Peristalsis is a nuanced mechanical stimulus comprised of multi-axial strain (radial and axial strain) and shear stress. Forces associated with peristalsis regulate diverse biological functions including digestion, reproductive function, and urine dynamics. Given the central role peristalsis plays in physiology and pathophysiology, we were motivated to design a bioreactor capable of holistically mimicking peristalsis. We engineered a novel rotating screw-drive based design combined with a peristaltic pump, in order to deliver multi-axial strain and concurrent shear stress to a biocompatible polydimethylsiloxane (PDMS) membrane “wall.” Radial indentation and rotation of the screw drive against the wall demonstrated multi-axial strain evaluated via finite element modeling. Experimental measurements of strain using piezoelectric strain resistors were in close alignment with model-predicted values (15.9 ± 4.2% vs. 15.2% predicted). Modeling of shear stress on the “wall” indicated a uniform velocity profile and a moderate shear stress of 0.4 Pa. Human mesenchymal stem cells (hMSCs) seeded on the PDMS “wall” and stimulated with peristalsis demonstrated dramatic changes in actin filament alignment, proliferation, and nuclear morphology compared to static controls, perfusion, or strain, indicating that hMSCs sensed and responded to peristalsis uniquely. Lastly, significant differences were observed in gene expression patterns of calponin, caldesmon, smooth muscle actin, and transgelin, corroborating the propensity of hMSCs toward myogenic differentiation in response to peristalsis. Collectively, our data suggest that the peristalsis bioreactor is capable of generating concurrent multi-axial strain and shear stress on a “wall.” hMSCs experience peristalsis differently than perfusion or strain, resulting in changes in proliferation, actin fiber organization, smooth muscle actin expression, and genetic markers of differentiation. The peristalsis bioreactor device has broad utility in the study of development and disease in several organ systems.

The involuntary contraction and relaxation of smooth muscle layers causes peristalsis, or wave-like movements through various organs, including the gastrointestinal tract, uterus, and ureters [Huizinga, 1999; Kunz and Leyendecker, 2002; Hosseini et al., 2018]. Forces associated with peristalsis regulate diverse biological functions including nutrient absorption, transport during reproductive processes, and urine dynamics. Peristalsis plays a central role in physiology and consists of two main mechanical components: multi-axial strain, and fluid shear (online suppl. Fig. 1; for all online suppl. material, see www.karger.com/doi/10.1159/000521752) [Yokoyama and Ozaki, 1990; Gayer and Basson, 2009]. Multi-axial strain is composed of radial and axial strain; in the gastrointestinal system and ureters, strain is induced by the contractions of the circular and longitudinal muscle layers [Mittal, 2011; Vahidi and Fatouraee, 2012]. In the uterus, only the circular muscles are responsible for uterine contractions [Kunz and Leyendecker, 2002]. In most organ systems, peristalsis is meant to aid in the transport (propulsion) of material through the organ. Flow of fluid or luminal content that accompanies this propulsion creates the induced shear [Fung and Yih, 1968; Aranda et al., 2015; Waldrop and Miller, 2016]. Within the organs that experience peristalsis, the cells are therefore continuously exposed to both multi-axial strain and shear stress. These mechanical signals are integrated by the cells into biochemical responses, through mechano-transduction [Cei et al., 2016]. Aberrant peristalsis, or dysperistalsis, is a patho-physiological condition associated with functional gastrointestinal disorders and endometriosis [Leyendecker et al., 1996; Iwakiri et al., 2017; Patel and Thavamani, 2021]. Given the central role of peristalsis in homeostasis, and the role aberrant peristalsis plays in the development and sustenance of patho-physiology, it is advantageous to include peristalsis and its associated kinematics in in vitro studies of disease and development.

The kinematics of peristalsis is complex, and hence challenging to replicate holistically in vitro without the use of bioreactors or microfluidic approaches. Currently, there are several examples of peristalsis bioreactors associated with the gastrointestinal tract at the microfluidic and the macro bioreactor levels [Kim et al., 2012; Cei et al., 2016; Costello et al., 2017; Wang et al., 2018; Zhou et al., 2018]. However, several of these approaches to study peristalsis, along with ureter and uterine models, often simplify the kinematics of peristalsis into perfusion-based fluid shear alone, or cyclic strain or cyclic uniaxial strain alone [Basu et al., 2012; da Rocha and Smith, 2012; Wang et al., 2014; Vardar et al., 2015; Hellstrom et al., 2017; Elad et al., 2020a; Elad et al., 2020b; Kim et al., 2020]. Neither of these two simplified stimuli adequately capture the concurrent multi-axial strain and shear of peristalsis. Appropriately mimicking both forces concurrently is important, because several cell types respond differently to shear stress and strain [Jufri et al., 2015; Ostrowski et al., 2016; Meza et al., 2019].

The ability to apply concurrent multi-axial strain and shear stress was therefore a central motivator to our new peristalsis bioreactor design. Our approach circumvents the limitations of current systems that either rely heavily on perfusion or uniaxial tension. Here we present a novel bioreactor design where a biocompatible membrane experiences cyclic multi-axial strain produced by a rotating screw drive, while simultaneously experiencing pulsatile fluid flow-induced shear stress. We integrated predictive finite element modeling and rapid prototyping to generate reproducible peristalsis bioreactors. We validated the cellular effect from the mechanical forces of our system through the use of human mesenchymal stem cells (hMSCs) due to their inherently mechanosensitive nature [Hoey et al., 2012; Kim et al., 2015; Kuo et al., 2015; Liu et al., 2015]. Overall, the objective of this work was to develop a bioreactor capable of mimicking the kinematics of peristalsis and elucidate changes in the cellular response to this unique mechanical event (Fig. 1).

Fig. 1.

Workflow of experimental process. The peristaltic bioreactor was designed using the CAD program, SolidWorks. Finite element modeling was performed to determine expected shear stress and strain within the bioreactor, for its current geometric dimensions. The bioreactor was 3D printed with a Prusa iMK3 and a biocompatible membrane “wall” was manufactured using PDMS. hMSCs were seeded on to the wall, whereupon the bioreactor was assembled and operated via a pre-programmed Arduino. Following 24 h in the bioreactor, cells were harvested for downstream analysis including microscopy and gene expression to identify cellular responses to peristalsis.

Fig. 1.

Workflow of experimental process. The peristaltic bioreactor was designed using the CAD program, SolidWorks. Finite element modeling was performed to determine expected shear stress and strain within the bioreactor, for its current geometric dimensions. The bioreactor was 3D printed with a Prusa iMK3 and a biocompatible membrane “wall” was manufactured using PDMS. hMSCs were seeded on to the wall, whereupon the bioreactor was assembled and operated via a pre-programmed Arduino. Following 24 h in the bioreactor, cells were harvested for downstream analysis including microscopy and gene expression to identify cellular responses to peristalsis.

Close modal

Materials

Cell culture reagents utilized in this study including validated bone marrow-derived hMSCs and their nutrient medium was purchased from Lonza (Walkersville, MD, USA). According to the manufacturer’s certification of analysis, hMSCs were positive for CD73, CD90, and CD105, tested via flow cytometry analysis. Polydimethylsiloxane (PDMS) was purchased from DOW Chemical (Midland, MI, USA). All other chemical reagents were purchased from Sigma Aldrich (St. Louis, MO, USA), unless otherwise indicated. Antibodies used for cellular staining were purchased from Santa Cruz Biotechnology (Dallas, TX, USA), unless otherwise indicated. Custom-made oligos were purchased from Integrated DNA Technologies (Coralville, IA, USA). All other molecular biology-grade reagents were purchased from Thermo Fisher Scientific (Waltham, MA, USA).

Design and Prototyping of Bioreactor and PDMS

Computer aided design (CAD) models of the bioreactor prototype were designed and edited in SolidWorks (Dassault Systèmes, Waltham, MA, USA). These models were used for conceptual design and computational fluid dynamics (CFD) analysis. The bioreactor parts include: (1) bioreactor top; (2) bioreactor bottom; (3) rotating screw drive; and (4) motor support. These parts are annotated in Figure 2a.

Fig. 2.

Bioreactor schematic and assembled prototype. a Schematic representing novel 3D rapid prototypable peristalsis bioreactor showing its principal components: (1) bioreactor top, (2) bioreactor bottom, (3) rotating screw drive, (4) motor support, (5) PDMS membrane, (6) DC motor, (7) cell chamber (a raised portion within the bioreactor top), (8) inlet, (9) outlet. b SolidWorks models of assembled bioreactor. Scale bars, 1 cm. c Views of fully assembled peristalsis bioreactor prototype made of polyJet material for ease of visualization. Scale bars, 1 cm.

Fig. 2.

Bioreactor schematic and assembled prototype. a Schematic representing novel 3D rapid prototypable peristalsis bioreactor showing its principal components: (1) bioreactor top, (2) bioreactor bottom, (3) rotating screw drive, (4) motor support, (5) PDMS membrane, (6) DC motor, (7) cell chamber (a raised portion within the bioreactor top), (8) inlet, (9) outlet. b SolidWorks models of assembled bioreactor. Scale bars, 1 cm. c Views of fully assembled peristalsis bioreactor prototype made of polyJet material for ease of visualization. Scale bars, 1 cm.

Close modal

The bioreactor was designed in such a way that a screw drive was housed in the bioreactor bottom, connected to an actuating DC motor that caused rotation of the screw drive. A PDMS membrane was used to create the “wall” (component 5 of the bioreactor, Fig. 2a). The PDMS membrane was placed abutting the screw drive (component 3), and clamped down by the bioreactor top (component 1). The biocompatible membrane was held in place by the pressure exerted from clamping down the bioreactor top to the bioreactor bottom. When clamped, the screw drive caused a 1.6-mm indentation in the 3-mm thick PDMS. The bioreactor top had a raised cell chamber (component 7), and an inlet (component 8) and outlet (component 9) for nutrient medium flow, enabled by a peristaltic pump (illustrated in Fig. 1). Cells were seeded on a constant area in the PDMS “wall.” The rotation of the screw drive and simultaneous peristaltic flow of nutrient medium was used to mimic the kinematics of peristalsis on the cell seeded monolayer, examined via computational modeling in the following section.

Rapid prototyping and manufacturing were executed using a Prusa iMK3 fused deposition modeling 3D printer. By utilizing CAD modeling in conjunction with 3D printing, modifications to the bioreactor design were efficiently manufactured and implemented into testing. For the final prototype, polylactic acid filament was chosen as the main substrate material as it was less prone to warping during the printing process [Pei et al., 2015]. CAD models were sliced using the PrusaSlicer software using the “quality” print condition of 0.1 mm. The models were printed at 100% infill to remove interstitial space and eliminate leakage. CAD renderings created on SolidWorks are demonstrated in Figure 2b, which resulted in a final print product, photographed in Figure 2c.

The “wall” (component 5) of the bioreactors was fabricated using PDMS membranes, manufactured (10:1 pre-polymer base to crosslinker) by pouring into custom membrane molds and degassed for 1 h. Custom membrane molds were designed to produce PDMS membranes of uniform dimensions (7 × 4 × 0.3 cm). Membranes were cured at 60°C for 4 h and allowed to cool to room temperature overnight. Once PDMS membranes were cured, they were prepared for cell seeding using protocols outlined in the following sections.

Computational Modeling and Validation of Shear Stress and Strain

The cell chamber (component 7, a part of component 1 in Fig. 2a) was designed as a CAD model on SolidWorks, as outlined in the section above. The model was then uploaded into ANSYS (Canonsburg, PA, USA) Fluent CFD software, where the CAD model was meshed and boundaries were defined as follows: (i) inlet, (ii) outlet, (iii) wall, and (iv) cell seeding area on the wall. The inlet (component 8) and outlet (component 9) represented the vertical channels from which media was pumped into and out of the cell chamber (component 7) on a constant mass-flow basis from the peristaltic pump. The “wall” was defined as all surfaces in contact with the fluid flow, excluding the wall membrane (shown in green, online suppl. Fig. 2a). The “cell seeding area” was defined as the luminal surface of the PDMS membrane “wall” that interfaced with the fluid flow, as well as the area where cells would be seeded during in vitro cell studies (shown in red, online suppl. Fig. 2a). This area remained constant across all experiments to ensure uniformity (2.79 cm2).

In order to determine the mass-flow rate in the CFD analysis, a theoretical flow rate was calculated for a fluid shear of 0.4 Pa through a rectangular cross-section. The shear of 0.4 Pa was chosen based on shear stresses experienced by various peristaltic organs, as reported in the literature [Kublickiene et al., 2000; Avvisato et al., 2007; Kimura et al., 2018]. The calculated optimized mass-flow rate to produce a minimum shear of 0.4 Pa was 28 mL/min.

The flow rate was integrated into the ANSYS software to model the flow across the “wall” itself and with the cell characteristics exhibited in the bioreactor. The cell was modeled as a spherical object with 30% embedded into the membrane wall while maintaining a perfect spherical conformation. The 30% embedding was intended to represent a more accurate cell anatomy experiencing the flow while approximating the cell-membrane interaction as rigid bonding. Using this method, the average shear over the artificial surface was assessed and quantified.

A finite element method was used to estimate the strain distribution induced in the membrane as the results of the screw indenting the membrane followed by its rotation. The simulations were performed using Abaqus/Standard Software (Abaqus, Johnston, RI, USA) implementing the implicit solution method. Deformations were considered quasi-static and inertia forces were assumed to be negligible. The screw was modeled as a rigid body. The membrane was modeled as a third-order Ogden hyperplastic material with material properties listed in online supplementary Table 1, derived from tensile testing of 10:1 PDMS membranes (online Suppl. Fig. 3).

The boundary condition was considered as follows: (i) the surface-to-surface contact interactions were applied via penalty friction formulation; (ii) the membrane geometry included a “wall” (shown in green) and the cell seeding area (shown in red; online suppl. Fig. 2a). The water seal ridge (yellow) and the outermost edge of the bioreactor were replaced by a clamped boundary condition in the simulation. The margin around the “wall” (green) was constrained by fixing the z-direction (preventing out-of-plane displacements), but remained free to have in-plane deformations (online suppl. Fig. 2b). The cell seeding area (red) was allowed to freely deform. The loading consisted of two steps: (i) the screw indenting the membrane in the z-axis for 1.6 mm (denoted by t = 0 s), and (ii) the screw rotating at 12 RPM around the x-axis while being held at the indented depth. One complete cycle, with the physical time duration of 5 s, was simulated.

Experimental Determination of Strain

Strain was experimentally measured using a custom polyvinyl alcohol-polydopamine (PVA-PDA) hydrogel strain sensor following protocols established for strain measurement [Liu et al., 2018; Wang et al., 2019]. Briefly, 10% (w/v) PVA solution, and a 160-mg/mL PDA solution were mixed and sonicated to produce a homogenous mixture. Then, 0.04 g/mL of sodium tetraborate solution was added to induce gelation of the PVA-PDA mixture, to result in a PVA-PDA hydrogel. The PVA-PDA hydrogel was then desiccated for 2 h and put through 2 cycles of a 12-h interval freeze-thaw process. This hydrogel was then inlayed into a strain sensor mold made from 3-mm thick VHB tape (3M, Maplewood, MN; USA) and copper foil placed along the X-direction. The piezoelectric, PVA-PDA strain sensor was then inserted into a modified bioreactor that would securely fit the strain sensor in the area representative of the cell seeding area on the PDMS membrane. It was also secured using cable ties to better simulate the in vitro bioreactor setup conditions. In the strain calibration set up, the screw drive was replaced with fixed, calibrated strain inducers to induce 0, 10, 26, and 41.7% strain maximally in the radial direction. Changes in electrical resistivity were measured using a digital multimeter, and a calibration curve was generated (online suppl. Fig. 4). To calculate the actual strain experienced by the screw drive, the hydrogel strain sensor was strained using the screw drive in the bioreactor and the resistivity was measured. The resulting strain was extrapolated from the calibration curve according to the measured resistivity (star, online suppl. Fig. 4).

hMSC Cell Culture

hMSCs (Lonza Bioscience) were cultured in mesenchymal stem cell growth medium supplemented with MSCGMTM SingleQuotsTM Supplement Kit (Lonza Bioscience). hMSCs were evaluated for trilineage differentiation capacity according to the manufacturer (online suppl. Fig. 5). Cells were cultured on tissue culture dishes and used up to Passage 5 as recommended by the manufacturer to avoid spontaneous differentiation. In order to maximize seeding of hMSCs on PDMS, Collagen I was used to coat the cell seeding area of the PDMS (1.8 cm2) at 200 μg/mL. Collagen-coated PDMS was stored at 4° until use. hMSCs were seeded on the PDMS at 200,000 cells/mL and incubated overnight to allow for attachment to the PDMS.

Bioreactor Assembly and Operation

All items used in bioreactor set-up were sterilized using ethanol and exposure to ultraviolet light for at least 10 min. Prior to full assembly, the remaining media from hMSC seeding on each PDMS membrane was removed and 25–30 mL of media was added to each bioreactor’s media reservoir. The bioreactor bottom was assembled by placing the screw drive, motor, and motor support in their respective locations (see Fig. 2a). Silicone grease was added to the top of the screw drive to minimize friction between the PDMS membrane and screw drive. The PDMS membrane was then transferred atop the screw drive and placed cell side up in the bioreactor bottom. The bioreactor top was then clamped down on to the membrane, and sealed using commercially available zip ties. The peristaltic pump tubing was connected to the bioreactor and media bottle to create a closed loop between the media, pump, and bioreactor. The assembled bioreactor, nutrient medium reservoir, and pump were then placed into the incubator. The motor and pump were connected to the Arduino that ran the pre-programmed code. A schematic of the full assembly is provided in Figure 1, step 4. The bioreactor ran for 24 h and cells were then collected for various downstream analyses. Three conditions were programmed with the Arduino: strain, shear, and peristalsis. The corresponding motor and pump speeds can be seen in Table 1. The differences seen between the motor speeds and flow rates across conditions were not significant. For the control condition, hMSC-seeded PDMS membranes were incubated statically for 24 h.

Table 1.

Operational parameters of the bioreactor listing pump flow rates and motor speeds

 Operational parameters of the bioreactor listing pump flow rates and motor speeds
 Operational parameters of the bioreactor listing pump flow rates and motor speeds

Immunofluorescent Staining and Fluorometry

hMSC-seeded PDMS membranes from static and mechanically stimulated conditions were cut along the cell seeding area for ease of staining. Membranes were rinsed with PBS and fixed in 4% formalin for 45 min. Samples were blocked using 0.15% Triton X and 5% fetal bovine serum at room temperature for 1 h. Cells were incubated with fluorescently tagged primary antibodies for 1 h at room temperature. The primary antibodies used included Ki67 (Alexa Fluor 647 conjugated), phalloidin (Alexa Fluor 488 conjugated), and smooth muscle actin (FITC conjugated). A nuclear counterstain (DAPI) was also included in the primary antibody incubation. Unbound antibodies were rinsed using PBS, and cell-seeded PDMS membranes were mounted using an antifade mounting reagent. Fluorescence was observed using a Leica SP8 Confocal Microscope, with 5 independent, non-overlapping regions for analysis. NIH ImageJ was utilized to perform fluorometry.

Ki67+ cells were identified by the presence of red fluorescence within DAPI counterstained nuclei, and were quantified manually. Analysis was performed to determine the number of actively proliferating cells (Ki67 antigen in the nuclei) versus the total cell number (number of nuclei) to determine the percentage of proliferating cells.

hMSCs were stained with fluorescently conjugated phalloidin to visualize cytoskeletal networks and actin alignment at low and high magnifications. To analyze actin filament angle orientation, green channels from the higher magnification images were isolated in ImageJ. A single-pixel-line filter was applied at sequential rotational intervals (0°, 1°, 2°, etc.) and the brightness of the remaining pixels were summed [Püspöki et al., 2016]. Then, the summed brightness at each angle was computed into a fraction. The totality of the 180 fractions equaled 1. Angles were then grouped into increments of 30° by adding the fractions from 1° to 30°, 30° to 60°, etc. to assess the angle distribution trends.

All high-magnification actin filament images were also assessed for nuclei circularity using the DAPI channel. The blue channel was isolated in ImageJ to remove any Phalloidin stains from interfering with the circularity measurement. A filter was applied to the selected DAPI image to mask out everything except the cell nuclei. The circularity of each nucleus is measured by Eq. 1.

Circularity = 4 * π * (area/perimeter2)

A perfectly circular nuclei would result in a value of 1, and an elongated nuclei would return a value approaching zero.

Using fluorescently conjugated smooth muscle actin, the alpha isotype of smooth muscle was visualized in hMSCs exposed to the bioreactor. To analyze the presence of smooth muscle actin, the amount of green fluorescence was compared to the total blue fluorescence, normalized to a constant imaging area. This resulted in a normalized value for the relative expression of smooth muscle actin for each condition.

Trilineage Differentiation Staining

Alizarin Red S, Alcian Blue, and Oil Red O stains were used to visualize Osteocyte, Chondrocyte, and Adipocyte differentiation, respectively, following well-established protocols [Eggerschwiler et al., 2019]. Membranes were rinsed with PBS and fixed in 4% formalin for 15 min. Alizarin Red S and Alcian Blue were rinsed with DI water and incubated at 37°C for 45 min and at room temperature for 15 min, respectively. Oil Red O was rinsed with PBS and incubated at room temperature for 20 min. Unbound stain was rinsed using DI water or PBS, and cell-seeded PDMS membranes were mounted using an antifade mounting reagent. All stains were observed using a Zeiss Axio Observer D1 (Carl Zeiss, Jena, Germany) inverted microscope with Achroplan 10×/0.32 objective. Images were captured using a Tucsen IS500 camera.

RNA Extraction and qPCR

RNA extraction was performed on harvested hMSCs using the RNeasy Mini Kit (Qiagen). RNA concentration and purity were evaluated using a NanoDrop OneC (ThermoFisher Scientific), and stored at −80°C until ready to use. Reverse transcription was performed following manufacturer’s protocols using the High-Capacity cDNA Reverse Transcription Kit (Applied Biosystems). qPCR was performed with a CFX96 Real-Time System (Bio-Rad) using the Applied Biosystems PowerSYBR Green PCR Mastermix (Thermofisher Scientific) for detection. Genes that were investigated include CNN1 (calponin), CALD1 (caldesmon), ACTA2 (smooth muscle actin), SM22α (transgelin), BGLAP (osteocalcin), ALPL (alkaline phosphatase), SPP1 (osteopontin), ACAN (aggrecan I), COL2A1 (collagen II), SOX9 (SRY-box transcription factor 9), COMP (cartilage oligomeric matrix protein), ADIPOQ (adiponectin), and LEP (leptin). The primer sequences used for each gene are shown in Table 2. Genes are categorized as follows: myocyte: CNN1, CALD1, ACTA2, and SM22α; osteocyte: BGLAP, ALPL, and SPP1; chondrocyte: ACAN, COL2A1, SOX9, and COMP; and adipocyte: ADIPOQ and LEP. Changes in gene expression were calculated using the 2ΔΔCt method, with GAPDH as the housekeeping control [Livak and Schmittgen, 2001]. qPCR experiments were run in triplicates, with 3 independent biological replicates. Comparisons were drawn between control hMSCS that were not exposed to mechanical stimulus and each of the three applied stimuli.

Table 2.

List of genes and corresponding forward and reverse primer sequences used in qPCR amplification

 List of genes and corresponding forward and reverse primer sequences used in qPCR amplification
 List of genes and corresponding forward and reverse primer sequences used in qPCR amplification

Statistical Analysis

Statistical analysis was performed on GraphPad Prism 9. All reported values are means ± SEM, resulting from 3 to 5 biological replicates. All qPCR data were normalized to control conditions within each experimental set and performed in triplicates over at least 3 biological replicates. Image analysis and morphometry included 5 non-overlapping fields of view from 3 to 5 biological replicates. ANOVA-based hypothesis testing was performed where appropriate, and statistical significance is indicated within each experimental data set, with associated p values.

Shear Stress Computational Modeling Validation

The design of the peristalsis bioreactor as envisioned, along with its geometry and dimensions, resulted in shear stresses that fell within the range of reported values for peristalsis. The mass flow rate of 28 mL/min set on the pump produced a homogeneous velocity profile with a laminar flow over the PDMS membrane surface (Fig. 3a). When modeled solely with the characteristics of the PDMS membrane, i.e., the absence of cells, 28 mL/min produced an average shear of 0.3627 Pa on the surface of the membrane (Fig. 3b). When the cell was modeled individually as a spherical object embedded within the PDMS membrane, the computational model predicted an average surface shear of 0.4387 Pa (Fig. 3c). Based on the shear maps produced within the ANSYS Fluent software, both the membrane “wall” and cell model experienced a similar shear as indicated by the color mapping. The cumulative data from the shear simulations provided sufficient evidence of the bioreactor’s capability of producing at least the minimum target shear value of 0.4 Pa with the given design and flow rate of 28 mL/min.

Fig. 3.

Results from strain and shear computational modeling. Shear computational modeling. a 3D fluid velocity visualization, showing a laminar flow through the cell chamber. b Fluid shear on the surface of the PDMS membrane without cells adhered, indicating an average wall shear of 0.36Pa. c Fluid shear on the surface of a 10-μ cell, 30% embedded into the luminal surface of the PDMS membrane indicating an average shear stress of 0.438Pa. Strain computational modeling. d Strain contours on cell deposition area at the maximum indentation point (t= 0 s) in three axes. e Strain variation along the length of the cell area at mid-plane across the membrane thickness at two time points (i) t= 0, and (ii) t= 2.5 s (i.e., after the half-cycle rotation of the screw) in all three axes. The black marker indicates the experimental strain measured in the E11 direction by the hydrogel strain sensor with the screw drive indentation.

Fig. 3.

Results from strain and shear computational modeling. Shear computational modeling. a 3D fluid velocity visualization, showing a laminar flow through the cell chamber. b Fluid shear on the surface of the PDMS membrane without cells adhered, indicating an average wall shear of 0.36Pa. c Fluid shear on the surface of a 10-μ cell, 30% embedded into the luminal surface of the PDMS membrane indicating an average shear stress of 0.438Pa. Strain computational modeling. d Strain contours on cell deposition area at the maximum indentation point (t= 0 s) in three axes. e Strain variation along the length of the cell area at mid-plane across the membrane thickness at two time points (i) t= 0, and (ii) t= 2.5 s (i.e., after the half-cycle rotation of the screw) in all three axes. The black marker indicates the experimental strain measured in the E11 direction by the hydrogel strain sensor with the screw drive indentation.

Close modal

Corroboration of Strain Computational Modeling and Experimental Methods

The strain induced within the bioreactor was validated through both experimental and computational methods. The PVA-PDA piezoelectric hydrogel measured resistivity changes when indented with calibration strain inducers, in the range of 0–41.7%. The change in resistivity was fit to a linear regression line with R2 = 0.93 (online suppl. Fig. 4). Using this calibration curve, the hydrogel strain sensor was used to measure experimental strain arising from the screw drive. Changes in resistivity values when indented within the bioreactor with the screw drive correspond to 15.9%, in the direction of measurement X, or E11.

A whole range of multi-axial strains were computationally predicted using finite element simulations. In the computational determination of strain, proper peristaltic patterning was observed. Contours for the normal logarithmic strains E11, E22, and E33, at the maximum indentation point, right before any rotation (i.e., t = 0 s), were estimated at the cell seeding side of the membrane (Fig. 3d). The strain fields indicated a periodic pattern resembling peristalsis waves. In particular, E11 was greater at two ends of the screw (along the x-axis) and less pronounced in the middle, while E22 showed a wave-like distribution along the perpendicular (y) direction, qualitatively capturing a peristaltic motion along x.

The computational determination of the strain also showed that maximum strain magnitudes consistently occurred at the half-rotation cycle (t = 2.5 s; Fig. 3e). The strains along the length of cell area (x-axis) were always maximum at the contact points of two crests with the membrane. The contact points monotonically moved away from each other and generated deeper indentations to t = 2.5 s (Fig. 3e) followed by an opposite pattern in the second half-cycle. The axial strain (E11) measured by the hydrogel piezoelectric strain sensor at the mid-plane was in excellent agreement with the finite element estimation as indicated by the black marker in the E11 graph (Fig. 3e).

Changes in Proliferation with Mechanical Stimulation

In order to understand the effect of the kinematics produced in the peristalsis bioreactor on cells, we utilized hMSCs. hMSCs were seeded on to a defined area within the bioreactor (with maximal contact with the screw drive, and hence its associated strain). hMSCs subjected to static control, perfusion, strain, or peristalsis were processed for immunofluorescence following 24 h of stimulation. Sections of PDMS were stained with a proliferation marker (Ki67, red fluorescence; Fig. 4a). Ki67 antigen expression within the DAPI counterstained nucleus was considered a positive marker for proliferation [Ohta and Ichimura, 2000]. After 24 h of exposure to the respective conditions, strain (33.21 ± 4.89%), perfusion (25.23 ± 4.99%), and peristalsis (41.98 ± 4.80%) had significantly lower proliferation compared to the control (****p < 0.0001, one-way ANOVA, Fig. 4b). Peristalsis also had a 16.74% decrease in proliferation compared to perfusion (**p < 0.01, one-way ANOVA, Fig. 4b), indicating that the combined application of shear stress and strain in peristalsis influenced hMSC proliferation uniquely.

Fig. 4.

Ki67 immunofluorescent staining of hMSCs exposed to mechanical stimuli. a Representative fluorescence micrographs of hMSCs stained with Ki67 antigen (red fluorescence) and counterstained with nuclear marker, DAPI (blue fluorescence) from four conditions – control (static), strain, perfusion, and peristalsis. Scale bars, 200 or 20 µm. hMSCs were maintained as static controls, or exposed to 24 h of mechanical stimulation in the bioreactor with patterns of perfusion, strain, or peristalsis. b Box and whisker plots of percentage of proliferating cells for each condition. The percentage of proliferating cells was determined by manual counting of Ki67+ nuclei, compared to all nuclei present in the field of view. Statistical significance is indicated following one-way ANOVA analysis, with ****p< 0.0001 and **p< 0.001 between conditions. hMSCs subject to peristalsis demonstrated significantly lower proliferation compared to static controls, or perfusion conditions.

Fig. 4.

Ki67 immunofluorescent staining of hMSCs exposed to mechanical stimuli. a Representative fluorescence micrographs of hMSCs stained with Ki67 antigen (red fluorescence) and counterstained with nuclear marker, DAPI (blue fluorescence) from four conditions – control (static), strain, perfusion, and peristalsis. Scale bars, 200 or 20 µm. hMSCs were maintained as static controls, or exposed to 24 h of mechanical stimulation in the bioreactor with patterns of perfusion, strain, or peristalsis. b Box and whisker plots of percentage of proliferating cells for each condition. The percentage of proliferating cells was determined by manual counting of Ki67+ nuclei, compared to all nuclei present in the field of view. Statistical significance is indicated following one-way ANOVA analysis, with ****p< 0.0001 and **p< 0.001 between conditions. hMSCs subject to peristalsis demonstrated significantly lower proliferation compared to static controls, or perfusion conditions.

Close modal

Changes in Actin Filament Orientation and Nuclear Morphology

hMSCs were maintained as static controls or subject to strain, perfusion, or peristalsis in the bioreactor (Table 1). After 24 h, the PDMS was collected from bioreactors and stained with an actin filament marker to assess the actin alignment within the hMSCs (phalloidin, green fluorescence, Fig. 5a). The nuclei of the stained phalloidin images were assessed for circularity. All three mechanical conditions significantly (*p < 0.05 and ***p < 0.001, one-way ANOVA) differed from the control with increased circularities, implying that mechanical stimuli increase nuclear circularity. Strain and perfusion increased from the control by 1.38- to 1.6-fold, respectively (Fig. 5b). hMSCs subject to peristalsis were the most circular (compare 0.71 ± 0.017 in peristalsis to 0.4 ± 0.062 in static controls; ****p < 0.0001, one-way ANOVA, Fig. 5b). Overall, the cellular nuclei were more circular in conditions with increased mechanical stimuli, specifically peristalsis.

Fig. 5.

Phalloidin immunofluorescent staining of hMSCS exposed to mechanical stimuli. a Representative micrographs of hMSCs stained with Phalloidin (green fluorescence) and counterstained with nuclear marker DAPI (blue fluorescence). hMSCs were exposed to 24 h of perfusion, strain, or peristalsis in the bioreactor, or maintained as static controls. Scale bars, 200 µm or 20 µm. b Box and whisker plot of nuclear circularity of the hMSCs after 24 h of exposure to each condition. A perfect circle has a value of 1.0. Significant differences in nuclear circularity are indicated following one-way ANOVA analysis, with ****p< 0.0001, ***p< 0.001, **p< 0.01, and *p< 0.05. hMSCs exposed to peristalsis had the higher increase in nuclear circularity. c Violin plot for the phalloidin actin filament angle distribution percentages across each condition. Angle distribution was determined with the use of a single-pixel-line-filter at 1° increments and then percentages were grouped into 30° increments. While cells exposed to strain and shear predominantly showed groupings within specific angles, hMSCs experiencing peristalsis were broadly and evenly distributed along all the angular bins evaluated from 0 to 180°.

Fig. 5.

Phalloidin immunofluorescent staining of hMSCS exposed to mechanical stimuli. a Representative micrographs of hMSCs stained with Phalloidin (green fluorescence) and counterstained with nuclear marker DAPI (blue fluorescence). hMSCs were exposed to 24 h of perfusion, strain, or peristalsis in the bioreactor, or maintained as static controls. Scale bars, 200 µm or 20 µm. b Box and whisker plot of nuclear circularity of the hMSCs after 24 h of exposure to each condition. A perfect circle has a value of 1.0. Significant differences in nuclear circularity are indicated following one-way ANOVA analysis, with ****p< 0.0001, ***p< 0.001, **p< 0.01, and *p< 0.05. hMSCs exposed to peristalsis had the higher increase in nuclear circularity. c Violin plot for the phalloidin actin filament angle distribution percentages across each condition. Angle distribution was determined with the use of a single-pixel-line-filter at 1° increments and then percentages were grouped into 30° increments. While cells exposed to strain and shear predominantly showed groupings within specific angles, hMSCs experiencing peristalsis were broadly and evenly distributed along all the angular bins evaluated from 0 to 180°.

Close modal

Phalloidin images were also evaluated based on actin filament angle orientation to quantify cellular alignment changes in response to mechanical stimulation compared to static controls. hMSCs in static controls were randomly oriented, with no apparent frequency distribution in any of the 30° increment bins (Fig. 5c). In fact, random distributions of orientations ranging from 9.7 to 19.7% was observed in all the bins in static controls. In contrast, the application of perfusion resulted in greater alignment along the 60° and 120° bins (holding between 17.7 and 19.5% of cells). The application of strain promoted approximately 18.7% fiber alignment along the 30° axis, in line with documented evidence that cells were likely to align perpendicularly in response to strain [Ghazanfari et al., 2009]. Interestingly, the application of peristalsis resulted in an even distribution of fiber alignment uniformly across 0–180°, with the range of cells falling into each bin between 13.2 and 16.5%.

Gene Expression and Differentiation Changes in hMSCs Exposed to Mechanical Stimuli

In order to further investigate the changes observed through image analysis, we used qPCR to identify underlying changes in gene expression. hMSCs exposed to mechanical stimuli were compared and normalized to static controls maintained for the same duration (24 h), indicated by the dotted line at 1 (Fig. 6a). Notably, peristalsis had increased expression compared to strain in CNN1, CALD1, ACTA2, SM22α, and ALPL (****p < 0.0001), and BGLAP and SPP1 (*p < 0.05; two-way ANOVA, Fig. 6a). Peristalsis also had increased expression compared to perfusion in CNN1, CALD1, ACTA2, SM22α, and ALPL (****p < 0.0001) and BGLAP (***p < 0.001; two-way ANOVA, Fig. 6a). Along with those genes, hMSCs stimulated with strain showed increased expression relative to peristalsis in ACAN, COL2A1, SOX9, ADIPOQ, and LEP (****p < 0.0001) and COMP (***p < 0.001; two-way ANOVA, Fig. 6a). Perfusion demonstrated an increase in expression in COMP (****p < 0.0001) compared to both peristalsis and strain (two-way ANOVA, Fig. 6a).

Fig. 6.

Gene expression and alpha smooth muscle actin and tri-lineage staining of hMSCs exposed to mechanical stimuli. a Gene expression analysis of hMSCs exposed to each mechanical stimulus at the 24-h time point. The dotted line at 1 indicates the normalized level of gene expression in static control hMSCs at the same time point. Genes are categorized as follows: myocyte: CNN1, CALD1, ACTA2, and SM22α; osteocyte: BGLAP, ALPL, and SPP1; chondrocyte: ACAN, COL2A1, SOX9, and COMP; and adipocyte: ADIPOQand LEP. Statistical comparisons from two-way ANOVAs are indicated as: ****p< 0.0001, ***p< 0.001, **p< 0.01, and *p< 0.05). b Representative micrographs of hMSCs stained with alpha smooth muscle actin (green fluorescence) and counterstained with the nuclear marker DAPI (blue fluorescence). hMSCs were exposed to 24 h of perfusion, strain, or peristalsis in the bioreactor, or maintained as static controls. Scale bars, 50 µm. c Fluorometric quantification of α-smooth muscle actin demonstrates significant differences in expression across all conditions compared to peristalsis. Significant differences in actin expression are indicated following one-way ANOVA analysis, with ****p< 0.0001 and ***p< 0.001. d Representative micrographs of hMSCs stained with Alizarin Red S (osteogenic), Alcian Blue (chondrogenic), or Oil Red O (adipogenic). hMSCs were exposed to 24 h of perfusion, strain, or peristalsis in the bioreactor, or maintained as static controls and stained for markers of trilineage differentiation. At 24 h, robust staining was not observed for trilineage markers tested. Scale bars, 400 µm.

Fig. 6.

Gene expression and alpha smooth muscle actin and tri-lineage staining of hMSCs exposed to mechanical stimuli. a Gene expression analysis of hMSCs exposed to each mechanical stimulus at the 24-h time point. The dotted line at 1 indicates the normalized level of gene expression in static control hMSCs at the same time point. Genes are categorized as follows: myocyte: CNN1, CALD1, ACTA2, and SM22α; osteocyte: BGLAP, ALPL, and SPP1; chondrocyte: ACAN, COL2A1, SOX9, and COMP; and adipocyte: ADIPOQand LEP. Statistical comparisons from two-way ANOVAs are indicated as: ****p< 0.0001, ***p< 0.001, **p< 0.01, and *p< 0.05). b Representative micrographs of hMSCs stained with alpha smooth muscle actin (green fluorescence) and counterstained with the nuclear marker DAPI (blue fluorescence). hMSCs were exposed to 24 h of perfusion, strain, or peristalsis in the bioreactor, or maintained as static controls. Scale bars, 50 µm. c Fluorometric quantification of α-smooth muscle actin demonstrates significant differences in expression across all conditions compared to peristalsis. Significant differences in actin expression are indicated following one-way ANOVA analysis, with ****p< 0.0001 and ***p< 0.001. d Representative micrographs of hMSCs stained with Alizarin Red S (osteogenic), Alcian Blue (chondrogenic), or Oil Red O (adipogenic). hMSCs were exposed to 24 h of perfusion, strain, or peristalsis in the bioreactor, or maintained as static controls and stained for markers of trilineage differentiation. At 24 h, robust staining was not observed for trilineage markers tested. Scale bars, 400 µm.

Close modal

Notably, four myogenic markers (CNN1, CALD1, ACTA2, and SM22α) were significantly increased in hMSCs stimulated with peristalsis, providing strong evidence that hMSCs were differentiating toward the myogenic lineage. In order to validate qPCR findings pointing towards myogenic differentiation, hMSCs exposed to strain, perfusion, or peristalsis were stained with α-smooth muscle actin. These results indicated an increased expression of α-smooth muscle actin in hMSCs stimulated with peristalsis (6.409 AFU) compared to perfusion and control (3.841 AFU and 3.917 AFU, ****p < 0.0001) and strain (4.820 AFU, ***p < 0.001; one-way ANOVA, Fig. 6c).

In addition to α-smooth muscle actin staining, we also confirmed the absence of staining for trilineage osteogenic, adipogenic, and chondrogenic differentiation (Fig. 6d). Static controls served as negative controls, while online supplementary Figure 5b demonstrates positive controls for trilineage differentiation of hMSCs using widely validated protocols [Eggerschwiler et al., 2019]. hMSCs subject to mechanical stimulus in the bioreactor did not demonstrate positivity for osteogenic (Alizarin Red S), chondrogenic (Alcian Blue), or adipogenic (Oil Red O) staining (Fig. 6d).

Taken together, our data suggest that an hMSC differentiation event was occurring. Depending on the kind of mechanical stimulation the cells were exposed to within the bioreactor, the orientation of differentiation lineage varied, with peristalsis promoting myogenic differentiation in hMSCs.

Peristalsis is a nuanced mechanical stimulus, occurring across several smooth muscle organ systems including the intestines, uterus, airways, and ureters [Huizinga, 1999; Kunz and Leyendecker, 2002; Hosseini et al., 2018]. Peristalsis, and thereby dysperistalsis, is central to physiology and patho-physiology [Leyendecker et al., 1996; Iwakiri et al., 2017; Patel andThavamani, 2021]. With a growing body of evidence suggesting that mechanotransduction plays a vital role in development and disease, it is advantageous to include peristalsis in the study of cells, tissues, and organs [Jaalouk and Lammerding, 2009; Maurer and Lammerding, 2019]. Biomanufacturing of cells for cellular therapy or regenerative engineering of tissues and organs will all benefit from having received mechanical cues during in vitro bioprocessing. In several current in vitro studies, the kinematics of peristalsis is often simplified to fluid shear, or cyclic uniaxial strain. However, peristalsis is a summation of multi-axial strain that propels luminal content, which them simultaneously results in shear stress. The mechanics associated with peristalsis therefore involve concurrent multi-axial strain and shear stress.

Several current designs mimic peristalsis favorably, especially those associated with the gastrointestinal tract at the microfluidic and the macro bioreactor levels [Kim et al., 2012; Cei et al., 2016; Costello et al., 2017; Wang et al., 2018; Zhou et al., 2018]. However, a majority of the studies intended to study peristalsis-associated mechanotransduction simplify the kinematics of peristalsis to shear stress or cyclic strain alone. In one example, both longitudinal and circular contractions were incorporated to closely simulate a physiological intestine [Cei et al., 2016]. However, there was no additional shear stress added to the device, besides the fluid stress created by the contractions. Instances of peristalsis bioreactors within the reproductive tract are much fewer and far between [da Rocha and Smith, 2012; Hellstrom et al., 2017; Elad et al., 2020a], with models relying on fluid shear stress from perfusion [Elad et al., 2020b], or uniaxial cyclic tensile stretch [Kim et al., 2020]. In urothelial tissue engineering, bioreactors are used to precondition ureteral grafts prior to implantation, which results in urothelial cellular organization [Cattan et al., 2011; Janke et al., 2019]. The mechanical stimulus however is either fluid flow or cyclic stretching, neither of which are biomimetic of ureteric peristalsis [Basu et al., 2012; Wang et al., 2014; Vardar et al., 2015]. All of these devices incorporate very important aspects of shear or strain but still fall short of a complete peristaltic model. In comparing our device to those currently available, the full effect of peristalsis can be explored with the inclusion of the simultaneous administration of shear and strain in our novel model.

In order to assess the peristaltic capabilities of our device, the fluid shear was simulated via CFD software and the strain was computationally and experimentally determined. The shear simulations produced an average shear of 0.3627 Pa (membrane) and 0.4387 Pa (cell surface on the membrane). This falls well within the desired range of reported values for kinematics of peristalsis. Mainly, during peristalsis, intestinal shear ranges from 0.2 to 3.2 Pa, ureter shear ranges from 0.002 to 0.006 Pa, and uterine shear ranges from 0.7 to 2.9 Pa [Kublickiene et al., 2000; Avvisato et al., 2007; Kimura et al., 2018]. Currently, a flow rate of 28 mL/min produces a shear of 0.4 Pa. This flow rate is tunable, and can therefore produce a range of shear stresses that makes this peristalsis bioreactor.

The average of the minimum values from each of the three ranges previously reported is 0.3 Pa. The value of 0.4 Pa falls near the average of the shear experienced during peristalsis in the body, making it the target point for our device for initial proof of concept studies. The resulting computational analysis indicated that our peristalsis bioreactor was capable of mimicking shear stresses within the range of those reported in the literature. By modeling the target shear stress, our device can therefore be applied to a broad range of organ systems.

In addition to the shear analysis, the strain was computationally and experimentally analyzed. There is a large range of peristalsis-induced strain forces found in the body; strain can be anywhere from 1% to upwards of 20% depending on the organ system [Forouhar et al., 2006; Fujita et al., 2010; Kim et al., 2012, 2020]. The experimental strain value of 15.9% was within these reported ranges. The results of computational modeling also indicated that the screw drive acted as an effective method for controlling the magnitude and frequency of oscillating strain to generate the desired peristalsis waves.

With the evaluation of the shear and strain meeting physiological values, hMSCs were applied in the use of the bioreactor to determine the effect of various mechanical stimuli (perfusion, strain, and peristalsis) compared to static controls. hMSCs are known to respond to mechanical stimulation, lending them as good models to study changes in cellular response to mechanical stimulus. Direct mechanical stimulation, in the form of oscillatory fluid flow, of hMSCs in vitro has been documented to have a strong role in influencing stem cell behavior and differentiation [Hoey et al., 2012]. Specifically, hMSCs are well documented for how they respond to shear and strain forces [Simmons et al., 2003; Friedl et al., 2007; Zhang et al., 2012; Yuan et al., 2013]. When strain forces are applied to hMSCS, they have a tendency to differentiate toward an osteogenic, tenogenic, or myogenic phenotype and align perpendicularly to the applied strain [McClarren and Olabisi, 2018; Nam et al., 2019]. In terms of hMSC response to perfusion, hMSCs have an increase in gene expression related to differentiation and, at high rates of perfusion, an overall decrease in proliferation [Zhao et al., 2007; Becquart et al., 2016].

Our data indicate a shift in hMSCs toward differentiation when exposed to mechanical stimuli in both the immunofluorescence (Fig. 4, 5) and gene expression analyses (Fig. 6a). There is also a significantly different response to peristalsis compared to strain, perfusion, and static controls. Specifically, in the proliferation study, the different effects of each force, perfusion, strain, and the combined peristalsis, indicate a connection between mechanical stimuli and the proliferative capacity of hMSCs. In hMSC studies with strain or perfusion, extended time points, like 24 h, and increased shear rates have decreased overall proliferation, lending them toward differentiation [Song et al., 2007; Zhao et al., 2007; Yourek et al., 2010]. In our data, with the combination of both the perfusion and strain stimuli in peristalsis, there was a large decrease in proliferation, consistent with the findings in literature for strain and shear individually (Fig. 4b) [Songet al., 2007; Zhao et al., 2007]. With such a decrease in proliferation, we hypothesized that differentiation was likely happening and not proliferation. This idea was largely supported by the gene expression findings. With the increased expression of genes associated strongly with smooth muscle cells, like CNN1, CALD1, ACTA2, and SM22α [Naritaet al., 2008], in peristalsis, a clear differentiation toward myogenic lineage was occurring (Fig. 6a) [Lin and Lilly, 2014; Brunet al., 2015]. CNN1 and ACTA2 were also upregulated in perfusion, along with the osteocyte marker, SPP1, the chondrocyte marker, COL2A1, and the adipocyte marker, ADIPOQ. While there was no direct lineage differentiation in the perfusion condition, it is evident the hMSCs were expressing differentiation markers with a decreased proliferative capacity, consistent with the literature for a decrease in proliferation and maintenance of lineage potential [Franket al., 2016]. hMSCs exposed to strain also demonstrated an increase in gene expression in COL2A1, SOX9, COMP, ADIPOQ, and LEP. With the upregulation of two genes for both chondrocyte and adipocyte, there again was not a clear differentiation lineage. However, like with perfusion, these results corroborate a decrease in proliferation still allowing for lineage differentiation to occur. With each of the genes evaluated, the effect of, and subsequent response to, peristalsis, strain, and perfusion can be used to support the idea of increased differentiation and decreased proliferation.

Furthermore, the differentiation of hMSCs exposed to each condition was evaluated with a smooth muscle actin fluorescence stain and trilineage differentiation stains (Fig. 6b, d). Consistent with the qPCR data (Fig. 6c), the smooth muscle actin fluorescent images indicated an increase in actin expression compared to control, strain, and perfusion which is also consistent with myogenic differentiation reported in hMSCs [Liu et al., 2013].

Along with differentiation characteristics, cell shape was evaluated. The alteration in nuclear morphology can be seen in the actin fiber alignment results (Fig. 5c). The application of peristalsis resulted in an even distribution of fiber alignment uniformly across 0–180°, while strain and perfusion aligned across one or two axes. Specifically, hMSCs have been shown to align perpendicularly to strain which is evident in our results with the majority of strain cells aligning along the 30° axis [Nam et al., 2019]. In addition to actin alignment in response to mechanical stimulation, our results also indicate an increase in nuclear circularity, particularly in peristalsis (Fig. 5b). hMSCs in their stem-like state have an inherently elongated nuclei which would output a value less than 1 [Ankam et al., 2018], while differentiating hMSCs approach 1. Ultimately, in comparing our findings with the hMSCs to well-established data, we have shown that the mechanics within the peristalsis bioreactor impact cells differently, dependent on the mechanical stimuli (perfusion, strain, or peristalsis), in terms of gene expression, α-smooth muscle actin expression, alignment and nuclear morphology [Ito et al., 2007; Song et al., 2007; Zhao et al., 2007; Mishra et al., 2008; Robin et al., 2013; Elsafadi et al., 2016; Nam et al., 2019].

Peristalsis is central to many smooth muscle organs, including the gastrointestinal tract, uterus, and ureters, and is comprised of multi-axial strain (radial and axial strain) and shear stress. Our new peristalsis bioreactor design combining finite element modeling with rapid prototyping delivers multi-axial strain and concurrent shear stress to cells seeded on a biocompatible membrane. hMSCs, widely used in tissue engineering and regenerative medicine, sensed and responded to peristalsis within the bioreactor uniquely. A bioreactor such as ours will allow the kinematics of peristalsis to be replicated in several in vitro processes, including regenerative engineering and biomanufacturing. Furthermore, the bioreactor is tunable and amenable to the study of mechano-transduction-related processes involved in development and disease across several organ systems.

The results reported in this manuscript did not involve human or animal subjects.

The authors have no conflicts of interest to disclose.

This work was supported by the Texas A&M University Department of Biomedical Engineering and the Texas A&M Engineering Experiment Station (S.R.), and NIH R00HL138288 (R.A.).

S.R. and L.Z.C. designed the bioreactor, L.Z.C. and D.N. prototyped the bioreactor, performed computational models of shear stress and experimentally validated strain. A.J.C., L.Z.C., and D.N. contributed to design refinements, prototyping process parameter refinements supervised by S.B.M. D.N. performed all operational programming of microprocessors and bioreactor control, and additionally performed all automated cellular morphometry and analysis thereof. A.J.C. performed all biological experiments and downstream analysis of cells exposed to bioreactors. H.B. performed all the finite element modeling, supervised by R.A. S.R. conceived the project, designed experiments, and oversaw operational analysis, data acquisition and interpretation. A.J.C., L.Z.C., D.N., and S.R. wrote initial drafts of the manuscript, which were critically reviewed and revised by R.A., H.B., and S.B.M. All authors agreed on the final submission of this manuscript.

Data supporting the findings of this study are available from the corresponding author upon reasonable request.

1.
Ankam
S
,
Teo
BKK
,
Pohan
G
,
Ho
SWL
,
Lim
CK
,
Yim
EKF
.
Temporal changes in nucleus morphology, lamin A/C and histone methylation during nanotopography-induced neuronal differentiation of stem cells
.
Front Bioeng Biotechnol
.
2018
;
6
(
69
):
69
.
2.
Aranda
V
,
Cortez
R
,
Fauci
L
.
A model of Stokesian peristalsis and vesicle transport in a three-dimensional closed cavity
.
J Biomech
.
2015 2015/06/25/
;
48
(
9
):
1631
8
.
3.
Avvisato
CL
,
Yang
X
,
Shah
S
,
Hoxter
B
,
Li
W
,
Gaynor
R
,
.
Mechanical force modulates global gene expression and β-catenin signaling in colon cancer cells
.
J Cell Sci
.
2007
;
120
(
15
):
2672
82
.
4.
Basu
J
,
Jayo
MJ
,
Ilagan
RM
,
Guthrie
KI
,
Sangha
N
,
Genheimer
CW
,
.
Regeneration of native-like neo-urinary tissue from nonbladder cell sources
.
Tissue Eng Part A
.
2012
;
18
(
9‐10
):
1025
34
.
5.
Becquart
P
,
Cruel
M
,
Hoc
T
,
Sudre
L
,
Pernelle
K
,
Bizios
R
,
.
Human mesenchymal stem cell responses to hydrostatic pressure and shear stress
.
Eur Cell Mater
.
2016 Feb 19
;
31
:
160
73
.
6.
Brun
J
,
Lutz
KA
,
Neumayer
KMH
,
Klein
G
,
Seeger
T
,
Uynuk-Ool
T
,
.
Smooth muscle-like cells generated from human mesenchymal stromal cells display marker gene expression and electrophysiological competence comparable to bladder smooth muscle cells
.
PLoS One
.
2015
;
10
(
12
):
e0145153
.
7.
Cattan
V
,
Bernard
G
,
Rousseau
A
,
Bouhout
S
,
Chabaud
S
,
Auger
FA
,
.
Mechanical stimuli-induced urothelial differentiation in a human tissue-engineered tubular genitourinary graft
.
Eur Urol
.
2011 Dec
;
60
(
6
):
1291
8
.
8.
Cei
D
,
Costa
J
,
Gori
G
,
Frediani
G
,
Domenici
C
,
Carpi
F
,
.
A bioreactor with an electro-responsive elastomeric membrane for mimicking intestinal peristalsis
.
Bioinspir Biomim
.
2016
;
12
(
1
):
016001
.
9.
Costello
CM
,
Phillipsen
MB
,
Hartmanis
LM
,
Kwasnica
MA
,
Chen
V
,
Hackam
D
,
.
Microscale bioreactors for in situ characterization of GI epithelial cell physiology
.
Sci Rep
.
2017 Oct 2
;
7
(
1
):
12515
.
10.
da Rocha
AM
,
Smith
GD
.
Culture systems: fluid dynamic embryo culture systems (microfluidics)
.
Methods Mol Biol
.
2012
;
912
:
355
65
.
11.
Eggerschwiler
B
,
Canepa
DD
,
Pape
HC
,
Casanova
EA
,
Cinelli
P
.
Automated digital image quantification of histological staining for the analysis of the trilineage differentiation potential of mesenchymal stem cells
.
Stem Cell Res Ther
.
2019 2019/02/26
;
10
(
1
):
69
.
12.
Elad
D
,
Jaffa
AJ
,
Grisaru
D
.
Biomechanics of early life in the female reproductive tract
.
Physiology
.
2020a Mar 1
;
35
(
2
):
134
43
.
13.
Elad
D
,
Zaretsky
U
,
Kuperman
T
,
Gavriel
M
,
Long
M
,
Jaffa
A
,
.
Tissue engineered endometrial barrier exposed to peristaltic flow shear stresses
.
APL Bioeng
.
2020b
;
4
(
2
):
026107
.
14.
Elsafadi
M
,
Manikandan
M
,
Dawud
RA
,
Alajez
NM
,
Hamam
R
,
Alfayez
M
,
.
Transgelin is a TGFβ-inducible gene that regulates osteoblastic and adipogenic differentiation of human skeletal stem cells through actin cytoskeleston organization
.
Cell Death Dis
.
2016
;
7
(
8
):
e2321
.
15.
Forouhar
AS
,
Liebling
M
,
Hickerson
A
,
Nasiraei-Moghaddam
A
,
Tsai
H-J
,
Hove
JR
,
.
The embryonic vertebrate heart tube is a dynamic suction pump
.
Science
.
2006
;
312
(
5774
):
751
3
.
16.
Frank
V
,
Kaufmann
S
,
Wright
R
,
Horn
P
,
Yoshikawa
HY
,
Wuchter
P
,
.
Frequent mechanical stress suppresses proliferation of mesenchymal stem cells from human bone marrow without loss of multipotency
.
Sci Rep
.
2016
;
6
(
1
):
24264
.
17.
Friedl
G
,
Schmidt
H
,
Rehak
I
,
Kostner
G
,
Schauenstein
K
,
Windhager
R
.
Undifferentiated human mesenchymal stem cells (hMSCs) are highly sensitive to mechanical strain: transcriptionally controlled early osteo-chondrogenic response in vitro
.
Osteoarthritis Cartilage
.
2007
;
15
(
11
):
1293
300
.
18.
Fujita
H
,
Hida
M
,
Kanemoto
K
,
Fukuda
K
,
Nagata
M
,
Awazu
M
.
Cyclic stretch induces proliferation and TGF-beta1-mediated apoptosis via p38 and ERK in ureteric bud cells
.
Am J Physiol Renal Physiol
.
2010
;
299
(
3
):
F648
55
.
19.
Fung
YC
,
Yih
CS
.
Peristaltic transport
.
J Appl Mech
.
1968
;
35
(
4
):
669
75
.
20.
Gayer
CP
,
Basson
MD
.
The effects of mechanical forces on intestinal physiology and pathology
.
Cell Signal
.
2009 Aug
;
21
(
8
):
1237
44
.
21.
Ghazanfari
S
,
Tafazzoli-Shadpour
M
,
Shokrgozar
MA
.
Effects of cyclic stretch on proliferation of mesenchymal stem cells and their differentiation to smooth muscle cells
.
Biochem Biophys Res Commun
.
2009 2009/10/23/
;
388
(
3
):
601
5
.
22.
Hellstrom
M
,
Bandstein
S
,
Brannstrom
M
.
Uterine tissue engineering and the future of uterus transplantation
.
Ann Biomed Eng
.
2017 Jul
;
45
(
7
):
1718
30
.
23.
Hoey
DA
,
Tormey
S
,
Ramcharan
S
,
O'Brien
FJ
,
Jacobs
CR
.
Primary cilia-mediated mechanotransduction in human mesenchymal stem cells
.
Stem Cells
.
2012
;
30
(
11
):
2561
70
.
24.
Hosseini
G
,
Ji
C
,
Xu
D
,
Rezaienia
MA
,
Avital
E
,
Munjiza
A
,
.
A computational model of ureteral peristalsis and an investigation into ureteral reflux
.
Biomed Eng Lett
.
2018
;
8
(
1
):
117
25
.
25.
Huizinga
JD
.
Gastrointestinal peristalsis: joint action of enteric nerves, smooth muscle, and interstitial cells of Cajal
.
Microsc Res Tech
.
1999
;
47
(
4
):
239
47
.
26.
Ito
T
,
Sawada
R
,
Fujiwara
Y
,
Seyama
Y
,
Tsuchiya
T
.
FGF-2 suppresses cellular senescence of human mesenchymal stem cells by down-regulation of TGF-beta2
.
Biochem Biophys Res Commun
.
2007
;
359
(
1
):
108
14
.
27.
Iwakiri
K
,
Hoshino
S
,
Kawami
N
.
Mechanisms underlying excessive esophageal acid exposure in patients with gastroesophageal reflux disease
.
Esophagus
.
2017
;
14
(
3
):
221
8
.
28.
Jaalouk
DE
,
Lammerding
J
.
Mechanotransduction gone awry
.
Nat Rev Mol Cell Biol
.
2009
;
10
(
1
):
63
73
.
29.
Janke
HP
,
de Jonge
PKJD
,
Feitz
WFJ
,
Oosterwijk
E
.
Reconstruction strategies of the ureter and urinary diversion using tissue engineering approaches
.
Tissue Eng Part B Rev
.
2019 Jun
;
25
(
3
):
237
48
.
30.
Jufri
NF
,
Mohamedali
A
,
Avolio
A
,
Baker
MS
.
Mechanical stretch: physiological and pathological implications for human vascular endothelial cells
.
Vasc Cell
.
2015 2015/09/18
;
7
(
1
):
8
.
31.
Kim
HJ
,
Huh
D
,
Hamilton
G
,
Ingber
DE
.
Human gut-on-a-chip inhabited by microbial flora that experiences intestinal peristalsis-like motions and flow
.
Lab Chip
.
2012 Jun 21
;
12
(
12
):
2165
74
.
32.
Kim
J
,
Ushida
T
,
Montagne
K
,
Hirota
Y
,
Yoshino
O
,
Hiraoka
T
,
.
Acquired contractile ability in human endometrial stromal cells by passive loading of cyclic tensile stretch
.
Sci Rep
.
2020
;
10
(
1
):
9014
.
33.
Kim
TJ
,
Joo
C
,
Seong
J
,
Vafabakhsh
R
,
Botvinick
EL
,
Berns
MW
,
.
Distinct mechanisms regulating mechanical force-induced Ca2+ signals at the plasma membrane and the ER in human MSCs
.
Elife
.
2015 Feb 10
;
4
:
e04876
.
34.
Kimura
H
,
Nishikawa
M
,
Yanagawa
N
,
Nakamura
H
,
Miyamoto
S
,
Hamon
M
,
.
Effect of fluid shear stress on in vitro cultured ureteric bud cells
.
Biomicrofluidics
.
2018
;
12
(
4
):
044107
.
35.
Kublickiene
KR
,
Lindblom
B
,
Krüger
K
,
Nisell
H
.
Preeclampsia: evidence for impaired shear stress–mediated nitric oxide release in uterine circulation
.
Am J Obstet Gynecol
.
2000
;
183
(
1
):
160
6
.
36.
Kunz
G
,
Leyendecker
G
.
Uterine peristaltic activity during the menstrual cycle: characterization, regulation, function and dysfunction
.
Reprod Biomed Online
.
2002
;
4(suppl 3)
:
5
9
.
37.
Kuo
Y-C
,
Chang
T-H
,
Hsu
W-T
,
Zhou
J
,
Lee
H-H
,
Hui-Chun Ho
J
,
.
Oscillatory shear stress mediates directional reorganization of actin cytoskeleton and alters differentiation propensity of mesenchymal stem cells
.
Stem Cells
.
2015
;
33
(
2
):
429
42
.
38.
Leyendecker
G
,
Kunz
G
,
Wildt
L
,
Beil
D
,
Deininger
H
.
Uterine hyperperistalsis and dysperistalsis as dysfunctions of the mechanism of rapid sperm transport in patients with endometriosis and infertility
.
Hum Reprod
.
1996 Jul
;
11
(
7
):
1542
51
.
39.
Lin
CH
,
Lilly
B
.
Endothelial cells direct mesenchymal stem cells toward a smooth muscle cell fate
.
Stem Cells Dev
.
2014
;
23
(
21
):
2581
90
.
40.
Liu
S
,
Zheng
R
,
Chen
S
,
Wu
Y
,
Liu
H
,
Wang
P
,
.
A compliant, self-adhesive and self-healing wearable hydrogel as epidermal strain sensor
.
J Mater Chem C
.
2018
;
6
(
15
):
4183
90
.
41.
Liu
Y
,
Deng
B
,
Zhao
Y
,
Xie
S
,
Nie
R
.
Differentiated markers in undifferentiated cells: Expression of smooth muscle contractile proteins in multipotent bone marrow mesenchymal stem cells
.
Dev Growth Differ
.
2013
;
55
(
5
):
591
605
.
42.
Liu
Y-S
,
Liu
Y-A
,
Huang
C-J
,
Yen
M-H
,
Tseng
C-T
,
Chien
S
,
.
Mechanosensitive TRPM7 mediates shear stress and modulates osteogenic differentiation of mesenchymal stromal cells through Osterix pathway
.
Sci Rep
.
2015
;
5
(
1
):
16522
.
43.
Livak
KJ
,
Schmittgen
TD
.
Analysis of relative gene expression data using real-time quantitative PCR and the 2–ΔΔCT method
.
Methods
.
2001
;
25
(
4
):
402
8
.
44.
Maurer
M
,
Lammerding
J
.
The driving force: nuclear mechanotransduction in cellular function, fate, and disease
.
Annu Rev Biomed Eng
.
2019
;
21
(
1
):
443
68
.
45.
McClarren
B
,
Olabisi
R
.
Strain and vibration in mesenchymal stem cells
.
Int J Biomater
.
2018
;
2018
(
8686794
):
8686794
.
46.
Meza
D
,
Musmacker
B
,
Steadman
E
,
Stransky
T
,
Rubenstein
DA
,
Yin
W
.
Endothelial cell biomechanical responses are dependent on both fluid shear stress and tensile strain
.
Cell Mol Bioeng
.
2019
;
12
(
4
):
311
25
.
47.
Mishra
PJ
,
Mishra
PJ
,
Humeniuk
R
,
Medina
DJ
,
Alexe
G
,
Mesirov
JP
,
.
Carcinoma-associated fibroblast-like differentiation of human mesenchymal stem cells
.
Cancer Res
.
2008
;
68
(
11
):
4331
9
.
48.
RK
.
Peristalsis in the circular and longitudinal muscles of the esophagus
. In:
Rafael
S
, editor.
Motor function of the pharynx, esophagus, and its sphincters.
San Rafael:
Morgan & Claypool Life Sciences
;
2011
.
49.
Nam
HY
,
Pingguan-Murphy
B
,
Abbas
AA
,
Merican
AM
,
Kamarul
T
.
Uniaxial cyclic tensile stretching at 8% strain exclusively promotes tenogenic differentiation of human bone marrow-derived mesenchymal stromal cells
.
Stem Cells Int
.
2019
;
2019
(
9723025
):
9723025
.
50.
Narita
Y
,
Yamawaki
A
,
Kagami
H
,
Ueda
M
,
Ueda
Y
.
Effects of transforming growth factor-beta 1 and ascorbic acid on differentiation of human bone-marrow-derived mesenchymal stem cells into smooth muscle cell lineage
.
Cell Tissue Res
.
2008
;
333
(
3
):
449
59
.
51.
Ohta
Y
,
Ichimura
K
.
Proliferation markers, proliferating cell nuclear antigen, Ki67, 5-bromo-2'-deoxyuridine, and cyclin D1 in mouse olfactory epithelium
.
Ann Otol Rhinol Laryngol
.
2000
;
109
(
11
):
1046
8
.
52.
Ostrowski
MA
,
Huang
EY
,
Surya
VN
,
Poplawski
C
,
Barakat
JM
,
Lin
GL
,
.
Multiplexed fluid flow device to study cellular response to tunable shear stress gradients
.
Ann Biomed Eng
.
2016
;
44
(
7
):
2261
72
.
53.
Patel
KS
,
Thavamani
A
.
Physiology, peristalsis.
StatPearls Publishing
;
2021
.
54.
Pei
E
,
Shen
J
,
Watling
J
.
Direct 3D printing of polymers onto textiles: experimental studies and applications
.
Rapid Prototyping J
.
2015
;
21
(
5
):
556
71
.
55.
Püspöki
Z
,
Storath
M
,
Sage
D
,
Unser
M
.
Transforms and operators for directional bioimage analysis: a survey
.
Adv Anat Embryol Cell Biol
.
2016
;
219
:
69
93
.
56.
Robin
Y-M
,
Penel
N
,
Pérot
G
,
Neuville
A
,
Vélasco
V
,
Ranchère-Vince
D
,
.
Transgelin is a novel marker of smooth muscle differentiation that improves diagnostic accuracy of leiomyosarcomas: a comparative immunohistochemical reappraisal of myogenic markers in 900 soft tissue tumors
.
Mod Pathol
.
2013
;
26
(
4
):
502
10
.
57.
Simmons
CA
,
Matlis
S
,
Thornton
AJ
,
Chen
S
,
Wang
CY
,
Mooney
DJ
.
Cyclic strain enhances matrix mineralization by adult human mesenchymal stem cells via the extracellular signal-regulated kinase (ERK1/2) signaling pathway
.
J Biomech
.
2003
;
36
(
8
):
1087
96
.
58.
Song
G
,
Ju
Y
,
Soyama
H
,
Ohashi
T
,
Sato
M
.
Regulation of cyclic longitudinal mechanical stretch on proliferation of human bone marrow mesenchymal stem cells
.
Mol Cell Biomech
.
2007 Dec
;
4
(
4
):
201
10
.
59.
Vahidi
B
,
Fatouraee
N
.
A biomechanical simulation of ureteral flow during peristalsis using intraluminal morphometric data
.
J Theor Biol
.
2012
;
298
:
42
50
.
60.
Vardar
E
,
Engelhardt
E-M
,
Larsson
HM
,
Mouloungui
E
,
Pinnagoda
K
,
Hubbell
JA
,
.
Tubular compressed collagen scaffolds for ureteral tissue engineering in a flow bioreactor system
.
Tissue Eng Part A
.
2015
;
21
(
17‐18
):
2334
45
.
61.
Waldrop
L
,
Miller
L
.
Large-amplitude, short-wave peristalsis and its implications for transport
.
Biomech Model Mechanobiol
.
2016
;
15
(
3
):
629
42
.
62.
Wang
M
,
Chen
Y
,
Khan
R
,
Liu
H
,
Chen
C
,
Chen
T
,
.
A fast self-healing and conductive nanocomposite hydrogel as soft strain sensor
.
Colloids Surf A
.
2019
;
567
:
139
49
.
63.
Wang
Y
,
Fu
Q
,
Zhao
RY
,
Deng
CL
.
Muscular tubes of urethra engineered from adipose-derived stem cells and polyglycolic acid mesh in a bioreactor
.
Biotechnol Lett
.
2014
;
36
(
9
):
1909
16
.
64.
Wang
Y
,
Kim
R
,
Hinman
SS
,
Zwarycz
B
,
Magness
ST
,
Allbritton
NL
.
Bioengineered systems and designer matrices that recapitulate the intestinal stem cell niche
.
Cell Mol Gastroenterol Hepatol
.
2018 Mar
;
5
(
3
):
440
e1
.
65.
Yokoyama
S
,
Ozaki
T
.
Contractions of the longitudinal and circular muscle of the small intestine
.
Prog Clin Biol Res
.
1990
;
327
:
483
92
.
66.
Yourek
G
,
McCormick
SM
,
Mao
JJ
,
Reilly
GC
.
Shear stress induces osteogenic differentiation of human mesenchymal stem cells
.
Regen Med
.
2010
;
5
(
5
):
713
24
.
67.
Yuan
L
,
Sakamoto
N
,
Song
G
,
Sato
M
.
High-level shear stress stimulates endothelial differentiation and VEGF secretion by human mesenchymal stem cells
.
Cel Mol Bioeng
.
2013
;
6
(
2
):
220
9
.
68.
Zhang
H
,
Kay
A
,
Forsyth
NR
,
Liu
KK
,
El Haj
AJ
.
Gene expression of single human mesenchymal stem cell in response to fluid shear
.
J Tissue Eng
.
2012
;
3
(
1
):
2041731412451988
.
69.
Zhao
F
,
Chella
R
,
Ma
T
.
Effects of shear stress on 3-D human mesenchymal stem cell construct development in a perfusion bioreactor system: experiments and hydrodynamic modeling
.
Biotechnol Bioeng
.
2007
;
96
(
3
):
584
95
.
70.
Zhou
W
,
Chen
Y
,
Roh
T
,
Lin
Y
,
Ling
S
,
Zhao
S
,
.
Multifunctional bioreactor system for human intestine tissues
.
ACS Biomater Sci Eng
.
2018 Jan 8
;
4
(
1
):
231
9
.