Sodium butyrate

Polyvinyl Butyrate Nanoparticles as Butyrate Donors for Colitis Treatment

ImageYunmei Mu, Yusuke Kinashi, Jinting Li, Takuma Yoshikawa, Akihiro Kishimura, Mitsuru Tanaka, Toshiro Matsui, Takeshi Mori,* Koji Hase,* and Yoshiki Katayama*

ACCESS
Metrics & More
Article Recommendations
*sı Supporting Information

ABSTRACT: Butyrate has been attracting attention for the suppression of inflammatory bowel disease (IBD). However, clinical trials of butyrate for IBD treatment have resulted in controversial outcomes, likely owing to the adverse effect of butyrate on the intestinal epithelium that was observed at high butyrate concentrations. Herein, we propose polyvinyl butyrate (PVBu) nanoparticles (NPs) as butyrate donors for delivery to the lower part of the intestine for the treatment of colitis. The PVBu NPs suppressed the inflammatory activation of macrophages in vitro, although sodium butyrate inversely further activated macrophages. Oral admin- istration of NPs did not change the luminal concentration of free butyrate; however, NPs showed a therapeutic effect on a colitis mouse model. In addition, incorporation of vitamin D3 into the NPs enhanced the therapeutic effect on colitis. Hence, PVBu NPs are a promising therapeutic for IBD treatment, not only as a butyrate donor but also as a carrier for hydrophobic drugs like vitamin D3.

KEYWORDS: butyrate, polyvinyl butyrate, nanoparticles, inflammatory bowel disease, vitamin D3

⦁ INTRODUCTION
Inflammatory bowel disease (IBD) is an autoimmune
diseaseincluding Crohn’s disease and ulcerative colitis that is characterized by long-term chronic inflammation in the gastrointestinal (GI) tract.1 Immune cells in the intestine that maintain the intestinal homeostasis are conversely activated in IBD.2 Alteration of gut microbiota and derived metabolites have been implicated in the pathogenesis of IBD.3 Despite the
efforts of researchers, there is currently no effective medicine for IBD.4
Butyrate, which is a metabolite of dietary fiber produced by intestinal bacteria, has attracted attention as a potential therapeutic for IBD. Butyrate exists in high concentration (11−25 mM) in human feces5−7 and works as an inhibitor of histone deacetylase (HDAc)8,9 or a ligand of G protein- coupled receptors10,11 in epithelium and immune cells including macrophagesto preserve gastrointestinal health,

Received: August 31, 2020
Accepted: January 25, 2021
Published: February 11, 2021
maintaining the intestinal barrier and the anti-inflammatory conditions in mucosal immunity.12 Patients with IBD have been reported to have a reduced intestinal concentration of butyrate.7,13 Oral supplementation of butyrate has been used for the treatment of an animal colitis model.12−16

The approaches used in these studies to avoid rapid absorption of butyrate in the upper part of the GI tract include feeding a Because of the unpleasant smell and taste of butyrate, rectal enema and oral application of capsules embedded with butyrate have been developed for use in clinical trials.18−22 The reported therapeutic effects of butyrate in the clinical trials have been controversial.5 A possible reason for the controversial results for butyrate is its conflicting effect on mucosal health. High concentrations of butyrate showed a negative effect as a result of suppressing proliferation of intestinal epithelia23 and inducing production of inflammatory cytokines from immune cells.24 It is difficult to control the luminal concentration of butyrate with the current rectal enema and oral capsule formulations, which results in the risk
of inducing adverse effects of butyrate. The ideal formulation of butyrate would allow for oral delivery to the lower part of the intestine with controlled luminal concentration.

Here, we propose polyvinyl butyrate (PVBu) as a novel butyrate donor for IBD treatment (Figure 1). Although PVBu is water-insoluble, its water solubility increases gradually with the hydrolysis of butyrate. Because the glass transition butyrate-mixed diet14,15 and feeding starch modified with
butyrate through ester bonds.16,17

Downloaded via RUTGERS UNIV on May 15, 2021 at 06:33:19 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
© 2021 American Chemical Society

2335

https://dx.doi.org/10.1021/acsabm.0c01105

ACS Appl. Bio Mater. 2021, 4, 2335−2341
Proposed mechanism of action of PVBu NPs for IBD treatment.

■temperature of PVBu is below the room temperature, it exists as a melt at room temperature. Owing to the tunable hydrophobicity and flexibility of the PVBu chain, nanoparticles (NPs) prepared from PVBu are expected to release butyrate at moderate speed by allowing access to hydrolysis enzymes. Accumulation in intestinal inflammatory lesions has been reported to be enhanced by tuning the NPs size to a few hundred nanometers.25,26 The small size allows the NPs to be engulfed by inflammatory cells such as resident macrophages, which are representative cells that respond to butyrate.8 In addition, we evaluate the suitability of PVBu NPs as a carrier for hydrophobic drug for delivery to the lower part of the intestine.

EXPERIMENTAL SECTION

Materials. Vinyl butyrate (VBu) was purchased from Tokyo Chemical Industry Co., Ltd. (Tokyo, Japan). Azobisisobutyronitrile (AIBN), acetone, methanol, and toluene were purchased from Wako Pure Chemical Industries, Ltd. (Tokyo, Japan). Lipase from porcine pancreas and Pluronic F-127 were purchased from Sigma-Aldrich (St.Louis, USA). Calcitriol was obtained from Cayman chemical (≥97%, Michigan, USA). 3,3′-Dioctadecyloxacarbocyanine perchlorate (DiO) was purchased from Takara-Clontech (Shiga, Japan). 1,1′-Dioctade- cyltetramethyl indotricarbocyanine iodide (DiR) was purchased from
Caliper Life Sciences (Hopkinton, USA). Dextran sulfate sodium (DSS), colitis grade (36,000−50,000 Mw) was obtained from MP Biomedicals (Irvine, USA).
Synthesis of PVBu.

PVBu was synthesized by radical polymer- ization. To a 200 mL round-bottom flask was added vinyl butyrate (20 g, 175 mmol), AIBN (0.28 g, 1.7 mmol), and acetone (80 g). Then, the solution was septa sealed and purged with nitrogen for 30 min. The flask was then transferred to a preheated water bath at 50 °C to polymerize for 48 h. The polymerization solution was dialyzed against methanol for 48 h. Following dialysis, the methanol was removed via rotary evaporation and then lyophilized. The obtained polymer was confirmed by 1H NMR (CDCl3, 400 MHz) spectros- copy.
Gel Permeation Chromatography Analysis. Weight average and number average molecular weights (Mw and Mn) of the polymer were determined on a multi detector SEC system fitted with a Malvern Viscotek TDA 305 system and three Tosoh TSK-gel columns (TSKgel GMHHR-H × 2, G2000HHR). Tetrahydrofuran was used as the eluent with a flow rate of 1.0 mL/min at 40 °C, and the system was calibrated to polystyrene standards from Agilent Technologies.
Preparation of PVBu NPs. 2 mL of toluene solution containing PVBu (0.4 g/mL) was prepared [in the cases of incorporation, vitamin D3 (1,25(OH)2D3) (0.63 μg/mL) or DiO or DiR (10 μg/ mL) was added to the toluene solution]. This solution was added to an aqueous solution (20 mL) containing Pluronic F-127 (5%) and homogenizer (T25 digital Ultra turax, IKA, Germany) at 12,000 rpm for 10 min and sonicated with a probe sonicator [UD-211 (TOMY) equipped with a TP-040 tip] for 5 min (20% power, 20 kHz, 20 W). Toluene was then evaporated from the solution by stirring (300 rpm) overnight. The NP dispersions were stored as prepared at 4 °C. Entrapment efficacy of vitamin D3 in NPs was determined by HPLC following our previous report.27
Measurement of the NP Size and ζ-potential. The size, ζ- potential, and polydispersity index (PDI) of the obtained NPs were measured by a Zeta Sizer Nano Series (Malvern Instrument, UK) analyzer at 25 °C following our previous report.27

Butyrate Release from NPs. 100 mg of lipase was suspended in 40 mL of 2.5 mM Tris−HCl buffer (pH 7.4) containing NPs (34 mM butyrate conc.) or VBu (34 mM). The enzymatic reaction was carried out at 37 °C with the pH-stat method using an automatic titrator (DKK-TOA Co., Japan) as reported previously.28 Butyric acid released from the NPs was quantitatively titrated with 20 mM NaOH to maintain a constant pH (pH 7.4). During the enzymatic reaction, the dispersion was stirred vigorously with a magnetic stirrer. Cell Culture. RAW 264.7 macrophages and RAW 264.7 macrophages transfected with secreted alkaline phosphatase (SEAP) as a reporter gene under the transcriptional control of an NF-κB response element were cultured in complete Dulbecco’s modified Eagle’s medium (DMEM) containing 10% fetal bovine serum, 4.5 mg/mL glucose, 100 U/mL penicillin, and 100 U/mL streptomycin, supplemented with 500 μg/mL G418. The cells were maintained in a humidified environment containing 5% CO2 and 95% air at 37 °C.
Cellular Uptake of NPs.

RAW264.7 macrophages (1 × 105 cells/ mL) were seeded in a 96-well glass surface plate. The cells were washed with Dulbecco’s phosphate-buffered saline (PBS) after incubation for 24 h. NP-DiO (5 mM butyrate conc.) dispersed in DMEM was added to the wells. After incubation for 24 h at 37 °C, the cells were observed using fluorescence microscopy (BZ-X700, Keyence Co., USA).
Suppressive Effect of NPs on Macrophage Activation. SEAP-transfected RAW 264.7 macrophages (1 × 106 cells/mL) were cultured with DMEM in a 96-well plate overnight. Then, DMEM was replaced with DMEM containing sodium butyrate (SB) (0.1−5 mM) or NPs (5 or 20 mM butyrate conc.). After incubation for 6 h, lipopolysaccharide (LPS) (final conc. 20 ng/mL) was added, and the plate was incubated for 18 h. The supernatants were collected for measurement of the SEAP level as described in a previous study.29

Observation of Retention of NPs in the GI Tract. Five-week- old C57BL/6J male mice were purchased from Kyudo Co., Ltd (Saga, Japan). Mice were housed in a facility maintained at 25 °C with a 12 h light/dark cycle. All animal experiments were performed using protocols approved by the Animal Ethics Committees of Kyushu University. The mice were fed with AIN 76A purified diet (Research Diets, Inc., USA) for 1 week. The mice were administrated NP-DiR [0.27 g/mL PVBu (2.4 M butyrate conc.), 10 mL/kg] by oral gavage. The mice were euthanized after 1, 3, or 4 h, and the intestine was resected and soaked in PBS for fluorescence measurement using an IVIS Lumina Imaging system (Xenogen Co., USA).

Measurement of Butyrate Concentration in Cecum. The butyrate concentration of the cecum contents was determined by gas chromatography (GC) according to the reported method.14 The 5- week-old C57BL/6J mice were fed with the AIN93G-formula diet (Oriental Yeast Co., Ltd, Japan) for 1 week. The mice were then orally administered the NPs [0.27 g/mL PVBu (2.4 M butyrate conc.), 10 mL/kg]. After 4 h, the mice were euthanized, and 100 mg of the cecum contents was disrupted and homogenized in 1 mL of water containing 5 ppm of isobutyric acid as an internal standard. The homogenates were mixed with extraction solution (mixture of 500 μL of 2.5 M hydrochloric acid and 10 mL of ether) followed by vigorous shaking and overnight storage at 4 °C. After centrifugation of the mixture, the ether layer was collected and dehydrated with anhydrous sodium sulfate. The samples were measured using a GC-18A (Shimadzu, Japan) chromatograph equipped with a DB-FFAP column (30 m × 0.32 mm × 0.25 μm) and a flame ionization detector. The

GC experiment was run under the following conditions: carrier gas, helium at a flow rate of 0.95 mL/min (50 kPa); split ratio, 1:20; temperature program (40−240 °C, 5 °C/min); and detector temperature 240 °C.
DSS-Induced Colitis Model Study. The 5-week-old C57BL/6J male mice were fed a low-fiber diet (AIN93G-formula diet). Five mice in each group received drinking water containing NPs (94 or 190 mM butyrate conc.), NP-D3 NPs (94 mM butyrate conc., 5.6 ng/mL vitamin D3 conc.), or SB (190 mM) for 10 days. On day 3, the mice were treated with 2% wt/v DSS in drinking water for 7 days, until day
9. On the last day, the mice received water without DSS. Body weights and clinical scores based on the severity of diarrhea and colonic hemorrhage30 were monitored daily. After euthanizing the mice on day 10, the colon was collected to measure its length.

■Statistical Data Analyses. Statistical data analyses were performed using analysis of variance followed by Dunnett’s test. A p value of <0.05 is the minimal level of statistical significance.

RESULTS AND DISCUSSION
NP Preparation. PVBu was prepared by radical polymer- ization using AIBN as an initiator. PVBu was purified by dialysis and obtained as a viscous liquid. 1H NMR of the obtained PVBu showed the successful removal of impurities (Figure S1). Mw and Mw/Mn were determined by gel permeation chromatography (GPC) analysis to be 1.6 × 104 and 1.8, respectively (Figure S2). PVBu NPs were prepared by the O/W emulsion method using Pluronic F-127 as a stabilizer, followed by evaporation of the organic solvent by stirring overnight. To incorporate vitamin D3 (dihydroxycho- lecalciferol; 1,25(OH)2D3) as a hydrophobic drug or fluorescent molecules for labeling, these compounds were dissolved in the organic phase for emulsification. The entrapment efficiency of vitamin D3 in the NPs (NP-D3) was determined to be 65.6 ± 4.0%. The characteristics of the NPs are summarized in Table 1. The NP sizes were tuned to

Table 1. NP Propertiesa

name size (nm) PDI ζ-potential (mv)
NP 193 ± 1.6 0.20 ± 0.01 −0.6 ± 0.14
NP-D3 203 ± 2.1 0.22 ± 0.01 −0.7 ± 0.32
NP-DiO 190 ± 0.2 0.21 ± 0.01 −0.5 ± 0.03
NP-DiR 191 ± 0.6 0.21 ± 0.01 −0.6 ± 0.09
aData are given as mean ± SD (n = 3).
around 200 nm, which is suitable for accumulation in inflammatory lesions in the intestine.25,31−33 The surface charge was almost neutral owing to the surface coating of Pluronic F-127. Because the glass transition temperature of PVBu is below the room temperature, the NPs were found to fuse with each other upon freeze drying. The stability of the aqueous dispersion of NPs during storage at 4 °C was therefore examined (Figure 2). The size and PDI of the NPs were almost constant for at least 56 days, indicating the high stability of the NP dispersion.
When the NPs are administered orally, there are several pathways for butyrate release: pancreatic enzymes in the lumen of the upper intestine, intestinal bacterial enzymes, and intracellular enzymes of intestinal cells. We examined the release of butyrate from NPs using pancreatic lipase, which is one of the release pathways. As shown in Figure 3, butyrate release from the NPs was much slower than that from the VBu monomer and rapidly saturated at a lower butyrate release. This release profile is thought to be due to the limited access of Size and PDI of NPs during storage as an aqueous dispersion at 4 °C. 0.27 mg/mL NPs in 10 mM HEPES buffer (pH 7.4).

Release of butyrate from vinyl butyrate (VBu) and NPs by pancreatic lipase at 37 °C. 100 mg of lipase was suspended in 40 mL of 2.5 mM Tris−HCl buffer (pH 7.4) containing NPs (34 mM butyrate conc.) or VBu (34 mM). the lipase to the hydrophobic PVBu chains in aqueous media. The resistance of the NPs to lipase indicates that orally administered NPs may provide protection from the rapid hydrolysis and consumption of butyrate in the upper intestine. Cellular Uptake and Anti-Inflammatory Effect of NPs in Vitro. Macrophages play an important role in the progression of the inflammation in colitis.8 First, we assessed the cellular uptake of the NPs by the macrophage cell line RAW264.7. As shown in Figure 4A, NP-DiO was actively taken up by macrophages. Then, the anti-inflammatory effect of the NPs was examined in terms of suppression of NF-κB signaling
in the macrophages induced by LPS. LPS is a ligand of toll-like receptor 4, which induces an inflammatory response mainly through the NF-κB pathway.34 Here, we used RAW264.7 transfected with SEAP, which secret SEAP as a reporter gene via NF-κB signaling. As shown in Figure 4B, low doses of SB (≤1 mM) had no effect on NF-κB signaling in LPS-stimulated macrophages, whereas 5 mM SB activated NF-κB signaling. Activation of the inflammatory response of macrophages by high concentrations of SB has been reported previously.35 In contrast, the NPs showed significant suppression of NF-κB signaling at high doses (5 and 20 mM).

The different effects of the NPs and SB on the NF-κB signaling are thought to result from the difference in their mode of action on the macrophages. Free butyrate binds both cell surface butyrate receptors and intracellular HDAc following uptake via endocytosis or transporters.8 In contrast, NPs do not stimulate cell surface butyrate receptors but slowly release butyrate when subjected to intracellular esterase or lipases after endocytotic uptake to inhibit HDAc. The selectivity of NPs for HDAc over the cell surface butyrate receptors may be the reason for their suppressive effect on NF-κB signaling. This is an ideal characteristic for a butyrate donor intended for treating intestinal inflammation by targeting inflammatory cells such as macrophages.

Cellular uptake and anti-inflammatory effect of NPs in RAW264.7. (A) Cellular uptake of NP-DiO by RAW264.7 after incubation for 24
Imageh. (B) Suppressive effect of SB and NPs on the LPS-induced NF-κB activation in RAW264.7. 6 h after the addition of SB or NPs, LPS (20 ng/mL) was added, and the samples were incubated for a further 18 h. The SEAP level in the supernatant was measured using a substrate. Data are given as mean ± SD (n = 3). ***p < 0.001 vs control; ###p < 0.001 and ns denotes not significant vs non.

NP retention in the GI tract and butyrate concentration in cecum. (A) Following oral gavage administration of NP-DiR [0.27 g/mL PVBu (2.4 M butyrate conc.), 10 mL/kg body weight], intestines were collected from euthanized mice at each time point and observed by IVIS.
(B) Butyrate concentration in the cecum contents 4 h after oral gavage of NPs (same dose as panel A). Data are given as mean ± SD (n = 5). ns denotes not significant.

Retention of NPs in the GI Tract and Butyrate Concentration in the Cecum. We examined the retention of NP-DiR in the GI tract after oral gavage administration in mice. As shown in Figure 5A, at 1 h, most of the DiR fluorescence was observed in the stomach and intestine; then at 3 h, the fluorescence had moved to the lower part of the intestine, cecum, and large intestine. At 4 h, most of the fluorescence could no longer be detected. The retention time of NP-DiR in the GI tract is similar to those of particles with a similar diameter.36,37
We measured the increment of butyrate concentration in theGI tract by oral administration of NPs. We followed a reported protocol in which the increment of butyrate concentration was detected for butyrate-modified starch fed mice. Mice were fed a low-fiber diet to minimize diet-originating butyrate.14 4 h after oral gavage administration of NPs at a relatively high dose (∼0.5 mmol/mouse), the amount of free butyrate in the cecum was measured.

As shown in Figure 5B, the NPs showed no significant difference in butyrate concentration compared with the nontreated group. Compared with butyrate-modified starch, the main chain of PVBu is not biodegradable, which may be the reason for the slow release of butyrate in the GI
tact. It is notable that the NPs did not increase the free butyrate concentration in the cecum even at such a high dose. NP-Ameliorated DSS-Induced Colitis. We assessed the therapeutic effect of NPs on a colitis mouse model induced by DSS. The treatment scheme is shown in Figure 6A. NPs or SB was applied in drinking water, and two butyrate concentrations were examined; low (94 mM) and high (190 mM). DSS was added to the drinking water from day 3 to induce colitis. During the experiment, the daily intakes of water and feed were almost constant among all of the treatment groups (Figure S3). Figure 6B shows a representative colon of each group after 10 days of treatment, and Figure 6C summarizes the average colon length of five mice for each group, which reflects the severity of the colon inflammation. The colon length of healthy mice was 69.6 ± 4.0 mm. The NPs suppressed the shortening of the colon length more than SB at the same high dose, showing the advantage of the NPs in enhancing the efficacy of butyrate delivery to the lower
intestine.

It is notable that the NPs did not raise the butyrate concentration in the cecum (Figure 5B) but showed the suppression of the inflammatory response both in vitro (Figure 4B) and in vivo (Figure 6). These results showed that the NPs Therapeutic effect of NPs on a colitis mouse model induced by DSS. (A) Mouse treatment scheme. The application of samples (NP, NP-D3, SB) as drinking water was commenced 2 days before the application of 2% DSS. The three samples in drinking water contained the same concentration of butyrate (94 and 190 mM for low and high dose, respectively). Representative colon image (B) and average colon length (C) of mice from each treatment group following treatment for 10 days. Changes in body weight (D) and clinical score (E) during the treatment. Data are given as mean ± SD (n = 5). *p < 0.05, **p < 0.01, ***p < 0.001 vs the vehicle control.
††p < 0.01 for comparison as indicated. have moderate tolerance to hydrolysis enzymes in the intestinal lumen, while they provided enough intracellular concentration of butyrate in intestinal cells to induce anti-inflammatory effect. Because a rapid increase of luminal concentration of butyrate induced toxicity,23,24 these characteristics of the NPs would be ideal as the butyrate donor.

The NPs containing vitamin D3 (NP-D3) showed a superior effect to that of the NPs when compared at the same low dose, indicating the synergetic effect of the NPs and vitamin D3. The reduction in body weight due to the progression of DSS was suppressed at a high NP dose and a low NP-D3 dose (Figure 6D). The superior effects for these two treatment groups were also observed in the clinical scores evaluated based on the severity of diarrhea and colonic hemorrhage (Figure 6E). Vitamin D3 is the ligand for the nuclear receptor (vitamin D3 receptor), which regulates the expression of various genes in many cell types.38 Recent experiments based on in vitro and IBD model animal experiments have revealed that vitamin D3 contributes to suppressing the progression of IBD through multiple pathways including the suppression of activation of immune cells.39 However, without using an effective carrier to the lower part of the intestine, oral application of free vitamin D3 shows no therapeutic effect on DSS-induced colitis mice.29

Thus, the synergy of NP-D3 showed that NPs work not only as a donor of butyrate but also as an effective carrier for vitamin D3 delivery to the lower intestine.In this study, we examined PVBu NPs as a butyrate donor for delivery to the lower part of the intestine for the treatment of colitis induced by DSS. The NPs showed a superior suppressive effect on the inflammatory response of macro- phages in vitro compared with SB. The superior effect of the NPs is attributed to their selectivity in working more as an HDAc inhibitor than as a ligand for cell surface receptors of butyrate. Although oral administration of the NPs did not increase the luminal concentration of butyrate in the cecum, the NPs showed a therapeutic effect on the colitis model mice, indicating that the NPs may target inflammatory cells such as resident macrophages by increasing the intracellular concen- tration of butyrate to suppress their activation. These characteristics of PVBu NPs make them suitable for avoiding the adverse effects associated with high concentrations of butyrate. We also demonstrated that PVBu NPs are an effective carrier for vitamin D3 which showed a synergistic effect on the therapeutic effect of butyrate. In conclusion, PVBu NPs are a promising therapeutic for IBD treatment, having both a suitable butyrate release profile and the ability to accommodate hydrophobic drugs such as vitamin D3.

⦁ CONCLUSIONS

■ASSOCIATED CONTENT
*sı Supporting Information
The Supporting Information is available free of charge at

https://pubs.acs.org/doi/10.1021/acsabm.0c01105.

1H NMR spectrum of PVBu; GPC profile of PVBu; and daily intake of diet and water (PDF)

⦁ AUTHOR INFORMATION

Corresponding Authors
Takeshi Mori − Graduate School of Systems Life Sciences, Department of Applied Chemistry, Faculty of Engineering, and Center for Future Chemistry, Kyushu University, Fukuoka 819-0395, Japan; orcid.org/0000-0002-1821- 5427; Email: [email protected]
Koji Hase − Division of Biochemistry, Faculty of Pharmacy, Keio University, Tokyo 105-8512, Japan; Division of Mucosal Barrierology, International
Research and Development Center for Mucosal Vaccines, The Institute of Medical Science, The University of Tokyo, Tokyo 108-8639, Japan; Email: [email protected]

Yoshiki Katayama − Graduate School of Systems Life Sciences, Department of Applied Chemistry, Faculty of Engineering,
ImageCenter for Future Chemistry, and International Research Center for Molecular System, Kyushu University, Fukuoka 819-0395, Japan; Department of Biomedical Engineering, Chung Yuan Christian University, Taoyuan 32023, Taiwan; orcid.org/0000-0002-4957-6241; Email: ykatatcm@
mail.cstm.kyushu-u.ac.jp

Authors
Yunmei Mu − Graduate School of Systems Life Sciences, Kyushu University, Fukuoka 819-0395, Japan
Yusuke Kinashi − Division of Biochemistry, Faculty of Pharmacy, Keio University, Tokyo 105-8512, Japan
Jinting Li − Graduate School of Systems Life Sciences, Kyushu University, Fukuoka 819-0395, Japan
Takuma Yoshikawa − Department of Applied Chemistry, Faculty of Engineering, Kyushu University, Fukuoka 819- 0395, Japan

Akihiro Kishimura − Graduate School of Systems Life Sciences, Department of Applied Chemistry, Faculty of Engineering, Center for Future Chemistry, and International Research Center for Molecular System, Kyushu University, Fukuoka 819-0395, Japan; orcid.org/0000-0002-0503- 1418
Mitsuru Tanaka − Research and Development Center for Five- Sense Devices, Kyushu University, Fukuoka 819-0395, Japan
ImageToshiro Matsui − Department of Bioscience and Biotechnology, Faculty of Agriculture, Graduate School of Kyushu University, Fukuoka 819-0395, Japan; orcid.org/ 0000-0002-9137-8417
Complete contact information is available at: https://pubs.acs.org/10.1021/acsabm.0c01105

Notes
The authors declare no competing financial interest.

⦁ ACKNOWLEDGMENTS
We thank Dr. Masaru Tanaka and Dr. Shingo Kobayashi for
the GPC analysis of PVBu. We thank Dr. Yukio Nagasaki for critical suggestions for animal experiment protocol.

⦁ REFERENCES
(1) Fiocchi, C. (1) Inflammatory Bowel Disease: Etiology and
Pathogenesis. Gastroenterology 1998, 115, 182−205.
(2) Cader, M. Z.; Kaser, A. Recent advances in inflammatory bowel
disease: Mucosal immune cells in intestinal inflammation. Gut 2013,
62, 1653−1664.
(3) Lavelle, A.; Sokol, H. Gut microbiota-derived metabolites as key
actors in inflammatory bowel disease. Nat. Rev. Gastroenterol. Hepatol.
2020, 17, 223−237.
(4) Neurath, M. F. Current and emerging therapeutic targets for Sodium butyrate IBD. Nat. Rev. Gastroenterol. Hepatol. 2017, 14, 269−278.
(5) Gill, P. A.; van Zelm, M. C.; Muir, J. G.; Gibson, P. R. Review
article: short chain fatty acids as potential therapeutic agents in human gastrointestinal and inflammatory disorders. Aliment. Pharmacol. Ther. 2018, 48, 15−34.
(6) Venegas, D. P.; De la fuente, MK.; Landskron, G.; Gonzaĺez, M.
J.; Quera, R.; Dijkstra, G.; Harmsen, H.; Faber, K.; Hermoso, M. A. Short chain fatty acids (SCFAs)-mediated gut epithelial and immune regulation and its relevance for inflammatory bowel diseases. Front. Immunol. 2019, 10, 277.
(7) Vernia, P.; Caprilli, R.; Latella, G.; Barbetti, F.; Magliocca, F. M.; Cittadini, M. Fecal Lactate and Ulcerative Colitis. Gastroenterology 1988, 95, 1564−1568.
(8) Chang, P. V.; Hao, L.; Offermanns, S.; Medzhitov, R. The
microbial metabolite butyrate regulates intestinal macrophage function via histone deacetylase inhibition. Proc. Natl. Acad. Sci. U.S.A. 2014, 111, 2247−2252.
(9) Flint, H. J.; Scott, K. P.; Louis, P.; Duncan, S. H. The role of the
gut microbiota in nutrition and health. Nat. Rev. Gastroenterol. Hepatol. 2012, 9, 577−589.
(10) Macia, L.; Tan, J.; Vieira, A. T.; Leach, K.; Stanley, D.; Luong,
S.; Maruya, M.; Ian McKenzie, C.; Hijikata, A.; Wong, C.; Binge, L.; Thorburn, A. N.; Chevalier, N.; Ang, C.; Marino, E.; Robert, R.; Offermanns, S.; Teixeira, M. M.; Moore, R. J.; Flavell, R. A.; Fagarasan, S.; Mackay, C. R. Metabolite-sensing receptors GPR43 and GPR109A facilitate dietary fibre-induced gut homeostasis through regulation of the inflammasome. Nat. Commun. 2015, 6, 7734.
(11) Voltolini, C.; Battersby, S.; Etherington, S. L.; Petraglia, F.; Norman, J. E.; Jabbour, H. N. A novel antiinflammatory role for the short-chain fatty acids in human labor. Endocrinology 2012, 153, 395− 403.
(12) Liu, H.; Wang, J.; He, T.; Becker, S.; Zhang, G.; Li, D.; Ma, X. Butyrate: A Double-Edged Sword for Health? Adv. Nutr. 2018, 9, 21− 29.
(13) Julian, R. M.; Holmes, E.; Fatima, K.; Sunil, K.; Pauline, S.; Fergus, S.; Wilson, I. D.; Yulan, W. Rapid and Noninvasive Metabonomic Characterization of Inflammatory Bowel Disease. J. Proteome Res. 2007, 6, 546−551.
(14) Furusawa, Y.; Obata, Y.; Fukuda, S.; Endo, T. A.; Nakato, G.;
Takahashi, D.; Nakanishi, Y.; Uetake, C.; Kato, K.; Kato, T.; Takahashi, M.; Fukuda, N. N.; Murakami, S.; Miyauchi, E.; Hino, S.; Atarashi, K.; Onawa, S.; Fujimura, Y.; Lockett, T.; Clarke, J. M.; Topping, D. L.; Tomita, M.; Hori, S.; Ohara, O.; Morita, T.; Koseki, H.; Kikuchi, J.; Honda, K.; Hase, K.; Ohno, H. Commensal microbe- derived butyrate induces the differentiation of colonic regulatory T cells. Nature 2013, 504, 446−450.
(15) Lucas, S.; Omata, Y.; Hofmann, J.; Böttcher, M.; Iljazovic, A.;
Sarter, K.; Albrecht, O.; Schulz, O.; Krishnacoumar, B.; Krönke, G.; Herrmann, M.; Mougiakakos, D.; Strowig, T.; Schett, G.; Zaiss, M. M.
Short-chain fatty acids regulate systemic bone mass and protect from pathological bone loss. Nat. Commun. 2018, 9, 55.
(16) Isobe, J.; Maeda, S.; Obata, Y.; Iizuka, K.; Nakamura, Y.; Fujimura, Y.; Kimizuka, T.; Hattori, K.; Kim, Y.-G.; Morita, T.; Kimura, I.; Offermanns, S.; Adachi, T.; Nakao, A.; Kiyono, H.; Takahashi, D.; Hase, K. Commensal-bacteria-derived butyrate promotes the T-cell-independent IgA response in the colon. Int. Immunol. 2020, 32, 243−258.
(17) Annison, G.; Illman, R. J.; Topping, D. L. Acetylated,
propionylated or butyrylated starches raise large bowel short-chain fatty acids preferentially when fed to rats. J. Nutr. 2003, 133, 3523− 3528.
(18) Hamer, H. M.; Jonkers, D. M. A. E.; Bast, A.; Vanhoutvin, S. A.
L. W.; Fischer, M. A. J. G.; Kodde, A.; Troost, F. J.; Venema, K.; Brummer, R.-J. M. Butyrate modulates oxidative stress in the colonic mucosa of healthy humans. Clin. Nutr. 2009, 28, 88−93.
(19) Bouter, K.; Bakker, G. J.; Levin, E.; Hartstra, A. V.; Kootte, R.
S.; Udayappan, S. D.; Katiraei, S.; Bahler, L.; Gilijamse, P. W.; Tremaroli, V.; Stahlman, M.; Holleman, F.; van Riel, N.; Verberne, H. J.; Romijn, J. A.; Dallinga-Thie, G. M.; Serlie, M. J.; Ackermans, M. T.; Kemper, E. M.; Willems van Dijk, K.; Backhed, F.; Groen, A. K.; Nieuwdorp, M. Differential metabolic effects of oral butyrate treatment in lean versus metabolic syndrome subjects article. Clin. Transl. Gastroenterol. 2018, 9, 155.
(20) Hamer, H. M.; Jonkers, D. M. A. E.; Vanhoutvin, S. A. L. W.; Troost, F. J.; Rijkers, G.; de Bruïne, A.; Bast, A.; Venema, K.; Brummer, R.-J. M. Effect of butyrate enemas on inflammation and antioxidant status in the colonic mucosa of patients with ulcerative colitis in remission. Clin. Nutr. 2010, 29, 738−744.
(21) Cleophas, M. C. P.; Ratter, J. M.; Bekkering, S.; Quintin, J.;
Schraa, K.; Stroes, E. S.; Netea, M.; Joosten, L. A. B. Effects of oral butyrate supplementation on inflammatory potential of circulating peripheral blood mononuclear cells in healthy and obese males. Sci. Rep. 2019, 9, 775.
(22) Raqib, R.; Sarker, P.; Mily, A.; Alam, N. H.; Arifuzzaman, A. S. M.; Rekha, R. S.; Andersson, J.; Gudmundsson, G. H.; Cravioto, A.; Agerberth, B. Efficacy of sodium butyrate adjunct therapy in shigellosis: a randomized, double-blind, placebo-controlled clinical trial. BMC Infect. Dis. 2012, 12, 111.
(23) Kaiko, G. E.; Ryu, S. H.; Koues, O. I.; Collins, P. L.; Solnica- Krezel, L.; Pearce, E. J.; Pearce, E. L.; Oltz, E. M.; Stappenbeck, T. S. The Colonic Crypt Protects Stem Cells from Microbiota-Derived Metabolites. Cell 2016, 165, 1708−1720.
(24) Kespohl, M.; Vachharajani, N.; Luu, M.; Harb, H.; Pautz, S.;
Wolff, S.; Sillner, N.; Walker, A.; Schmitt-Kopplin, P.; Boettger, T.; Renz, H.; Offermanns, S.; Steinhoff, U.; Visekruna, A. The microbial metabolite butyrate induces expression of Th1- associated factors in cD4+ T cells. Front. Immunol. 2017, 8, 1036.
(25) Hua, S.; Marks, E.; Schneider, J. J.; Keely, S. Advances in oral nano-delivery systems for colon targeted drug delivery in inflamma-

tory bowel disease: Selective targeting to diseased versus healthy tissue. Nanomed. Nanotechnol. Biol. Med. 2015, 11, 1117−1132.
(26) Dacoba, T. G.; Olivera, A.; Torres, D.; Crecente-Campo, J.;
Alonso, M. J. Modulating the immune system through nano- technology. Semin. Immunol. 2017, 34, 78−102.
(27) Mu, Y.; Kinashi, Y.; Kishimura, A.; Mori, T.; Hase, K.;
Katayama, Y. Effect of polyvinyl butyrate nanoparticles incorporated with immune suppressing vitamins on alteration of population of intestinal immune cells. Prog. Nat. Sci. 2020, 30, 707−710.
(28) Chahinian, H.; Vanot, G.; Ibrik, A.; Rugani, N.; Sarda, L.;
Comeau, L.-C. Production of
Extracellular
Lipases byPenicillium
cyclopiumPurification and Characterization of a Partial Acylglycerol Lipase. Biosci. Biotechnol. Biochem. 2000, 64, 215−222.
(29) Zai, K.; Hirota, M.; Yamada, T.; Ishihara, N.; Mori, T.;
Kishimura, A.; Suzuki, K.; Hase, K.; Katayama, Y. Therapeutic effect of vitamin D3-containing nanostructured lipid carriers on inflamma- tory bowel disease. J. Controlled Release 2018, 286, 94−102.
(30) Chassaing, B.; Aitken, J. D.; Malleshappa, M.; Vijay-Kumar, M.
Dextran sulfate sodium (DSS)-induced colitis in mice. Curr. Protoc. Immunol. 2014, 104, 15.25.1−15.25.14.
(31) Lamprecht, A.; Schaf̈er, U.; Lehr, C. M. Size-dependent
bioadhesion of micro- and nanoparticulate carriers to the inflamed colonic mucosa. Pharm. Res. 2001, 18, 788−793.
(32) Bekerman, T.; Golenser, J.; Domb, A. Cyclosporin Nano-
particulate Lipospheres for Oral Administration. J. Pharm. Sci. 2004,
93, 1264−1270.
(33) Collnot, E.-M.; Ali, H.; Lehr, C.-M. Nano- and microparticulate drug carriers for targeting of the inflamed intestinal mucosa. J. Controlled Release 2012, 161, 235−246.
(34) Palsson-McDermott, E. M.; O’Neill, L. A. J. Signal transduction
by the lipopolysaccharide receptor, Toll-like receptor-4. Immunology
2004, 113, 153−162.
(35) Foey, A. D. Butyrate regulation of distinct macrophage subsets: Opposing effects on M1 and M2 macrophages. Int. J. Probiotics Prebiotics 2011, 6, 147−158.
(36) Hu, X.; Fan, W.; Yu, Z.; Lu, Y.; Qi, J.; Zhang, J.; Dong, X.;
Zhao, W.; Wu, W. Evidence does not support absorption of intact solid lipid nanoparticles via oral delivery. Nanoscale 2016, 8, 7024− 7035.
(37) Li, D.; Zhuang, J.; He, H.; Jiang, S.; Banerjee, A.; Lu, Y.; Wu, W.; Mitragotri, S.; Gan, L.; Qi, J. Influence of Particle Geometry on Gastrointestinal Transit and Absorption following Oral Adminis- tration. ACS Appl. Mater. Interfaces 2017, 9, 42492−42502.
(38) Bookout, A. L.; Jeong, Y.; Downes, M.; Yu, R. T.; Evans, R. M.;
Mangelsdorf, D. J. Anatomical Profiling of Nuclear Receptor
Expression Reveals a Hierarchical Transcriptional Network. Cell
2006, 126, 789−799.
(39) Fletcher, J.; Cooper, S. C.; Ghosh, S.; Hewison, M. The role of vitamin D in inflammatory bowel disease: Mechanism to manage- ment. Nutrients 2019, 11, 1019.