Previous Article in Journal
New Insights on Y, La, Nd, and Sm Extraction with Bifunctional Ionic Liquid Cyphos IL 104 Incorporated in a Polymer Inclusion Membrane
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Simulating the Permeation of Toxic Chemicals through Barrier Materials

by
Alex Bicket
1,
Vivian Lau
2 and
Jules Thibault
1,*
1
Department of Chemical and Biological Engineering, University of Ottawa, Ottawa, ON K1N 6N5, Canada
2
Defence Research and Development Canada, Ralston, AB T1A 8K6, Canada
*
Author to whom correspondence should be addressed.
Membranes 2024, 14(9), 183; https://doi.org/10.3390/membranes14090183 (registering DOI)
Submission received: 2 August 2024 / Revised: 21 August 2024 / Accepted: 23 August 2024 / Published: 24 August 2024
(This article belongs to the Section Membrane Applications for Other Areas)

Abstract

:
Chemical warfare agents that are liquids with low vapor pressure pose a contact hazard to anyone who encounters them. Personal protective equipment (PPE) is utilized to ensure safe interaction with these agents. A commonly used method to characterize the permeability of PPE towards chemical weapons is to apply droplets of the liquid agent to the surface of the material and measure for chemical breakthrough. However, this method could produce errors in the estimated values of the transport properties. In this paper, we solved numerically the three-dimensional cylindrical Fick’s second law of diffusion for a liquid permeating through a non-porous rubbery membrane to determine the time the permeating species will emerge on the other side of the polymer membrane. Simulations of different amounts of surface area coverage and the geometries of permeate on the membrane surface indicated that incomplete surface area coverage affects the estimation of the transport properties, making the experimentally determined transport properties unsuitable for predictive use. We simulated different permeation values to determine the factors that most influenced the estimation error and if the error was consistent over different permeate–membrane combinations. Finally, a method to correct the experimentally determined permeability is suggested.

1. Introduction

When a hazardous chemical is released, it often requires people to access the area and put themselves at risk of exposure to the hazard to rescue injured people, investigate the event and its impact, or conduct a clean-up of the hazardous contamination. Elastomeric membranes are used as barrier materials in personal protective equipment (PPE) to protect against skin contact with toxic chemicals, including chemical warfare agents [1]. Non-porous rubbery polymers such as nitrile rubber, silicone rubber, and butyl rubber are common PPE materials, especially for gloves [2] where their flexibility gives the wearer dexterity while still being afforded a good degree of protection. In addition to functional properties, such as flexibility and weight, barrier materials are selected based on their chemical resistance, which is evaluated by measuring breakthrough, permeation, and degradation of the material when exposed to chemicals of interest [1]. Breakthrough time is one measure of how well a given barrier will protect against a chemical hazard, measuring the time it takes a chemical to completely permeate through the material [3], and is used to rate how long a piece of protective equipment will protect the wearer against that chemical.
However, transport properties are difficult to predict and are reliant on the individual system (permeate, membrane, and temperature), requiring physical experiments to derive the protective properties of the materials. In the case of diffusion through non-porous polymer membranes, the solution diffusion model is used to describe the motion of molecules of the permeate into and through the membrane [4,5,6]. Experimental procedures such as those described by ASTM F739, ASTM D6978, EN 374, EN 16523, ISO 6529, or US Army TOP 08-2-501A [7,8] are used to characterize the permeation of liquids through protective materials and determine transport properties in combination with the time-lag method of determining diffusivity using the steady-state flux of permeate through a membrane, as described by Dalton, Frisch and others [6,8,9,10].
Test methods used to evaluate PPE materials for chemical warfare defense applications involve placing individual droplets of a chemical agent on a circular sample of the test fabric and sampling the clean side of the membrane to detect the breakthrough of the chemical [8,11]. These tests can be resource-intensive and difficult to perform with highly toxic chemical permeates. However, simulation of the diffusion of molecules through dense polymer membranes is computationally intensive and difficult to use to predict transport properties [6,12].
An alternative approach is to apply experimentally obtained transport properties to computational models of membrane permeation. However, the experimentally obtained transport properties measured in standardized tests may not be appropriate. Rivin et al. [11] suggested that test methods using droplets of chemical permeant produce errors in the estimation of the transport properties derived through such testing suspected to be caused by the system not reaching steady-state diffusion or leading to transport properties different from those that would be obtained in the case of flooded permeation cells. Indeed, in this work, it is shown that transport properties obtained from experiments with partial coverage of the test membrane may be inaccurate. As a result, transport properties obtained from such experiments need to be interpreted with caution when used for PPE selection to protect against hazardous chemicals.
The objective of this study is to perform a series of numerical experiments to assess the impact of a partially covered membrane surface on the estimation of the membrane transport properties determined using the time-lag method. The impacts of partial coverage were assessed in conjunction with changing membrane thickness, changing the diffusivity of the permeate in the membrane, and changing the coverage geometry of the permeate on the surface of the membrane. The motivation for using partial membrane coverage is to minimize the amount of toxic liquid used in a permeation experiment. It is anticipated that the series of numerical simulations will allow us to find a correction factor that could be used to recover the actual membrane transport properties from those obtained experimentally. With a good estimate of the transport properties, the numerical model could then be used to examine various permeation scenarios.

2. Materials and Methods

2.1. Experimental Permeation System

The numerical experiments performed in this investigation are designed to model a typical diffusion cell for testing the permeation of hazardous chemicals [7,8,11,13]. A circular membrane swatch of the test membrane material is used and permeate is applied to one side of the membrane. As the permeate diffuses through the membrane, the amount of permeate found on the receptor side is quantified as a function of time to determine the transport properties of the membrane. The numerical experiments conducted were used to simulate the permeation of liquids through non-porous polymer membranes. Figure 1 shows a representative diffusion cell which is used to determine experimentally the transport properties of the liquid moving across a membrane [7,8,11,13,14].

2.2. Numerical Simulations

In this investigation, the three-dimensional Fick’s second law of diffusion Equation (1) was solved by finite differences using a program, written in Fortran programming language, developed by the authors for this application to determine the rate of permeation through a cylindrical dense membrane (Figure 2a), as typically incorporated within a permeation cell.
c t = D [ 2 c z 2 + 2 c r 2 + 1 r c r + 1 r 2 2 c ϕ 2 ]
Equation (1) may be reduced to a one-dimensional partial differential equation in the permeation direction z when the feed side of the membrane is completely flooded. When the feed side of the membrane is partially covered, the three-dimensional diffusion equation must usually be solved. In this investigation, the finite difference in the three-directional mass transfer equation will always be solved to determine the temporal variation in the concentration within the membrane.
Equation (1) was discretized into a sufficiently large number of mesh points to generate small three-dimensional elements as shown in Figure 2b to achieve mesh independence. The discretized equation for a central mesh point is given in Equation (2) in its explicit form.
c i , j , k t + Δ t     c i , j , k t Δ t = D [ c i 1 , j , k t     2 c i , j , k t + c i + 1 , j , k t Δ z 2 + c i , j 1 , k t     2 c i , j , k t + c i , j + 1 , k t Δ r 2 + 1 ( j 1 ) Δ r c i , j + 1 , k t     c i , j 1 , k t 2 Δ r + 1 ( ( j 1 ) Δ r ) 2 c i , j , k 1 t     2 c i , j , k t + c i , j , k + 1 t Δ ϕ 2 ]
The indices (i,j,k) refer to (z,r,ϕ) directions divided into (NI,NJ,NK) meshes. This equation calculates the internal concentration ci,j,k at mesh point (i,j,k) at the next time step Δt, i.e., at time t + Δt, based on the concentration at mesh point (i,j,k) and its six neighboring mesh points (Figure 2b) at time t. Equation (2) is valid for all internal mesh points describing the membrane. It is assumed that the dense membrane is isotropic, implying the membrane has a uniform composition structure throughout with constant diffusivity D and solubility S. In addition, it is assumed that the diffusivity and solubility are not functions of concentration and the system is isothermal.
To completely define the problem, it is necessary to provide the initial condition and six boundary conditions, two for each dimension. It is assumed that initially, the concentration of the migrating species throughout the membrane is zero (Equation (3)).
At   t = 0 , c i , j , k t = 0 = 0 i , j , k
At time zero, the liquid permeating species is added to the top surface of the membrane where it progressively dissolves into the dense membrane and migrates to the permeate side. The boundary conditions for the z-direction are given in Equations (4) and (5). Equation (4) simply states that the species concentration (c) at the top surface within the membrane is in instantaneous equilibrium with the concentration (C) of the liquid coverage outside the membrane. S is the membrane solubility of the migrating species. Equation (5) assumes that the concentration of permeate at the surface of the membrane on the receptor side is zero or close to zero. While this will not be strictly the case, we can assume that in a permeation cell where a liquid receptor solution is used instead of a sweep gas, the solvent has a high enough affinity for the permeate that the concentration at the membrane surface is equivalent to the concentration of permeate in the bulk solvent. In a diffusion cell where the permeant is sampled on the receptor side as a vapor, a constant gas flow is used to carry the permeant vapor to the detector [11]. As the rates of diffusion in the modeled system are generally very low and the levels of permeate that indicate breakthrough are generally very small, then the assumption is still valid.
At   z = 0 , c i = 1 , j , k t = S C i = 1 , j , k t j , k
At   z = L , c i = N I , j , k t = 0 j , k
The boundary conditions for the radial direction are given in Equations (6) and (7). Equation (6) states that symmetry prevails at the center (r = 0) of the membrane. The second term in the r-direction of Equation (1) is undetermined at the center since both the partial derivative and the radius are both equal to zero. However, using L’Hôpital’s rule, this term is easily equated to the second partial derivative concerning r and the radial term can be easily discretized. It is important to note that for the symmetry condition at the center of the membrane to prevail, the liquid must be distributed symmetrically on the surface of the membrane, especially when a multiple-drop pattern is used. Equation (7) simply states that at the outside circumference of the membrane, an impermeability condition prevails.
At   r = 0 , c i , j = 1 , k t r = 0 i , k
At   r = R , c i , N J , k t r = 0 i , k
The boundary conditions for the angular direction (ϕ) are given in Equations (8) and (9). Both boundary conditions assumed symmetry at ϕ equal to zero and some other angle where symmetry prevails. This angle could be 2π if symmetry does not prevail at a smaller angle. In this investigation, π/2 and 2π were used depending on the case study. One particular aspect of the angular term of Equation (1) is the singularity at the center point of the membrane. Few techniques have been proposed to deal with this singularity [15]. In this investigation, at the center point (j = 1), the zero radius was replaced with 0.1% of the first radial mesh point Δr.
At   ϕ = 0 , c i , j , k = 1 t ϕ = 0 i , j
At   ϕ = ϕ 1 , c i , j , k = N K t ϕ = 0 i , j
The average upstream and downstream fluxes were obtained from the concentration gradients at the upstream and downstream fluid–membrane interfaces using Equations (10) and (11). Convergence to the steady permeation rate was assumed to be reached when the average fluxes (J) at the upstream and downstream surfaces were the same, within 0.01%. The downstream average flux allows calculation of the rate of accumulation of the migrating species on the permeate side. Under steady state, the rate of accumulation becomes constant and the extrapolation to the time axis of the constant accumulation curve as a function of time provides the time lag (θ) from which the effective diffusivity can be determined, see Equation (12).
J z = 0 = 1 π R 2 0 R 0 2 π D c z | z = 0 r d r d ϕ ( Upstream )
J z = L = 1 π R 2 0 R 0 2 π D c z | z = L r d r d ϕ ( Downstream )
θ = L 2 6 D e f f
where R is the radius of the membrane, Deff is the effective diffusivity of the migrating species in the membrane, and L is the membrane thickness. Note that an apparent or effective diffusivity (Deff) has been used, Equation (12), instead of the intrinsic diffusivity (D) because the measured diffusivity via the time-lag method may not be equal to the intrinsic diffusivity in the case of an incomplete liquid coverage on the upstream surface of the membrane. It is important to stress that the intrinsic diffusivity of the membrane remains constant and is independent of the fraction of the liquid coverage. For complete liquid coverage, the measured diffusivity should be equal to the intrinsic diffusivity.

2.3. Different Liquid Patterns and Surface Area Coverage

In this paper, several cases are examined to determine which variables influenced the values of the transport properties calculated for partial coverage scenarios. The fraction of coverage of the feed surface area of the membrane was varied to examine and quantify previous work carried out by Rivin et al. [11] which suggested that incomplete coverage leads to errors in the determined transport properties. The simulation also examined if different patterns of permeate added to the membrane surface would impact these measurements. Several patterns with different drop sizes and geometries were simulated and labelled 1 to 4 as illustrated in Figure 3. The droplet patterns were adapted from the test patterns described in US Army TOP 08-2-501A [8] to a droplet pattern with radial symmetry. Other drop patterns were also investigated.

3. Results and Discussion

This section presents the simulation results for various case studies to examine the impact of various variables on the estimation of the membrane properties, namely the liquid coverage pattern, fraction surface coverage, membrane intrinsic properties, and membrane thickness. The results show that the membrane properties are estimated accurately, as expected, in the case of a flooded cell and these results will serve as a basis of comparison for the other case studies.

3.1. Validation of the Numerical Simulations

Before presenting the results of various numerical simulations, it is important to examine the accuracy of the numerical model with a known benchmark solution. Figure 4 shows the comparison between the analytical and numerical solutions for the case where the feed side of the membrane is completely flooded at time t = 0. It is the only benchmark solution that is available for this problem, which is reduced in this case to a one-dimensional solution. Equation (13) gives the analytical solution of the concentration profile c(z,t) throughout the membrane as a function of time (t) and relative membrane thickness (z/L) for this case study. This analytical solution was obtained by the method of separation of variables [16]. The results of Figure 4 clearly show that the numerical solution for the concentration profile as a function of time corresponds exactly to the analytical concentration profile. An overall mass balance evaluated the accuracy of the simulations in the case of the partial liquid coverage of the membrane. In all cases, the overall mass balance was shown to be very accurate.
c ( z , t ) = C S ( 1 z L ) ( 2 C S π ) n = 1 [ 1 n sin ( n π z L ) e ( D n 2 π 2 t L 2 ) ]

3.2. The Impact of the Membrane Diffusivity

To examine the impact of the membrane diffusivity (D) on the estimated effective diffusivity (Deff) for different membrane thicknesses, a series of numerical experiments were conducted over a wide range of membrane diffusivity values and for a constant liquid fraction coverage. The estimated diffusivity was obtained with the time-lag method. The results of these numerical experiments are presented in Figure 5 and show that, for a constant membrane thickness, the estimated effective diffusivity is directly proportional to the actual membrane diffusivity as the ratio Deff/D remains constant over the whole range of diffusivity. On the other hand, the thickness of the membrane has a major impact on the diffusivity ratio. For thin membranes, the diffusivity ratio is close to unity and the estimated diffusivity is relatively well estimated, where it approaches the value that is obtained with a flooded cell. It is important to state that the actual diffusivity (D) of the membrane is independent of the coverage pattern, and it is the value that we are attempting to recover. However, because of the fractional liquid coverage, the three-dimensional species permeation leads to an estimated value, termed effective diffusivity (Deff), that is different from the intrinsic membrane diffusivity.

3.3. The Impact of Membrane Thickness

Figure 6 presents the results of a series of simulations to examine the impact of the membrane thickness on the estimation of the membrane diffusivity for a single drop of liquid centrally located on the membrane and covering 18% of the membrane surface. The results show that, as the thickness of the membrane increases, the difference between the effective diffusivity and the actual diffusivity increases over a wide range of thicknesses. The diffusivity ratio (Deff/D) reaches a minimum at a thickness of approximately 6 mm with a relative diffusivity of approximately equal to 0.55. At higher membrane thicknesses, the effective diffusivity starts to increase. It is hypothesized that the diffusivity ratio will asymptotically increase back to unity for a much larger, albeit unrealistic, membrane thickness. Similar trends were observed for all the partial coverage patterns tested and the local minimum was in the same order of magnitude. This minimum occurs at smaller thicknesses for smaller coverage areas. These numerical simulations demonstrate that a significant estimation error of the intrinsic diffusivity should be expected for some common protective equipment [17,18,19,20], given the typical thickness of these protective garments. These errors typically lead to an underestimation of the actual diffusivity for a given system.
To explore the reason for the observed minimum in the diffusivity ratio and its subsequent increase as the thickness increases, the radial concentration profiles under steady state were obtained at five different relative depths (z/L) within the membrane and are presented in Figure 7. These results were obtained for a centered droplet covering 30% of the membrane surface area. Figure 7a presents the concentration profiles for a very thin membrane, L = 100 µm, showing very negligible radial diffusion such that the inlet and exit permeation areas are nearly identical with a flat concentration profile. The membrane diffusivity is very well estimated, with a ratio of Deff/D of 0.984. In Figure 7b, a larger radial diffusion was observed for a membrane having a thickness ten times larger (1.0 mm), which delayed the attainment of the steady state and the time lag. As a result, the estimated diffusion coefficient is smaller than the actual diffusion coefficient with a diffusivity ratio Deff/D of 0.809. When the membrane thickness is yet ten times larger (10 mm) and equal to the radius of the membrane, the migrating species have a significantly larger radial diffusion as shown in Figure 7c. The thickness is sufficiently large for the migrating species to diffuse to the impermeable diffusion cell boundary. As a result, a nearly flat concentration profile is observed close to the permeate side of the membrane. The estimated diffusion coefficient is smaller than the actual diffusion coefficient with a diffusivity ratio Deff/D of 0.643. This data point for this relatively thick membrane is in the increasing portion of the Deff/D versus L plot and further increasing the membrane thickness, although unrealistic, would lead to a relative diffusivity even closer to the actual diffusivity as a larger relative portion of the membrane reaches a flat concentration profile.

3.4. The Impact of Permeate Geometry

Figure 8 presents the results of the series of simulations performed to examine the variation in the relative diffusivity as a function of the thickness for two different drop patterns covering 12% of the surface membrane area. The droplet diameter, the diffusivity, and the number of droplets were held constant while, as shown in the two insets of Figure 8, the location of the outer layer of droplets was changed. The results show that, for thin membranes, the relative diffusivity ratio is approaching unity. However, as the thickness of the membrane increases, the relative diffusivity decreases rapidly to reach a minimum at a thickness between 1 and 2 mm. For both drop patterns, the minimum value of the relative diffusivity is nearly the same with a value of approximately 0.51 and 0.49 for the droplet pattern with the outer ring, shown in red, closer to the center and for the droplet pattern with the outer ring, shown in blue, further from the center, respectively. Beyond the minimum value, the relative diffusivity increases more smoothly with an increase in thickness. Even though the relative diffusivity is affected by the distribution of droplets on the surface of the membrane, the main factors affecting the diffusivity estimated by the time-lag method remain the thickness of the membrane and the liquid fraction coverage.

3.5. The Impact of Membrane Surface Area Covered by the Permeate

Figure 9 presents the results of a series of simulations performed to assess the variation in the relative diffusivity for two different liquid patterns, namely patterns 2 and 3, as a function of the surface coverage fraction. The results show that, as the fraction of the membrane surface covered by the liquid permeate increases, the relative diffusivity ratio approaches unity. Indeed, the intrinsic diffusivity of the membrane is obtained when the membrane surface is flooded by the liquid permeate. Results clearly show the effect of the liquid fraction coverage on the relative diffusivity as determined by the time-lag method. Both liquid patterns compared in this series of simulations are in close agreement with each other, suggesting that the effects of coverage are much less impacted by the geometry of the permeate on the membrane surface than the effects of thickness and the coverage fraction. Practically, in this case, as the fraction of the membrane covered by permeate increases, the patterns tested become increasingly similar.
Figure 10 shows a comparison of the relative diffusivity varying with the fractional coverage of the membrane surface for pattern 3 and four different membrane thicknesses. The results of Figure 10 clearly show the effects of both the coverage fraction and the membrane thickness for a constant coverage pattern. The value of the relative diffusivity Deff/D asymptotically tends to unity in all cases when the fraction coverage of the membrane surface area approaches unity, that is a flooded cell (pattern 1). These results corroborate the work of Rivin et al. [11] who state that the “flooded cell” scenario is the ideal method to estimate the intrinsic diffusivity of the membrane, as it corresponds to the case of one-dimensional diffusion.

3.6. Receding Liquid Coverage

Previous results clearly show that with flooded cell experiments, the membrane diffusivity is estimated accurately, whereas, for partial liquid surface coverage, its estimation needs special attention. In addition, with partial coverage, the liquid coverage area may recede as a function of time because of permeant sorption and evaporation, thereby affecting the dynamic permeation rate. Indeed, the rate of recession of the liquid typically depends on the thickness of the liquid cover, the liquid surface tension, the rate of evaporation, and its diffusivity and solubility in the membrane. Figure 11 presents the comparison of the dynamic permeation results for an initial central circular liquid pattern for a 2 mm thick membrane. A series of simulations were performed with different exponential decays of the liquid coverage radius as expressed by Equation (13), with the radius R of the liquid cover decreasing from the initial liquid radius R0 as a function of a receding factor α, time t, and time lag θ. The results shown in Figure 11 were obtained with an initial radius of approximately 0.007 m and a receding factor of 0.05 chosen to examine a slowly receding liquid.
R = R 0 e α t / θ
Figure 11a presents the inlet and outlet average fluxes of the migrating species as a function of time through the membrane. For a constant liquid cover (α = 0), the two average fluxes progressively become equal as they approach the steady state. When the liquid cover recedes as a function of time, as expected, the average fluxes become smaller than fluxes under constant coverage. For a longer permeation time, the liquid will completely disappear, and the two fluxes will eventually approach zero. For higher values of the receding factor, simulating higher diffusivities or lower vapor pressures, the average fluxes will decrease even faster. The corresponding amount of the migrating species that has permeated through the membrane is plotted in Figure 11b for the two values of the receding factor. For constant coverage, the rate of accumulation becomes relatively constant, and the time lag can be determined, while for receding coverage, the rate initially increases before progressively decreasing. For the latter case, it is more difficult to determine the time lag as the slope of the accumulation curve is never constant and does not entirely satisfy the underlying assumptions of the time-lag method. Figure 11c presents the estimation of the effective diffusivity determined from the instantaneous time lag as a function of time that was calculated from the slope of the plots of the accumulated amount that had permeated through the membrane (Figure 11b). In the case of constant liquid coverage, the instantaneous time lag progressively assumes a nearly constant value. A small decrease is still observed in Figure 11c because of the slow lateral diffusion that is required to reach a perfect steady state. For a receding liquid coverage, the instantaneous effective diffusivity initially decreases, then passes through a minimum before increasing again as a function of time. It is interesting to note that the minimum value of the effective diffusivity is very close to the actual value of the membrane diffusivity. In practice, a permeation experiment will probably last only for the first part of the curve when the concentration curve of the accumulated permeate would attain an almost linear trend, corresponding to an experimental time of between 1 and 2 × 106 s (Figure 11b) for this relatively thick membrane. Finally, Figure 11d presents the variation in the radius of the central liquid coverage as a function of time for both cases considered in this section.
In practice, in the case of a receding liquid coverage to use the synergy between the experimental and numerical determinations of the effective diffusivity, it would be necessary to take images of the feed surface pattern as a function of time. It will then be possible to estimate the intrinsic diffusivity of the migrating species within the tested membrane.

3.7. Predictive Model for the Estimation of the Diffusivity Ratio

The results of the previous sections have clearly shown that a partial liquid coverage on the surface of the membrane may lead to an incorrect estimation of the diffusion coefficient of the liquid being tested. Ideally, a complete surface coverage should be used whenever possible. However, in the case of a very toxic liquid, it is advisable to use a drop pattern distribution to minimize the quantity of liquid used. The results obtained in this investigation provide the necessary data to develop a model that could predict the diffusion coefficient ratio Deff/D. The correction factor depends on numerous factors. Nevertheless, it is desired to find the simplest model that could provide an acceptable estimation of Deff/D. Having determined the effective diffusion coefficient Deff experimentally, the actual diffusion coefficient D can therefore be estimated.
In this investigation, an artificial neural network was used to estimate the diffusion coefficient ratio as a function of the two most important contributing factors: the thickness of the membrane and the coverage fraction. Artificial neural networks are now commonly used successfully for a myriad of engineering applications. The high degree of plasticity in its structure is the main reason for its ability to efficiently represent the underlying causal relationship between input and output data. In this investigation, a three-layer feedforward neural network (FFNN) was used to predict Deff/D as a function of some input process variables. The FFNN is comprised of an input layer, a hidden layer, and an output layer, as shown in Figure 12. The input layer contains three neurons corresponding to the two independent variables plus the bias neuron. The two independent variables of Figure 12 are the scaled values of the logarithmic value of the membrane thickness and the liquid coverage fraction. The input information is simply transferred to the neurons of the hidden layer and the output layer. Appendix A provides the detailed development of the FFNN used to predict Deff/D. The results of 797 numerical simulations were used to determine the weights of the FFNN by minimizing the sum of squares of the differences between the diffusivity ratio Deff/D obtained numerically and the one predicted by the neural network. The 797 data points are comprised of 226 records for pattern 2, 254 for pattern 3, and 317 for pattern 4. Figure 13 presents the parity plot of the prediction of the diffusion coefficient ratio Deff/D for three different liquid coverage patterns. The coefficient of regression R2 is provided in Figure 13 for each liquid coverage pattern. The fit is not perfect since only the two main factors were considered as inputs to the model. The average relative error is 2.44% ± 2.36 and the maximum relative error is 15.4%. It is believed that the prediction error of this model is acceptable and allows us to estimate with good accuracy the diffusion coefficient ratio Deff/D.

4. Conclusions

This paper illustrates that the distribution of permeate on a membrane surface, which includes both the coverage of the surface area, and the pattern of species migration applied to the membrane, significantly influences the transport properties derived from permeation experiments. The effective diffusivity of the system varies with the fraction of the surface area of the membrane covered by the migrating species, the thickness of the membrane, and the geometry of the permeate applied to the membrane surface. For a flooded permeation cell, the intrinsic diffusivity is perfectly estimated. For membranes with partial liquid coverage, the effective diffusivity, as measured by the time-lag method, is lower than the intrinsic diffusivity of the membrane due to the multi-directional nature of the diffusion process. The experimental test methods used to evaluate PPE materials for protection against chemical warfare agents use partial liquid coverage. As a result, it may not be appropriate to use directly the transport properties obtained from these experimental methods to compare or generalize the protective performance of materials tested under different parameters such as material thickness, liquid coverage, and droplet pattern.
In instances where conducting experiments with flooded cells is either not feasible or impractical, applying a correction factor to the experimentally obtained transport properties can yield a more precise estimate of the intrinsic diffusivity. Subsequent experiments may be simulated using these intrinsic transport properties to investigate the impact of alterations in geometry and membrane thickness. Future research will involve conducting experiments to both assess and enhance the precision of our model, as well as to compare it against empirical data. Moreover, it would be interesting to investigate experimentally and numerically mixed-matrix or hybrid membranes to determine the estimation errors in their transport properties when characterized with the time-lag method. These newer membranes incorporate a reactive filler capable of neutralizing toxic chemicals as a means of detoxification [21,22,23].

Author Contributions

Conceptualization, A.B. and J.T.; methodology, A.B., V.L. and J.T.; software, J.T.; validation, A.B. and J.T.; formal analysis, A.B. and J.T.; investigation, A.B., V.L. and J.T.; data curation, A.B. and J.T.; writing—original draft preparation, A.B. and J.T.; writing—review and editing, A.B., V.L. and J.T.; supervision, V.L. and J.T.; project administration, J.T.; funding acquisition, J.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science and Engineering Research Council (NSERC) Canada, Discovery Grant (DG), grant number RGPIN-2018-04433 (JT).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Data are available upon request.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

Appendix A

When you want to find an accurate model quickly that could be used for predicting a dependent variable, such as the diffusion coefficient ratio (Deff/D), an artificial neural network can be easily used. Artificial neural networks are now commonly used for a large number of engineering applications, making use of the high degree of plasticity in their structure to encapsulate the underlying causal relationship between input and output data. In this investigation, a three-layer feedforward neural network (FFNN) was used to predict the diffusion coefficient ratio (Deff/D) as a function of two input process variables. It is believed that a three-layer neural network is more than adequate to model the current problem. Cybenko [24] showed that a three-layer FFNN was sufficient to encapsulate any input–output relationship if a sufficient number of neurons were used. This appendix provides all the details to develop and use the final FFNN for the prediction of (Deff/D).
The first step is to choose the set of independent variables that have a high probability of accurately inferring the dependent variable. In this case, two variables with the most impact, namely the thickness of the membrane and the liquid coverage fraction, were chosen to predict (Deff/D) as the output of the FFNN. These choices fix the number of neurons in the input and output layers: three neurons in the input layer including the bias and one neuron in the output layer as shown in Figure 12. The number of neurons in the hidden layer is typically determined by trial and error to ensure a good prediction while avoiding overtraining. The selected neural network of Figure 12 has seven neurons in the hidden layer, including the bias neuron. The input and the output variables were scaled in the range of [0,1] using the minimum and maximum of each variable as shown in Equations (A1) and (A2). Because the thickness (L) varied from 0.0001 to 0.02 m, the logarithmic function was scaled instead of the actual thickness. The coverage fraction (CF) did not need scaling because it already varies between zero and one, but the minimum and maximum were still retained in Equation (A1) for completeness. For the output value, the minimum and maximum values were chosen as 0.35 and 1.15, respectively.
| X 1 X 2 | = | log 10 ( L ) + 4 1.7 + 4 C F 0 1 0 |
Y = D e f f D 0.35 1.15 0.35
The scaled inputs of the first layer are fanned out to all functional neurons in the hidden layer. Each functional neuron in the hidden layer simply transforms the weighted sum of the inputs of the first layer through a nonlinear function to calculate its output. In this investigation, a sigmoid function was used as the nonlinear function Equation (A3). The weights between the neurons of the two adjacent layers are the parameters of the model. Each output Hj of the functional layer of the hidden layer is given by Equation (A4).
f ( Ψ ) = 1 1 + e Ψ
H j = 1 1 + e ( W 1 j X 1 + W 2 j X 2 + W 3 j ) = 1 1 + E x p ( i = 1 3 W i j X i )
The outputs of the hidden layer, including the bias neuron of the hidden layer, are then sent to the output neuron. The output neuron performs the same task as the neurons of the second layer to generate the final output of the FFNN. A sigmoid function was also used as the transfer function of the output neuron. The output Y ^ of the last layer is given by Equation (A5).
Y ^ = 1 1 + e ( W 11 H 1 + W 21 H 2 + W 31 H 3 + W 41 H 4 + W 51 H 5 + W 61 H 6 + W 71 ) = 1 1 + E x p ( j = 1 7 W j 1 H j )
The fitting or training of a neural network boils down to minimizing the sum of squared errors between the actual values and predicted values by modifying the weights of the connections between the neurons of two adjacent layers. Typically, the data set is divided into three data subsets: training, validation, and testing. The neural model is initially fitted using the training data set whereas the validation data are used in parallel with the training data set to provide an unbiased evaluation of the model fit, a necessary step to avoid overtraining. The testing data set provides a further unbiased evaluation of the final model with data that the model has never seen. The FFNN structure of Figure 12 was chosen as the final model and can be represented mathematically in matrix form with Equations (A6) and (A7).
  W i j   | H 1 H 2 H 3 H 4 H 5 H 6 | = f ( |   8 . 126 34.272 10.038 4.847 15.057 1.522 2.454 16.699 4.180 4.192 1.247 3.337 3.884 0.764 2.669 6.086   49.982 4.629 | | X 1 X 2 1 | )
W jk Y ^ = f ( | 1.703 4.008 3.263 23.176 24.287 1.097 23.972 | | H 1 H 2 H 3 H 4 H 5 H 6 1 | )
The only hyperparameter that was changed in the development of the neural network was the number of neurons in the hidden layer. The model could be simplified with a loss of accuracy in the predictions. The relative percentage prediction errors were 5.33% ± 5.04, 3.58% ± 3.18 and 2.44% ± 2.36 for three, five, and seven hidden neurons (including the bias), respectively. The maximum values of the relative percentage errors were 24.9%, 16.5%, and 15.4%, respectively.

References

  1. Dickson, E.F.G. Personal Protective Equipment for Chemical, Biological, and Radiological Hazards: Design, Evaluation and Selection, 1st ed.; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2012. [Google Scholar] [CrossRef]
  2. Daugherty, M.L.; Watson, A.P.; Vo-Dinh, T. Currently Available Permeability and Breakthrough Data Characterizing Chemical Warfare Agents and Their Simulants in Civilian Protective Clothing Mater. J. Hazard. Mater. 1992, 30, 243–267. [Google Scholar] [CrossRef]
  3. Government of Canada, C.C. for O. H. and S. CCOHS: Chemical Protective Clothing-Glove Selection. Available online: https://www.ccohs.ca/oshanswers/prevention/ppe/gloves.html (accessed on 14 June 2023).
  4. Paul, D.R. The Solution-Diffusion Model for Swollen Membranes. Sep. Purif. Methods 1976, 5, 33–50. [Google Scholar] [CrossRef]
  5. Yampolskii, Y.; Pinnau, I.; Freeman, B. (Eds.) Materials Science of Membranes for Gas and Vapor Separation, 1st ed.; John Wiley & Sons, Ltd.: Chichester, UK, 2006. [Google Scholar] [CrossRef]
  6. Baker, R.W. Membrane Technology and Applications, 3rd ed.; John Wiley & Sons, Ltd.: Chichester, UK, 2012. [Google Scholar] [CrossRef]
  7. Banaee, S.; Hee, S.S.Q. Glove Permeation of Chemicals: The State of the Art of Current Practice, Part 1: Basics and the Permeation Standards. J. Occup. Environ. Hyg. 2019, 16, 827–839. [Google Scholar] [CrossRef] [PubMed]
  8. Test Operations Procedure (TOP) 08-2-501A, Permeation Testing of Materials with Chemical Agents or Simulants (Swatch Testing); Final TOP 08-2-501A; U.S. Army Dugway Proving Ground: Dugway, UT 84022-5000, 2013. Available online: https://apps.dtic.mil/sti/pdfs/ADA586844.pdf (accessed on 19 March 2023).
  9. Frisch, H.L. The Time Lag in Diffusion. J. Phys. Chem. 1957, 61, 93–95. [Google Scholar] [CrossRef]
  10. Rutherford, S.W.; Do, D.D. Review of Time Lag Permeation Technique as a Method for Characterisation of Porous Media and Membranes. Adsorption 1997, 3, 283–312. [Google Scholar] [CrossRef]
  11. Rivin, D.; Lindsay, R.S.; Shuely, W.J.; Rodriguez, A. Liquid Permeation Through Nonporous Barrier Materials. J. Membr. Sci. 2005, 246, 39–47. [Google Scholar] [CrossRef]
  12. Burganos, V.N. Modeling and Simulation of Membrane Structure and Transport Properties. In Fundamentals of Transport Phenomena in Membranes; Elsevier B V: Amsterdam, The Netherlands, 2010. [Google Scholar]
  13. Franz, T.J. Percutaneous Absorption. On the Relevance of in Vitro Data. J. Investig. Dermatol. 1975, 64, 190–195. [Google Scholar] [CrossRef] [PubMed]
  14. Mellstrom, G.A. Comparison of Chemical Permeation Data Obtained with ASTM and ISO Permeation Test Cells—I. The ASTM Standard Test Procedure. Ann. Occup. Hyg. 1991, 35, 153–166. [Google Scholar] [CrossRef] [PubMed]
  15. Thibault, J.; Bergeron, S.; Bonin, H.W. On Finite-Difference Solutions of the Heat Equation in Spherical Coordinates. Numer. Heat Transf. 1987, 12, 457–474. [Google Scholar] [CrossRef]
  16. Wu, H.; Al-Qasas, N.; Kruczek, B.; Thibault, J. Simulation of Time-lag Permeation Experiments Using Finite Differences. J. Fluid Flow Heat Mass Transf. 2015, 2, 14–25. [Google Scholar] [CrossRef]
  17. Jones, M.R. Personal Protective Equipment (PPE): Practical and Theoretical Considerations. In Chemical Warfare Agents: Biomedical and Psychological Effects, Medical Countermeasures, and Emergency Response; Lukey, B.J., Romano, J.A., Jr., Salem, H., Eds.; CRC Press: Boca Raton, FL, USA, 2019. [Google Scholar] [CrossRef]
  18. Ouvry. CBRN Butyl Gloves OG07®. Ouvry-CBRN Protective System. Available online: https://ouvry.com/en/produit/cbrn-butyl-gloves/ (accessed on 9 May 2023).
  19. Phalen, R.N.; Wong, W.K. Chemical Resistance of Disposable Nitrile Gloves Exposed to Simulated Movement. J. Occup. Environ. Hyg. 2012, 9, 630–639. [Google Scholar] [CrossRef] [PubMed]
  20. Kimberly Clark Professional. KleenGuardTM G10 Flex Nitrile Gloves (54333)-M Packaging, 100 Gloves/Box, 10 Boxes/Case, 1000 Gloves/Case. KC Professional. Available online: https://www.kcprofessional.com/en-ca/products/safety-and-personal-protection-equipment/hand-protection-and-gloves/general-purpose-gloves/54331/54333 (accessed on 9 May 2023).
  21. Lee, J.; Kim, E.-Y.; Chang, B.-J.; Han, M.; Lee, P.-S.; Moon, S.-Y. Mixed-matrix membrane reactors for the destruction of toxic chemicals. J. Membr. Sci. 2020, 605, 118112. [Google Scholar] [CrossRef]
  22. Motamedhashemi, M.M.Y.; Monji, M.; Egolfopoulos, F.; Tsotsis, T. A hybrid catalytic membrane reactor for destruction of a chemical warfare simulant. J. Membr. Sci. 2015, 473, 1–7. [Google Scholar] [CrossRef]
  23. Luo, H.B.; Lin, F.-R.; Liu, Z.-Y.; Kong, Y.-R.; Idrees, K.B.; Liu, Y.; Zou, Y.; Farha, O.K.; Ren, X.M. MOF−Polymer Mixed Matrix Membranes as Chemical Protective Layers for Solid-Phase Detoxification of Toxic Organophosphates. ACS Appl. Mater. Interfaces 2023, 15, 2933–2939. [Google Scholar] [CrossRef] [PubMed]
  24. Cybenko, G. Approximation by Superpositions of a Sigmoidal Function. Math. Control Signals Syst. 1989, 2, 303–314. [Google Scholar] [CrossRef]
Figure 1. Simplified diffusion cell diagram depicting chemical permeant placed as droplets (top) or as flooded cell (bottom).
Figure 1. Simplified diffusion cell diagram depicting chemical permeant placed as droplets (top) or as flooded cell (bottom).
Membranes 14 00183 g001
Figure 2. Discretized three-dimensional cylindrical membrane: (a) schematic diagram of a discretized membrane, and (b) central mesh point with its six neighboring mesh points used in Equation (2).
Figure 2. Discretized three-dimensional cylindrical membrane: (a) schematic diagram of a discretized membrane, and (b) central mesh point with its six neighboring mesh points used in Equation (2).
Membranes 14 00183 g002
Figure 3. Examples of liquid patterns of the migrating species (in red) used at the membrane feed surface as boundary conditions for numerical experiments.
Figure 3. Examples of liquid patterns of the migrating species (in red) used at the membrane feed surface as boundary conditions for numerical experiments.
Membranes 14 00183 g003
Figure 4. Comparison of analytical and numerical solutions in the case of one-dimensional diffusion occurring when flooded cell experiments are performed.
Figure 4. Comparison of analytical and numerical solutions in the case of one-dimensional diffusion occurring when flooded cell experiments are performed.
Membranes 14 00183 g004
Figure 5. The ratio of the effective diffusivity and the actual diffusivity (Deff/D) as a function of the membrane diffusivity for four different membrane thicknesses with constant fraction coverage (pattern 4).
Figure 5. The ratio of the effective diffusivity and the actual diffusivity (Deff/D) as a function of the membrane diffusivity for four different membrane thicknesses with constant fraction coverage (pattern 4).
Membranes 14 00183 g005
Figure 6. The diffusivity ratio (Deff/D) as a function of the membrane thickness for a central liquid drop (pattern 2) covering 18% of the membrane surface.
Figure 6. The diffusivity ratio (Deff/D) as a function of the membrane thickness for a central liquid drop (pattern 2) covering 18% of the membrane surface.
Membranes 14 00183 g006
Figure 7. Concentration vs. r/R0 at different values of z/L for a single-centered droplet for three membrane thicknesses: (a) 100 μm, (b) 1.0 mm, and (c) 10 mm.
Figure 7. Concentration vs. r/R0 at different values of z/L for a single-centered droplet for three membrane thicknesses: (a) 100 μm, (b) 1.0 mm, and (c) 10 mm.
Membranes 14 00183 g007
Figure 8. Diffusivity ratio Deff/D as a function of the membrane thickness for two different drop patterns for a liquid surface coverage of 12%.
Figure 8. Diffusivity ratio Deff/D as a function of the membrane thickness for two different drop patterns for a liquid surface coverage of 12%.
Membranes 14 00183 g008
Figure 9. The ratio of the effective diffusivity over the actual diffusivity as a function of the liquid fraction coverage for patterns 2 and 3.
Figure 9. The ratio of the effective diffusivity over the actual diffusivity as a function of the liquid fraction coverage for patterns 2 and 3.
Membranes 14 00183 g009
Figure 10. The ratio of the effective diffusivity over the actual diffusivity as a function of the coverage fraction for pattern 3, for four different membrane thicknesses.
Figure 10. The ratio of the effective diffusivity over the actual diffusivity as a function of the coverage fraction for pattern 3, for four different membrane thicknesses.
Membranes 14 00183 g010
Figure 11. Comparison of the dynamic permeation results with (α = 0.05) and without (α = 0.0) receding liquid pattern: (a) upstream and downstream molar fluxes, (b) total amount permeated, (c) effective diffusivity estimated by the time-lag method, and (d) variation in the radius of the liquid coverage for the two receding factors.
Figure 11. Comparison of the dynamic permeation results with (α = 0.05) and without (α = 0.0) receding liquid pattern: (a) upstream and downstream molar fluxes, (b) total amount permeated, (c) effective diffusivity estimated by the time-lag method, and (d) variation in the radius of the liquid coverage for the two receding factors.
Membranes 14 00183 g011
Figure 12. Feedforward neural network used for the estimation of the diffusion coefficient ratio Deff/D.
Figure 12. Feedforward neural network used for the estimation of the diffusion coefficient ratio Deff/D.
Membranes 14 00183 g012
Figure 13. Parity plot of the prediction of the diffusion coefficient ratio Deff/D for three different liquid coverage patterns and the FFNN of Figure 12.
Figure 13. Parity plot of the prediction of the diffusion coefficient ratio Deff/D for three different liquid coverage patterns and the FFNN of Figure 12.
Membranes 14 00183 g013
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Bicket, A.; Lau, V.; Thibault, J. Simulating the Permeation of Toxic Chemicals through Barrier Materials. Membranes 2024, 14, 183. https://doi.org/10.3390/membranes14090183

AMA Style

Bicket A, Lau V, Thibault J. Simulating the Permeation of Toxic Chemicals through Barrier Materials. Membranes. 2024; 14(9):183. https://doi.org/10.3390/membranes14090183

Chicago/Turabian Style

Bicket, Alex, Vivian Lau, and Jules Thibault. 2024. "Simulating the Permeation of Toxic Chemicals through Barrier Materials" Membranes 14, no. 9: 183. https://doi.org/10.3390/membranes14090183

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop