Numerical Study on Convection-Diffusion and Wall Adsorption of Products in the Small Intestine
under the Synergy of Peristalsis and Segmentation
1. Model Overview and Core Assumptions
This model is built based on COMSOL Multiphysics 6.3, focusing on the small intestine
(including the duodenum and jejunum), stomach, and pyloric region. It aims to establish a
comprehensive numerical model for simulating peristaltic wave motion, segmental motion,
fluid flow, and transport of dilute substances (product bacterial solution) in the human
small intestine over a simulation duration of 600 seconds. Through multi-physics coupling
(moving mesh, laminar flow, dilute substance transport, and boundary ordinary differential
equations), the model quantifies the convection-diffusion behavior of the product bacterial
solution in the digestive tract and its adhesion characteristics to the intestinal wall. The
core assumptions are as follows:
1.1 Model Introduction
As the core site for digestion and absorption of nutrients in the human digestive
system, the peristaltic and segmental motions of the small intestine play a crucial role
in promoting the mixing of intestinal contents, accelerating substance transport, and
improving absorption efficiency. Accurate simulation of the dynamic environment of the
small intestine and the product transport process provides guidance for predicting
product usage and action sites. This study develops a multi-physics coupled numerical
model of the human small intestine. The specific objectives are:
(1) Construct a geometric model of the small intestine (including the duodenum, jejunum,
and pylorus)
based on physiological parameters to simulate the synergistic effect of peristaltic
waves and segmental motions;
(2) Integrate fluid flow, dilute substance transport, and
wall adsorption-desorption kinetics into the model to realize coupled simulation of
multi-physical processes;
(3) Verify the reliability of the model through theoretical
analysis and comparison with indirect experimental data, providing a basis for
subsequent studies on bacterial colony adhesion and microplastic degradation.
1.2 Geometric and Motion Assumptions
(1) Axisymmetric Simplification: The digestive tract (stomach, pylorus, duodenum,
jejunum) is constructed as a 2D axisymmetric model, ignoring circumferential
(φ-direction) changes and only considering radial (r) and axial (z) motions and
transport.
(2) Peristaltic Wave Simplification: The radial contraction of peristaltic waves in the
jejunum is described using a Gaussian distribution function. The gastric peristaltic
waves are simulated by superimposing a sine-squared function with a smooth step
function. The opening and closing of the pylorus and changes in the duodenal radius are
achieved through piecewise functions to ensure smooth spatial and temporal transitions.
(3) Segmentation Wave Assumption: The segmentation of substances in the jejunum and
duodenum is realized through wall displacement in the form of a sine wave, and the
segmentation region is limited to a specific axial range (constrained by a smooth
function).
1.3 Fluid and Transport Assumptions
(1) Incompressible Laminar Flow: The fluid density (ρ = 1040 kg/m³) and dynamic
viscosity (μ = 0.0014 Pa·s) are constant[3]. The flow satisfies the
Navier-Stokes
equations, and turbulence (turbulence model is turned off) and gravity effects are
ignored.
(2) Dilute Substance Assumption: The initial concentration of the product
bacterial solution is low (initial concentration C₀ = 1 mol/m³), which does not affect
the fluid viscosity and density. Substance transport only considers convection and
molecular diffusion (turbulent diffusion is ignored). Since Colony-Forming Units (CFU)
are not the dependent variable unit used in COMSOL, mol is used instead of CFU in this
study. This is only a difference in measurement units, and there is no numerical
difference between [mol/m³] and [CFU/m³] in numerical calculations.
(3) Wall Adhesion Kinetics: The adsorption-desorption process of the bacterial
solution on the intestinal wall conforms to the Langmuir adsorption model, and the
desorption rate is enhanced by the wall shear stress (regulated by the shear sensitivity
coefficient α).
1.4 Numerical Calculation
(1) Moving Mesh Compatibility:The computational mesh required for solving all
control equations of the fluid flow model is created using the built-in mesh generation
code in COMSOL Multiphysics® software. The mesh deforms synchronously with the wall
peristalsis, and the "deformed domain" physics field is used to update the mesh
displacement. The mesh smoothing type is Yeoh, and the boundary layer mesh is refined to
capture the wall shear effect. There are 16,972 mesh vertices, including 22,279
triangular elements, 4,578 quadrilateral elements, and 26,857 domain elements.
(2) Time Step and Convergence:An adaptive time step scheme is adopted to
automatically adjust the solver time step size to maintain an absolute tolerance of ≤
0.001. The total duration of the transient calculation is 600 s, and the output results
are stored with a time step of 0.1 s. The PARDISO linear solver is used, and Anderson
acceleration is applied for nonlinear iterations to stabilize convergence.
(3) Physics Field Coupling:Four physics fields are set up: laminar flow, boundary
ordinary differential and algebraic equations (describing the adsorption-desorption
process), dilute substance transport, and moving mesh. All four physics fields are
involved in coupling during numerical calculations.
2. Global Definitions and Model Parameters
2.1 Core Physical Parameters
This numerical model is constructed using COMSOL Multiphysics® 6.3 software, covering
four core physics fields: moving mesh (simulating intestinal wall motion), laminar flow
(simulating fluid flow), dilute substance transport (simulating substance diffusion and
convection), and boundary ordinary differential equations (simulating wall
adsorption-desorption kinetics). The model adopts 2D axisymmetric geometric
simplification (ignoring circumferential changes) to balance computational efficiency
and simulation accuracy. The key parameters are detailed in the following tables.
Parameters for Geometry, Fluid Properties, Laminar Flow, Moving Mesh, and Dilute
Substance Transport
| Name |
Expression |
Unit |
Value |
Description |
| L |
1.225 |
[m] |
1.225 m |
Tube length (changed to 1.80 [m] for the ileum) |
| rho |
1040 |
[kg/m^3] |
1040 kg/m³ |
Fluid density |
| mu |
0.0014 |
[Pa*s] |
0.0014 Pa·s |
Dynamic viscosity |
| amp |
0.73 |
- |
0.73 |
Relative contraction amplitude |
| tc |
7 |
[s] |
7 s |
Period |
| ta |
1.75 |
[s] |
1.75 s |
Contraction gating start time |
| tb |
5.25 |
[s] |
5.25 s |
Relaxation gating start time |
| thalf |
1.75 |
[s] |
1.75 s |
Gating smoothing time |
| z0 |
L/2 |
[m] |
0.6125 |
Peristalsis center (0.90 [m] for the ileum) |
| up |
7.50E-04 |
[m/s] |
7.5E-4 m/s |
Peristaltic propulsion velocity |
| sigma2 |
2.00E-04 |
[m^2] |
2E-4 m² |
Gaussian variance (= 2 cm²) |
| R0 |
0.00865 |
[m] |
0.00865 m |
Intestinal radius |
| zc |
0.05*L |
- |
0.06125 m |
Axial position z of the bolus center |
| rc |
0 |
[m] |
0 m |
Radial position r of the bolus center (m); set to 0 when placed near the axis
|
| az |
0.02 |
[m] |
0.02 m |
Semi-axis of the bolus in the z-direction (m) |
| ar |
0.8*R0 |
- |
0.00692 m |
Semi-axis of the bolus in the r-direction (m) |
| sm |
0.5*min(az,ar) |
- |
0.00346 m |
Smoothing half-width to avoid oscillations caused by hard edges |
| lambda |
0.07 |
[m] |
0.07 m |
Segmentation wave wavelength |
| xigema |
0.25*R0 |
- |
0.0021625 m |
Segmentation wave amplitude |
| f |
10 |
[1/min] |
0.16667 1/s |
Segmentation frequency |
| a |
8.5*lambda-d |
- |
0.59 m |
Segmentation wave smoothing start point |
| b |
9*lambda+d |
- |
0.635 m |
Segmentation wave smoothing end point |
| d |
0.005 |
[m] |
0.005 m |
Segmentation smoothing transition half-width |
| dPdz |
-0.04 |
- |
-0.04 |
Target axial pressure drop [Pa·s] |
| target |
5.64E-08 |
[m^3] |
5.6415E-8 m³ |
Target cumulative volume per second |
| L1 |
0.245 |
[m] |
0.245 m |
Duodenal length |
| L2 |
0.159 |
[m] |
0.159 m |
Stomach length |
| R1 |
0.0386 |
[m] |
0.0386 m |
Stomach fundus width |
| L1.5 |
0.015 |
[m] |
0.015 m |
Pyloric region length |
| A |
0.9 |
[cm] |
0.009 m |
Gastric peristaltic wave amplitude |
| v |
2.34 |
[mm/s] |
0.00234 m/s |
Gastric peristaltic wave propagation velocity |
| L4 |
2.04 |
[cm] |
0.0204 m |
Gastric peristaltic wave width |
| delta |
0.1*L4 |
- |
0.00204 m |
Spatial smoothing parameter, L is half the period length |
| delta_t |
0.5 |
- |
0.5 |
Temporal smoothing parameter |
| tanth |
L2/(R1-R0) |
- |
5.3088 |
Tangent value |
| sinth |
abs(tanth)/sqrt(1+tanth^2) |
- |
0.98272 |
Sine value |
| R2 |
1.05 |
[cm] |
0.0105 m |
Duodenal radius |
| t0 |
0 |
- |
0 |
Time when the duodenum starts to change |
| tau_on |
2 |
[s] |
2 s |
Temporal smoothing activation duration |
| zw |
0.05*L1 |
- |
0.01225 m |
Spatial transition length at both ends |
| L3 |
0.002 |
[m] |
0.002 m |
Thin domain length |
Adhesion Parameters
| Name |
Expression |
Unit |
Value |
Description |
| D |
1.00E-13 |
[m^2/s] |
1E-13 m²/s |
Bulk diffusion coefficient |
| alpha |
1.00E-07 |
[1/Pa] |
1E-7 1/Pa |
Shear sensitivity coefficient |
| kon |
1.00E-05 |
[m/s] |
1E-5 m/s |
Adsorption coefficient |
| koff0 |
1.00E-04 |
[1/s] |
1E-4 1/s |
Basal desorption rate |
| Gmax |
1.00E-05 |
[mol/m^2] |
1E-5 mol/m² |
Maximum adsorption capacity |
| C0 |
1 |
[mol/m^3] |
1 mol/m³ |
Initial concentration in the bolus |
Geometric Model
For numerical purposes, the geometry of the small intestine is assumed to be a perfect
cylinder with flexible walls, and the entire geometry is assumed to be filled with
liquid intestinal contents. The flow in the cylindrical tube is assumed to be
axisymmetric, and the gastric antrum is assumed to have a funnel-shaped feature. All
geometric parameters are based on real physiological data. The geometric model includes
four parts: the gastric antrum (only including the starting point of gastric peristalsis
to the pylorus), the pylorus, the duodenum, and the jejunum (the ileum model is
constructed separately for comparison). The "Form Union" operation in COMSOL is used to
construct a unified 2D axisymmetric computational domain[1][3][5].
(1) Gastric Antrum: The diameter of the upper surface is 1.73 cm, the diameter of the
lower surface is 7.71 cm, and the length is 15.9 cm. Its volume in 3D is consistent with
the typical volume of the gastric antrum.
(2) Pyloric Region: The diameter is 1.73 cm, and the length is 1.5 cm.
(3) Duodenum: The diameter in the relaxed state is 2.1 cm, and the length is 24.5 cm.
(4) Jejunum: The diameter is 1.73 cm, and the length is 120 cm[1][3][5].
Fig 1. 2D and 3D Axisymmetric Model Geometry (1 - Stomach, 2 -
Pylorus, 3 - Duodenum, 4 - Jejunum)
Fig 2. Ideal Gastric Geometry[1]
2.2 Laminar Flow
Model Assumptions
The intestinal contents are assumed to be incompressible Newtonian fluids. Since the
expected Reynolds number (Re) is ~3, the velocity field of the intestinal contents in
the numerical models of the jejunum and ileum generated by peristaltic waves is assumed
to be laminar. The intestinal contents are assumed to be Newtonian fluids with a dynamic
viscosity of 0.0014 Pa·s and a density of 1040 kg/m³[3]. Due to the complex
spatial
position of the small intestine in the human body, the effect of gravity is relatively
small compared to the peristaltic driving force, so the effect of gravity is ignored.
The flow is axisymmetric, and the circumferential (φ) velocity component and
concentration gradient are zero.
Governing Equations of the Laminar Flow Model
The continuity equation and Navier-Stokes equations are used to solve the laminar fluid
flow in the entire domain. The opening side of the gastric antrum and the end of the
jejunum are set as open boundaries. A moving boundary condition is applied to all
boundaries representing the digestive tract wall except the axis of symmetry boundary.
(1) Navier-Stokes Equations
The Navier-Stokes equations solve the gradients of velocity with respect to pressure,
viscous forces, and gravity.In Equation,t is time, p is pressure (Pa), μ is dynamic
viscosity (Pa·s), ∇u is the velocity gradient tensor, (∇u)t is the transpose of ∇u
(s⁻¹), and g is the gravity vector (m/s²). In this study, the effect of gravity on the
flow is ignored. The intestinal contents are assumed to be Newtonian fluids with a
dynamic viscosity (μ) of 0.0014 Pa·s and a density (ρ) of 1040 kg/m³.
(2) Open Boundary Condition
The open edges at both ends of the axisymmetric geometry are defined as open boundaries
to describe the boundaries in contact with a large amount of fluid. The given open
boundary condition is applied to these two edges, indicating that the sum of pressure
and viscous forces is zero. The fluid can flow freely in and out of the domain without
external stress. In the equation, represents the boundary normal pointing outward of the
domain.
(3) Wall Condition
This is the most commonly used wall condition in fluid mechanics. It assumes that at the
contact surface between the fluid and the solid wall, the relative velocity of the fluid
is zero. That is, the fluid molecules close to the wall will "stick" to the wall without
relative sliding with the wall.
Moving Mesh
The dynamic mesh module is used to simulate the deformation of the digestive system wall
caused by peristalsis and segmentation. The motion modes simulated in the model include:
gastric peristalsis, intermittent opening and closing of the pylorus and gastric
emptying, duodenal segmentation, jejunal peristalsis, and jejunal segmentation.
2.3 Gastric Peristalsis
Introduction to Gastric Peristalsis
In general, the motion pattern of the gastric wall can be characterized by two types of
muscle contractions, as described below. The first type of motor activity originates
from the upper part of the stomach and propagates. It is characterized by slow and weak
wave activity, which only produces slight indentations in the gastric wall. The second
type of motor activity is characterized by a series of regular antral contraction waves
(ACWs) that originate from the middle of the stomach and propagate circumferentially
toward the pylorus. These contractions are born as shallow indentations but deepen as
they move downward (nearly blocking the gastric lumen when they approach the pylorus).
Their frequency is approximately 3 cycles per minute, and they are controlled by
electrical stimulation generated by the gastric pacemaker, which is located in the
middle body region of the greater curvature of the stomach.
In the current modeling stage, we have been able to identify most of the parameters
based on literature data. The peristaltic pattern of the gastric wall is characterized
as follows: ACWs are initiated every 20 seconds at a location 15.9 cm from the pylorus,
with a lifespan of 61 seconds[1].
Fig 3. Schematic Diagram of Gastric Peristaltic Waves
The relative occlusion degree is defined as RO=A/D, where D is the distance from the
gastric wall to the central axis, and A is the amplitude of the gastric peristaltic
wave. The variation of RO is specified as follows: from 0% to 40% linearly from 0 s to
19.5 s, remains constant at 40% from 19.5 s to 35.5 s, increases linearly again from
35.5 s to 60 s, reaches a maximum of 80% when advancing toward the pylorus, and finally
decreases to 0 gradually from 60 s to 61 s.
Governing Equations of Gastric Peristalsis
(1) The generation of gastric peristaltic waves is controlled by setting the
displacement of the gastric wall. The normal displacement of the gastric wall is:
L4 is the wavelength of a single peristaltic wave, Δ is the wave interval with a
value
of 20 s,v is the wave velocity, and k is the wave number.
Fig 4. The four figures are respectively the Aof - wave amplitude
function, the Sa - occlusion degree function, the spatiotemporal diagram of gastric
peristaltic waves, and the schematic diagram of 2D planar gastric peristaltic waves. It
can be seen that the amplitude of the peristaltic wave reaches the maximum in the middle
of the gastric antrum.
2.4 Pylorus
Introduction to Pyloric Mechanism
As the outlet of the stomach, under the set parameters, the motor activities of the
gastric antrum, gastric outlet, and duodenum are synchronized. The pylorus acts as a
"soft valve" that opens and closes periodically over time, acting only in a short axial
range (pyloric segment). When the valve contracts, the effective radius of this pyloric
segment decreases, almost blocking the passage from the stomach to the duodenum; when
the valve relaxes, the radius returns to a larger value, and the gastric contents flow
out in a short period of time. This cycle repeats, forming a typical "stepwise" gastric
emptying rhythm.
The pyloric cycle is divided into 4 phases:
(1) Contraction Rising Edge: Gradually shrinks from fully open to the target minimum
radius.
(2) Contraction Holding Phase: Maintains a small radius for a long time (the valve is
"almost closed"), with minimal emptying and strong inhibition of reflux.
(3) Relaxation Falling Edge: Quickly returns from a small radius to a larger radius
(fast opening).
(4) Relaxation Holding Phase: Maintains an open state for a short time, allowing a wave
of contents to pass through quickly.
Fig 5. The inhibitory effect of the pylorus on reflux between the stomach and
duodenum in the 2D plane. Streamlines of fluid flow in the planar region of the
stomach at t = 406.2 s, arrows indicate the direction of velocity, and streamlines
are colored by velocity magnitude (m/s).
Fig 6. The z-component velocity field of the model at t = 499 s.
It can be seen that there is an obvious positive large-velocity flow at the pylorus. The
results shown by the model are consistent with the classic function of physiological
gastric emptying.
Governing Equations of the Pylorus
We model the pylorus as a local radius modulation with a length of L1.5, expressed by a
sine-squared function. Its instantaneous radius is simulated by defining the movement of
the boundary mesh to realize the opening and closing of the pylorus:
The contraction ratio c(a) and its smooth step are:
The axial spatial weight is:
The initial state of the pylorus is relaxed, with a period of 20 s: it contracts
gradually from 0 to 4 seconds, maintains contraction from 4 to 12 seconds, relaxes
gradually from 12 to 18 seconds, and maintains relaxation from 18 to 20 seconds. Under
the set parameters, the motor activities of the gastric antrum, gastric outlet, and
duodenum are synchronized[2]. The geometry of the computational domain
changes
periodically.
Fig 7. Spatiotemporal Diagram of Pyloric Gating
2.5 Jejunal Motion
Jejunal Peristaltic Waves
Three different motion patterns are observed in the small intestine[6] to
promote food
digestion and absorption. Peristalsis is responsible for the propulsion of chyme along
the intestine; segmental contractions involve local contraction and relaxation without
overall propulsion along the intestine, and the model studies their role in the mixing
of chyme with surrounding fluids; pendular contractions involve the "back-and-forth"
motion of the intestinal wall. The model mainly simulates the peristalsis and segmental
contraction processes in the small intestine.
Governing Equations of Jejunal Peristaltic Waves
The peristaltic waves generated in the jejunum are simulated by defining the
displacement of the intestinal wall. The jejunal peristaltic parameters are modeled as a
Gaussian distribution function. The intestinal wall displacement and the step function
used are defined as follows:
The relative amplitude amp of the peristaltic curve and the contraction period tc are
determined based on literature data[3].
Fig 8. Shows a complete typical peristaltic wave in the 2D axisymmetric plane and
the change in the z-component velocity field caused by peristalsis.
Segmental Motion of the Jejunum and Duodenum
Segmental motion refers to the quasi-steady, nearly local reciprocating contraction of
the small intestinal wall on a short axial scale[5]. Its main functions are
to stir and
mix the luminal contents and enhance mass transfer near the wall. Unlike peristaltic
waves that mainly promote propulsion, segmental motion contributes little to the net
propulsive flow but can significantly affect local shear, boundary layer thickness, and
absorption/reaction environment. In the model, segmentation only acts in a preset short
axial interval, and segmentation is regarded as a small radial displacement (the
negative sign indicates contraction); its spatial shape is described by a sine function
in the axial direction and a harmonic function in time; the intensity is clipped by a
soft window to ensure smooth transitions at the ends and avoid numerical oscillations.
In the model settings, segmental motion occurs simultaneously in the jejunum and
duodenum, while peristaltic waves only occur in the jejunum.
(1) Governing Equations of Segmental Motion
Similar to peristaltic waves, the governing equation of segmental motion is the wall
displacement defined on the wall. Segmental motion is regarded as a small radial
displacement field oscillating sinusoidally with time in the interval[5]:
Where Zs is the reference phase origin of the segmental wave in
this segment, and W(z) is a spatial smoothing function (soft window). The soft window is
defined by the difference between smooth steps at both ends:
The image characteristic of the segmentation function is a column
of standing waves with phase changing uniformly with time.
Fig 9. Spatiotemporal diagram of duodenal segmental motion and
schematic diagram of the change in the z-component velocity field caused by segmental
motion in the 2D symmetric plane.
2.6 Dilute Substance Transport
Model Introduction
The dilute substance transport model is the core module for quantifying the
convection-diffusion behavior of the bacterial solution (using mol instead of CFU as the
measurement unit, with consistent numerical calculation logic) in the intestine. Its
core goal is to couple fluid flow (laminar flow) with the dynamic deformation of the
intestine (moving mesh) to simulate the process of the bacterial solution from initial
release, propulsion with intestinal contents, to radial diffusion. The model is based on
the "dilute substance assumption" (the concentration of the product bacterial solution
is extremely low, which does not change the fluid density and viscosity), only
considering molecular diffusion and convective transport, and realizing the coordinated
calculation with laminar flow and moving mesh through multi-physics coupling. Finally,
it outputs evaluation indicators such as the spatiotemporal distribution of product
concentration, centroid trajectory, and centroid velocity.
Model Assumptions and Core Objectives
(1) Model Assumptions
a. Dilute Substance Characteristics: The initial concentration of the bacterial solution
is 1 mol/m³, which is much lower than the fluid saturation concentration. Its presence
does not affect the density (1040 kg/m³) and dynamic viscosity (0.0014 Pa·s) of the
intestinal contents, meeting the physical simplification conditions of "dilute
substances".
b. Transport Mechanism: The transport of the bacterial solution in the intestine is
dominated only by molecular diffusion and convection, and turbulent diffusion is ignored
(since the Reynolds number of intestinal flow is Re ≈ 3, which belongs to the laminar
flow category).
c. Initial Distribution: The bacterial solution is initially released in the form of an
"elliptical bolus" at the jejunal axial position zc = 0.06125 m and radial position rc =
0 (near the axis). The semi-axes of the bolus in the axial (z) and radial (r) directions
are az = 0.02 m and ar = 0.00692 m, respectively. The smooth function flc2hs is used to
avoid numerical oscillations caused by hard edges (smoothing half-width sm = 0.00346 m).
d. Physics Field Coupling: The convective velocity field of the bacterial solution
directly references the solution results of the laminar flow module (u_fluid, w_fluid),
and the influence of mesh deformation on the concentration field is corrected by the
displacement field (d(r, t), d(z, t)) of the moving mesh module to ensure the accuracy
of concentration calculation in the dynamic intestinal environment.
(2) Core Objectives
a. Quantify the spatiotemporal concentration distribution of the bacterial solution
under the synergy of small intestinal peristalsis, segmental motion, and gastric
intermittent emptying mechanism, and evaluate the influence of intestinal motion on the
mixing and propulsion of the product bacterial solution.
b. Calculate the axial centroid trajectory and axial dispersion coefficient of the
bacterial solution, and verify the reproducibility of the model for the physiological
function of "peristaltic propulsion - segmental mixing".
c. Provide the bacterial solution concentration boundary condition at the wall for the
wall adsorption-desorption kinetics (boundary ODE module) to realize the full-process
coupling of "transport - adsorption".
Governing Equations and Transport Mechanisms
The transport of the bacterial solution in the intestine follows the
convection-diffusion equation. Considering the influence of moving mesh deformation on
the concentration field (introducing the time derivative term of mesh displacement), its
general form is as follows:
Where, Ji is the diffusion flux, c is the concentration of the bacterial solution, D is
the
molecular diffusion coefficient of the bacterial solution in the intestinal contents
(m²/s), u is the velocity vector, and Rc is the reaction term. There is no chemical
reaction in the main domain, and the adsorption-desorption reaction at the wall will be
set separately through the boundary reaction, see Section 2.7 Boundary Ordinary
Differential and Algebraic Equations.
Mesh Deformation Correction Term u
Due to the dynamic deformation of the mesh caused by the peristalsis and segmental
motion of the intestinal wall, when COMSOL handles substance transport in the moving
mesh, it will automatically introduce a mesh velocity term into the concentration
equation to correct the calculation of the time derivative. The corrected concentration
time change rate is:
ct is the intrinsic time derivative of concentration, cr and cz are the spatial
gradients of concentration in the radial and axial directions, respectively. (∂r/∂t) and
(∂z/∂t) are the radial and axial displacement rates of the mesh output by the moving
mesh module. This correction ensures that the mesh deformation does not cause "spurious
diffusion" in the concentration calculation.
Parameter Definition
See Section 2.1.1
Initial Concentration Distribution
It is assumed that the product passes through the stomach and duodenum under the
protection of the outer coating and is released exactly at the entrance of the jejunum.
The initial concentration of the bacterial solution is defined by an analytical
function, adopting the form of "elliptical constraint + smooth step" to avoid numerical
instability caused by sudden changes in the initial concentration. The expression at t =
0 s is as follows:
flc2hs is a built-in smooth Heaviside function in COMSOL. When the value in the
parentheses is ≥ 0 (within the elliptical range), the function value approaches 1, and
the concentration of the bacterial solution is C0 (C0 = 1 [mol/m³]); when the value is <
0 (outside the elliptical range), the function value smoothly transitions to 0,
ensuring that there is no hard boundary in the initial concentration field.
Fig 10. Spatial distribution of product concentration in
the 2D symmetric plane at t = 0, 17, 30 s.
Boundary Condition Settings
According to the physiological function of the intestine and the coupling
requirements of the model physics field, the boundary conditions of the dilute
substance transport module are divided into 3 categories, covering all geometric
boundaries.
(1) No-Flux Boundary
A no-flux boundary is set on the solid walls without bacterial solution
adsorption/release, such as the gastric wall, duodenal wall (non-adsorption
region), and pyloric wall. The diffusion flux is 0, that is, the concentration
gradient perpendicular to the wall is 0.
This means that the bacterial solution cannot diffuse through these walls and
only moves with the fluid along the direction parallel to the wall or diffuses
radially.
(2) Open Boundary
Open boundary conditions are set at the inlet and outlet of the overall
digestive tract model, i.e., the end of the jejunum and both sides of the
gastric antrum opening.
The external concentration is set to 0.
To avoid reflux errors caused by sudden concentration changes at the boundary,
the "upwind" numerical scheme is coupled, and the concentration value in the
inflow direction is preferentially used for calculation to ensure flux
continuity.
(3) Surface Reaction Boundary
A surface reaction boundary is set on the jejunal wall to simulate the
intestinal mucosa region where the bacterial solution can be adsorbed. The flux
of the bacterial solution is determined by the "adsorption-desorption kinetics".
The dilute substance transport module only provides the bacterial solution
concentration cw at the wall, and the flux calculation is dominated by the
boundary ordinary differential and algebraic equation module.
The net flux of the bacterial solution at the wall is equal to the difference
between the adsorption flux (Jads) and the desorption flux (Jdes). The specific
kinetic equation is detailed in Section 2.7 (Boundary ODE Module).
Multi-Physics Coupling
The dilute substance transport module needs to be directly coupled with the
laminar flow and moving mesh modules, and the coupling with the boundary
ordinary differential and algebraic equations will describe the bacterial
solution adsorption-desorption process (see Section 2.7). The coupling logic is
realized through the "Reactive Flow, Dilute Species" in the "Multiphysics" of
COMSOL. The specific methods are as follows:
(1) Coupling with Laminar Flow
The "Fluid Flow (spf)" interface is adopted, and the velocity field (u_fluid,
w_fluid) solved by the laminar flow module is directly substituted into the
convection term u*(c∇) of the convection-diffusion equation.
(2) Coupling with Moving Mesh
d(z, t) of the moving mesh module is substituted into the concentration time
derivative correction term:
Numerical Calculation
The "linear Lagrangian element" is used to match the velocity discretization of
the laminar flow module to ensure the stability of the convection term
calculation. Consistent with the overall transient settings, an adaptive time
step is adopted, with a maximum step size of 0.1 s and a minimum step size of
1×10⁻⁴ s. The convergence criterion is "concentration absolute tolerance ≤
0.001". The PARDISO direct solver is used to handle the sparse matrix of the
concentration field to ensure the solution accuracy in the high-gradient region
near the boundary layer.
Indicator Settings
The transport characteristics of the bacterial solution are analyzed from three
dimensions: concentration spatiotemporal distribution, centroid trajectory, and
axial dispersion coefficient to verify the physiological rationality of the
model. See Section 3. Results and Analysis for the calculation result analysis.
2.7 Boundary Ordinary Differential and Algebraic Equations
Overview of the ODE/DAE Module
The boundary ODE/DAE module is the core module for quantifying the interaction between
the bacterial solution and the intestinal wall. Its core function is to describe the
dynamic equilibrium process of "adsorption-desorption" of the bacterial solution at the
wall adsorption sites after reaching the intestinal wall in the dilute substance
transport module, based on the Langmuir adsorption model and shear-enhanced desorption
mechanism. This module only acts on the preset adsorption boundary (jejunal wall, see
Section 2.6.5.1 Surface Reaction Boundary). Through coupling with the dilute substance
transport module and laminar flow module, it realizes the quantitative transfer of
"bacterial solution concentration → adsorption flux → wall adhesion amount", and finally
outputs key indicators such as total wall adhesion amount, coverage area ratio, and net
flux, providing a direct basis for evaluating the retention and distribution effect of
the product bacterial solution in the intestine and a theoretical basis for the study of
drug dosage and action sites.
Variables, Core Objectives, and Key Assumptions
(1) Variable Settings
| Name |
Expression |
Description |
| Gamma |
- |
Wall adsorption amount |
| tauw_alt |
spf.mu*spf.sr |
Viscosity × shear rate (Pa) |
| koff |
koff0*(1 + alpha*abs(tauw_alt)) |
Shear-enhanced desorption |
| theta |
min(1,max(0,Gamma/Gmax)) |
Adsorption site occupancy rate |
| Jads |
kon*c*(1 - theta) |
Adsorption rate |
| Jdes |
koff*Gamma |
Desorption rate |
| M_Gamma |
intop3( 2*pi*r*Gamma ) |
Total wall adhesion amount |
| CoverAreaFrac |
intop3( 2*pi*r*flc2hs(Gamma/Gmax - 0.2, 1e-3) ) / intop3( 2*pi*r ) |
Coverage area ratio: the proportion of the area where Γ reaches the threshold
(the threshold can be changed, e.g., 20%) |
| z_bar |
intop3( 2*pi*r*z*Gamma ) / intop3( 2*pi*r*Gamma ) |
Coverage center position (where it is most concentrated) |
| J_net |
intop3( 2*pi*r*(Jads - Jdes) ) |
Net flux (current intensity of deposition - desorption) |
| M0 |
intopD( 2*pi*r*an3(r,z) )[mol/m^3] |
Initial total dosage (mol) |
| AttachYield |
M_Gamma / M0 |
Adhesion yield (how much of the current dose remains) |
(2) Core Objectives
Establish a kinetic model for the adsorption-desorption of the bacterial solution on the
intestinal wall, and quantify the adsorption rate (Jads), desorption rate (Jdes), and
net flux (Jnet); reveal the enhancing effect of intestinal wall shear stress on the
desorption process, and clarify the influence of the shear sensitivity coefficient (α)
on the adhesion stability; calculate the spatiotemporal distribution of the wall
adsorption amount (Γ) in the "wall domain", and evaluate the coverage range and
retention efficiency of the bacterial solution in the intestine.
(3) Key Assumptions
a. Monolayer Adsorption: The adsorption sites on the intestinal wall are limited and
uniformly distributed. The adsorption of the bacterial solution follows the Langmuir
model, with a maximum adsorption capacity of Gmax (no multi-layer adsorption).
b. Shear-Enhanced Desorption: The wall shear stress (τw) generated by the flow of
intestinal contents will break the binding between the bacterial solution and the wall,
leading to an increase in the desorption rate with the increase of shear stress. The
quantitative relationship is related through the shear sensitivity coefficient α.
c. Non-Renewable Adsorption Sites: Within the model simulation duration (600 s), it is
assumed that the adsorption sites on the intestinal wall do not regenerate or desorb,
and only the cycle of "unoccupied sites → adsorption → desorption → empty sites" is
considered.
d. Coupling Dependence: The adsorption rate depends on the bacterial solution
concentration at the wall (cw) provided by the dilute substance transport module (see
Section 2.6.5.3), the desorption rate depends on the wall shear rate (sr) provided by
the laminar flow module, and mesh deformation does not affect the distribution of
adsorption sites (the moving mesh module only corrects the geometric position and does
not change the wall adsorption characteristics).
Governing Equations and Kinetic Mechanisms
(1) Core Equation of Adsorption-Desorption Kinetics
The adsorption-desorption process of the bacterial solution on the intestinal wall
satisfies "dynamic equilibrium". The time change rate of the wall adsorption amount Γ
(Gamma) is equal to the net flux (Jnet), which is the core governing equation of the
boundary ODE:
Where, Jads is the adsorption flux, describing the migration rate
of the bacterial solution from the fluid phase to the wall phase; Jdes is the desorption
flux, describing the detachment rate of the bacterial solution from the wall phase to
the fluid phase.
(2) Adsorption Flux: Langmuir Adsorption Model
The adsorption process is controlled by both the "bacterial solution concentration at
the wall" and the "proportion of unoccupied adsorption sites".
Where, kon is the adsorption rate constant (m/s), reflecting the
binding ability between the bacterial solution and the wall sites; cw is the
concentration of the bacterial solution at the wall, provided in real-time by the dilute
substance transport module, i.e., the calculation result of the bacterial solution
concentration c at the jejunal wall; θ is the adsorption site occupancy rate; Gmax is
the maximum adsorption capacity.
(3) Desorption Flux: Shear-Enhanced Mechanism
The desorption process is divided into two parts: "basal desorption" and "shear-enhanced
desorption". Mechanical shear will enhance desorption (shear force promotes the
detachment of the bacterial solution from the intestinal wall).
Where, koff0 is the effective desorption rate, which changes with
shear stress; koff0 is the basal desorption rate, i.e., the desorption ability without
shear; is the shear sensitivity coefficient, quantifying the enhancing effect of shear
on desorption; τw is the wall shear stress (Pa), which is approximately calculated in
the laminar flow module in the model.
Where, μ is the fluid dynamic viscosity, and sr is the shear rate.
Initial Conditions
At the initial moment of the model (t = 0 s), there is no bacterial solution adsorption
on the intestinal wall, and the initial values of the adsorption amount and adsorption
rate are set to 0, which is consistent with the physiological initial state of "the
bacterial solution has not reached the wall".
Boundary Selection and Multi-Physics Coupling Mechanism
(1) Applicable Boundary
Only the jejunal wall is selected as the boundary. The jejunum is the core region for
intestinal nutrient absorption and bacterial colonization. Selecting this segment of the
wall to simulate adsorption is consistent with the product's target site requirements.
(2) Coupling Logic
It needs to be directly coupled with the dilute substance transport and laminar flow
modules. The coupled data is realized through the "variable transfer" and "multi-physics
interface" of COMSOL. The specific logic is as follows:
| Coupled Module |
Transferred Data |
Transfer Direction |
Physical Meaning |
Implementation Method |
| Dilute Substance Transport (tds) |
Wall bacterial solution concentration cw |
Dilute Substance Transport → ODE/DAE |
Provide the concentration driving term for the adsorption flux Jads |
Reference the concentration calculation result c at the jejunal wall boundary of
the dilute substance transport module |
| Dilute Substance Transport (tds) |
Net flux Jnet |
ODE/DAE → Dilute Substance Transport |
Provide the wall flux boundary condition for the dilute substance transport |
Substitute Jnet into the jejunal surface boundary of the dilute substance
transport |
| Laminar Flow (spf) |
Wall shear rate sr |
Laminar Flow → ODE/DAE |
Provide the shear driving term for the effective desorption rate koff |
Reference the shear rate variable sr of the laminar flow module, and
calculateτw in combination |
(3) Verification of Coupling Equations
After coupling the dilute substance transport module, the complete format of the
bacterial solution flux balance equation at the jejunal wall boundary is:
The left side is the wall diffusion flux of the dilute substance transport module
(positive when pointing to the wall), and the right side is the net adsorption flux of
the ODE/DAE module. The equality of the two ensures the mass conservation of the "fluid
phase - wall phase", verifying the rationality of the coupling logic. n*(-D∇c)represents
the diffusion of the bacterial solution through the jejunal wall, i.e., entering the
"wall domain" from the "intestinal domain" or vice versa.
Numerical Calculation Settings
(1) Discretization: The shape function adopts the "discontinuous Lagrangian" quadratic
element.
(2) Linear Solver: The PARDISO direct solver of the overall model is used to handle the
sparse matrix equation of the adsorption amount Γ, avoiding the convergence risk of
iterative solution.
Indicator Settings
Based on the transient simulation results of COMSOL Multiphysics 6.3 (600 s, output time
step of 0.1 s, solution time: 13188 s), combined with the physiological function of the
intestine, the analysis is carried out from three core dimensions: the hydrodynamic
characteristics of the model, the transport behavior of dilute substances, and the wall
adsorption kinetics. The influence of the synergistic effect of peristalsis and
segmentation on product migration, mixing, and wall retention is quantified, the
reproducibility of the model for physiological processes is verified, and data support
is provided for the prediction of product action sites and dosage optimization.
3. Results and Analysis
Based on the transient simulation results of COMSOL Multiphysics 6.3 (600 s, output time
step of 0.1 s, solution time: 13188 s), combined with the physiological function of the
intestine, the analysis is carried out from three core dimensions: the hydrodynamic
characteristics of the model, the transport behavior of dilute substances, and the wall
adsorption kinetics. The influence of the synergistic effect of peristalsis and segmentation
on product migration, mixing, and wall retention is quantified, the reproducibility of the
model for physiological processes is verified, and data support is provided for the
prediction of product action sites and dosage optimization.
3.1 Hydrodynamic Characteristics of the Model
Definition of Global Flow Rate and Monitoring Position
To test the ability of the model to reflect the basic characteristics of fluid transport
in the digestive system, global flow rate variables such as inflow (jejunal inflow),
outflow (jejunal outflow), infloww (gastric inflow into the pylorus), and outfloww
(pyloric inflow into the duodenum) are defined in the intestinal model. These variables
are the core indicators for quantifying fluid retention, propulsion efficiency, and
inter-regional mass exchange in the intestine. These variables are defined through the
"Global Equations" and "Integration Coupling" functions of COMSOL, directly related to
the fluid flow (laminar flow module) and geometric boundaries (such as the jejunal
inlet/end, pyloric interface). Their dynamic changes can verify the mass conservation of
the model, reflect the regulatory effect of peristalsis and segmentation on the flow
rate, and provide data support for the "fluid-driven mechanism" of product transport.
| Global Variable Name |
Expression |
Monitoring Boundary |
Physical Meaning |
| inflow (Jejunal Inflow) |
intop1(2pir*w_fluid) |
Jejunal Inlet |
Instantaneous volume flow rate of fluid entering the jejunum from the
duodenum/pylorus |
| outflow (Jejunal Outflow) |
intop2(2pir*w_fluid) |
Jejunal Terminal Outlet |
Instantaneous volume flow rate of fluid flowing out of the model from the
jejunum |
| infloww (Gastric Inflow into Pylorus) |
intop5(2pir*w_fluid) |
Stomach-Pylorus Interface |
Instantaneous volume flow rate of fluid entering the pylorus from the gastric
antrum |
| outfloww (Pyloric Inflow into Duodenum) |
intop4(2pir*w_fluid) |
Pylorus-Duodenum Interface |
Instantaneous volume flow rate of fluid entering the duodenum from the pylorus
|
| netflow (Jejunal Net Flow) |
netflowt - (outflow + inflow)/2 |
Global Integration (Jejunal Domain) |
Cumulative net retention of fluid in the jejunum (inflow - outflow) |
| netfloww (Pyloric Net Flow) |
netflowwt - (outfloww + infloww)/2 |
Global Integration (Pyloric Domain) |
Cumulative net retention of fluid in the pylorus (gastric inflow - duodenal
outflow) |
Flow Direction: w_fluid is the axial (z-direction) velocity component. A positive value
indicates the positive direction along the z-axis (from stomach → pylorus → duodenum →
jejunum). Therefore, a positive value of inflow/infloww represents fluid "entering" the
target region, and a positive value of outflow/outfloww represents "flowing out".
Dynamic Change Law of Flow Rate
(1) Flow Rate Change in the Duodenal Domain
Fig 11. Cumulative Flow Rate in the Duodenum
Further, the mass conservation of the model and the physiological rationality of the
flow pattern are verified through the cumulative inflow and outflow curves at the
jejunal site. The cumulative volume curve of the jejunum shows a typical "step + fine
line" shape: the stepwise jump synchronized every 20 s reveals the intermittent emptying
driven by the pyloric gating; the 6 s fine lines inside the steps result from the
high-frequency modulation of the instantaneous flux by segmental motion. Reflux only
occurs at the moment of valve switching with a limited amplitude, indicating that the
valve control has a significant inhibitory effect on reflux. The slow fluctuation of the
step height is consistent with the amplitude time course of the gastric antral
peristaltic wave train (lifespan ≈ 61 s). The three curves satisfy the net volume
relationship and rise approximately in parallel, indicating that the mass conservation
is consistent with the boundary flux evaluation.
(2) Flow Rate Change in the Pyloric Domain
Fig 12. Cumulative Flow Rate at the Pylorus
The cumulative efflux (blue), cumulative reflux (green), and cumulative net emptying
(red) at the pylorus show a clear stepwise rise within 0–600 s. The step rhythm is
completely in phase with the 20 s cycle "soft valve" gating in the model: the gating is
applied to the pyloric segment with a length of L1.5 through the boundary displacement
function, and the axial weight of sin2(ℼZ/L1.5) is used to limit the action range. Thus,
the cumulative curve is nearly horizontal during the "contraction holding" period of the
valve and triggers a stepwise jump during the "relaxation holding" period, reproducing
the classic rhythm of intermittent gastric emptying. Fine undulations of approximately 6
s can be seen inside each step, which is consistent with the set segmentation frequency
of the model (f=10 min-1), indicating that the duodenal/jejunal segmental motion causes
high-frequency modulation of the instantaneous flux and near-wall shear rather than
decisive propulsion. It can be obtained that at 600 s, the cumulative efflux volume is
2.3*10-4 m3, the cumulative reflux is 0.65*10-4 m3, and the cumulative net emptying is
1.45*10-4 m3. The reflux only accounts for 30% of the efflux, indicating that the gating
and geometric contraction have a significant inhibitory effect on reflux; the estimated
average net emptying rate is 2.4*10-7 m3s-1 approximately . The three curves maintain
the relationship of "red ≈ blue − green" for a long time and rise approximately in
parallel, indicating that the boundary integral operator used has good mass conservation
and numerical consistency; at the same time, the slow envelope of the jump amplitude of
each step fluctuates with time, which is consistent with the gastric antral peristaltic
wave train lifespan (≈ 61 s) and the time course modulation of the amplitude envelope
Aof, reflecting the synergistic effect of gastric antral peristalsis, pyloric gating,
and segmental motion on multiple time scales.
Dynamic Change Law of Velocity Field
Fig 13. Streamline diagram of the duodenal region near the pylorus at t = 415.2 s,
colored by velocity magnitude.
At t = 415.2 s (pyloric relaxation window), the local velocity field shows a typical
"central jet + bilateral reflux vortex" structure: the jet is ejected from the narrowest
part of the pylorus and forms a long and narrow high-velocity zone at the duodenal
inlet, then decays rapidly along the radial direction; a pair of near-wall
counter-rotating vortex cells are induced by the shear layer on both sides of the jet,
and the streamlines are closed in the vortex cells, indicating that this region is
mainly for mixing and recirculation and contributes little to net propulsion. The large
shear gradient at the lip of the jet means that the wall shear stress reaches a
relatively high value, which is consistent with the aforementioned division of labor of
"segmentation enhances mixing on a short scale, and gating triggers jet transport during
the window period", and explains the source of stepwise emptying and high-frequency fine
lines in the cumulative curve from the perspective of the instantaneous field.
Fig 14. Streamline diagram of fluid flow in the planar region of the stomach at t =
421.5 s (peristaltic wave approaching the pylorus), colored by velocity magnitude.
The flow field predicted by this model is highly consistent with the classic description
of gastric motion when the contents are Newtonian fluids (0.0014 Pa·s, 1040 kg/m³). In
particular, the flow field predicted in the pyloric antrum region shows two main flow
patterns. When the gastric antral contraction wave moves toward the pylorus, the gastric
contents are forced to reflux. When the gastric occlusion degree reaches 56%, a
retrograde jet-like motion is formed. To quantify the development process of this jet
motion, the maximum retrograde flow velocity was tracked within a 20-second flow cycle.
As the gastric antral contraction wave approaches the pyloric sphincter, the increase in
contraction wave velocity and occlusion degree causes the retrograde jet to accelerate
continuously, reaching a maximum retrograde flow velocity of 8.8 cm/s at the narrowest
part of the pyloric canal. Another main flow pattern identified in the pyloric antrum
region is the annular vortex formed between two adjacent gastric antral contraction
waves (Figure 14). Through strong mixing, friction, and grinding of the gastric
contents, retrograde jet motion and vortex structures have long been considered the main
drivers of gastric digestion of food[1].
3.2 Transport Behavior of Dilute Substances
Spatiotemporal Distribution of Concentration
Fig 15. Shows the spatial distribution of product concentration in
the 2D symmetric plane at t = 0, 150, 300, 450, 600 s.
Over time, under the combined action of peristaltic propulsion and segmental mixing, the
concentration field gradually expands along the axial direction and diffuses radially.
It shows monotonic dilution under the combined action of convection dominance,
diffusion, and segmental mixing. At the same time, the attenuation of the bulk
concentration is also continuously absorbed by the wall adsorption-desorption flux.
Centroid Position
Fig 16. Product Centroid Position
This figure shows the time evolution of the axial centroid of the bolus. The curve shows
a typical "step + fine line" characteristic: the large-scale steps are strictly in phase
with the 20 s pyloric gating, indicating that dzc/dt increases significantly during the
relaxation window and almost stagnates during the contraction holding period; the
high-frequency fine teeth inside the steps are synchronized with the set segmentation
frequency, reflecting the periodic modulation of segmental motion on near-wall mixing
and instantaneous convection rather than decisive propulsion. The overall monotonic rise
and no continuous retraction indicate that the net propulsion is dominated by gating,
and reflux only occurs at the moment of opening and closing switching (consistent with
the aforementioned "cumulative volume step"); the slight envelope fluctuation of the
curve on a longer time scale is consistent with the modulation of the gastric antral
peristaltic wave train lifespan and amplitude envelope Aof. In summary, the behavior of
the centroid is highly consistent with the rhythm of volume efflux: gating determines
the migration "jump distance", segmentation accelerates spreading and mixing, and
adsorption/desorption only changes the bulk depletion and spatial distribution, but does
not change the forward direction of the centroid.
Axial Dispersion Coefficient
Fig 17. Axial Dispersion Coefficient
The axial dispersion coefficient shows "a slowly increasing envelope + obvious periodic
oscillation": within 0–600 s, its peak-to-peak amplitude and positive half-cycle average
value continue to increase, indicating that the bolus advances into a region with
stronger shear/stronger stirring (the central jet generated by pyloric windowing and
downstream segmentation jointly enhance axial spreading); the large-scale oscillation in
groups of 20 s is in phase with the pyloric gating: positive peaks appear in the
relaxation window, and it is close to zero or negative valleys appear in the contraction
holding period. The high-frequency fine lines (6 s) superimposed in the steps come from
the modulation of segmental motion on instantaneous mixing. The negative value in the
figure is not "physically negative diffusion", but "the instantaneous axial spreading
rate is negative, which is consistent with the aforementioned "step + fine line" rhythm
of centroid stepwise forward movement and cumulative volume.
3.3 Wall Adsorption Kinetics
Wall adsorption is the core link for the product bacterial solution to achieve retention
and action in the intestine. Based on the Langmuir adsorption model and shear-enhanced
desorption mechanism, key indicators such as total adhesion amount, coverage range, and
net flux are used to quantify the dynamic behavior of the bacterial solution at the
jejunal adsorption boundary, reveal the coupling law of "fluid shear - wall
concentration - adsorption equilibrium", and verify the reproducibility of the model for
the intestinal adsorption process. See Section 2.7.2.1 for evaluation indicators and
calculations.
Total Wall Adhesion Amount
Fig 18. Wall Adhesion Amount
Within 0–600 s, the total wall adhesion amount MΓ increases monotonically with
superimposed
20 s serrated oscillations; by 600 s, it has reached approximately 5.6×10−7mol,
indicating that under the current dosage and parameter set, net adsorption is dominant
for a long time, and the bulk phase is continuously "extracted" to the wall without
obvious global saturation (Г/Gmax has not approached 1). The serrations are in phase
with the pyloric gating: the relaxation window triggers a jet → the near-wall
concentration c in the bulk phase increases sharply and the supply is sufficient, but at
the same time, the shear increases to make koff larger; the shear decreases during the
contraction holding period, the recirculation is enhanced, koff decreases, and Jads
continues to work in the medium shear zone. The result of the alternation of the two
phases is "scouring - redeposition" in each 20 s cycle, but the net increase is
positive, resulting in the "sawtooth-like rise" shown in the figure. On a longer time
scale, the slope rises slowly, which is consistent with the synergistic enhancement of
near-wall supply and mixing by the gastric antral ACW wave train (61 s) +
duodenal/jejunal segmentation: the stepwise forward movement of the bolus centroid
periodically sends the high-concentration zone to the new downstream wall, and the
product of the available area × near-wall concentration × contact time expands with
time, so MΓ shows a nearly linear - weakly superlinear growth.
Fig 19. Coverage Area Ratio
The coverage area ratio is "the proportion of the intestinal wall area that reaches the
effective coverage threshold. Its formula is:
Where Ajej is the total area of the jejunal wall, Γ is the wall adhesion amount, Gmax is
the maximum wall capacity, andθthr is the set "effective coverage threshold" (usually
0.2, i.e., coverage ≥ 20% is regarded as "covered", that is, the adhesion amount
reaching 20% of the maximum capacity is regarded as covered). Therefore, this indicator
emphasizes the geometric coverage range rather than "how much quantity", which is
suitable for answering where the drug/bacterial solution is distributed and how large
the coverage area is, and is complementary to the total adhesion amount MГ ("how much
quantity"). Figure 17 shows that the coverage area ratio increases monotonically within
0–600 s and presents a "step + fine line" shape: by 600 s, it has reached approximately
0.93, indicating that under the current administration and kinetic parameters, most
areas of the jejunal wall have reached effective coverage. The steps of the curve are in
phase with the 20 s pyloric gating: the jet pulse during the relaxation period quickly
pushes the near-wall high-concentration zone downstream, triggering a large number of
new walls to cross the threshold, and the coverage ratio jumps; the propulsion weakens
during the contraction holding period, and the coverage ratio rises slowly/platforms.
From the perspective of spatial distribution, the curve indicates that the model
successfully reproduces the entire process of "pulsatile transport (gating determines
the jump distance) - local mixing (segmentation determines spreading and near-wall
supply) - wall capture".
Net Flux
Fig 20. Net Flux
The figure shows the global net flux Jnet, which can be understood as the instantaneous
net rate of "adsorption - desorption" on the entire segment of the intestinal wall; it
is the time derivative of the total wall adhesion amount MΓ, so Jnet >0 indicates that
the wall is increasing net. The curve in the figure shows three phases and is
superimposed with periodic textures in phase with physiological driving forces:
(1) 0-60 s: Jnet jumps rapidly from zero, which comes from the initial bolus + the
near-wall high-concentration jet supply generated by the first pyloric relaxation.
(2) 60-350 s: The 20 s sawtooth is in phase with the pyloric gating, and the 6 s fine
teeth come from segmental mixing. The gating "batches" the high-concentration zone to
the downstream, and the segmental motion enhances the near-wall supply. However, with
the increase of the covered area and the continuous extraction of the bulk phase, the
growth of Jads=konc(1-θ) is partially offset, thus entering a quasi-steady state limited
by transport.
(3) 350-600 s: The bulk concentration further decreases, more walls approach the
capacity limit, and the proportion of the desorption term koff=koff0(1+α|ﬢw|) in the
high-shear window of the gating increases relatively. The combination of these factors
leads to a decline in the net adsorption rate.
Adhesion Yield
The total initial administered dose (sum of the bulk phase and initial wall value) is
denoted as M0. The adhesion yield is defined as the normalized proportion of the
cumulative retention on the wall, which measures "the proportion of the initial dose
that has been captured by the wall by time t".
Fig 21. Adhesion Yield
At t=600 s, AttachYield=0.14, indicating that approximately 14% of the initial dose has
been transferred to the wall.
Fig 22. 3D Product Concentration Distribution(t=0-30s)
Fig 23. 3D Product concentration Distribution(t=30-60s)
Fig 24. 3D Product concentration Distribution(t=540-570s)
Fig 25. 3D Product concentration Distribution(t=570-600s)
5. Appendix B
In the process of CFD (Computational Fluid Dynamics) calculations, the setting of the
initial state is highly relevant to the convergence of the computation. To ensure stable
computation, the following adjustments have been made to the initial geometric settings of
the model:
1.The initial relaxed radius of the pyloric region is set to R0 in the design, but in the
actual initial setup, it is set to 0.5R0. Thus, during the first pyloric cycle, the pyloric
radius will contract from R0 to 0.1R0 and then return to 0.5R0 at the start of the next
cycle.
2.The initial radius of the duodenum is set to R0, and it gradually returns to the designed
radius R2 after the computation begins. This adjustment is made to ensure numerical
convergence at the start of the calculation.
[1]Ferrua, M.J. and Singh, R.P. (2010), Modeling the Fluid Dynamics in a Human Stomach to
Gain Insight of Food Digestion. Journal of Food Science, 75:
R151-R162. https://doi.org/10.1111/j.1750-3841.2010.01748.x
[2]Trusov, P. V., Zaitseva, N. V., Kamaltdinov, M. R., A Multiphase Flow in the
Antroduodenal Portion of the Gastrointestinal Tract: A Mathematical
Model, Computational and Mathematical Methods in Medicine, 2016, 5164029, 18
pages, 2016. https://doi.org/10.1155/2016/5164029
[3]J.S. Karthikeyan, Deepti Salvi, Mukund V. Karwe,
Modeling of fluid flow, carbohydrate digestion, and glucose absorption in human small
intestine,Journal of Food Engineering,Volume 292,2021,110339,ISSN
0260-8774,https://doi.org/10.1016/j.jfoodeng.2020.110339.
[4]T. Ishikawa, T. Sato, G. Mohit, Y. Imai, T. Yamaguchi,
Transport phenomena of microbial flora in the small intestine with peristalsis,Journal of
Theoretical Biology,Volume 279, Issue 1,2011,Pages 63-73,ISSN
0022-5193,https://doi.org/10.1016/j.jtbi.2011.03.026.
[5]Jinping Zha, Siyu Zou, Jianyu Hao, Xinjuan Liu, Guillaume Delaplace, Romain Jeantet,
Didier Dupont, Peng Wu, Xiao Dong Chen, Jie Xiao,
The role of circular folds in mixing intensification in the small intestine: A numerical
study,Chemical Engineering Science,Volume 229,2021,116079,ISSN
0009-2509,https://doi.org/10.1016/j.ces.2020.116079.
[6]Palmada, N.; Hosseini, S.; Avci, R.; Cater, J.E.; Suresh, V.; Cheng, L.K. A Systematic
Review of Computational Fluid Dynamics Models in the Stomach and Small Intestine. Appl.
Sci. 2023, 13, 6092. https://doi.org/10.3390/app13106092