Fluid flow and flow-induced shear stress are critical components of the vascular microenvironment commonly studied using microfluidic cell culture models. Microfluidic vascular models mimicking the physiological microenvironment also offer great potential for incorporating on-chip biomolecular detection. In spite of this potential, however, there are few examples of such functionality. Detection of biomolecules released by cells under flow-induced shear stress is a significant challenge due to severe sample dilution caused by the fluid flow used to generate the shear stress, frequently to the extent where the analyte is no longer detectable. In this work, we developed a computational model of a vascular microfluidic cell culture model that integrates physiological shear flow and on-chip monitoring of cell-secreted factors. Applicable to multilayer device configurations, the computational model was applied to a bilayer configuration, which has been used in numerous cell culture applications including vascular models. Guidelines were established that allow cells to be subjected to a wide range of physiological shear stress while ensuring optimal rapid transport of analyte to the biosensor surface and minimized biosensor response times. These guidelines therefore enable the development of microfluidic vascular models that integrate cell-secreted factor detection while addressing flow constraints imposed by physiological shear stress. Ultimately, this work will result in the addition of valuable functionality to microfluidic cell culture models that further fulfill their potential as labs-on-chips.
INTRODUCTION
The forces exerted by flowing blood on vascular cells in vivo play critical roles in regulating vascular cell biology.1 Traditional static in vitro cell culture models fail to replicate fluid flow-induced shear stresses, motivating the development of microfluidic platforms that better mimic the vascular microenvironment.2,3 Indeed, microfluidic platforms have been extensively applied to study the effects of flow-induced shear stress on various cell types.4
Among the many cell behaviours affected by flow-induced shear stress is the secretion of biomolecules that impact cellular signaling and function. For example, nitric oxide, an important mediator of vasodilation, is released by endothelial cells in a shear-dependent manner.5 However, in situ microfluidic biosensing of cell-secreted biomolecules is typically performed under static fluid conditions.6 Assay integration is an oft-cited advantage of microfluidic cell culture devices,7–9 but in situ detection of secreted biomolecules under flow conditions presents significant design challenges due to physiological shear stress constraints, and the coupled nature of fluid flow and mass transport to the biosensor surface. Moreover, flow significantly dilutes cell-secreted factor concentrations, hindering or even preventing detection. Predicting mass transport of secreted factors and the biosensor response is therefore critical to inform the design of microfluidic cell culture systems integrating biosensing capabilities.
Computational modeling is a useful tool for studying physical phenomena, assessing feasibility and guiding design choices in microfluidics with integrated quantitation assays.10,11 The behaviour of surface-based biosensors has been well-characterized using theoretical and computational modeling, which provide excellent insight into convection-diffusion-reaction design considerations.12–14 This analysis, however, is not sufficient to address the constraints of incorporating physiological levels of shear stress in the device.
Computational modeling of biosensing considerations is also generally limited to the scenario of a planar biosensor embedded within a single flow channel. This single layer design requires that the biosensor be directly adjacent to the cell layer, posing practical challenges to isolating cell culture and biosensor regions within the device. An alternative configuration is a bilayer device in which cells are cultured in the upper channel on a porous membrane support (e.g., Boyden chamber) while a surface-based biosensor is located in the lower channel underneath the membrane (Fig. 1A); the model, however, is generally applicable to other common multilayer configurations such as a hydrogel layer between two channels.15,16 The bilayer membrane configuration is widely used in microfluidic cell culture applications,17–20 and is particularly well-suited for microfluidic vascular models, where it has been used to study cancer cell21 and monocyte22 adhesion to an endothelial monolayer, endothelial permeability,23,24 the vascular/valvular microenvironment25 and the blood-brain barrier.26,27 Bilayer culture devices have also been coupled with mass spectrometry for on-line monitoring of drug permeability.28 For integrated biosensing applications, the bilayer configuration is advantageous as it positions the cells and biosensor closely together while isolating the biosensor from the elevated flow rates in the upper channel, allowing cell-secreted factors to accumulate in the lower channel.
Here we present a computational model of fluid flow, mass transport and biosensor kinetics to develop design guidelines that enable both on-chip detection of cell-secreted factors and application of physiological flow-induced shear stress to cultured cells in a bilayer microfluidic device. These guidelines detail the selection of critical design parameters of the bilayer device that will ensure optimal device functionality, which is contingent on meeting several important criteria. First, shear stress in the upper channel must be within physiological limits. Secondly, secreted biomolecule transport rates across the membrane must be maximized in order to efficiently supply analyte to the biosensor. Finally, the biosensor signal itself must also equilibrate rapidly in order to perform measurements within a practical timescale. Meeting these criteria enables the rapid and sensitive detection of secreted biomolecules from cells under a wide range of designer-specified shear stresses, adding novel and valuable functionality to microfluidic vascular models, further fulfilling their potential as lab-on-a-chip systems.
MATERIALS AND METHODS
Computational model description
A schematic of the computationally-modeled bilayer microfluidic device is shown in Fig. 1B. Based on a model previously developed by Young et al.,23 the cell monolayer and porous membrane support were modeled as a uniform porous medium. The microchannels, porous membrane and cell monolayer were layered in the y-direction. A small height-to-width aspect ratio was imposed on the upper and lower microchannel geometry and therefore, coupled with the low Reynolds numbers in the device, no variance of the model in the z-direction was assumed, resulting in a two dimensional model in the xy-plane. For simplicity, the biosensor geometry was modeled as a rectangular strip on the bottom surface of the lower channel. COMSOL Multiphysics 4.2 finite element analysis software was used to model and numerically simulate fluid flow, mass transport and reaction kinetics in the microfluidic device. A more detailed description of the implementation may be found in the ESI.
Governing equations & boundary conditions
The governing equations and boundary conditions used to computationally model the bilayer microfluidic device are described below. A complete list of constant and simulation parameter values may be found in Table I. In order to achieve the desired device functionality, namely the integration of both biosensing and shear stress stimulation in a single device, proper selection of critical design parameter values is necessary. Table II contains a list of the design parameters considered in the guidelines developed in this paper.
Simulation parameter name . | Symbol . | Values . | References . |
---|---|---|---|
Dynamic viscosity | μ | 6.7326×10-4 (Pa·s) at 37°C | |
Cell culture medium density | ρ | 1000 (kg m-3) | |
Hydrostatic backpressure head | Phead | 98.1 (Pa) | |
Cell monolayer concentration | CS | KD = koff/kon (M) | |
Cell monolayer + membrane thickness | δm | 11 (μm) | 23 |
Porosity | εp | 0.0169 | 23 |
Darcy hydraulic permeability | κ | 10-16-10-20 (m2) | 23,32,33 |
Diffusion coefficient | DS | 10-10 (m2 s-1) | 29 |
Biosensor length | Ls | 1000 μm | |
Surface receptor concentration | bmax | 10-7-10-9 (mol m-2) | 34,35 |
Association constant | kon | 104-1012 (M-1s-1) | 36 |
Dissociation constant | koff | 10-3-10-4 (s-1) | 36 |
Simulation parameter name . | Symbol . | Values . | References . |
---|---|---|---|
Dynamic viscosity | μ | 6.7326×10-4 (Pa·s) at 37°C | |
Cell culture medium density | ρ | 1000 (kg m-3) | |
Hydrostatic backpressure head | Phead | 98.1 (Pa) | |
Cell monolayer concentration | CS | KD = koff/kon (M) | |
Cell monolayer + membrane thickness | δm | 11 (μm) | 23 |
Porosity | εp | 0.0169 | 23 |
Darcy hydraulic permeability | κ | 10-16-10-20 (m2) | 23,32,33 |
Diffusion coefficient | DS | 10-10 (m2 s-1) | 29 |
Biosensor length | Ls | 1000 μm | |
Surface receptor concentration | bmax | 10-7-10-9 (mol m-2) | 34,35 |
Association constant | kon | 104-1012 (M-1s-1) | 36 |
Dissociation constant | koff | 10-3-10-4 (s-1) | 36 |
Design parameter name . | Symbol . |
---|---|
Upper channel dimensions | ht, wt, Lt |
Lower channel dimensions | hb, wb, Lb |
Membrane dimensions | wm, Lm |
Transmembrane pressure | ΔPT |
Péclet number | Pep |
Biot number | Bi |
Design parameter name . | Symbol . |
---|---|
Upper channel dimensions | ht, wt, Lt |
Lower channel dimensions | hb, wb, Lb |
Membrane dimensions | wm, Lm |
Transmembrane pressure | ΔPT |
Péclet number | Pep |
Biot number | Bi |
Fluid flow
Flow in the device was divided into three regions: laminar flow in the upper channel generated by an external source (e.g., pump); porous media flow through the uniform porous medium representing the cell monolayer and porous membrane; and laminar flow in the lower channel generated by a positive pressure difference between the upper and lower channel, and where no external flow was applied. The fluid, i.e., cell culture media, was assumed to have the density and dynamic viscosity of water.
Free flow in the upper and lower channels is governed by the Navier-Stokes equations for incompressible flow:
where the density ρ (kg m-3) and dynamic viscosity μ (Pa·s) of the fluid were that of water, u is the velocity field of the fluid within the microchannels and p is the pressure. The upper channel flow boundary conditions consisted of specifying pressure values at the inlet and outlet boundaries. The lower channel backpressure was assumed to be greater than the upper channel backpressure, leading to a hydrostatic pressure head Phead between the upper and lower channel outlets. The lower channel flow boundary condition consisted of specifying Phead at the two lower channel outlet boundaries (i.e., both the left and right side boundaries are outlets), which was selected to be equivalent to 98.1 Pa, the backpressure from an upper and lower channel outlet height difference of 10 mm. A no-slip condition was assumed at all microchannel walls and the biosensor surface.
Porous media flow in the uniform porous layer representing the cell monolayer and porous membrane support is governed by the Brinkman equations:
where εp is the porosity, u are vectors of the velocity field within the porous medium and κ is the Darcy permeability of the porous medium (m2).
Cell layer secretion
No universal model of cell secretion currently exists. Therefore, secretion was modeled as a constant local concentration on the upper boundary of the uniform porous layer in order to study the evolution of the concentration profile within the device and the biosensor response. Biomolecules were assumed to be secreted apically from the cell monolayer, pass across the cell monolayer first, followed by the porous membrane before entering the lower channel.
The concentration of secreted analyte was held constant on the surface of the permeable membrane:
where CS is the apical concentration of analyte over the cell monolayer, Lm is the length of the porous membrane support, hb is the lower channel height and δm is the combined thickness of the cell layer and porous membrane support.
Transport of secreted analyte
Transport of secreted analyte was assumed to be equivalent to the transport of a diluted species since cellular secretion generally results in low secreted species concentrations. The concentration profile within the device is given by the following equation:
where ca is the concentration of analyte (M), and DS is the diffusion coefficient of a typical secreted protein (10-10 m2s-129). The secreted biomolecules were assumed to be smaller in size than endothelial cell tight and leaky junctions (∼2-20 nm30). Cell-biomolecule interactions such as charge, steric and surface receptor binding effects were neglected, and thus solute reflection at the upper cell monolayer surface was also neglected. Oncotic pressure-driven transport was assumed to be negligible due to the protein concentration in the medium being nearly identical on both sides of the porous membrane support and cell monolayer.
No flux passes through the walls of the microfluidic device, leading to the following boundary condition on the device walls where n is the outward unit normal vector:
Biosensor binding kinetics
The biosensor on the bottom surface of the lower channel was modeled as a second order reaction representing surface receptor-biomolecule binding interactions given by the following rate equation:12,31
where bs is the surface concentration of bound receptors (mol m-2), bmax is the total surface concentration of receptors, kon is the association rate constant (M-1s-1) and koff is the dissociation rate constant (s-1).
The binding of analyte to the biosensor necessitates a boundary condition describing the flux of biomolecules to the biosensor surface:
RESULTS AND DISCUSSION
Design consideration: Physiological shear stress in upper channel
Fig. 2A shows the upper channel shear stress at transmembrane pressures from 0 to 1,800 Pa for different upper channel height-to-length ratios ht/Lt. The lower the ratio, the wider the range of transmembrane pressures that may be applied while remaining within physiological shear stress limits. For a given ht/Lt ratio, the presence of backpressure Phead = 98.1 Pa in the lower channel results in a non-zero upper channel shear stress and flow rate being required to produce ΔPT = 0, and eliminate backflow from the lower to upper channel. The threshold shear stress τ0 was defined as the minimum shear stress for which ΔPT > 0 (derivation found in ESI). Fig. 2B represents the plane of possible upper channel shear stresses that may be generated over a range of ht/Lt ratios. The threshold shear stress (Eq. S3 of the supplementary material) separates this plane into a region where conditions for biosensing are met since ΔPT > 0, and a region where biosensing is not possible since ΔPT < 0 and no analyte transport to the lower channel occurs. By referring to Fig. 2B and selecting an appropriate upper channel ht/Lt ratio, a wide range of physiological shear stresses may be generated from low 1-5 dyn cm-2 post-capillary venule shear stress37 to higher 40-60 dyn cm-2 arteriole/capillary shear stress2,37 while also ensuring ΔPT > 0 and enabling biosensing within the device.
These findings lead to the following design guideline that ensures analyte transport to the biosensor and physiological shear stresses: the minimum desired upper channel shear stress, i.e., threshold shear stress τ0, determines the maximum device ht/Lt ratio. Coupled with the constraint wt/ht = 10, all three upper channel microchannel dimensions may therefore be selected. For example, for a minimum desired upper channel shear stress of τ0 = 5 dyn cm-2, the maximum device ht/Lt is ∼0.005 (Fig. 2B). For a typical Lt = 20,000 μm, this ht/Lt ratio yields ht = 100 μm and wt = 1,000 μm. Furthermore, Fig. 2B may be used to establish the ΔPT for different levels of physiological shear stress for ht/Lt between 10-4 to 5×10-3. For example, an arterial stress of 35 dyn cm-2 in a channel with ht/Lt = 0.005 requires ΔPT ∼ 600 Pa (Fig. 2B). Upper channel shear stresses within arteriole shear stress values or lower were used in all subsequent simulations and therefore meet the requirement of physiological shear stress.
Design consideration: Permeability & length of cell monolayer/porous membrane support
The properties of the cell monolayer and porous membrane support have a significant effect on apically-secreted biomolecule transport from the upper to lower channel, as the biomolecules must travel through the cell monolayer. In particular, the permeability of this combined layer to fluid and biomolecules is a critical consideration in the context of analyte transport to the biosensor surface. The hydraulic permeability for a bare porous membrane support is ∼6×10-16 m223 whereas that of a fully confluent endothelial cell monolayer32,33 with tight cell-cell junctions is ∼10-18-10-20 m2.
Fig. 3A illustrates the numerically-simulated laminar flow profile in the upper channel and vertical flow through the cell monolayer-porous membrane support layer in the microfluidic device. Analyte transported from upper to lower channel is indicated by flux lines as well as the device concentration profile where secreted biomolecules are observed to accumulate in the lower channel. Fig. 3B shows simulated time constants for a range of Darcy permeabilities κ and transmembrane pressures ΔPT for a membrane length Lm of 10 mm. The time constant is the time required for the analyte concentration to reach 63% of the final saturating concentration, CS = 1 nM, in the region of interest (ROI). The ROI is defined as the surface of the lower channel as indicated in the inset of Fig. 3B. The biosensor reaction was not modeled in these simulations to examine analyte transport independent from biosensor effects on the local analyte concentration.
For a given ΔPT, there was a significant monotonic increase in time constant value as the permeability decreased (Fig. 3B). Across the range of ΔPT simulated, greater ΔPT led to lower time constants. However, at Darcy permeabilities lower than ∼10-19 m2, increasing ΔPT became a less effective strategy of reducing the time constant even if increased to 1,600 Pa, which corresponds to the upper end of physiologically-relevant shear stresses.
Fig. 3C illustrates the effect of the porous membrane length Lm on the time constants at low and high transmembrane pressures. Larger porous membrane lengths resulted in lower time constants due to the additional surface area available for analyte transport to the lower channel and the larger ΔPT gradient lengthwise across the membrane. This effect is more apparent at low ΔPT where the ΔPT gradient is negligible for smaller Lm, leading to higher time constants. For example, at ΔPT = 0 Pa, Lm = 1 mm leads to larger time constants across all permeabilities. However, as ΔPT increases, the ΔPT across all Lms become sufficiently large to transport analyte to the lower channel ROI at comparable rates, leading to similar time constants. At ΔPT = 1,600 Pa, the time constants for different membrane lengths converge as permeability increases.
These analyses are equally applicable to other multichannel configurations used in microfluidic cell culture, including configurations in which a hydrogel layer (with or without seeded cells) is used instead of a porous membrane support to separate channels.15,16 Furthermore, the model assumes secretion in the apical direction, but could also be modified to study basal secretion where biomolecule transport is not across the entire cell monolayer thickness, but instead mostly across the porous membrane support.
The simulated range of Darcy permeabilities was determined from hydraulic conductivity values, which are used in well-known models of transendothelial transport such as the Kedem-Katchalsky equations38 to describe convective transport of solute across the endothelium. The simulated Darcy permeabilities, however, do not inherently account for further reductions in solute permeability arising from endothelium size-selectivity and semi-permeability, a limitation of the model presented here. Proteins like albumin (69 kDa) can nevertheless still cross the endothelium and are similar in size to the tight junctions between endothelial cells30 (∼2 nm). Indeed, biomolecules secreted in the vascular environment3 are often smaller than albumin, including endothelin-1 (2.5 kDa), nitric oxide (30 Da) and transforming growth factor-β (25 kDa). Moreover, Young et al. previously experimentally validated an analytical model of transendothelial transport based on Darcy’s law for endothelial permeability measurements23 where the cell monolayer and porous membrane support were also considered a uniform porous medium.
Design consideration: Rapid analyte transport
We studied the relative effects of convective and diffusive transport rates through the porous membrane by considering the pore Péclet number Pep, which combines the critical mass transport parameters for porous media flow into a single dimensionless quantity:
A pore Péclet number Pep << 1 corresponds to diffusion-dominated transport and Pep >> 1 corresponds to convection-dominated transport. A realistic range of possible Pep values was determined to be ∼10-3-102 based on the parameter space explored in Fig. 3. Fig. 4 shows the time constants for the aforementioned range of Pep values and different membrane lengths Lm between 0.5-10 mm. Transport is diffusion-limited for Pep < 10-2, and further decreases in Pep did not significantly increase the time constant. For convection-dominated transport (Pep > 10), further increases in Pep did not significantly decrease the time constants. Thus, the greatest reduction in time constant values for a given membrane length occurred in the transition region, with Pep values of 10-2 to 101. In order to design for rapid analyte transport, Pep > 10 in the convection-dominated region is ideal. However, given the constraints imposed by physiological upper channel shear stress and cell monolayer permeability, it is more realistic to design for Pep in the transition region. For example, using Eq. 10, Pep = 1, which corresponds to a time constant of ∼700 s, requires ΔPT ∼1,100 Pa for κ = 10-18 m2, which leads to an upper channel shear stress of 36 dyn cm-2 for ht/Lt = 3×10-3. In terms of selecting a membrane length Lm, for Pep > 1, time constants are relatively independent of Lm while for Pep < 1, larger Lm lead to lower time constants. Thus, longer Lm improve analyte transport times when lower Pep are required.
Design consideration: Factors affecting the biosensor response
We next considered the effect of a surface-based biosensor on the local analyte concentration in the lower channel on the complete convection-diffusion-biosensor system. The porous membrane length, surface concentration of receptors and dissociation rate constant were fixed at Lm = 10 mm, bmax = 10-7 mol m-2 and koff = 10-3 s-1, respectively. The cell monolayer concentration CS was set as the equilibrium dissociation constant KD value where, upon biosensor signal equilibration, 50% of the surface receptors are bound. KD is an important measure of binding affinity where a lower KD represents a higher affinity. The biosensor was positioned in the lower channel under the membrane edge nearest the upper channel inlet.
In Fig. 5A, the effect of different Pep on the % signal equilibration over time is shown (KD = 100 nM, koff = 10-3 s-1). The % signal equilibration represents the percentage of surface receptors bound at a given time relative to the total number of receptors bound at equilibrium. For larger Pep = 0.5 and 1, representing increased convective transport to the biosensor surface, the % signal equilibration rises to completion more rapidly than when Pep < 0.1, again highlighting the important role of analyte transport to the biosensor surface.
One mechanism by which rapid response times may be achieved even at lower Pep is shown in Fig. 5B. The % signal equilibration increases rapidly even at a low Pep of 10-2 for ca/KD = 10 relative to ca/KD = 1, which are ratios of the local concentration relative to the equilibrium constant. Thus, the concentration of secreted analyte in the lower channel, if larger than the KD value of the receptor, may significantly lower the biosensor signal equilibration time.
Design consideration: Rapid biosensor equilibration times
The affinity-based biosensor’s response depends on numerous parameters including the biosensor surface-bound receptor properties, analyte transport to the surface and local analyte concentration. We studied the effect of convective-diffusive analyte transport and the surface receptor binding rates on the biosensor response, as represented respectively by the pore Péclet number (Eq. 10) and the Biot number:
A Biot number << 1 indicates a reaction-limited system where the reaction timescale greatly exceeds the diffusive transport timescale. Conversely, a Biot number >> 1 indicates a diffusive transport-limited system, which typically occurs with surface receptors that exhibit a high affinity towards its ligand.
Fig. 6 shows the percentage of signal equilibration across a Biot number range spanning six orders of magnitude (Lm = 10 mm, κ = 10-18 m2). A timescale of 1 h was selected to compare the kinetics of the system across different conditions. A higher percentage of equilibration corresponds to a lower response time, a particularly important consideration due to the effect of the cell monolayer in reducing convective transport of the analyte to the biosensor surface. A complete list of parameter values used to compute Bi is detailed in Table S1 of the supplementary material. For each Bi, analyte transport rates to the sensor surface were also varied by examining four different Pep values between 10-3 to 100. Furthermore, for each Biot number, simulations were performed for koff = 10-3 s-1 (Fig. 6A) and koff = 10-4 s-1 (Fig. 6B).
With koff = 10-3 s-1, the % equilibration monotonically decreased as Bi increased for a given Pep value, with significant decreases occurring at Bi ∼ 10-1-100. At high Bi, the % equilibration was very low implying long biosensor equilibration times. Furthermore, the % equilibration increased as the Pep value increased for a given Bi, indicating that larger analyte transport rates resulted in faster biosensor signal equilibration. For example, the biosensor signal equilibration is at 96% completion for Pep = 100 at Bi < 100, but only 40-50% for Pep = 10-3. At Bi > 100, the % equilibration decreases rapidly as Bi increases where, at high Bi, the % equilibration after 1 h is very low even at Pep = 100, the highest value simulated. The same trends were also observed when koff = 10-4 s-1 except that the % equilibrations for a given Bi were lower. For example, at Bi = 100 and Pep = 100, signal equilibration was at 24% completion in Fig. 6B compared to 96% for the equivalent parameters in Fig. 6A, a ∼70% difference. koff thus has a significant effect on biosensor equilibration times.
In designing the device-integrated biosensor, a balance is required between biosensor affinity, surface receptor concentration and equilibration time. Constraints are imposed on the surface receptors and analyte transport within the membrane microfluidic device, described by koff, Bi, and Pep. Given that a surface receptor with koff = 10-4 s-1 yields much slower response times than a receptor with koff = 10-3 s-1, one can infer that koff values ≤ 10-5 s-1 will yield prohibitively lengthy response times even at high Pep. It should be stressed that these equilibration times are an inherent property of receptors with low koff values, and are not unique to this system, as the reaction-limited equilibration time scales proportionally to koff.12 Unlike a single channel device configuration, however, it is not possible to simply increase the flow rate over the biosensor surface in order to increase convective transport and thereby minimize equilibration times. Moreover, the flow rate is ultimately constrained by the need to apply physiological shear stress. Thus, surface-bound receptors with koff >10-4 s-1 are recommended. Based on the significant drop in % signal equilibration after 1 h for Bi > ∼100 in Fig. 6, Bi within 10-1 to 100 or less are recommended to reduce biosensor response times in the microfluidic cell culture model. For a lower channel height hb = 0.1 mm, this allows for combinations of kon and bmax between 104-106 M-1 s-1 and 10-7-10-9 mol/m2, respectively, for both koff = 10-3 and 10-4 s-1, which compare favourably to receptor association constant and surface density values found in the literature.34–36 Receptors with KD = koff/kon values between 1-100 nM and 0.1-10 nM may therefore be used for koff = 10-3 and 10-4 s-1, respectively. Antibodies possess, on average, KD values in the nanomolar range and are considered to exhibit a high level of affinity toward the analyte.39,40 Aptamers for important cell-secreted factors in the vascular environment such as transforming growth factor-beta 1 and tumour necrosis factor-alpha have also been shown to possess KD values in the nanomolar range (1.07 nM41 and 7.0 nM,42 respectively). Thus, it is possible to use receptors that exhibit high affinity towards their ligands while also achieving rapid biosensor equilibration times. Further reductions in the Biot number may be achieved by reducing hb. However, unlike kon and bmax, it is not possible to vary hb by multiple orders of magnitude and thus only limited reductions in Bi are possible. Similarly, the diffusion coefficient Ds for secreted biomolecules remains on the same order of magnitude as the simulated value of 10-10 m2/s. In conjunction with the guidelines previously described, it is therefore possible to design for a specific physiological shear stress range while also maximizing cell-secreted factor transport to the lower channel and minimizing the biosensor response time.
CONCLUSION
We developed a computational model of a microfluidic vascular model integrating flow-induced shear stress and the monitoring of cell-secreted biomolecules using an integrated biosensor. The model was applied to a bilayer configuration that isolates the biosensor from significant analyte dilution, thereby maximizing the concentration of analyte detected, but is readily extended to other multilayer configurations. Using the computational model, we developed guidelines that address both physiological and mass transport constraints as well as enable device operation to be tailored to designer-specified physiological shear stresses while ensuring rapid transport of secreted biomolecules to the biosensor. Special consideration was also given to minimizing the biosensor response time within the mass transport limits of the device configuration. These guidelines may ultimately be used to incorporate physiological shear flow and integrated biosensing into microfluidic vascular models, adding valuable functionality for use in fundamental cell biology and drug screening applications.
SUPPLEMENTARY MATERIAL
See supplementary material for a detailed description of the computational model mesh, derivation of the critical shear stress and biosensor reaction kinetic simulation parameters.
ACKNOWLEDGMENTS
The authors have no conflicts of interest to declare. This work was partially funded by the Natural Sciences and Engineering Research Council of Canada and the Queen Elizabeth II Graduate Scholarship in Science and Technology Program.