Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Received: 27 July 2018 | Revised: 29 September 2018 | Accepted: 11 October 2018

DOI: 10.1111/jfbc.12715

FULL ARTICLE

Study on the ability of partially hydrolyzed guar gum to


modulate the gut microbiota and relieve constipation

Xiaodan Fu1* | Rong Li2* | Tan Zhang1 | Meng Li1 | Haijin Mou1

1
College of Food Science and
Engineering, Ocean University of China, Abstract
Qingdao, China The aim of this study was to evaluate the effects of high‐ (HHGG, Mw 10,000–
2
Qingdao Women and Children Hospital,
30,000 Da) and medium‐molecular‐weight (MHGG, Mw 2,000–10,000 Da) partially
Qingdao, China
hydrolyzed guar gum (PHGG) on modulation of gut microbiota and relief of constipa‐
Correspondence
tion in mice. Mice were treated with galacto‐oligosaccharide (GOS) and xylo‐oligo‐
Haijin Mou, College of Food Science and
Engineering, Ocean University of China, saccharide (XOS) at a dose of 1 g/kg bw as positive controls. Low‐ and high‐dose
No. 5 Yushan Road, Qingdao 266003,
HHGG and MHGG groups received 250 mg or 1 g/kg bw, respectively. Treatment
Shandong, China.
Email: mousun@ouc.edu.cn was administered intragastrically for 15 days, and constipation model was induced by
loperamide lavage at d 16. PHGG could increase fecal moisture and small intestinal
Funding information
Key Research and Development Project transit and shortened the time to first black stool defecation after constipation. The
Foundation of Shandong Province, Grant/
highest short‐chain fatty acid production was observed in the high‐dose MHGG
Award Number: 2017YYSP003; Shandong
Natural Science Fund, Grant/Award group. Additionally, PHGG, GOS, and XOS predominantly promoted the accumula‐
Number: ZR2017MD006; Research and
tion of Bacteroidetes and inhibited the growth of Desulfovibrio. This study suggested
Development, Grant/Award Number:
2017YYSP003 and ZR2017MD006; that MHGG treatment could elicit constipation relief in mice.
Shandong Province
Practical applications
In this study, partially hydrolyzed guar gum (PHGG) produced by mannanase hydroly‐
sis was applied for the relieving constipation in mice. The medium‐molecular‐weight
product (Mw 2,000–10,000 Da) could elicit constipation relief and modulate the gut
microbiota in mice, which shows the potential to act as dietary fiber for constipation
treatment.

KEYWORDS
16S rRNA analysis, constipation relief, gut microbiota, partially hydrolyzed guar gum (PHGG),
short‐chain fatty acids

1 | I NTRO D U C TI O N commonly regarded as a metabolic “organ,” which affects the host’s


metabolism, physiology, nutrition, and immune function (Tropini,
Abundant microbial communities inhabit the mammalian gastroin‐ Earle, Huang, & Sonnenburg, 2017). Indeed, the gut microbiota has
testinal tract, especially the rumen and large intestine, where bacte‐ been reported to be closely related to pathological intestinal condi‐
rial density can exceed 10 11
per gram (Flint, Bayer, Rincon, Lamed, & tions such as malnutrition (Kau, Ahern, Griffin, Goodman, & Gordon,
White, 2008). Their symbiotic and mutualistic relationship with the 2011), obesity (Zhang et al., 2009), and chronic inflammatory dis‐
host establishes a complex and dynamic ecosystem, which plays a eases such as inflammatory bowel disease (IBD) (Frank et al.., 2007).
central role in host health (Zhao et al., 2013). This vast community is Chronic constipation is a common functional gastrointestinal dis‐
order characterized by prolonged gastrointestinal emptying times,
*
These authors contributed equally to this work

J Food Biochem. 2018;e12715. wileyonlinelibrary.com/journal/jfbc © 2018 Wiley Periodicals, Inc. | 1 of 14


https://doi.org/10.1111/jfbc.12715
2 of 14 | FU et al.

difficult or low frequency defecation, and dry, hard feces (Zhao & of intestinal function as a result of PHGG treatment was assessed
Yu, 2016). In recent years, with changes in diet structure and the based on SCFA production, fecal moisture, and variation of the fecal
influence of psychological and social factors, a clear upward trend microbiota after intragastric administration. Additionally, the effect
has been observed for the incidence of constipation, which seri‐ of PHGG on the promotion of gut transit was evaluated in mice after
ously affects patients of all ages and gender and affects their health inducing constipation by loperamide administration.
and quality of life (Bharucha, 2007; Khalif, Quigley, Konovitch, &
Maximova, 2005). The etiology and pathophysiology of chronic con‐
2 | M ATE R I A L S A N D M E TH O DS
stipation is complicated and remains poorly understood, however, a
growing number of studies have revealed the potential constipation
2.1 | Materials
relief effect of modulation of the intestinal microbiota (Li, Lu, & Yang,
2013). β‐mannanase (50,000 U/g) was provided by Genyuan
Furthermore, the treatment of chronic constipation remains a Biotechnology Co., Ltd. (Qingdao, Shandong. China). Food grade
challenge, and possible microbiota‐based therapy has attracted con‐ guar gum (mannose/galactose (M/G) ratio = 1.68, galactomannan
siderable interest recently. Dietary fiber has been recommended as ≥90%, w/w) was provided by Q ingzhou Ronmer Biology Technology
a basic treatment for chronic constipation, which can promote the Co., Ltd. (Qingzhou, Shandong, China). XOS, a commercial prod‐
excretion of intestinal mucin (Derrien et al., 2010), increase stool uct, are mixtures of oligosaccharides formed by 2–7 xylose resi‐
bulk and promote colonic transit (López Román et al., 2008), stimu‐ dues linked through β‐(1→4)‐linkages (Mw 300–1,000 Da), with
late the growth of intestinal microbiota, and promote the accumula‐ a purity of over 95%, was purchased from Shandong Longlive
tion of bacterial fermentation products (Jennings, Davies, Costarelli, Bio‐Technology Corp. Ltd (Shandong, China). GOS, a commercial
& Dettmar, 2009). Moreover, specific non‐digestible carbohydrates preparation of lactose, comprised of DP 2–8 galactose units linked
are increasingly used as prebiotics to manipulate the gut microbiota through β‐(1→4)‐linkages (Mw 300–1,500 Da), with a purity over
which may have beneficial effects on host health (Kruse, Kleessen, & 95%, was provided by New Francisco Biotechnology Corp. Ltd
Blaut, 1999). Among them, galacto‐oligosaccharides (GOS) and xylo‐ (Guangdong, China). The drug loperamide hydrochloride (2 mg per
oligosaccharide (XOS) have been reported to have potential benefi‐ capsule) was purchased from Xian Janssen Pharmaceutical Ltd.
cial effects on constipation and promote intestinal transit (Li et al., (Xi’an, Shanxi, China). Gum acacia and charcoal were purchased
2013; Li, Zong, Qi, & Liu, 2011). from Sigma (USA). High‐purity SCFA standards, including acetate,
Guar gum is a seed galactomannan obtained from the endosperm propionate, butyrate, iso‐butyrate, valerate, iso‐valerate, and 4‐
portion of the guar plant Cyamopsis tetragonolobus, a member of the methylvaleric acid were purchased from Sigma Aldrich Chemical.
Leguminosae family. It mainly consists of high molecular weight ga‐ All other chemicals and reagents used in this study were of ana‐
lactomannans with β‐D‐1,4‐linked mannose units as linear chains lytical grade.
and α‐D‐1,6‐linked galactose units as side chains (Mudgil, Barak, &
Khatkar, 2014). Guar gum is commonly used as food additive because
2.2 | Preparation of PHGG
of its thickening and viscosity control properties (Morris, 2010).
However, an aqueous solution of guar gum is extremely viscous even PHGG with different molecular weights were obtained using ul‐
at low concentrations, which limits its incorporation at higher levels trafiltration after enzymatic hydrolysis by commercial acidic β‐
into food products without affecting the food’s nutritional or tex‐ mannanase (Genyuan Biotechnology Co., Ltd., Qingdao, China).
tural properties (Srivastava & Kapoor, 2005). Hence, partially hydro‐ Briefly, a solution of guar gum (100 g/L) at pH = 2.5 was added
lyzed guar gum (PHGG) that has a smaller molecular weight and lower to acidic β‐mannanase (0.4 g/L) and incubated at 55°C for 12 hr.
viscosity with resemblance to the basic chemical structure of native The reaction was terminated by heating the solution at 90°C for
guar gum, has potential applications in food nutrition as a fiber sup‐ 10 min. The hydrolysates were centrifuged at 10,000 × g for
plement (Mudgil et al., 2014; Slavin & Greenberg, 2003). 15 min to separate out the non‐hydrolyzed parts. Spiral wound
PHGG is generally recognized as safe and has been reported to PES membranes with MWCO 30,000 Da (PW8040F30, General
reduce symptoms of IBS (Niv et al., 2016), improve glycemic con‐ Electric Company, the USA) and PVDF hollow fiber membranes
trol (Goulder, Alberti, & Jenkins, 1978), and relieve diarrhea (Spapen with MWCO 10,000 Da (Zhongshan Lonkee Memerane Industry
et al., 2001). Moreover, PHGG has also been studied in constipated Co. Ltd, Guangdong, China) were used to intercept the larger frag‐
subjects, and has been shown to reduce the fecal pH and increase ments with molecular weights between 10,000–30,000 Da from
defecation frequency and fecal moisture (Hidehisa et al., 1994). the supernatants. Then, the filtrate passed through spiral wound
However, these studies focused on a PHGG mixture with an average PES membranes with MWCO 2,000 Da (Microdyn‐Nadir GmbH,
molecular weight of 20,000 Da. The beneficial effects of lower mo‐ Germany) to intercept the fragments with molecular weights be‐
lecular weight (Mw 2,000–10,000 Da) guar gum on the modulation tween 2,000–10,000 Da. Subsequently, the hydrolyzed guar gum
of constipation have not been reported. of medium‐molecular‐weight (MHGG, Mw 2,000–10,000 Da) and
Therefore, we prepared PHGG with different molecular weights high‐molecular‐weight (HHGG, Mw 10,000–30,000 Da) were ob‐
using ultrafiltration after enzymatic hydrolysis. The enhancement tained after freeze‐drying.
FU et al. | 3 of 14

2.3 | Monosaccharide composition of PHGG normal control, were treated with loperamide via gavage at day 16 to
induce constipation, and subsequently, the GI transit (the rate of car‐
PHGG was hydrolyzed using 2 mol/L trifluoroacetic acid (TFA) at
bon powder propulsion through the small intestine), and the time to
110°C for 4 hr. The sample derivatization and further HPLC analy‐
the first black stool defecation were measured. The protocol of ani‐
sis procedures applied in this study were determined according to
mal use was reviewed and approved by of the ethical committee of
method reported by Bai et al. (2015). The derived samples were ana‐
experimental animal care at Ocean University of China (Certificate
lyzed with an Agilent 1200 Infinity HPLC system equipped with an
No. SYXK20120014) according to the guidelines.
Agilent Eclipse XDB‐C18 column (5 µm, 250 mm × 4.6 mm i.d.). The
mobile phase consisted of 83% A and 17% B. Eluent A was a phos‐
phate buffer containing 0.05 mol/L KH2PO 4 (pH = 6.9), and eluent B 2.6 | Induction of constipation in mice and
was an acetonitrile of chromatographic grade. The temperature of measurement of gut transit
the column was 25°C and the flow rate was 0.8 ml/min. Calibration
Loperamide was dissolved in saline, and constipation was induced in
was obtained with a series of standard sugar solutions, including
the mice by oral administration at a dose of 5 mg/kg bw, using a vol‐
D‐mannose, D‐glucose, D‐xylose (Sinopharm Chemical Reagent
ume of 10 ml/kg of mouse body weight (Moezi, Pirsalami, & Inaloo,
Co., Ltd, Shanghai, China), and L‐arabinose and D‐galactose (Beijing
2015). Mice were fasted 18 hr before experiments in order to make
Solarbio Science & Technology Co., Ltd., China).
the intestines free of stool, freely drinking in cages with a floor made
of metal grids. On the test day, 30 min after the administration of
2.4 | Animals and experimental diets loperamide, all the mice received 0.25 ml of an intragastric suspen‐
sion of 10% vegetable charcoal in 5% gum acacia (Sigma Chemical
Four‐week‐old male Kunming mice (SPF strain) were purchased from
Co., St. Louis, MO) which contains requisite doses of dissolved test
Jinan Peng Yue experimental animal breeding Co., Ltd. (Jinan, China,
samples. Intestinal transit was measured with the charcoal meal
SCXK: 20140007). They were housed in a standard polycarbonate
method described by Puig and Pol (Pol, Planas, & Puig, 1995; Puig,
cages in an air‐conditioned room (temperature 21°C–22°C, relative
Pol, & Warner, 1996). Half of the mice in each group (n = 6) were
humidity 55%–65%) under a 12 hr (06:00–18:00) light cycle with
sacrificed 25 min later by cervical dislocation, and the stomach and
free access to food and water. All the mice were fed normal chow
small intestines were totally removed. The mesentery was com‐
which was purchased from Sebiona Biological Technology Co. Ltd.
pletely separated to allow the intestine be fully straightened while
(Guangzhou, China). According to the manufacturer, the main nutri‐
avoiding stretching. Subsequently, the total length of the small in‐
tional composition of the feed was recorded as the following: crude
testine (INTS) from the pyloric sphincter to the ileocecal junction
protein (14.5%), crude lipid (4.00%), crude fiber (4.5%), crude ash
and the distance advanced by the charcoal (CHARC) from the begin‐
(4.7%), calcium (0.72%), and phosphorous (0.60%).
ning of the pyloric sphincter to the forefront of the charcoal bulk
were measured, and the intestinal transit (% intestinal transit) was
calculated as the percentage of the length of the CHARC relative to
2.5 | Experimental design
the INTS.
After the acclimatization for 7 d, 96 mice were weighed and divided ( )
% transit = CHARC∕INTS × 100%
into 8 groups of 12 with approximately equal total body weight
(300.14–308.01 g). The normal and constipation control groups
were treated only with saline via gavage once per day for 15 con‐ The time to the first black stool defecation was measured follow‐
secutive days. The test groups were divided into four groups, in‐ ing the methods of Nagakura, Naitoh, Kamato, Yamano, and Miyata
cluding HHGG and MHGG, each of sample covering two doses. The (1996). After the administration of vegetable charcoal, mice were re‐
high‐ and low‐dose HHGG (HHGG‐H/‐L) and MHGG (MHGG‐H/‐L) turned to their cages. A white sheet was placed under the cages to
groups received 250 mg/kg body weight (bw) or 1 g/kg bw once per allow the researchers to distinguish black stools from normal ones.
day for 15 consecutive days, respectively. An additional two positive The observation was performed for 6 hr after the administration of
control groups were treated with 1 g/kg bw XOS and GOS via gav‐ the marker.
age once per day for 15 consecutive days. All materials were freshly
prepared in physiologic saline solution to such concentrations that
2.7 | Fecal and colonic SCFA measurements
requisite doses were administered in a volume of 10 ml/kg of the
murine body weight. Feces samples were collected at 0, 5, 10, and For fecal SCFA analysis, 0.2–0.5 g of frozen samples was thawed
15 d, and the fecal water content was measured immediately after and homogenized in 1 ml water and kept under ice for 20–30 min
excretion. Partial fresh fecal samples, collected at 15 d, were sub‐ with occasional mixing with a vortex. After that, the clear superna‐
sequently submerged in liquid nitrogen, then immediately froze at tant was obtained after a centrifugation (2000 × g, 10 min at 4°C).
−80°C until further SCFA and gut microbiota analysis. All of the Acetate, propionate, iso‐butyrate, butyrate, iso‐valerate, valer‐
animals were fasted overnight for approximately 18 hr before induc‐ ate, and 4‐methylvaleric acid were used as standards and 2‐eth‐
ing constipation (water was not restricted). All groups, except the ylbutyric acid was used as an internal standard. The analysis was
4 of 14 | FU et al.

carried out using an Agilent 6890 N gas chromatograph equipped standard of sequence number corresponding to the sample with the
with a flame ionization detector (Agilent Technologies, USA) and a least sequences. The changes in the relative abundance of the bac‐
polar HP‐FFAP capillary column (0.25 µm × 320 µm × 30 m). The terial phyla and genera were displayed using heat map, which was
initial oven temperature was 80°C, maintained for 0.5 min, and constructed using the gplot package of R software. Cluster tree was
then raised to 150°C at 8°C/min and held for 1.0 min, then in‐ generated with the ape package of R using the average method to
creased to 200°C at 20°C/min, and finally held at this temperature show the differences among samples. A Spearman correlation analy‐
for 4 min. The temperature of the FID and the injection port was sis of intestinal transit and fecal indices was constructed using SPSS
250 and 200°C, respectively. Nitrogen was used as the carrier gas 17.0 software. The Spearman correlation coefficient of the top 40
at a flow of 10 ml/min, and the flow rates of hydrogen and air were abundant bacterial genera and the environmental factors including
30 and 300 ml/min, respectively. A supernatant sample of 1 µl was the intestinal transit rate, the fecal water content, and the total con‐
inserted in the glass liner of the splitless injection for GC analysis. tent of SCFAs was also calculated using the pheatmap package of R,
Data handling was carried out with a HP Chem Station Plus soft‐ which was displayed by heat map.
ware (A.09.xx, Agilent).

2.9 | Statistical analysis


2.8 | DNA extraction, MiSeq sequencing, and
The results are presented as the mean ± the standard deviation
data processing
(X ± SD). Data from the control and other experimental groups were
Total genome DNA from fecal samples (d 15) was extracted using analyzed with the SPSS 17.0 program using the Student’s t test.
the CTAB method (Rouhibakhsh, Priya, Periasamy, Haq, & Malathi, Differences among groups were evaluated for significance by the
2008). DNA concentration and purity were monitored on 1% aga‐ one‐way ANOVA (Tukey post hoc test). Correlations were consid‐
rose gels. The amplification sequencing of the V4–V5 region of ered significant when p < 0.05.
the bacterial 16 S rRNA genes was performed using universal
primers (515F 5′ GTGCCAGCMGCCGCGGTAA‐3′ and 907R 5′‐
3 | R E S U LT S A N D D I S CU S S I O N
CCGTCAATTCCTTTGAGTTT‐3′). PCR reactions were carried out
with Phusion® High‐Fidelity PCR Master Mix (New England Biolabs).
3.1 | Monosaccharide composition of guar gum and
PCR products were mixed with same volume of 1X loading buffer
PHGG
(contained SYB green), and then electrophoresed on a 2% agarose
gel. The main bright strip between 400–450 bp was chosen for fur‐ During preparation of PHGG, four fragments were obtained by
ther experiments. The PCR amplification products were purified centrifugation and ultrafiltration, and the Man/GOS (M/G) ratio
using the Qiagen Gel Extraction Kit (Qiagen, Germany) and quan‐ was analyzed (Table 1). Lower M/G ratios were observed in HHGG
tified using QuantiFluor™‐ST (Promega, the USA) according to the and precipitates compared to that of the guar gum, and the HHGG
manufacturer’s protocol. Sequencing libraries were conducted by/ was the lowest. However, higher M/G ratios were observed in the
Nevogene Technology Co., Ltd. (Beijing, China) using TruSeq® DNA fragments (MHGG) with lower molecular weights. The distribution
PCR‐Free Sample Preparation Kit (Illumina, USA) and sequenced of galactose substituents in guar gum can be classified into three
on an Illumina HiSeq 2,500 platform. Operational taxonomic units types: ordered, random, and blockwise (Daas, Schols, & de Jongh,
(OTUs) were clustered with minimum identity threshold of 97% using 2000). Galactoses are usually observed to assemble compactly and
UPARSE software (version7.0.100, https://drive5.com/uparse/) regularly at the ordered area of guar gum, which hinders enzymoly‐
(Edgar, 2013). The taxonomy of each 16S rRNA gene sequence was sis (Srivastava & Kapoor, 2014), resulting the unhydrolyzed fractions
analyzed using RDP Classifier (Version 2.2, https://sourceforge. with high Mw value and low M/G ratio (HHGG). The random‐ and
net/projects/rdp-classifier/) against GreenGene Database (https:// non‐substituted mannan backbone area contain abundant enzyme‐
greengenes.lbl.gov/cgi-bin/nph-index.cgi) (DeSantis et al., 2006; binding sites (Daas et al., 2000; Kurakake, Sumida, Masuda, Oonishi,
Wang, Garrity, Tiedje, & Cole, 2007). Follow‐up analyses were per‐ & Komaki, 2006), and mainly contributed to MHGG with a higher
formed after OTU abundance information was normalized using a M/G ratio and lower Mw value.

TA B L E 1 Content of mannose and


Sample Mannose (w/w%) Galactose (w/w%) Man‐Gala ratio
galactose in guar gum and its hydrolysate
Guar gum 51.7 35.2 1.47
Precipitates after 36.1 32.5 1.11
enzymolysis
HHGGb 43.0 43.5 0.99
c
MHGG 55.5 30.0 1.85
a
Man (Mannose), Gal (Galactose); bHHGG (high‐molecular‐weight partially hydrolyzed guar gum);
c
MHGG (medium‐molecular‐weight partially hydrolyzed guar gum).
FU et al. | 5 of 14

3.2 | Determination of fecal water content to that of the normal group after constipation was induced by lop‐
eramide administration. Compared with the constipation control
Constipation is a common gastrointestinal disorder characterized
group, the fecal water content increased significantly in all the pos‐
by symptoms such as hard stool, straining, and infrequent defeca‐
itive‐ and PHGG‐treated groups (Figure 1b, p < 0.05). Furthermore,
tion (Moezi et al., 2015), which can be assessed principally by a
the fecal water content of the high‐dose MHGG group was the
decrease in the water content of the fecal pellet. Hard stool is a
highest of all the treatment groups and the value was close to that
typical symptom, and stool softening is a physician’s first step in
of the normal group, followed by that of the high‐dose HHGG
the management of chronic constipation by increasing stool water
group. It was reported that gel‐forming dietary fibers, such as guar
content (Mcrorie et al., 1998). During 15 d of administration, fecal
gum and pectin, possess good water‐holding capacity and lead to
water content was increased in all groups with administration time,
an increase in fecal bulking action (Stephen & Cummings, 1979).
however, no significant difference was seen between the experi‐
PHGG can become a gelatinous and viscous substance after ab‐
mental groups (Figure 1a). As the result show in Figure 1b, fecal
sorbing water and thus may lead to an increase in fecal moisture.
water content decreased in all of the treatment groups compared

F I G U R E 1 Fecal water content


change in all groups (a); Post‐induction
of constipation by loperamide (b). Mice
were treated with GOS and XOS at a
dose of 1 g/kg bw. The low‐ and high‐
dose HHGG (HHGG‐L/‐H) and MHGG
(MHGG‐L/‐H) groups received 250 mg/
kg bw or 1 g/kg bw, respectively. Water
content of defecation was calculated as
the difference between the wet and dry
weights of collected pellets. Values are
mean ± SD, n = 6. **p < 0.01 compared
to control group (Student’s t test).
*
p < 0.05 compared to control group
(Student’s t test). Abbreviation: XOS
(xylo‐oligosaccharide), GOS (galacto‐
oligosaccharide), HHGG (high‐molecular‐
weight partially hydrolyzed guar gum),
MHGG (medium‐molecular‐weight
partially hydrolyzed guar gum)
6 of 14 | FU et al.

time to the first black stool defecation was observed in the normal
3.3 | Intestinal transition after loperamide‐induced
group with a significant difference (p < 0.01) compared to that of the
constipation
control. High‐dose MHGG and HHGG showed significant improve‐
Difficult or infrequent passage of stool is one of the symptoms of ment (p < 0.01) in the time to first black defecation compared to that
chronic constipation (Bharucha, 2007). Recently, some research has of the control, and the best time was observed for the high‐dose
investigated the effects of possible microbiota‐based therapy on MHGG group, which was the closest to that of the normal group.
chronic constipation, such as dietary fiber, prebiotics, and probiotics However, no significant difference between the normal, low‐dose
(Zhao & Yu, 2016). Dietary fiber can be fermented by intestinal mi‐ MHGG, and HHGG groups was observed. Furthermore, a significant
crobiota to produce hydrogen, methane, and carbon dioxide, which improvement (p < 0.05) was observed in the GOS and XOS groups.
can increase stool bulk and promote colonic transit (López Román et These results indicated that PHGG could shorten the time to defeca‐
al., 2008). Additionally, prebiotics such as GOS have been reported tion and this effect was concentration‐dependent. The water‐hold‐
to have potential effects on the promotion of intestinal peristalsis ing capacity is one of the important mechanisms by which dietary
and relief of constipation (Li et al., 2013). In this study, we evalu‐ fibers increase fecal bulk and modulate gastrointestinal transit time
ated the constipation relieving effects of two fragments of PHGG (Stephen & Cummings, 1980). Thus, the increase in fecal moisture
with different molecular weights compared to that of XOS and by PHGG may result in the formation of softer feces, leading to an
GOS in mice. As the results show in Figure 2, the intestinal tran‐ increase in the frequency of defecation.
sit rate was significantly (p < 0.01) reduced in the loperamide‐only
group (35.40% ± 3.07%) compared to that of the normal group
3.4 | Changes of SCFAs in feces of mice
(78.94% ± 5.40%), indicating the establishment of an acceptable
model of constipation. Higher intestinal transit rates were observed Undigested carbohydrates can be fermented by gut microbiota
for all of the PHGG‐treated groups, though there was no statisti‐ under anaerobic conditions in the large intestine, mainly in the form
cal difference between the low‐dose PHGG groups and the control of SCFA production (Flint, Scott, Louis, & Duncan, 2012). SCFAs have
group. However, great enhancement was observed for the high‐dose a complex effect on the host, and can be used as energy sources
PHGG groups compared to that of the control (p < 0.01). Importantly, and important signaling molecules in physiological metabolism of
high‐dose MHGG showed the highest effects on promoting intesti‐ the host (Pomare, Branch, & Cummings, 1985). SCFAs have been re‐
nal function with the highest intestinal transit rate (73.79% ± 5.17%), ported to stimulate ileal propulsive contractions and colonic smooth
followed by the high‐dose HHGG group (60.73% ± 8.34%). GOS muscle contractility, which may help promote the bowel movement
significantly improved intestinal peristalsis with a higher intestinal (Zhao & Yu, 2016). The concentration of SCFAs in feces of all the
transit rate (43.40% ± 5.58%), but there was no significant differ‐ experimental groups at day 15 is shown in Figure 4. Compared with
ence between the XOS and control groups. that of the control and normal groups, higher concentrations of
The time to first black stool defecation of mice in all treated total SCFAs were observed in all the treated groups. Moreover, a
groups after inducing constipation is shown in Figure 3. The shortest significant increase (p < 0.05) was observed in the high‐dose MHGG

F I G U R E 2 Small intestinal transit rate


of mice after inducing constipation with
loperamide. Values are mean ± SD, n = 6.
**
p < 0.01 compared to control group
(Student’s t test). *p < 0.05 compared
to control group (Student’s t test).
Abbreviation: XOS (xylo‐oligosaccharide),
GOS (galacto‐oligosaccharide), HHGG
(high‐molecular‐weight partially
hydrolyzed guar gum), MHGG (medium‐
molecular‐weight partially hydrolyzed
guar gum)
FU et al. | 7 of 14

F I G U R E 3 Time to first black stool defecation of mice after establishing constipation model using loperamide. Values are mean ± SD,
n = 6. **p < 0.01 compared to control group (Student’s t test). *p < 0.05 compared to control group (Student’s t test). Abbreviation: XOS (xylo‐
oligosaccharide), GOS (galacto‐oligosaccharide), HHGG (high‐molecular‐weight partially hydrolyzed guar gum), MHGG (medium‐molecular‐
weight partially hydrolyzed guar gum)

F I G U R E 4 SCFA production in feces


from all experimental groups. Values are
mean ± SD, n = 6. Differences among
groups were evaluated for significance
by the one‐way ANOVA (Tukey post
hoc test), and the variations of the total
SCFA production were signed outside
the column, and different letters indicate
a significant difference (p < 0.05).
Abbreviation: XOS (xylo‐oligosaccharide),
GOS (galacto‐oligosaccharide), HHGG
(high‐molecular‐weight partially
hydrolyzed guar gum), MHGG (medium‐
molecular‐weight partially hydrolyzed
guar gum)

group, followed by that of the XOS, GOS, and low‐dose MHGG GPR41, affecting intestinal transit and protecting against the de‐
groups. Besides, both the low‐ and high‐dose HHGG groups showed velopment of inflammatory diseases (Brown et al., 2003; Maslowski
low SCFA production capacity which indicates that lower molecu‐ et al., 2009). In this study, MHGG promoted the accumulation of
lar weight carbohydrates may be more easily fermented by the gut acetate compared to that observed for the untreated groups, and
microbiota. the production was dependent on the concentration of MHGG. A
Acetate, comprising around 62.31%–73.59% of the total SCFA significant increase in acetate (p < 0.05) was observed for the high‐
production, was the major SCFA in all the experimental groups, dose MHGG and GOS groups, followed by the low‐dose MHGG
which was followed by propionate (18.25%–22.81%). Acetate and group. Additionally, an increasing trend of acetate was also observed
propionate have been reported to act as agonists of GPR43 and in the XOS and HHGG groups. Compared to that of the untreated
8 of 14 | FU et al.

groups, XOS significantly promoted the accumulation of propio‐ may be attributed to the additional oligosaccharides and polysac‐
nate (p < 0.05) and increased the relative ratio of propionate in total charides. Moreover, three phyla decreased in all the treated groups,
SCFAs. Moreover, a significant increase in propionate was observed including Firmicutes, Proteobacteria, and Tenericutes. Furthermore,
for the high‐dose MHGG and GOS groups, followed by the low‐dose a decreased Firmicutes/Bacteroidetes ratio was observed in the
MHGG group (p < 0.05). Additionally, the HHGG had no obvious ef‐ XOS, GOS, and PHGG‐treated groups compared to that of the
fect on promoting the production of propionate. untreated groups, and the lowest Firmicutes/Bacteroidetes ratio
Butyrate is considered a main energy substrate and factor was observed in the high‐dose MHGG group, followed by the low‐
for the stimulation and differentiation of colonocytes (Rowe & dose HHGG, XOS, and GOS groups. The decreased Firmicutes/
Bayless, 1992). Some studies have reported that butyrate can Bacteroidetes ratio was reported to be associated with weight loss
elicit potential therapeutic effects on constipation, such as re‐ (Ley, Turnbaugh, Klein, & Gordon, 2006), however, its relationship
lieving the pain during defecation (Vanhoutvin et al., 2009). with constipation remains poorly understood. Differences in the
Additionally, the anti‐inflammatory properties of butyrate could composition of the intestinal microbiota between constipated pa‐
potentially promote bowel movement by reducing intestinal in‐ tients and healthy controls have been demonstrated by Zhu et al.
flammation (Kristjánsson, Venge, Wanders, Lööf, & Hällgren, (2014), and the data showed a significant increase in Bacteroidetes
2004). In this study, XOS and GOS showed good butyrogenic ef‐ and a decreased tendency of Firmicutes in the control group com‐
fect compared to that of the untreated groups, followed by MHGG pared with that of the constipated patients. Therefore, the variation
treatment groups. However, no butyrogenic effect was observed of microflora of Bacteroidetes and Firmicutes may contribute to
in the HHGG groups. modulating constipation with PHGG.
The top 40 dominant genera in all the experimental groups were
selected to construct a heat map (Figure 6). The fecal microbiota
3.5 | Structural changes in the fecal
in all the samples was dominated by Bacteroides and Alloprevotella,
microbiota of mice
followed by Lactobacillus, Prevotellaceae_UCG‐001, Alistipes, and
Chronic constipation is a prevalent gastrointestinal disorder, how‐ Odoribacter. Compared to that of the normal group, increases in
ever it’s etiology is most likely multifactorial and is not well studied Bacteroides were observed in all the treated groups except GOS.
(Zhao & Yu, 2016). The gut microbiome plays important roles in gas‐ The increased abundance of Bacteroides may be associated with
trointestinal functions such as modulating immune responses and the fermentation of carbohydrates, which may contribute to the
promoting gastrointestinal motility (Maurice, Haiser, & Turnbaugh, production of organic acids (Zhao & Yu, 2016). Lactobacilli are the
2013; Mazmanian, Liu, Tzianabos, & Kasper, 2005), and increasing generally recognized beneficial species for regulating intestinal
data show that regulation of intestinal microbiota may have benefi‐ function and previous studies have reported that an abundance of
cial effects on relieving constipation. However, the potential effect Lactobacillus was either lower or higher in constipated models than
of gut microbiota in constipation remains poorly understood. To as‐ in that of normal groups (Chassard et al., 2012; Kashyap et al., 2013;
sess the influence of PHGG on the gut microbiome, feces of all eight Monteagudo‐Mera et al., 2016). In this study, low‐dose MHGG
experimental groups were collected at d 15. High‐quality 16S rRNA promoted the accumulation of Lactobacillus in all treated groups.
gene sequences were generated through HiSeq sequencing analysis. Husebye, Hellström, Sundler, Chen, and Midtvedt (2001) trans‐
The average sequence read was 57,241 per sample and follow‐up planted conventional intestinal microbiota into germ‐free rats to
analyses were performed after subsampling of all samples accord‐ reveal that Lactobacillus acidophilus could accelerate small intestinal
ing to the minimum number of sequences. OTUs were generated at transit. The potential effect of gut microbiota in constipation is not
≥97% sequence identity and the number of OTUs per sample ranged well understood, but production of H2S by fermentation of sulphate‐
between 229 and 300. reducing bacteria (SRB) could modulate peripheral pain‐related sig‐
The heat map showed that at phyla level, the fecal micro‐ nals and may influence gut sensory motor functions (Kawabata et
biome in all the samples was dominated by Bacteroidetes, fol‐ al., 2007). Moreover, Chassard et al. (2012) reported that a high
lowed by the Firmicutes, Actinobacteria, Tenericutes, and number of H2‐utilizing sulphate‐reducing bacteria were observed
Proteobacteria (Figure 5). Compared with that of the untreated in their constipation predominant‐irritable bowel syndrome (IBS‐C)
groups, Bacteroidetes showed a higher relative abundance in the group compared to that of healthy controls, which could in turn in‐
XOS, GOS, and all PHGG‐treated groups, of which the highest was fluence colonic motility and generate IBS symptoms. Desulfovibrio is
observed in the high‐dose MHGG group. Bacteroidetes are among the dominant SRB in the flora of the human colon with a capacity for
the major members of the gut microbiota of animals, they are in‐ intestinal colonization through the mucus layer, and its metabolites
creasingly regarded as degraders of high molecular weight organic are potential pathogenic factors of chronic gastrointestinal diseases
matter, such as proteins and carbohydrates (Thomas, Hehemann, (Devereux, Delaney, Widdel, & Stahl, 1989; Gibson, Macfarlane, &
Rebuffet, Czjzek, & Michel, 2011), and the presence of numerous Cummings, 1993). In this study, an obvious decrease in Desulfovibrio
carbohydrate‐active enzymes were found through sequencing of was observed in all the PHGG‐treated groups and positive groups.
the Bacteroidetes genomes (Davies, Gloster, & Henrissat, 2005). The lowest abundance was observed in the XOS group, followed by
Therefore, the increase in Bacteroidetes in all the treated groups the high‐dose MHGG group.
FU et al. | 9 of 14

F I G U R E 5 Heat map of the bacterial abundance in phyla level at day 15 in all experimental groups. The numbers of reads of each phylum
were assigned to each color block. The abundance of each phylum is plotted in red–blue color scale. The red and blue color indicates
high and low abundance, respectively. Abbreviation: XOS (xylo‐oligosaccharide), GOS (galacto‐oligosaccharide), HHGG (high‐molecular‐
weight partially hydrolyzed guar gum), MHGG (medium‐molecular‐weight partially hydrolyzed guar gum)

In addition to modulate the gut microbiota associated


3.6 | Factors influencing intestinal transit and
with constipation as we discussed above, dietary fiber can
structuring bacterial community in feces
also be utilized by specific bacteria to produce SCFAs, which
may be beneficial to normalize defecation (Brown et al., 2003; The Spearman correlation analysis showed that the intestinal ­transit
Maslowski et al., 2009). Alloprevotella, which has been rec‐ rate was significantly positively correlated with fecal water content
ognized previously to produce moderate amounts of acetate (p < 0.01), but was not related to the content of SCFAs (Table2).
during fermentation (Downes, Dewhirst, Tanner, & Wade, These results are in accordance with the conclusion mentioned above,
2013), was enriched in high‐dose MHGG group. Bacteroides, pointing that the water‐holding capacity of PHGG is one of the
which has been reported to produce acetate and propionate major mechanisms for constipation relief. A Spearman correlation
(Louis, Hold, & Flint, 2014), was stimulated by the treatment coefficient of the top 40 abundant bacterial genera and environmen‐
of PHGG and XOS. Moreover, MHGG, XOS, and GOS promoted tal factors was displayed by the heat map (Figure 7). Among them,
the proliferation of Odoribacter and Alistipes, which are known Odoribacter, a butyrate‐producing bacterium (Vital et al., 2017) and
for their butyrate‐producing capabilities in intestinal tract Rikenella, an acetate and propionate‐producing bacterium (Kaneuchi
(Vital, Karch, & Pieper, 2017). & Mitsuoka, 1978), showed significantly positive correlation with
10 of 14 | FU et al.

F I G U R E 6 Heat map of the relative abundance of the top 40 genera at day 15 in all experimental groups. The abundance of each
genus is plotted in red–blue color scale. The red and blue color indicates high and low abundance, respectively. Abbreviation: XOS (xylo‐
oligosaccharide), GOS (galacto‐oligosaccharide), HHGG (high‐molecular‐weight partially hydrolyzed guar gum), MHGG (medium‐molecular‐
weight partially hydrolyzed guar gum)

TA B L E 2 Correlation analysis of intestinal transit rate and fecal indices

Index Fecal water content Acetate Propionate Iso‐butyrate Butyrate Total SCFAs
**
Intestinal transit 0.952 0.143 0.071 −0.214 0.024 0.024
rate

Note. The correlation coefficient and significance were obtained using Spearman correlation analysis. Significant values are shown as:
*
p < 0.05; **p < 0.01.

the total content of SCFAs (0.01 < p ≤ 0.05). The abundances of correlation with any bacterial genus, proving that the change of intestinal
Lachnospiraceae_NK4A136_group (p ≤ 0.001), Ruminiclostridium_6 microbiota contributed little to the transit rate. Fecal water con‐
(p ≤ 0.001), and Desulfovibrio (0.001 < p ≤ 0.01) decreased in all the tent in the modeling mice also showed no correlation with the in‐
fiber‐treated groups, showing significantly negative correlation with testinal microbiota except for Ruminococcaceae_NK4A214_group
the total content of SCFAs. However, intestinal transit rate showed no (0.01 < p ≤ 0.05).
FU et al. | 11 of 14

F I G U R E 7 Correlation heat map of the top 40 genera and environmental factors, including the intestinal transit rate, the fecal water
content, and the total content of SCFAs. The Spearman correlation coefficient was calculated using the pheatmap package of R. X and Y axis
are environmental factors and genera, respectively. R was marked in different colors, and the legend in the right side is the color range of
different R values. ***p ≤ 0.001, **0.001 < p ≤ 0.01, *0.01 < p ≤ 0.05

4 | CO N C LU S I O N S This in vivo test suggests that MHGG perform a constipation reliev‐


ing effect by promoting small intestinal transit and increasing the
In conclusion, PHGG significantly increased the water content of fecal water content. Furthermore, since the preparation of MHGG
feces after constipation modeling, and MHGG at a high‐dose ex‐ by enzymatic hydrolysis is low‐cost and eco‐friendly, it may provide
hibited the best effect for constipation relief. MHGG could sig‐ an option for the development of novel dietary fiber.
nificantly promote small intestinal transit, followed by HHGG and
GOS, and this effect was concentration‐dependent. The same ten‐
AC K N OW L E D G M E N T S
dency was observed for the time to the first black stool defecation.
Furthermore, MHGG possessed better fermentability in the intes‐ The authors are grateful for the financial support by the Key
tine of mice than HHGG, showing the highest SCFA production of Research and Development Project Foundation of Shandong
all the test groups. Moreover, PHGG promoted the accumulation of Province (2017YYSP003) and Shandong Natural Science Fund
Bacteroidetes and inhibited harmful species, such as Desulfovibrio. (ZR2017MD006).
12 of 14 | FU et al.

C O N FL I C T O F I N T E R E S T S Frank, D. N., St. Amand, A. l., Feldman, R. A., Boedeker, E. C., Harpaz,
N., & Pace, N. R. (2007). Molecular‐phylogenetic characterization of
The authors declare no conflict of interest. microbial community imbalances in human inflammatory bowel dis‐
eases. Proceedings of the National Academy of Sciences of the United
States of America, 104(34), 13780–13785. https://doi.org/10.1073/
pnas.0706625104
ORCID
Gibson, G. R., Macfarlane, G. T., & Cummings, J. H. (1993). Sulphate re‐
ducing bacteria and hydrogen metabolism in the human large intes‐
Xiaodan Fu http://orcid.org/0000-0001-6969-7369
tine. Gut, 34(4), 437–439. https://doi.org/10.1136/gut.34.4.437
Goulder, T. J., Alberti, K. G. M. M., & Jenkins, D. A. (1978). Effect of
added fiber on the glucose and metabolic response to a mixed meal
REFERENCES
in normal and diabetic subjects. Diabetes Care, 1(6), 351. https://doi.
Bai, W., Fang, X., Zhao, W., Huang, S., Zhang, H., & Qian, M. (2015). org/10.2337/diacare.1.6.351
Determination of oligosaccharides and monosaccharides in Hakka Hidehisa, T., Nobuhisa, W., Tsutomu, O., Noriyuki, I., Junzo, Y., & Takehiko,
rice wine by precolumn derivation high‐performance liquid chroma‐ Y. (1994). Influence of partially hydrolyzed guar gum on constipation
tography. Journal of Food and Drug Analysis, 23(4), 645–651. https:// in women. Journal of Nutritional Science and Vitaminology, 40(3), 251–
doi.org/10.1016/j.jfda.2015.04.011 259. https://doi.org/10.3177/jnsv.40.251
Bharucha, A. E. (2007). Constipation. Best Practice & Research Clinical Husebye, E. E., Hellström, P. M., Sundler, F., Chen, J., & Midtvedt, T.
Gastroenterology, 21(4), 709–731. https://doi.org/10.1016/j.bpg.2007. (2001). Influence of microbial species on small intestinal myoelectric
07.001 activity and transit in germ‐free rats. American Journal of Physiology.
Brown, A. J., Goldsworthy, S. M., Barnes, A. A., Eilert, M. M., Tcheang, Gastrointestinal and Liver Physiology, 280(3), G368–G380. https://doi.
L., Daniels, D., … Dowell, S. J. (2003). The orphan G protein‐coupled org/10.1152/ajpgi.2001.280.3.G368
receptors GPR41 and GPR43 are activated by propionate and other Jennings, A., Davies, G. J., Costarelli, V., & Dettmar, P. W. (2009). Dietary
short chain carboxylic acids. Journal of Biological Chemistry, 278(13), fibre, fluids and physical activity in relation to constipation symp‐
11312–11319. https://doi.org/10.1074/jbc.M211609200 toms in pre‐adolescent children. Journal of Child Health Care, 13(2),
Chassard, C., Dapoigny, M., Scott, K. P., Crouzet, L., Del’homme, C., 116–127. https://doi.org/10.1177/1367493509102469
Marquet, P., … Bernalier‐Donadille, A. (2012). Functional dysbio‐ Kaneuchi, C., & Mitsuoka, T. (1978). Bacteroides microfusus, a new spe‐
sis within the gut microbiota of patients with constipated‐irritable cies from the intestines of calves, chickens, and Japanese quails.
bowel syndrome. Alimentary Pharmacology & Therapeutics, 35(7), International Journal of Systematic and Evolutionary Microbiology,
828–838. https://doi.org/10.1111/j.1365-2036.2012.05007.x 28(4), 478–481. https://doi.org/10.1099/00207713-28-4-478
Daas, P., Schols, H. A., & de Jongh, H. H. J. (2000). On the galactosyl Kashyap, P. C., Marcobal, A., Ursell, L. K., Larauche, M., Duboc, H.,
distribution of commercial galactomannans. Carbohydrate Research, Earle, K. A., … Sonnenburg, J. L. (2013). Complex interactions among
329(3), 609–619. https://doi.org/10.1016/S0008-6215(00)00209-3 diet, gastrointestinal transit, and gut microbiota in humanized
Davies, G. J., Gloster, T. M., & Henrissat, B. (2005). Recent structural mice. Gastroenterology, 144(5), 967–977. https://doi.org/10.1053/j.
insights into the expanding world of carbohydrate‐active enzymes. gastro.2013.01.047
Current Opinion in Structural Biology, 15(6), 637–645. https://doi. Kau, A. L., Ahern, P. P., Griffin, N. W., Goodman, A. L., & Gordon, J. I.
org/10.1016/j.sbi.2005.10.008 (2011). Human nutrition, the gut microbiome, and immune system:
Derrien, M., van Passel, M. W. J., van de Bovenkamp, J. H. B., Schipper, Envisioning the future. Nature, 474(7351), 327–336. https://doi.
R. G., de Vos, W. M., & Dekker, J. (2010). Mucin‐bacterial interac‐ org/10.1038/nature10213
tions in the human oral cavity and digestive tract. Gut Microbes, 1(4), Kawabata, A., Ishiki, T., Nagasawa, K., Yoshida, S., Maeda, Y., Takahashi,
254–268. https://doi.org/10.4161/gmic.1.4.12778 T., … Nishikawa, H. (2007). Hydrogen sulfide as a novel nocicep‐
DeSantis, T. Z., Hugenholtz, P., Larsen, N., Rojas, M., Brodie, E. L., Keller, tive messenger. Pain, 132(1–2), 74–81. https://doi.org/10.1016/j.
K., … Andersen, G. L. (2006). Greengenes, a chimera‐checked 16S pain.2007.01.026
rRNA gene database and workbench compatible with ARB. Applied Khalif, I. L., Quigley, E. M. M., Konovitch, E. A., & Maximova, I. D. (2005).
and Environmental Microbiology, 72(7), 5069–5072. https://doi. Alterations in the colonic flora and intestinal permeability and ev‐
org/10.1128/AEM.03006-05 idence of immune activation in chronic constipation. Digestive
Devereux, R., Delaney, M., Widdel, F., & Stahl, D. A. (1989). Natural relation‐ and Liver Disease, 37(11), 838–849. https://doi.org/10.1016/j.
ships among sulfate‐reducing eubacteria. Journal of Bacteriology, 171(12), dld.2005.06.008
6689–6695. https://doi.org/10.1128/jb.171.12.6689-6695.1989 Kristjánsson, G., Venge, P., Wanders, A., Lööf, L., & Hällgren, R. (2004).
Downes, J., Dewhirst, F. E., Tanner, A. C. R., & Wade, W. G. (2013). Clinical and subclinical intestinal inflammation assessed by the
Description of alloprevotella rava gen. nov., sp. nov., isolated from mucosal patch technique: Studies of mucosal neutrophil and eo‐
the human oral cavity, and reclassification of Prevotella tannerae sinophil activation in inflammatory bowel diseases and irritable
Moore et al. 1994 as alloprevotella tannerae gen. nov., comb. nov. bowel syndrome. Gut, 53(12), 1806–1812. https://doi.org/10.1136/
International Journal of Systematic and Evolutionary Microbiology, gut.2003.036418
63(Pt 4), 1214–1218. https://doi.org/10.1099/ijs.0.041376-0 Kruse, H. P., Kleessen, B., & Blaut, M. (1999). Effects of inulin on faecal
Edgar, R. C. (2013). UPARSE: Highly accurate OTU sequences from bifidobacteria in human subjects. British Journal of Nutrition, 82(5),
microbial amplicon reads. Nature Methods, 10, 996. https://doi. 375–382. https://doi.org/10.1017/S0007114599001622
org/10.1038/nmeth.2604 Kurakake, M., Sumida, T., Masuda, D., Oonishi, S., & Komaki, T. (2006).
Flint, H. J., Bayer, E. A., Rincon, M. T., Lamed, R., & White, B. A. (2008). Production of galacto‐manno‐oligosaccharides from guar gum by
Polysaccharide utilization by gut bacteria: Potential for new insights β‐mannanase from penicillium oxalicum SO. Journal of Agricultural
from genomic analysis. Nature Reviews Microbiology, 6, 121. https:// and Food Chemistry, 54(20), 7885–7889. https://doi.org/10.1021/
doi.org/10.1038/nrmicro1817 jf061502k
Flint, H. J., Scott, K. P., Louis, P., & Duncan, S. H. (2012). The role of the gut Ley, R. E., Turnbaugh, P. J., Klein, S., & Gordon, J. I. (2006). Human gut
microbiota in nutrition and health. Nature Reviews Gastroenterology & microbes associated with obesity. Nature, 444, 1022. https://doi.
Hepatology, 9, 577–589. https://doi.org/10.1038/nrgastro.2012.156 org/10.1038/4441022a
FU et al. | 13 of 14

Li, T., Lu, X., & Yang, X. (2013). Stachyose‐enriched α‐galacto‐oligosac‐ concentrations in venous blood. Journal of Clinical Investigation, 75(5),
charides regulate gut microbiota and relieve constipation in mice. 1448–1454. https://doi.org/10.1172/JCI111847
Journal of Agricultural and Food Chemistry, 61(48), 11825–11831. Puig, M. M., Pol, O., & Warner, W. (1996). Interaction of morphine and
https://doi.org/10.1021/jf404160e clonidine on gastrointestinal transit in mice. Anesthesiology, 85(6),
Li, Y., Zong, Y., Qi, J., & Liu, K. (2011). Prebiotics and oxidative stress in 1403–1412. https://doi.org/10.1097/00000542-199612000-00022
constipated rats. Journal of Pediatric Gastroenterology and Nutrition, Rouhibakhsh, A., Priya, J. J., Periasamy, M., Haq, Q. M., & Malathi, V.
53(4), 447–452. https://doi.org/10.1097/MPG.0b013e31821eed83 G. (2008). An improved DNA isolation method and PCR protocol
López Román, J., Martínez Gonzálvez, A. B., Luque, A., Pons Miñano, J. for efficient detection of multicomponents of begomovirus in le‐
A., Vargas Acosta, A., Iglesias, J. R., & Hernández, M. (2008). Efecto gumes. Journal of Virological Methods, 147(1), 37–42. https://doi.
de la ingesta de un preparado lácteo con fibra dietética sobre el es‐ org/10.1016/j.jviromet.2007.08.004
treñimiento crónico primario idiopático. Nutrición Hospitalaria, 23, Rowe, W. A., & Bayless, T. M. (1992). Colonic short‐chain fatty acids:
12–19. Fuel from the lumen? Gastroenterology, 103(1), 336–338. https://doi.
Louis, P., Hold, G. L., & Flint, H. J. (2014). The gut microbiota, bacterial org/10.1016/0016-5085(92)91133-O
metabolites and colorectal cancer. Nature Reviews Microbiology, 12, Slavin, J. L., & Greenberg, N. A. (2003). Partially hydrolyzed guar
661–672. https://doi.org/10.1038/nrmicro3344 gum: Clinical nutrition uses. Nutrition, 19(6), 549–552. https://doi.
Maslowski, K. M., Vieira, A. T., Ng, A., Kranich, J., Sierro, F., Di, Y., … org/10.1016/S0899-9007(02)01032-8
Mackay, C. R. (2009). Regulation of inflammatory responses by gut Spapen, H., Diltoer, M., Van Malderen, C., Opdenacker, G., Suys, E., &
microbiota and chemoattractant receptor GPR43. Nature, 461, 1282. Huyghens, L. (2001). Soluble fiber reduces the incidence of diarrhea
https://doi.org/10.1038/nature08530 in septic patients receiving total enteral nutrition: A prospective,
Maurice, C. F., Haiser, H. J., & Turnbaugh, P. J. (2013). Xenobiotics double‐blind, randomized, and controlled trial. Clinical Nutrition,
shape the physiology and gene expression of the active human 20(4), 301–305. https://doi.org/10.1054/clnu.2001.0399
gut microbiome. Cell, 152(1), 39–50. https://doi.org/10.1016/j. Srivastava, M., & Kapoor, V. P. (2005). Seed galactomannans: An overview.
cell.2012.10.052 Chemistry & Biodiversity, 2(3), 295–317. https://doi.org/10.1002/
Mazmanian, S. K., Liu, C. H., Tzianabos, A. O., & Kasper, D. L. (2005). An cbdv.200590013
immunomodulatory molecule of symbiotic bacteria directs matura‐ Srivastava, P. K., & Kapoor, M. (2014). Cost‐effective endo‐mannanase
tion of the host immune system. Cell, 122(1), 107–118. https://doi. from bacillus sp. CFR1601 and its application in generation of oli‐
org/10.1016/j.cell.2005.05.007 gosaccharides from guar gum and as detergent additive. Preparative
Mcrorie, J. W., Daggy, B. P., Morel, J. G., Diersing, P. S., Miner, Biochemistry & Biotechnology, 44(4), 392–417. https://doi.org/10.108
P. B., & Robinson, M. (1998). Psyllium is superior to docu‐ 0/10826068.2013.833108
sate sodium for treatment of chronic constipation. Alimentary Stephen, A. M., & Cummings, J. H. (1979). Water‐holding by dietary fibre
Pharmacology & Therapeutics, 12(5), 491–497. https://doi. in vitro and its relationship to faecal output in man. Gut, 20(8), 722–
org/10.1046/j.1365-2036.1998.00336.x 729. https://doi.org/10.1136/gut.20.8.722
Moezi, L., Pirsalami, F., & Inaloo, S. (2015). Constipation enhances the Stephen, A. M., & Cummings, J. H. (1980). Mechanism of action of di‐
propensity to seizure in pentylenetetrazole‐induced seizure models etary fibre in the human colon. Nature, 284, 283–284. https://doi.
of mice. Epilepsy & Behavior, 44, 200–206. https://doi.org/10.1016/j. org/10.1038/284283a0
yebeh.2015.01.013 Thomas, F., Hehemann, J.‐H., Rebuffet, E., Czjzek, M., & Michel, G.
Monteagudo‐Mera, A., Arthur, J. C., Jobin, C., Keku, T., Bruno‐Barcena, (2011). Environmental and gut bacteroidetes: The food connec‐
J. M., & Azcarate‐Peril, M. A. (2016). High purity galacto‐oligosac‐ tion. Frontiers in Microbiology, 2, 93. https://doi.org/10.3389/
charides (GOS) enhance specific bifidobacterium species and their fmicb.2011.00093
metabolic activity in the mouse gut microbiome. Beneficial Microbes, Tropini, C., Earle, K. A., Huang, K. C., & Sonnenburg, J. L. (2017). The
7(2), 247–264. https://doi.org/10.3920/BM2015.0114 gut microbiome: Connecting spatial organization to function.
Morris, J. B. (2010). Morphological and reproductive characterization Cell Host & Microbe, 21(4), 433–442. https://doi.org/10.1016/j.
of guar (Cyamopsis tetragonoloba) genetic resources regenerated in chom.2017.03.010
Georgia. USA. Genetic Resources and Crop Evolution, 57(7), 985–993. Vanhoutvin, S. A. L. W., Troost, F. J., Kilkens, T. O. C., Lindsey, P. J.,
https://doi.org/10.1007/s10722-010-9538-8 Hamer, H. M., Jonkers, D. M. A. E., … Brummer, R. J. M. (2009). The
Mudgil, D., Barak, S., & Khatkar, B. S. (2014). Guar gum: Processing, effects of butyrate enemas on visceral perception in healthy volun‐
properties and food applications—A review. Journal of Food teers. Neurogastroenterology & Motility, 21(9), 952–e976. https://doi.
Science and Technology, 51(3), 409–418. https://doi.org/10.1007/ org/10.1111/j.1365-2982.2009.01324.x
s13197-011-0522-x Vital, M., Karch, A., & Pieper, D. H. (2017). Colonic butyrate‐producing
Nagakura, Y., Naitoh, Y., Kamato, T., Yamano, M., & Miyata, K. (1996). communities in humans: An overview using omics data. mSystems,
Compounds possessing 5‐HT3 receptor antagonistic activity in‐ 2(6). https://doi.org/10.1128/mSystems.00130-17
hibit intestinal propulsion in mice. European Journal of Pharmacology, Wang, Q., Garrity, G. M., Tiedje, J. M., & Cole, J. R. (2007). Naïve bayes‐
311(1), 67–72. https://doi.org/10.1016/0014-2999(96)00403-7 ian classifier for rapid assignment of rRNA sequences into the new
Niv, E., Halak, A., Tiommny, E., Yanai, H., Strul, H., Naftali, T., & Vaisman, bacterial taxonomy. Applied and Environmental Microbiology, 73(16),
N. (2016). Randomized clinical study: Partially hydrolyzed guar 5261–5267.
gum (PHGG) versus placebo in the treatment of patients with irri‐ Zhang, H., DiBaise, J. K., Zuccolo, A., Kudrna, D., Braidotti, M., Yu, Y., …
table bowel syndrome. Nutrition & Metabolism, 13, 10. https://doi. Krajmalnik‐Brown, R. (2009). Human gut microbiota in obesity and
org/10.1186/s12986-016-0070-5 after gastric bypass. Proceedings of the National Academy of Sciences
Pol, O., Planas, E., & Puig, M. M. (1995). Peripheral effects of naloxone of the United States of America, 106(7), 2365–2370. https://doi.
in mice with acute diarrhea associated with intestinal inflamma‐ org/10.1073/pnas.0812600106
tion. Journal of Pharmacology and Experimental Therapeutics, 272(3), Zhao, Y., Wu, J. F., Li, J. V., Zhou, N.‐Y., Tang, H., & Wang, Y. L. (2013).
1271–1276. Gut microbiota composition modifies fecal metabolic profiles in
Pomare, E. W., Branch, W. J., & Cummings, J. H. (1985). Carbohydrate mice. Journal of Proteome Research, 12(6), 2987–2999. https://doi.
fermentation in the human colon and its relation to acetate org/10.1021/pr400263n
14 of 14 | FU et al.

Zhao, Y. M., & Yu, Y.‐B. (2016). Intestinal microbiota and chronic
constipation. SpringerPlus, 5(1). 1130. https://doi.org/10.1186/ How to cite this article: Fu X, Li R, Zhang T, Li M, Mou H.
s40064-016-2821-1 Study on the ability of partially hydrolyzed guar gum to
Zhu, L., Liu, W., Alkhouri, R., Baker, R. D., Bard, J. E., Quigley, E. M., &
modulate the gut microbiota and relieve constipation. J Food
Baker, S. S. (2014). Structural changes in the gut microbiome of con‐
stipated patients. Physiological Genomics, 46(18), 679–686. https:// Biochem. 2018;e12715. https://doi.org/10.1111/jfbc.12715
doi.org/10.1152/physiolgenomics.00082.2014

You might also like