends using exonuclease and polymerase activities. Three-prime ends of DNA fragments were adenylated and adaptors with hairpin loop structures were ligated for hybridization preparation. Library fragments were purified with AMPure XP System (Beckman Coulter Inc., Brea, CA) and 370–420 bp cDNA fragments were selected for PCR amplification. Polymerase chain reaction products were purified and assessed with the 2100 Bioanalyzer Instrument (Agilent Technologies Inc.). Strand-specific RNA sequencing was conducted NovaSeq 6000 Sequencing System (Illumina Inc., San Diego, CA) to generate 150-bp paired-end reads at an average depth of 55,666,406 reads per sample.

Bioinformatic and statistical analyses

Raw reads were filtered to remove sequencing adaptors, low-complexity reads, and reads containing low-quality bases using fastp software [49]. Quality control scores (Q20 = 96.8 ± 0.370%; Q30 = 91.41 ± 0.831%) and GC content (47.9 ± 1.25%) were calculated. Paired-end clean reads were aligned to the Bos taurus reference genome (ARS-UCD1.2) using HISAT 2 (version 2.0.5) [50]. Mapped reads of each sample were assembled using StringTie (version 1.3.3b) [51] in a reference-based approach. There was an average of 40,644,458 uniquely mapped reads per sample (73.0% of total reads). Reads per gene were counted using featureCounts (version 1.5.0-p3) [52]. Quantification of each gene was calculated as the number of fragments per kilobase of transcript sequence per million base pairs sequenced (FPKM). Visualization of box plots, violin plots, and FPKM density plots confirmed that gene expression was distributed equally across each steer and that the FPKM density conformed to a negative binomial distribution. Differential gene expression (DEG) analysis was quantified using the DESeq2 R-package (version 1.20) [53].

The DEG analysis used the negative binomial generalized linear model to fit gene expression level as a negative binomial distribution and Wald statistics to perform hypothesis testing. Multiple testing adjustment of the P-values was performed using the Benjamini-Hochberg procedure for controlling the false discovery rate (FDR). Genes were identified as differentially expressed if the FDR-adjusted P-value (padj) was ≤ 0.05. Gene ontology and KEGG pathway enrichment analyses of DEGs were conducted using the clusterProfiler R-package [54]. Enrichment analyses were considered significant if the FDR-adjusted P-value was ≤ 0.05. Tendencies were declared when 0.05 < FDR-adjusted P ≤ 0.10. Changes in gene expression were classified as upregulated or downregulated based on the sign of the log2 fold change.

Initially, pairwise comparisons between each of the four treatments were assessed. However, casein did not generate DEG that collectively affected KEGG pathway or gene ontology enrichment compared with control. In addition, there were minimal effects that resembled a casein × GLP-2 treatment interaction when GLP-2 vs. casein + GLP-2 treatments were compared. Therefore, all subsequent data analyses were conducted to compare the main effects of postruminal casein infusion (n = 11 for water, n = 12 for casein) and exogenous GLP-2 administration (n = 11 for BSA, n = 12 for GLP-2).

Results

A complete description of differentially expressed genes, KEGG pathways, and gene ontologies can be found in supporting information (S1S8 Files).

Summary of results

A summary of the effects of exogenous GLP-2 administration on differentially expressed genes, KEGG pathways, and gene ontologies is presented in Table 1. Exogenous GLP-2 administration upregulated (padj < 0.05) 667 DEGs, 26 KEGG pathways, 198 biological processes, 56 cellular components, and 60 molecular functions. Exogenous GLP-2 downregulated (padj < 0.05) 1,101 DEGs, 14 KEGG pathways, 270 biological processes, 105 cellular components, and 46 molecular functions. Postruminal casein infusion did not result (padj > 0.05) in any enriched KEGG pathways or gene ontologies of the jejunal mucosa.

thumbnail
Table 1. Summary of the effects of exogenous GLP-2 administration on the number of differentially expressed genes, KEGG pathways, and gene ontologies of the jejunal mucosa.

https://doi.org/10.1371/journal.pone.0308983.t001

Differentially expressed genes

Exogenous GLP-2 administration resulted in 1,768 DEGs (padj < 0.05) with 667 upregulated genes and 1,101 downregulated genes (Fig 1). Postruminal casein infusion resulted in 7 DEGs (padj < 0.05) with 0 upregulated genes and 7 downregulated genes in the jejunal mucosa. Transcripts of the jejunal mucosa that were downregulated (padj < 0.05) with postruminal casein infusion include ENSBTAG00000047783 (log2 fold change = −4.00), ENSBTAG00000040367 (log2 fold change = −6.07), ENSBTAG00000047029 (log2 fold change = −1.13), CCR7 (C-C motif chemokine receptor 7; log2 fold change = −2.41), SFRP5 (secreted frizzled related protein 5; log2 fold change = −4.11), FCER2 (FC epsilon receptor II; log2 fold change = −3.48), and C1QTNF2 (C1q and TNF related 2; log2 fold change = −5.41) (S4 File).

thumbnail
Fig 1. Volcano plot of jejunal mucosal genes influenced by exogenous GLP-2 administration.

Data points highlighted in red were differentially expressed (padj < 0.05).

https://doi.org/10.1371/journal.pone.0308983.g001

Gene ontologies

The top 10 biological processes that were upregulated (padj < 0.05) with exogenous GLP-2 administration were the alcohol metabolic process, sphingolipid metabolic process, organic hydroxy compound metabolic process, anion transport, cellular lipid catabolic process, monocarboxylic acid metabolic process, carbohydrate metabolic process, transition metal ion transport, small molecule catabolic process, and lipid catabolic process (Fig 2). The top 10 cellular components that were upregulated (padj < 0.05) with exogenous GLP-2 administration were apical junction complex, apical part of the cell, brush border, Golgi subcompartment, Golgi apparatus part, apical plasma membrane, the cluster of actin-based cell projections, cell-cell junction, occluding junction, and Golgi membrane. The top 10 molecular functions that were upregulated (padj < 0.05) with exogenous GLP-2 administration were hydrolase activity (hydrolyzing O-glycosyl compounds), hydrolase activity (hydrolyzing glycosyl bonds), anion transmembrane transporter activity, coenzyme binding, transition metal ion transmembrane transporter activity, lipid transporter activity, organic anion transmembrane transporter activity, active transmembrane transporter activity, UDP-glycosyltransferase activity, and cofactor binding.

thumbnail
Fig 2. Top enriched gene ontology terms of the jejunal mucosa that were upregulated with exogenous GLP-2 administration.

GeneRatio is the ratio of upregulated DEGs to all genes in a gene ontology. Padj is the probability value for hypergeometric test adjusted for Benjamini-Hochberg FDR correction.

https://doi.org/10.1371/journal.pone.0308983.g002

The top 10 biological processes that were downregulated (padj < 0.05) with exogenous GLP-2 administration were ribonucleoprotein complex biogenesis, ribosome biogenesis, rRNA metabolic process, rRNA processing, ncRNA processing, ribosomal large subunit biogenesis, ribonucleoprotein complex assembly, ribosomal small subunit biogenesis, and ribonucleoprotein complex subunit organization (Fig 3). The top 10 cellular components that were downregulated (padj < 0.05) with exogenous GLP-2 administration were ribosomal subunit, cytosolic ribosome, ribosome, cytosolic part, cytosolic large ribosomal subunit, preribosome, large ribosomal subunit, cytosolic small ribosomal subunit, and nucleolar part. The top 10 molecular functions that were downregulated (padj < 0.05) with exogenous GLP-2 administration were the structural constituent of ribosome, rRNA binding, translation factor activity (RNA binding), catalytic activity (acting on RNA), translation initiation factor activity, catalytic activity (acting on DNA), chromatin binding, snoRNA binding, histone binding, and helicase activity.

thumbnail
Fig 3. Top enriched gene ontology terms of the jejunal mucosa that were downregulated with exogenous GLP-2 administration.

GeneRatio is the ratio of upregulated DEGs to all genes in a gene ontology. Padj is the probability value for hypergeometric test adjusted for Benjamini-Hochberg FDR correction.

https://doi.org/10.1371/journal.pone.0308983.g003

KEGG pathways

The KEGG pathways that were upregulated (padj < 0.05) with exogenous GLP-2 administration were sphingolipid metabolism, adherens junction, bile secretion, galactose metabolism, vitamin digestion and absorption, mineral absorption, peroxisome, carbon metabolism, tight junction, other glycan degradation, starch and sucrose metabolism, retinol metabolism, lysosome, insulin resistance, ascorbate and aldarate metabolism, N-glycan biosynthesis, valine leucine and isoleucine degradation, glycerolipid metabolism, PPAR signaling pathway, and lysine degradation, protein digestion and absorption, AMPK signaling pathway, fat digestion and absorption, propanoate metabolism, gastric cancer, and various types of N-glycan biosynthesis (Table 2).

thumbnail
Table 2. Enriched KEGG pathways of the jejunal mucosa that were upregulated with exogenous GLP-2 administration1.

https://doi.org/10.1371/journal.pone.0308983.t002

The KEGG pathways that tended to be upregulated (padj < 0.10) with exogenous GLP-2 administration were citrate cycle (TCA cycle), amino sugar and nucleotide sugar metabolism, tryptophan metabolism, arginine biosynthesis, serotonergic synapse, pentose and glucoronate interconversions, fructose and mannose metabolism, glycolysis/gluconeogenesis, Hippo signaling pathway, ether lipid metabolism, ABC transporters, drug metabolism–cytochrome P450, pyruvate metabolism, porphyrin and chlorophyll metabolism, and protein processing in endoplasmic reticulum (Table 3).

thumbnail
Table 3. Enriched KEGG pathways of the jejunal mucosa that tended to be upregulated with exogenous GLP-2 administration1.

https://doi.org/10.1371/journal.pone.0308983.t003

Downregulated KEGG pathways (padj < 0.05) in response to exogenous GLP-2 administration include ribosome, RNA transport, DNA replication, ribosome biogenesis in eukaryotes, spliceosome, cell cycle, base excision repair, mismatch repair, nucleotide excision repair, primary immunodeficiency, RNA polymerase, Fanconi anemia pathway, B cell receptor signaling pathway, and leishmaniasis (Table 4).

thumbnail
Table 4. Enriched KEGG pathways of the jejunal mucosa that were downregulated with exogenous GLP-2 administration1.

https://doi.org/10.1371/journal.pone.0308983.t004

mRNA expression of jejunal carbohydrases, hexose transporters, and transcription factors

Exogenous GLP-2 administration did not influence (padj = 0.17) SI mRNA expression in the jejunum but tended to increase (padj = 0.07) jejunal MGAM mRNA expression (Table 5). Jejunal LCT and TREH mRNA expression were increased (padj < 0.05) with exogenous GLP-2 administration. Exogenous GLP-2 administration increased (padj = 0.01) the jejunal mRNA expression of SLC5A1 and tended to increase (padj = 0.09) the jejunal mRNA expression of SLC2A2. Jejunal SLC2A5 and CDX2 mRNA expression were not influenced (padj > 0.15) by exogenous GLP-2 administration. Exogenous GLP-2 increased (padj < 0.01) or tended to increase (padj = 0.08) the mRNA expression of genes encoding the transcriptional coactivators, HNF1A and GATA4. Jejunal KAT2B mRNA expression was upregulated (padj = 0.01) by exogenous GLP-2 administration.

thumbnail
Table 5. Effects of exogenous GLP-2 administration on mRNA expression of genes encoding brush border carbohydrases, intestinal glucose transporters, and transcriptional regulators of carbohydrate-responsive genes1.

https://doi.org/10.1371/journal.pone.0308983.t005

Discussion

Influence of postruminal casein infusion on the jejunal transcriptome

To our knowledge, this is the first study that has evaluated the effects of postruminal casein infusion on the jejunal mucosal transcriptome in ruminants. Several plausible explanations exist as to why postruminal casein infusion for 7 d did not affect the jejunal transcriptome in the current study. The basal diet was formulated to exceed net energy and metabolizable protein requirements, whereas effects of postruminal casein infusion might have been detected if dietary energy or protein was limiting. For example, small intestinal starch disappearance increased by 37.6 percentage units (from 54.9% to 92.5%) when postruminal casein was added to a protein-free diet [55]. By comparison, postruminal casein infusion only increased small intestinal starch disappearance by a maximum of 10.2 percentage units when dietary metabolizable protein requirements were met [12]. Supporting this concept, a recent review summarized the effects of postruminal casein infusion from 51 studies and concluded that casein increased DM intake when metabolizable protein supply was deficient and that casein had no effect or decreased DM intake when metabolizable protein supply was in excess [56]. Therefore, it is possible that the lack of transcriptomic response to postruminal casein infusion was due to the positive energy and metabolizable protein balance of steers in the current study.

The method of nutrient administration, time to collect jejunal samples after infusion termination, and approach used to measure gene expression could have influenced the outcome of the current experiment. In the current study, casein was delivered continuously via postruminal infusion for 7 d. It is possible that the continuous postruminal infusion of casein masked effects that would be observed with a single pulse dose or multiple pulse doses per day. Previous research demonstrated that the timing of sampling in relation to luminal nutrient flow is important for detecting DEGs in the intestinal mucosa of cattle [57]. Li [57] found that the number of DEGs in duodenal mucosa decreased from 1490 DEGs to 482 DEGs 24 h after terminating the postruminal infusion of partially hydrolyzed starch. Small intestinal transit time is approximately 3 h in cattle [12] and all steers in the current study were slaughtered within a maximum of 3 to 4 h after termination of the 7-d postruminal infusion. Therefore, it is possible that the final portion of luminal casein had passed the jejunum by the time samples were collected for analysis. Because the intestinal mucosa is a heterogeneous tissue with numerous cell types [24], approaches using single-cell RNA sequencing technology could potentially detect DEGs of lowly expressed cell populations, such as enteroendocrine cells, that were likely underrepresented in the current study [58].

A common finding between the current study and Li [57] is that 7-d postruminal infusion of partially hydrolyzed starch or raw corn starch with casein does not influence many functions of the intestinal epithelium. Although previous research detected 1490 DEGs, the enrichment analyses showed that DEGs found after 7-d of postruminal partially hydrolyzed starch infusion only affected a few, generic gene ontologies of the duodenal mucosa (such as biological process, cellular process, molecular function, binding) [57]. Likewise, postruminal infusion of casein did not result in any enriched KEGG pathways or gene ontologies in the jejunal mucosa in the current study. These findings collectively suggest that continuous postruminal infusion of starch or starch with casein for 7 d does not influence the transcriptome of the intestinal mucosa in cattle.

Increased small intestinal starch disappearance, pancreatic mass, and α-amylase activity have been consistent responses to postruminal casein infusion in ruminants [515, 17, 23, 55, 59]. Notably, the lack of jejunal transcriptomic response in the current study indicates that increasing pancreatic α-amylase activity via 7-d postruminal casein infusion does not increase the mRNA expression of jejunal brush border carbohydrases. These findings support the previously proposed concept that carbohydrases of the bovine pancreas and small intestinal mucosa do not coordinate a response to luminal nutrient flows [23]. Factors mediating the positive response of increased small intestinal starch digestion to postruminal casein may include increasing pancreatic α-amylase activity [23], increasing the digestion of other nutrients [7, 13, 16], stimulating microbial activity [6062], changing passage rate [62], or fostering chemical interactions with starch in the small intestinal lumen [63, 64].

Effects of exogenous GLP-2 on transcriptional mechanisms associated with jejunal mucosal growth

A schematic of the functional and transcriptomic effects of exogenous GLP-2 on the jejunum in cattle is summarized (Fig 4). Exogenous GLP-2 administration for 7 d increased jejunal mucosal mass, mucosal DNA content, and mucosal protein content of steers used in the current study by 45.6%, 32.6%, and 42.8%, respectively [23]. Exogenous GLP-2 administration also increases jejunal villus height, crypt depth, and crypt cell proliferation in cattle [32]. After GLP-2 binds to its receptor, it is generally thought that independent actions of insulin-like growth factor 1 (IGF-1) and epidermal growth factor (EGF) are critical downstream mediators of increased cellular proliferation [6568]. These mediators are thought to lead to activation of Wnt/β-catenin and PI3K/Akt signaling pathways to increase cellular proliferation and differentiation [65, 69]. Currently, there are no annotations of GLP-2 receptor signaling pathways in the KEGG pathway or gene ontology databases. Nevertheless, several transcripts associated with EGF receptor signaling (ERBB2, ERBB3), insulin signaling (INSR, IRS2), Wnt/β-catenin signaling (WNT1, WNT8B, WNT11, FZD3, FZD5, DVL1, GSK3B, CTNNB1, TCF7L2) and PI3K/Akt signaling (PDPK1) pathways were upregulated with exogenous GLP-2 administration. These data suggest that part of the increase in jejunal mucosal growth of steers exposed to exogenous GLP-2 in the current study might have occurred through the activation of GLP-2 receptor signaling pathways.

thumbnail
Fig 4. Schematic of the selected functional and transcriptomic effects of exogenous GLP-2 administration on jejunal mucosal growth and small intestinal starch assimilation capacity.

Functional effects were summarized from studies in cattle [23, 32]. Transcriptomic effects were summarized from the current study. Figure was created with BioRender.com.

https://doi.org/10.1371/journal.pone.0308983.g004

Environmental and cellular cues between conserved signaling pathways (Wnt/β-catenin, TGF-β, Notch, Hedgehog) can lead to cross-talk influencing the activation of the Hippo signaling pathway [70]. The Hippo signaling pathway tended to be upregulated in response to exogenous GLP-2 administration in the current study. In addition to the upregulated transcripts associated with cross-talk from Wnt/β-catenin signaling, key transcriptional regulators of the Hippo signaling pathway (MOB1B, YAP1, SMAD3, SMAD9) controlling cellular proliferation and apoptosis were upregulated in response to exogenous GLP-2 administration. Protein kinases of the Hippo signaling pathway can be regulated by several upstream inputs from intracellular, extracellular, and noncellular components including G protein-coupled receptors, metabolism, mechanical stimuli, cell polarity and density, and stress [70, 71]. Notably, upregulation of transcripts encoding for cadherin 1 (CDH1), an adherens junction protein, and its mammalian ajuba family member (LIMD1) suggest that protein kinase activity (MST1/2, LATS1/2) regulating the Hippo signaling pathway was inhibited, allowing YAP1 translocation to the nucleus to function as a transcriptional co-activator of pro-proliferation and anti-apoptosis genes [7274].

Net tissue accretion can be due to a combination of increasing cellular proliferation and survival and decreasing cellular apoptosis and proteolysis [19, 21, 32, 35, 7577] and activation of these pathways by GLP-2 could vary based on physiological state, the presence of luminal nutrient supply, and with mucosal inflammation/injury [1921]. Because increased mucosal mass can occur via multiple mechanisms of increasing proliferation and decreasing apoptosis, there can be multiple mediators of GLP-2-induced mucosal growth [65]. In the current study, many KEGG pathways associated with genetic information processing were downregulated with exogenous GLP-2 administration including ribosome, RNA transport, DNA replication, ribosome biogenesis in eukaryotes, spliceosome, cell cycle, base excision repair, mismatch repair, nucleotide excision repair, and RNA polymerase. In addition, several transcripts associated with ubiquitin-mediated proteolysis (30/166) were downregulated with exogenous GLP-2 administration, which suggests that part of the increase in jejunal mucosal growth with exogenous GLP-2 administration occurred through decreased proteolysis in the current study. Steers received subcutaneous GLP-2 injections twice daily for 7 d in the current study, and tissues were collected on day 8. Therefore, jejunal mucosa was collected approximately 14–15 h after the last GLP-2 injection was given on day 7. Glucagon-like peptide 2 has a short half-life (7 min) in plasma in humans [78] but, subcutaneous GLP-2 administration can sustain elevated pharmacological plasma GLP-2 concentrations for more than 7 h in cattle [32]. It is possible that changes in the mRNA expression of genes in genetic information processing pathways are transient and that desensitization of the GLP-2 receptor could have occurred with chronic GLP-2 treatment for 7 d [32]. It is possible that the effects of exogenous GLP-2 for 7 d on genetic information processing pathways occurred because of decreased cellular apoptosis and proteolysis, chronic exposure causing GLP-2 receptor desensitization, timing of sampling, or because the pharmacological dose given could potentially differ from responses that would occur under normal physiological conditions.

Effects of exogenous GLP-2 on transcriptional mechanisms associated with jejunal starch assimilation

Epithelial cells are the predominant cell types (> 80%) of the jejunal mucosa [79] and the brush border membrane is estimated to contain approximately 10% to 11% of the total protein content of epithelial cells [80]. Sucrase-isomaltase and maltase-glucoamylase comprise 8.2% and 2.7% of the total protein content of the brush border membrane in humans [81]. Therefore, increasing the growth of the jejunal mucosa via the administration of intestinotrophic hormones, like GLP-2, could potentially result in increased expression of genes encoding brush border membrane proteins [82] (hydrolases, channels and transporters, cytoskeletal components, myosin motor proteins, and adhesion proteins) that are important for adaptation of the intestinal mucosa to high-grain diets [24]. In addition to the brush border of absorptive epithelial cells, the jejunal mucosa sampled in the current study for high-throughput RNA sequencing contains multiple cell types (goblet cells, immune cells, and enteroendocrine cells) that could have specialized functions in response to changes in luminal nutrient flows and exogenous GLP-2 administration [24].

Several prior studies spanning different animal models have demonstrated that exogenous GLP-2 can increase the mRNA expression and enzyme activity of brush border carbohydrases [22, 23, 76, 83, 84]. Brubaker [76] was the first to demonstrate that exogenous GLP-2 increased duodenal maltase activity in mice, but did not influence sucrase or lactase activities. In parenterally-fed rats, intravenous administration of GLP-2 increased jejunal and ileal sucrase-isomaltase mRNA expression [83]. Exogenous GLP-2 administration has been shown to increase sucrase-isomaltase and maltase-glucoamylase mRNA expression and activity in pre-term parenterally-fed neonatal piglets [22]. However, effects of exogenous GLP-2 on sucrase-isomaltase mRNA expression and activity were not observed when administered to full-term parenterally-fed neonatal piglets [84] and enteral nutrition decreased sucrase, maltase, and lactase activities compared with parenteral nutrition [22]. These studies indicate that effects of exogenous GLP-2 on brush border carbohydrase expression and activity are influenced by species, age, and luminal nutrient flow.

We demonstrated that exogenous GLP-2 administration increased jejunal maltase, isomaltase, and glucoamylase activities in the current study [23]. In accordance with increased jejunal α-glucosidase activity, genes associated with brush-border carbohydrate digestion including maltase-glucoamylase, lactase, and trehalase (MGAM, LCT, TREH) were upregulated or tended to be upregulated in the jejunal mucosa of steers treated with exogenous GLP-2 in the current study. It had been suggested that GLP-2-induced suppression of proteolysis and apoptosis may contribute to increased MGAM expression and activity but, more specific mechanisms may be involved [22]. Caudal type homeobox 2 (CDX2) has long been thought to be an important transcription factor contributing to the upregulation of MGAM [22, 83]. However, other authors did not find a strong association between CDX2 and MGAM mRNA expression in their studies [22, 83] and we did not find evidence of this association in our current study, as the mRNA expression of CDX2 was not influenced by exogenous GLP-2 administration. Recent experiments have demonstrated that carbohydrate-responsive genes (SI, MGAM, SLC5A1) are transcriptionally regulated by histone H3 modifications at lysine residues (acetylation/methylation) and increased coactivator binding (KAT2B, CDX2, HNF1, GATA4) in the promoter/enhancer and transcribed regions of those genes in rodents fed high-starch diets [8589]. In accordance with these findings in rodents, we found that genes associated with transcriptional regulation of MGAM, SI, and SLC5A1 (KAT2B, HNF1, GATA4) were upregulated by exogenous GLP-2 administration. Our data suggests that exogenous GLP-2 administration increased MGAM transcription by increasing histone H3 acetylation and increasing interactions of transcriptional coactivators in the promoter and transcribed mRNA regions [90].

Of note, exogenous GLP-2 administration did not increase jejunal SI mRNA expression in the current study. In humans and mice, approximately 80% of the apparent maltase activity is derived from sucrase-isomaltase and the remaining 20% is derived from maltase-glucoamylase [91, 92]. In humans, maltase-glucoamylase is the primary protein exhibiting α-glucosidase activity at low substrate concentrations while sucrase-isomaltase is the primary protein exhibiting α-glucosidase activity at high substrate concentrations [93]. Others have suggested that the opposite response may occur in ruminants, because of the lower Km value for isomaltase compared to maltase in jejunal homogenates from cattle [94]. The upregulation of jejunal MGAM mRNA expression and activity in the current study may support this concept. Although ruminants do not contain sucrase activity [37], it should be acknowledged that small intestinal SI mRNA expression can be transcriptionally regulated by luminal nutrient flows in cattle, as increasing dietary fructose consumption in milk replacer decreased small intestinal SI mRNA expression in calves [95]. The absence of intestinal sucrase activity [96], missing stalk region of sucrase-isomaltase [37], lack of transcriptional response of SI to exogenous GLP-2, and lack of effect of GLP-2 on postruminal starch disappearance [23] leaves many questions about the regulation of α-glucosidase activity in ruminants to be answered. Limitations in the extent of small intestinal starch digestion in ruminants might be similar to humans with phenotype V of congenital sucrase-isomaltase deficiency [37] and may not be completely overcome by exogenous GLP-2 administration.

Exogenous GLP-2 administration has been shown to increase the mRNA expression, protein abundance, and activity of intestinal glucose transporters in rodents [97100] and net portal glucose absorption in parenterally-fed piglets [101, 102]. A prior study had demonstrated that intravenous GLP-2 infusion increased hepatic portal plasma flow but, did not influence net portal glucose absorption in cattle fed a 50:50 alfalfa hay and calf starter diet and fasted for 12 h prior to measurements [103]. To date, the effects of exogenous GLP-2 administration on portal glucose absorption in fed animals has not been evaluated. Although portal glucose absorption was not measured in the current study, we found that exogenous GLP-2 administration increased or tended to increase the expression of genes encoding SGLT1 (SLC5A1) and GLUT2 (SLC2A2) in cattle postruminally infused with starch, suggesting that the capacity for apical and basolateral glucose transport might have increased with exogenous GLP-2 administration. Like MGAM, SLC5A1 mRNA expression is transcriptionally regulated by histone H3 acetylation and co-activator binding (HNF1, GATA4, and CDX2) in both the promoter and transcribed regions [8789, 104107] and we found evidence of these associations in the current study. It has also been demonstrated that activation of the sweet taste receptor complex (T1R2-T1R3) of enteroendocrine L-cells with artificial sweeteners results in increased villus height and crypt depth, GLP-2 secretion, maltase activity, SGLT1 protein abundance, and glucose uptake activity in small intestinal tissue from ruminants [108]. Further investigations using GLP-2 receptor knockout mice have revealed that GLP-2-induced increases in SGLT1 expression and glucose uptake activity are dependent on GLP-2 receptor binding [109]. After GLP-2 binds to its receptor, vasoactive intestinal peptide or pituitary adenylate cyclase-activating polypeptide bind to VIPR1, increasing intracellular cAMP, resulting in increased SGLT1 expression [109]. Notably, we found that exogenous GLP-2 administration increased jejunal VIPR1 mRNA expression in the current study, which may indicate a role of vasoactive intestinal peptide in the downstream regulation of SLC5A1 mRNA expression by GLP-2 in cattle.

Interestingly, exogenous GLP-2 administration has been shown to promote the translocation of GLUT2 from the basolateral membrane to the apical membrane in the rat jejunum to provide a low affinity, high capacity hexose transport mechanism when luminal hexose concentrations exceed transport capacity by the high affinity, low capacity glucose transporter, SGLT1 [100]. The apical GLUT2 translocation hypothesis [110] is highly controversial [111, 112] largely due to the methodology used to assess its validity under normal physiological conditions and because of contamination that occurs during the isolation of brush border membranes. Some evidence for GLUT2 localization on the apical membrane in lactating dairy cows and neonatal calves exists [113115]. In the current study, it is difficult to speculate whether a tendency to increase SLC2A2 mRNA expression resulted in greater GLUT2 protein abundance or activity or the amount of GLUT2 protein localized to each membrane. Differentially expressed genes associated with positive regulation of protein exit from the endoplasmic reticulum and localization to the plasma membrane were significantly enriched with exogenous GLP-2 administration. Sphingolipids contribute to lipid raft formation and intracellular translocation which aid in the trafficking of proteins to the membrane and numerous gene ontology terms related to sphingolipid metabolism and signaling were upregulated with exogenous GLP-2 administration. In the current study, exogenous GLP-2 upregulated the AMPK signaling KEGG pathway, and its activation has been associated with GLUT2 insertion into the apical membrane [116, 117]. In contrast, the expression of genes encoding proteins associated with suspected mechanisms of apical GLUT2 translocation (CACNA1D, TAS1R2, TAS1R3, PRKCB) [117120] were not upregulated by exogenous GLP-2 administration in the current study. Conflicting points of evidence from the current study provide justification for further investigation of the apical GLUT2 translocation hypothesis in cattle.

Tissues of the portal-drained viscera, including the small intestine, utilize large amounts of glucose to maintain metabolic homeostasis [39, 121123]. Increasing luminal starch flow in the small intestine results in increased glucose utilization by the portal-drained viscera [124]. With this knowledge from prior studies and the likely increases in α-glucoside hydrolysis and intestinal glucose uptake with exogenous GLP-2 administration in the current study, it would be expected that exogenous GLP-2 administration would increase glucose utilization by the jejunal mucosa. As expected, exogenous GLP-2 upregulated several genes associated with starch and sucrose metabolism and galactose metabolism KEGG pathways. In addition, exogenous GLP-2 upregulated several gene ontologies related to carbohydrate digestion, absorption, and metabolism including carbohydrate metabolic process, hydrolase activity, glucosidase activity, monosaccharide metabolic process, and hexose metabolic process. Glycolysis/gluconeogenesis (HKDC1, HK2, PGM1, PCK2, PDHA1) and the tricarboxylic acid cycle (ACO1, IDH1, SDHA, PC, DLD, SUCLG1, SUCLG2) KEGG pathways tended to upregulate with exogenous GLP-2 administration. These findings suggest that glucose utilization by the jejunal mucosa increased with exogenous GLP-2 administration. Overall, the current study provides a transcriptional overview of the mechanisms associated with increased jejunal starch assimilation capacity in response to exogenous GLP-2 administration.

Effects of exogenous GLP-2 on transcriptional mechanisms associated with intestinal barrier function

In addition to nutrient digestion and absorption, the gastrointestinal tract acts as a barrier, limiting permeability to potentially toxic or pathogenic organisms and compounds. Epithelial permeability of the intestinal mucosa can occur through transcellular transport associated with solute or water transport through a cell or paracellular transport in between cell-cell junctions [125]. In our study, exogenous GLP-2 administration tended to decrease postruminal fluid flow and increase fecal dry matter content, suggesting increased postruminal fluid absorption [23]. Exogenous GLP-2 administration has been shown to decrease jejunal conductance and ex vivo fluxes of Na+, CrEDTA, and horseradish peroxidase in mice [126]. Cameron and Perdue [127] found that exogenous GLP-2 administration decreased bacteria adhering to and penetrating the intestinal mucosa in stress-induced mice.

Others have found that exogenous GLP-2 or GLP-2 receptor agonist administration increased the mRNA expression of genes encoding tight junction proteins [128131]. In the current study, exogenous GLP-2 administration upregulated adherens junction and tight junction KEGG pathways. In these KEGG pathways, 36% and 25% of the total number of transcripts in each KEGG pathway term were upregulated, respectively. Several gene ontologies including transcripts encoding tight junction proteins were upregulated with exogenous GLP-2 administration including apical junction complex, apical, basolateral, and lateral plasma membrane; bicellular tight, cell-cell, occluding junction; cell, cell-cell junction organization; epithelial cell differentiation and morphogenesis; apical junction, cell, cell-cell, and bicellular tight junction assembly; and anion and chloride transport. Also, connections of adherens and tight junction proteins to the actin cytoskeleton through adaptor proteins may have been improved, as indicated by an upregulated cluster of actin-based cell projections, anchoring junction, actin-based cell projections, actin cytoskeleton, and desmosome gene ontologies. Upregulation of gene ontologies related to phospholipid and sphingolipid metabolism and terms related to apical membrane suggest that exogenous GLP-2 increased the physical barrier (plasma membrane) of individual cells of the jejunal mucosa. These data suggest that exogenous GLP-2 administration enhanced the physical barrier function of the jejunal mucosa by decreasing transcellular (plasma membrane) and paracellular (tight and adherens junction) permeability. Increased postruminal fluid absorption coupled with increased mRNA expression of cell-cell junction proteins suggest that the capacity for absorption was increased and that selective permeability of the jejunal epithelium was improved.

In addition to the physical barrier of the small intestine, intestinal permeability can be affected by exogenous factors, epithelial cell turnover, cytokines, and immune cells [125]. Notably, intestinal barrier dysfunction is associated with increased intestinal inflammation and initiation of immunoregulatory processes [132]. In the current study, exogenous GLP-2 administration downregulated primary immunodeficiency and B cell receptor signaling KEGG pathways. Gene ontologies related to immune function that were downregulated with exogenous GLP-2 administration include B cell activation, regulation of B cell activation, immunoglobulin-mediated immune response, B cell-mediated immunity, immunoglobulin production, regulation of lymphocyte activation, adaptive immune response, and others. These data suggest that the adaptive immune function of the jejunal mucosa was improved with exogenous GLP-2 administration. Collectively, next-generation RNA sequencing findings of the current study suggest that increased jejunal mucosal growth with exogenous GLP-2 administration may be associated with improved intestinal barrier function, which was supported by changes in KEGG pathways and gene ontologies associated with decreased intestinal permeability, inflammation, and immune system activation.

Implications for animal growth and feed efficiency

To our knowledge, this is the first report that has investigated the effects of exogenous GLP-2 on the jejunal mucosal transcriptome using an untargeted RNA-sequencing approach in mammals. Although our primary focus was to describe transcriptional mechanisms affected by exogenous GLP-2 and its relationship with phenotypic responses (growth, starch assimilation, barrier function), this dataset could be useful for identifying novel functions of exogenous GLP-2 and lead to future investigations detailing mechanisms of GLP-2-mediated responses in the small intestinal mucosa. Comparing the findings of the current study with other RNA-sequencing datasets shows that DEGs and KEGG pathways affected by exogenous GLP-2 are similar to DEGs and KEGG pathways identified in cattle with divergent average daily gain [133, 134]. Notably, the mRNA expression of APOB, CLCA4, CUBN, CYP2B6, and SLC9A3 were differentially expressed in the jejunal mucosa in all 3 studies. Cattle with greater average daily gain exhibit 9 upregulated KEGG pathways of the jejunal mucosa [133] that were also found to be upregulated in the current study in response to exogenous GLP-2 administration. These pathways include vitamin digestion and absorption, galactose metabolism, mineral absorption, bile secretion, retinol metabolism, starch and sucrose metabolism, protein digestion and absorption, PPAR signaling pathway, and insulin resistance.

It has been suggested that increasing proliferation of the small intestinal epithelium could result in an increased capacity for nutrient absorption that would typically outweigh increased energy expenditure for maintaining additional tissue [24]. Fitzsimons [135] hypothesized that beef cattle with greater gain:feed have a greater capacity for nutrient absorption from the small intestine. Supporting this hypothesis, others have found that jejunal mucosal density is positively correlated with gain:feed [136]. Also, the number of cells in the duodenal and ileal mucosa is positively correlated with improvements in feed efficiency [137]. Those authors suggested that the benefits of increased metabolic activity in the small intestinal mucosa due to increased cellularity would be greater than the increased energetic cost of maintaining the tissue, leading to improved feed efficiency [137]. Our phenotypic [23] and transcriptomic data from the current study suggest that some of the metabolic pathways and molecular functions of the jejunal mucosa affected by exogenous GLP-2 administration may also underlie physiological differences in feed efficiency in beef cattle. These observations warrant future research to investigate if potential changes to small intestinal function with exogenous GLP-2 administration could lead to improved growth, health, and/or feed efficiency of beef cattle.

Conclusions

Under the conditions of the current experiment, postruminal casein infusion for 7 d downregulated 7 genes of the jejunal mucosa but did not result in enriched KEGG pathways or gene ontologies. These findings might suggest that positive nutritional responses to postruminal casein infusion are mediated through alternative mechanisms or tissues. Results from the current study demonstrated that exogenous GLP-2 administration affected the expression of protein-coding genes that are associated with functions of the jejunal mucosa including mucosal growth, small intestinal starch assimilation, and intestinal barrier function. Differentially enriched pathways that were affected by exogenous GLP-2 administration may be related to phenotypes of improved nutrient assimilation capacity, animal growth, and feed efficiency. Next-generation RNA sequencing generated novel targets for future research to elucidate mechanisms of GLP-2-mediated responses in the jejunal mucosa.

Supporting information

S1 File. Upregulated DEG in response to exogenous GLP-2 administration.

https://doi.org/10.1371/journal.pone.0308983.s001

(XLSX)

S2 File. Downregulated DEG in response to exogenous GLP-2 administration.

https://doi.org/10.1371/journal.pone.0308983.s002

(XLSX)

S3 File. Upregulated DEG in response to postruminal casein infusion.

https://doi.org/10.1371/journal.pone.0308983.s003

(XLSX)

S4 File. Downregulated DEG in response to postruminal casein infusion.

https://doi.org/10.1371/journal.pone.0308983.s004

(XLSX)

S5 File. Upregulated gene ontology in response to exogenous GLP-2 administration.

https://doi.org/10.1371/journal.pone.0308983.s005

(XLSX)

S6 File. Downregulated gene ontology in response to exogenous GLP-2 administration.

https://doi.org/10.1371/journal.pone.0308983.s006

(XLSX)

S7 File. Upregulated KEGG pathway in response to exogenous GLP-2 administration.

https://doi.org/10.1371/journal.pone.0308983.s007

(XLSX)

S8 File. Downregulated KEGG pathway in response to exogenous GLP-2 administration.

https://doi.org/10.1371/journal.pone.0308983.s008

(XLSX)

Acknowledgments

The authors thank Kirk Vanzant and Megan Urig for their assistance with animal feeding, management, and surgery preparation. The authors thank Winston Lin and Suelen Avila for assistance with sample processing.

References

  1. 1. Ørskov ER. Starch digestion and utilization in ruminants. J Anim Sci. 1986;63(5):1624–33. Epub 1986/11/01. pmid:3793653.
  2. 2. Black JL. A theoretical consideration of the effect of preventing rumen fermentation on the efficiency of utilization of dietary energy and protein in lambs. Br J Nutr. 1971;25(1):31–55. Epub 1971/01/01. pmid:5539293.
  3. 3. Owens FN, Zinn RA, Kim YK. Limits to starch digestion in the ruminant small intestine. J Anim Sci. 1986;63(5):1634–48. Epub 1986/11/01. pmid:3539905.
  4. 4. Brake DW, Swanson KC. RUMINANT NUTRITION SYMPOSIUM: Effects of postruminal flows of protein and amino acids on small intestinal starch digestion in beef cattle. J Anim Sci. 2018;96(2):739–50. Epub 2018/02/01. pmid:29385466.
  5. 5. Swanson KC, Matthews JC, Woods CA, Harmon DL. Postruminal administration of partially hydrolyzed starch and casein influences pancreatic α-amylase expression in calves. J Nutr. 2002;132(3):376–81. pmid:11880558
  6. 6. Swanson KC, Matthews JC, Woods CA, Harmon DL. Influence of substrate and/or neurohormonal mimic on in vitro pancreatic enzyme release from calves postruminally infused with partially hydrolyzed starch and/or casein. J Anim Sci. 2003;81(5):1323–31. pmid:12772861
  7. 7. Richards CJ, Swanson KC, Paton SJ, Harmon DL, Huntington GB. Pancreatic exocrine secretion in steers infused postruminally with casein and cornstarch. J Anim Sci. 2003;81(4):1051–6. pmid:12723095
  8. 8. Trotta RJ, Sitorski LG, Acharya S, Brake DW, Swanson KC. Duodenal infusions of starch with casein or glutamic acid influence pancreatic and small intestinal carbohydrase activities in cattle. J Nutr. 2020;150(4):784–91. Epub 2019/12/26. pmid:31875476.
  9. 9. Wang X, Taniguchi K. Activity of pancreatic digestive enzyme in sheep given abomasal infusion of starch and casein. Anim Sci Technol (Jpn). 1998;69:870–4.
  10. 10. Swanson KC, Benson JA, Matthews JC, Harmon DL. Pancreatic exocrine secretion and plasma concentration of some gastrointestinal hormones in response to abomasal infusion of starch hydrolyzate and/or casein. J Anim Sci. 2004;82(6):1781–7. pmid:15217006
  11. 11. Brake DW, Titgemeyer EC, Anderson DE. Duodenal supply of glutamate and casein both improve intestinal starch digestion in cattle but by apparently different mechanisms. J Anim Sci. 2014;92:4057–67. pmid:25057031
  12. 12. Brake DW, Titgemeyer EC, Bailey EA, Anderson DE. Small intestinal digestion of raw cornstarch in cattle consuming a soybean hull-based diet is improved by duodenal casein infusion. J Anim Sci. 2014;92(9):4047–56. pmid:25023803
  13. 13. Acharya S, Petzel EA, Hales KE, Underwood KR, Swanson KC, Bailey EA, et al. Effects of long-term postgastric infusion of casein or glutamic acid on small intestinal starch digestion and energy balance in cattle. J Anim Sci. 2023;101:skac329. pmid:36592759
  14. 14. Blom EJ, Anderson DE, Brake DW. Increases in duodenal glutamic acid supply linearly increase small intestinal starch digestion but not nitrogen balance in cattle. J Anim Sci. 2016;94:5332–40. pmid:28046181
  15. 15. Richards CJ, Branco AF, Bohnert DW, Huntington GB, Macari M, Harmon DL. Intestinal starch disappearance increased in steers abomasally infused with starch and protein. J Anim Sci. 2002;80(12):3361–8. pmid:12542178
  16. 16. Mabjeesh SJ, Guy D, Sklan D. Na+/glucose co-transporter abundance and activity in the small intestine of lambs: enhancement by abomasal infusion of casein. Br J Nutr. 2003;89(5):573–80. pmid:12720577
  17. 17. Mendoza GD, Britton RA. Response of intestinal starch digestion to duodenal infusion of casein. J App Anim Res. 2003;24(2):123–8.
  18. 18. Taniguchi K, Huntington GB, Glenn BP. Net nutrient flux by visceral tissues of beef steers given abomasal and ruminal infusions of casein and starch. J Anim Sci. 1995;73(1):236–49. pmid:7601740
  19. 19. Burrin DG, Stoll B, Guan X, Cui L, Chang X, Holst JJ. Glucagon-like peptide 2 dose-dependently activates intestinal cell survival and proliferation in neonatal piglets. Endocrinology. 2005;146(1):22–32. pmid:15486229
  20. 20. Burrin DG, Stoll B, Guan X, Cui L, Chang X, Hadsell D. GLP-2 rapidly activates divergent intracellular signaling pathways involved in intestinal cell survival and proliferation in neonatal piglets. Am J Physiol Endocrinol Metab. 2007;292(1):E281–E91. pmid:16954336
  21. 21. Burrin DG, Stoll B, Jiang R, Petersen Y, Elnif J, Buddington RK, et al. GLP-2 stimulates intestinal growth in premature TPN-fed pigs by suppressing proteolysis and apoptosis. Am J Physiol Gastrointest Liver Physiol. 2000;279(6):G1249–G56. pmid:11093948
  22. 22. Petersen YM, Elnif J, Schmidt M, Sangild PT. Glucagon-like peptide 2 enhances maltase-glucoamylase and sucrase-isomaltase gene expression and activity in parenterally fed premature neonatal piglets. Pediatr Res. 2002;52(4):498–503. pmid:12357042
  23. 23. Trotta RJ, Swanson KC, Klotz JL, Harmon DL. Postruminal Casein Infusion and Exogenous Glucagon-Like Peptide 2 Administration Differentially Stimulate Pancreatic α-Amylase and Small Intestinal α-Glucosidase Activity in Cattle. J Nutr. 2023;153(10):2854–67. pmid:37573014
  24. 24. Steele MA, Penner GB, Chaucheyras-Durand F, Guan LL. Development and physiology of the rumen and the lower gut: Targets for improving gut health. J Dairy Sci. 2016;99(6):4955–66. Epub 2016/03/14. pmid:26971143.
  25. 25. Steele MA, Croom J, Kahler M, AlZahal O, Hook SE, Plaizier K, et al. Bovine rumen epithelium undergoes rapid structural adaptations during grain-induced subacute ruminal acidosis. Am J Physiol Regul Integr Comp Physiol. 2011;300(6):R1515–R23. pmid:21451145
  26. 26. Steele MA, Schiestel C, AlZahal O, Dionissopoulos L, Laarman AH, Matthews JC, et al. The periparturient period is associated with structural and transcriptomic adaptations of rumen papillae in dairy cattle. J Dairy Sci. 2015;98(4):2583–95. pmid:25682143
  27. 27. Liu JH, Xu TT, Liu YJ, Zhu WY, Mao SY. A high-grain diet causes massive disruption of ruminal epithelial tight junctions in goats. Am J Physiol Regul Integr Comp Physiol. 2013;305(3):R232–R41. pmid:23739344
  28. 28. Zhao K, Chen YH, Penner GB, Oba M, Guan LL. Transcriptome analysis of ruminal epithelia revealed potential regulatory mechanisms involved in host adaptation to gradual high fermentable dietary transition in beef cattle. BMC Genomics. 2017;18(1):976. pmid:29258446
  29. 29. Steele MA, Vandervoort G, AlZahal O, Hook SE, Matthews JC, McBride BW. Rumen epithelial adaptation to high-grain diets involves the coordinated regulation of genes involved in cholesterol homeostasis. Physiol Genomics. 2011;43(6):308–16. pmid:21245418
  30. 30. McLeod KR, Baldwin RLt, Solomon MB, Baumann RG. Influence of ruminal and postruminal carbohydrate infusion on visceral organ mass and adipose tissue accretion in growing beef steers. J Anim Sci. 2007;85(9):2256–70. Epub 2007/04/14. pmid:17431050.
  31. 31. NASEM. Nutrient Requirements of Beef Cattle: Eighth Revised Edition. Washington, DC: The National Academies Press; 2016. 494 p.
  32. 32. Taylor-Edwards CC, Burrin DG, Holst JJ, McLeod KR, Harmon DL. Glucagon-like peptide-2 (GLP-2) increases small intestinal blood flow and mucosal growth in ruminating calves. J Dairy Sci. 2011;94(2):888–98. Epub 2011/01/25. pmid:21257057.
  33. 33. Lopez LC, Frazier ML, Su CJ, Kumar A, Saunders GF. Mammalian pancreatic preproglucagon contains three glucagon-related peptides. Proc Natl Acad Sci USA. 1983;80(18):5485–9. pmid:6577439
  34. 34. Burrin DG, Stoll B, Guan X. Glucagon-like peptide 2 function in domestic animals. Domest Anim Endocrinol. 2003;24(2):103–22. pmid:12586312
  35. 35. Tsai CH, Hill M, Drucker DJ. Biological determinants of intestinotrophic properties of GLP-2 in vivo. Am J Physiol Gastrointest Liver Physiol. 1997;272(3):G662–G8. pmid:9124589
  36. 36. Harmon DL, Yamka RM, Elam NA. Factors affecting intestinal starch digestion in ruminants: A review. Can J Anim Sci. 2004;84(3):309–18.
  37. 37. Trotta RJ, Harmon DL, Matthews JC, Swanson KC. Nutritional and Physiological Constraints Contributing to Limitations in Small Intestinal Starch Digestion and Glucose Absorption in Ruminants. Ruminants. 2022;2(1):1–26.
  38. 38. Owens CE, Zinn RA, Hassen A, Owens FN. Mathematical linkage of total-tract digestion of starch and neutral detergent fiber to their fecal concentrations and the effect of site of starch digestion on extent of digestion and energetic efficiency of cattle. Prof Anim Sci. 2016;32(5):531–49.
  39. 39. Kreikemeier KK, Harmon DL, Brandt RT Jr, Avery TB, Johnson DE. Small intestinal starch digestion in steers: effect of various levels of abomasal glucose, corn starch and corn dextrin infusion on small intestinal disappearance and net glucose absorption. J Anim Sci. 1991;69(1):328–38. pmid:2005026
  40. 40. Krehbiel CR, Matthews JC. Absorption of amino acids and peptides. In: D’Mello JPF, editor. Amino Acids in Animal Nutrition. Wallingford, United Kingdom: CAB International; 2003. p. 41–70.
  41. 41. Binnerts WT, Van’t Klooster AT, Frens AM. Soluble chromium indicator measured by atomic absorption in digestion experiments. Vet Rec. 1968;82:470.
  42. 42. MacLeod NA, Corrigall W, Stirton RA, Ørskov ER. Intragastric infusion of nutrients in cattle. Br J Nutr. 1982;47(3):547–52. Epub 1982/05/01. pmid:6805503.
  43. 43. Ardalan M, Hussein AH, Titgemeyer EC. Effect of Post-Ruminal Casein Infusion on Milk Yield, Milk Composition, and Efficiency of Nitrogen Use in Dairy Cows. Dairy. 2022;3(1):163–73.
  44. 44. Kreikemeier KK, Harmon DL, Peters JP, Gross KL, Armendariz CK, Krehbiel CR. Influence of dietary forage and feed intake on carbohydrase activities and small intestinal morphology of calves. J Anim Sci. 1990;68(9):2916–29. Epub 1990/09/01. pmid:1698758.
  45. 45. Liao SF, Vanzant ES, Boling JA, Matthews JC. Identification and expression pattern of cationic amino acid transporter-1 mRNA in small intestinal epithelia of Angus steers at four production stages. J Anim Sci. 2008;86(3):620–31. pmid:17998425
  46. 46. Bauer ML, Harmon DL, Bohnert DW, Branco AF, Huntington GB. Influence of α-linked glucose on sodium-glucose cotransport activity along the small intestine in cattle. J Anim Sci. 2001;79(7):1917–24. pmid:11465380
  47. 47. Liao SF, Harmon DL, Vanzant ES, McLeod KR, Boling JA, Matthews JC. The small intestinal epithelia of beef steers differentially express sugar transporter messenger ribonucleic acid in response to abomasal versus ruminal infusion of starch hydrolysate. J Anim Sci. 2010;88(1):306–14. Epub 2009/10/13. pmid:19820061.
  48. 48. Connor EE, B RL VI, Capuco AV, Evock-Clover CM, Ellis SE, Sciabica KS. Characterization of glucagon-like peptide 2 pathway member expression in bovine gastrointestinal tract. J Dairy Sci. 2010;93(11):5167–78. pmid:20965332
  49. 49. Chen S, Zhou Y, Chen Y, Gu J. fastp: an ultra-fast all-in-one FASTQ preprocessor. Bioinformatics. 2018;34(17):i884–i90. pmid:30423086
  50. 50. Kim D, Paggi JM, Park C, Bennett C, Salzberg SL. Graph-based genome alignment and genotyping with HISAT2 and HISAT-genotype. Nat Biotechnol. 2019;37(8):907–15. pmid:31375807
  51. 51. Pertea M, Pertea GM, Antonescu CM, Chang TC, Mendell JT, Salzberg SL. StringTie enables improved reconstruction of a transcriptome from RNA-seq reads. Nat Biotechnol. 2015;33(3):290–5. pmid:25690850
  52. 52. Liao Y, Smyth GK, Shi W. featureCounts: an efficient general purpose program for assigning sequence reads to genomic features. Bioinformatics. 2014;30(7):923–30. pmid:24227677
  53. 53. Love MI, Huber W, Anders S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 2014;15(12):1–21. pmid:25516281
  54. 54. Yu G, Wang LG, Han Y, He QY. clusterProfiler: an R package for comparing biological themes among gene clusters. OMICS. 2012;16(5):284–7. pmid:22455463
  55. 55. Taniguchi K, Sunada Y, Obitsu T. Starch digestion in the small intestine of sheep sustained by intragastric infusion without protein supply. Anim Sci Technol (Jpn). 1993;64:892-.
  56. 56. Martineau R, Ouellet DR, Kebreab E, Lapierre H. Casein infusion rate influences feed intake differently depending on metabolizable protein balance in dairy cows: A multilevel meta-analysis. J Dairy Sci. 2016;99(4):2748–61. pmid:26851862
  57. 57. Li CJ, Lin S, Ranilla-García MJ, Baldwin RL. Transcriptomic profiling of duodenal epithelium reveals temporally dynamic impacts of direct duodenal starch-infusion during dry period of dairy cattle. Front Vet Sci. 2019;6:214. pmid:31312641
  58. 58. Haber AL, Biton M, Rogel N, Herbst RH, Shekhar K, Smillie C, et al. A single-cell survey of the small intestinal epithelium. Nature. 2017;551(7680):333–9. pmid:29144463
  59. 59. Swanson KC, Richards CJ, Harmon DL. Influence of abomasal infusion of glucose or partially hydrolyzed starch on pancreatic exocrine secretion in beef steers. J Anim Sci. 2002;80(4):1112–6. pmid:12002319
  60. 60. Harmon DL, Swanson KC. Review: Nutritional regulation of intestinal starch and protein assimilation in ruminants. Animal. 2020;14(S1):s17–s28. Epub 2020/02/07. pmid:32024574.
  61. 61. Gilbert MS, Pantophlet AJ, Berends H, Pluschke AM, van den Borne JJGC, Hendriks WH, et al. Fermentation in the small intestine contributes substantially to intestinal starch disappearance in calves. J Nutr. 2015;145(6):1147–55. pmid:25878206
  62. 62. Shen J, Mu C, Wang H, Huang Z, Yu K, Zoetendal EG, et al. Stimulation of gastric transit function driven by hydrolyzed casein increases small intestinal carbohydrate availability and its microbial metabolism. Mol Nutr Food Res. 2020;64(21):2000250. pmid:32945612
  63. 63. Chen P, Wang RM, Xu BC, Xu FR, Ye YW, Zhang B. Food emulsifier based on the interaction of casein and butyrylated dextrin for improving stability and emulsifying properties. J Dairy Sci. 2023;106(3):1576–85. pmid:36631321
  64. 64. Sun YL, Tang YF, Wang L, Li MZ, Chen ZX. Sodium caseinate medium promotes the in vitro digestion of starch: An insight into the macromolecular crowding effect. Food Chem. 2024;436:137763. pmid:37857200
  65. 65. Dubé PE, Brubaker PL. Frontiers in glucagon-like peptide-2: multiple actions, multiple mediators. Am J Physiol Endocrinol Metab. 2007;293(2):E460–E5. pmid:17652153
  66. 66. Fesler Z, Mitova E, Brubaker PL. GLP-2, EGF, and the intestinal epithelial IGF-1 receptor interactions in the regulation of crypt cell proliferation. Endocrinology. 2020;161(4). pmid:32147716
  67. 67. Yusta B, Holland D, Koehler JA, Maziarz M, Estall JL, Higgins R, et al. ErbB signaling is required for the proliferative actions of GLP-2 in the murine gut. Gastroenterology. 2009;137(3):986–96. pmid:19523469
  68. 68. Dubé PE, Forse CL, Bahrami J, Brubaker PL. The essential role of insulin-like growth factor-1 in the intestinal tropic effects of glucagon-like peptide-2 in mice. Gastroenterology. 2006;131(2):589–605. pmid:16890611
  69. 69. Shi X, Li X, Wang Y, Zhang K, Zhou F, Chan L, et al. Glucagon-like peptide-2-stimulated protein synthesis through the PI 3-kinase-dependent Akt-mTOR signaling pathway. Am J Physiol Endocrinol Metab. 2011;300(3):E554–E63. pmid:21177288
  70. 70. Ma X, Shah YM. The Role of Hippo Signaling in Intestinal Homeostasis. In: Said HM, editor. Physiology of the Gastrointestinal Tract. Sixth ed: Elsevier; 2018. p. 131–40.
  71. 71. Ahmad US, Uttagomol J, Wan H. The regulation of the Hippo pathway by intercellular junction proteins. Life. 2022;12(11):1792. pmid:36362947
  72. 72. Reddy BVVG Irvine KD. Regulation of Hippo signaling by EGFR-MAPK signaling through Ajuba family proteins. Dev Cell. 2013;24(5):459–71. pmid:23484853
  73. 73. Sun G, Irvine KD. Ajuba family proteins link JNK to Hippo signaling. Sci Signal. 2013;6(292): pmid:24023255
  74. 74. Jagannathan R, Schimizzi GV, Zhang K, Loza AJ, Yabuta N, Nojima H, et al. AJUBA LIM proteins limit hippo activity in proliferating cells by sequestering the hippo core kinase complex in the cytosol. Mol Cell Biol. 2016;36(20):2526–42 pmid:27457617
  75. 75. Drucker DJ, Erlich P, Asa SL, Brubaker PL. Induction of intestinal epithelial proliferation by glucagon-like peptide 2. Proc Natl Acad Sci USA. 1996;93(15):7911–6. pmid:8755576
  76. 76. Brubaker PL, Izzo A, Hill M, Drucker DJ. Intestinal function in mice with small bowel growth induced by glucagon-like peptide-2. Am J Physiol Endocrinol Metab. 1997;272(6):E1050–E8. pmid:9227451
  77. 77. Tsai CH, Hill M, Asa SL, Brubaker PL, Drucker DJ. Intestinal growth-promoting properties of glucagon-like peptide-2 in mice. Am J Physiol Endocrinol Metab. 1997;273(1):E77–E84. pmid:9252482
  78. 78. Hartmann B, Harr MB, Jeppesen PB, Wojdemann M, Deacon CF, Mortensen PB, et al. In vivo and in vitro degradation of glucagon-like peptide-2 in humans. J Clin Endocrinol Metab. 2000;85(8):2884–8. pmid:10946898
  79. 79. de Santa Barbara P, Van Den Brink GR, Roberts DJ. Development and differentiation of the intestinal epithelium. Cell Mol Life Sci. 2003;60:1322–32. pmid:12943221
  80. 80. Connock MJ, Elkin A, Pover WFR. The preparation of brush borders from the epithelial cells of the guineapig small intestine by zonal centrifugation. The Histochemical Journal. 1971;3(1):11–22. pmid:4329282
  81. 81. Amiri M, Naim HY. Characterization of mucosal disaccharidases from human intestine. Nutrients. 2017;9(10):1106. pmid:28994704
  82. 82. McConnell RE, Benesh AE, Mao S, Tabb DL, Tyska MJ. Proteomic analysis of the enterocyte brush border. Am J Physiol Gastrointest Liver Physiol. 2011;300(5):G914–G26. pmid:21330445
  83. 83. Kitchen PA, Fitzgerald AJ, Goodlad RA, Barley NF, Ghatei MA, Legon S, et al. Glucagon-like peptide-2 increases sucrase-isomaltase but not caudal-related homeobox protein-2 gene expression. Am J Physiol Gastrointest Liver Physiol. 2000;278(3):G425–G8. pmid:10712262
  84. 84. Petersen YM, Burrin DG, Sangild PT. GLP-2 has differential effects on small intestine growth and function in fetal and neonatal pigs. Am J Physiol Regul Integr Comp Physiol. 2001;281(6):R1986–R93. pmid:11705785
  85. 85. Honma K, Mochizuki K, Goda T. Carbohydrate/fat ratio in the diet alters histone acetylation on the sucrase–isomaltase gene and its expression in mouse small intestine. Biochem Biophys Res Commun. 2007;357(4):1124–9. pmid:17466947
  86. 86. Mochizuki K, Honma K, Shimada M, Goda T. The regulation of jejunal induction of the maltase–glucoamylase gene by a high-starch/low-fat diet in mice. Mol Nutr Food Res. 2010;54(10):1445–51. pmid:20425755
  87. 87. Honma K, Mochizuki K, Goda T. Inductions of histone H3 acetylation at lysine 9 on SGLT1 gene and its expression by feeding mice a high carbohydrate/fat ratio diet. Nutrition. 2009;25(1):40–4. pmid:18952408
  88. 88. Inoue S, Mochizuki K, Goda T. Jejunal induction of SI and SGLT1 genes in rats by high-starch/low-fat diet is associated with histone acetylation and binding of GCN5 on the genes. J Nutr Sci Vitaminol. 2011;57(2):162–9. pmid:21697636
  89. 89. Inoue S, Honma K, Mochizuki K, Goda T. Induction of histone H3K4 methylation at the promoter, enhancer, and transcribed regions of the Si and Sglt1 genes in rat jejunum in response to a high-starch/low-fat diet. Nutrition. 2015;31(2):366–72. pmid:25592016
  90. 90. Goda T, Honma K. Molecular regulations of mucosal maltase expressions. J Pediatr Gastroenterol Nutr. 2018;66:S14–S7. pmid:29762370
  91. 91. Galand G. Brush border membrane sucrase-isomaltase, maltase-glucoamylase and trehalase in mammals. Comparative development, effects of glucocorticoids, molecular mechanisms, and phylogenetic implications. Comp Biochem Physiol B. 1989;94(1):1–11. pmid:2513162
  92. 92. Lin AHM, Nichols BL, Quezada-Calvillo R, Avery SE, Sim L, Rose DR, et al. Unexpected high digestion rate of cooked starch by the Ct-maltase-glucoamylase small intestine mucosal α-glucosidase subunit. PLoS One. 2012;7(5):e35473. pmid:22563462
  93. 93. Quezada-Calvillo R, Robayo-Torres CC, Ao Z, Hamaker BR, Quaroni A, Brayer GD, et al. Luminal substrate “brake” on mucosal maltase-glucoamylase activity regulates total rate of starch digestion to glucose. J Pediatr Gastroenterol Nutr. 2007;45(1):32–43. pmid:17592362
  94. 94. Smith WN, Brake DW, Lindholm-Perry AK, Oliver WT, Freetly HC, Foote AP. Associations of mucosal disaccharidase kinetics and expression in the jejunum of steers with divergent average daily gain. J Anim Sci. 2020;98(9):skaa285. pmid:32860689
  95. 95. Trotta RJ, Ward AK, Swanson KC. Influence of dietary fructose supplementation on visceral organ mass, carbohydrase activity, and mRNA expression of genes involved in small intestinal carbohydrate assimilation in neonatal calves. J Dairy Sci. 2020;103(11):10060–73. Epub 2020/09/15. pmid:32921447.
  96. 96. Dollar AM, Porter JW. Utilization of carbohydrates by the young calf. Nature. 1957;179(4573):1299–300. Epub 1957/06/22. pmid:13440959.
  97. 97. Cheeseman CI, Tsang R. The effect of GIP and glucagon-like peptides on intestinal basolateral membrane hexose transport. Am J Physiol Gastrointest Liver Physiol. 1996;271(3):G477–G82. pmid:8843773
  98. 98. Cheeseman CI. Upregulation of SGLT-1 transport activity in rat jejunum induced by GLP-2 infusion in vivo. Am J Physiol Regul Integr Comp Physiol. 1997;273(6):R1965–R71. pmid:9435650
  99. 99. Cheeseman CI, O’Neill D. Basolateral D-glucose transport activity along the crypt-villus axis in rat jejunum and upregulation induced by gastric inhibitory peptide and glucagon-like peptide-2. Exp Physiol. 1998;83(5):605–16. pmid:9793781
  100. 100. Au A, Gupta A, Schembri P, Cheeseman CI. Rapid insertion of GLUT2 into the rat jejunal brush-border membrane promoted by glucagon-like peptide 2. Biochem J. 2002;367(1):247–54. pmid:12095416
  101. 101. Guan X, Stoll B, Lu X, Tappenden KA, Holst JJ, Hartmann B, et al. GLP-2-mediated up-regulation of intestinal blood flow and glucose uptake is nitric oxide-dependent in TPN-fed piglets. Gastroenterology. 2003;125(1):136–47. pmid:12851879
  102. 102. Cottrell JJ, Stoll B, Buddington RK, Stephens JE, Cui L, Chang X, et al. Glucagon-like peptide-2 protects against TPN-induced intestinal hexose malabsorption in enterally refed piglets. Am J Physiol Gastrointest Liver Physiol. 2006;290(2):G293–G300. pmid:16166344
  103. 103. Taylor-Edwards CC, Burrin DG, Kristensen NB, Holst JJ, McLeod KR, Harmon DL. Glucagon-like peptide-2 (GLP-2) increases net amino acid utilization by the portal-drained viscera of ruminating calves. Animal. 2012;6(12):1985–97. Epub 2012/10/04. pmid:23031436.
  104. 104. Vayro S, Wood IS, Dyer J, Shirazi-Beechey SP. Transcriptional regulation of the ovine intestinal Na+/glucose cotransporter SGLT1 gene: Role of HNF-1 in glucose activation of promoter function. Eur J Biochem. 2001;268(20):5460–70. pmid:11606209
  105. 105. Balakrishnan A, Stearns AT, Rhoads DB, Ashley SW, Tavakkolizadeh A. Defining the transcriptional regulation of the intestinal sodium-glucose cotransporter using RNA-interference mediated gene silencing. Surgery. 2008;144(2):168–73. pmid:18656622
  106. 106. Kekuda R, Saha P, Sundaram U. Role of Sp1 and HNF1 transcription factors in SGLT1 regulation during chronic intestinal inflammation. Am J Physiol Gastrointest Liver Physiol. 2008;294(6):G1354–G61. pmid:18339704
  107. 107. Martín MG, Wang J, Solorzano-Vargas RS, Lam JT, Turk E, Wright EM. Regulation of the human Na+-glucose cotransporter gene, SGLT1, by HNF-1 and Sp1. Am J Physiol Gastrointest Liver Physiol. 2000;278(4):G591–G603. pmid:10762614
  108. 108. Moran AW, Al-Rammahi M, Zhang C, Bravo D, Calsamiglia S, Shirazi-Beechey SP. Sweet taste receptor expression in ruminant intestine and its activation by artificial sweeteners to regulate glucose absorption. J Dairy Sci. 2014;97(8):4955–72. pmid:24881785
  109. 109. Moran AW, Al-Rammahi MA, Batchelor DJ, Bravo DM, Shirazi-Beechey SP. Glucagon-like peptide-2 and the enteric nervous system are components of cell-cell communication pathway regulating intestinal Na +/glucose co-transport. Front Nutr. 2018;5:101. Epub 2018/11/13. pmid:30416998.
  110. 110. Kellett GL, Helliwell PA. The diffusive component of intestinal glucose absorption is mediated by the glucose-induced recruitment of GLUT2 to the brush-border membrane. Biochem J. 2000;350(1):155–62.
  111. 111. Ferraris RP, Choe JY, Patel CR. Intestinal absorption of fructose. Annu Rev Nutr. 2018;38:41–67. pmid:29751733
  112. 112. Röder PV, Geillinger KE, Zietek TS, Thorens B, Koepsell H, Daniel H. The role of SGLT1 and GLUT2 in intestinal glucose transport and sensing. PLoS One. 2014;9(2):e89977. pmid:24587162
  113. 113. Lohrenz A-K, Duske K, Schönhusen U, Losand B, Seyfert HM, Metges CC, et al. Glucose transporters and enzymes related to glucose synthesis in small intestinal mucosa of mid-lactation dairy cows fed 2 levels of starch. J Dairy Sci. 2011;94(9):4546–55. pmid:21854927
  114. 114. Steinhoff-Wagner J, Zitnan R, Schönhusen U, Pfannkuche H, Hudakova M, Metges CC, et al. Diet effects on glucose absorption in the small intestine of neonatal calves: Importance of intestinal mucosal growth, lactase activity, and glucose transporters. J Dairy Sci. 2014;97(10):6358–69. pmid:25108868
  115. 115. Steinhoff-Wagner J, Schönhusen U, Zitnan R, Hudakova M, Pfannkuche H, Hammon HM. Ontogenic changes of villus growth, lactase activity, and intestinal glucose transporters in preterm and term born calves with or without prolonged colostrum feeding. PLoS One. 2015;10(5):e0128154. pmid:26011395
  116. 116. Walker J, Jijon HB, Diaz H, Salehi P, Churchill T, Madsen KL. 5-aminoimidazole-4-carboxamide riboside (AICAR) enhances GLUT2-dependent jejunal glucose transport: a possible role for AMPK. Biochem J. 2005;385(2):485–91. pmid:15367103
  117. 117. Kellett GL, Brot-Laroche E. Apical GLUT2: a major pathway of intestinal sugar absorption. Diabetes. 2005;54(10):3056–62. pmid:16186415
  118. 118. Mace OJ, Lister N, Morgan E, Shepherd E, Affleck J, Helliwell P, et al. An energy supply network of nutrient absorption coordinated by calcium and T1R taste receptors in rat small intestine. J Physiol. 2009;587(1):195–210. pmid:19001049
  119. 119. Morgan EL, Mace OJ, Affleck J, Kellett GL. Apical GLUT2 and Cav1. 3: regulation of rat intestinal glucose and calcium absorption. J Physiol. 2007;580(2):593–604. pmid:17272350
  120. 120. Leturque A, Brot-Laroche E, Le Gall M. GLUT2 mutations, translocation, and receptor function in diet sugar managing. Am J Physiol Endocrinol Metab. 2009;296(5):E985–E92. pmid:19223655
  121. 121. Reynolds CK, Huntington GB. Partition of portal-drained visceral net flux in beef steers: 1. Blood flow and net flux of oxygen, glucose and nitrogenous compounds across stomach and post-stomach tissues. Br J Nutr. 1988;60(3):539–51. pmid:3219322
  122. 122. Balcells J, Seal CJ, Parker DS. Effect of intravenous glucose infusion on metabolism of portal-drained viscera in sheep fed a cereal/straw-based diet. J Anim Sci. 1995;73(7):2146–55. pmid:7592103
  123. 123. Kreikemeier KK, Harmon DL. Abomasal glucose, maize starch and maize dextrin infusions in cattle: Small-intestinal disappearance, net portal glucose flux and ileal oligosaccharide flow. Br J Nutr. 1995;73(5):763–72. pmid:7626594
  124. 124. Harmon DL, Richards CJ, Swanson KC, Howell JA, Matthews JC, True AD, et al., editors. Influence of ruminal or postruminal starch on visceral glucose metabolism in steers. 15th Symposium on Energy Metabolism in Animals; 2001 11–16 September; Snekkersten, Denmark: Wageningen Press.
  125. 125. Groschwitz KR, Hogan SP. Intestinal barrier function: molecular regulation and disease pathogenesis. J Allergy Clin Immunol. 2009;124(1):3–20. pmid:19560575
  126. 126. Benjamin MA, McKay DM, Yang PC, Cameron H, Perdue MH. Glucagon-like peptide-2 enhances intestinal epithelial barrier function of both transcellular and paracellular pathways in the mouse. Gut. 2000;47(1):112–9. pmid:10861272
  127. 127. Cameron HL, Perdue MH. Stress impairs murine intestinal barrier function: improvement by glucagon-like peptide-2. J Pharmacol Exp Ther. 2005;314(1):214–20. pmid:15798004
  128. 128. Walker MP, Evock-Clover CM, Elsasser TH, Connor EE. Short communication: Glucagon-like peptide-2 and coccidiosis alter tight junction gene expression in the gastrointestinal tract of dairy calves. J Dairy Sci. 2015;98(5):3432–7. Epub 2015/03/03. pmid:25726101.
  129. 129. Reiner J, Thiery J, Held J, Berlin P, Skarbaliene J, Vollmar B, et al. The dual GLP-1 and GLP-2 receptor agonist dapiglutide promotes barrier function in murine short bowel. Ann N Y Acad Sci. 2022;1514(1):132–41. pmid:35580981
  130. 130. Moran GW, O’Neill C, McLaughlin JT. GLP-2 enhances barrier formation and attenuates TNFα-induced changes in a Caco-2 cell model of the intestinal barrier. Regul Pept. 2012;178(1–3):95–101. pmid:22809889
  131. 131. Cani PD, Possemiers S, Van de Wiele T, Guiot Y, Everard A, Rottier O, et al. Changes in gut microbiota control inflammation in obese mice through a mechanism involving GLP-2-driven improvement of gut permeability. Gut. 2009;58(8):1091–103. pmid:19240062
  132. 132. Turner JR. Intestinal mucosal barrier function in health and disease. Nat Rev Immunol. 2009;9(11):799–809. pmid:19855405
  133. 133. Foote AP, Keel BN, Zarek CM, Lindholm-Perry AK. Beef steers with average dry matter intake and divergent average daily gain have altered gene expression in the jejunum. J Anim Sci. 2017;95(10):4430–9. pmid:29108031
  134. 134. Lindholm-Perry AK, Butler AR, Kern RJ, Hill R, Kuehn LA, Wells JE, et al. Differential gene expression in the duodenum, jejunum and ileum among crossbred beef steers with divergent gain and feed intake phenotypes. Anim Genet. 2016;47(4):408–27. Epub 2016/05/27. pmid:27226174.
  135. 135. Fitzsimons C, McGee M, Keogh K, Waters SM, Kenny DA. Molecular physiology of feed efficiency in beef cattle. In: Scanes CG, Hill RA, editors. Biology of Domestic Animals. 1 ed. Boca Raton, FL: CRC Press; 2017. p. 122–65.
  136. 136. Meyer AM, Hess BW, Paisley SI, Du M, Caton JS. Small intestinal growth measures are correlated with feed efficiency in market weight cattle, despite minimal effects of maternal nutrition during early to midgestation. J Anim Sci. 2014;92(9):3855–67. pmid:25057033
  137. 137. Montanholi Y, Fontoura A, Swanson K, Coomber B, Yamashiro S, Miller S. Small intestine histomorphometry of beef cattle with divergent feed efficiency. Acta Vet Scand. 2013;55:1–6. pmid:23379622