key: cord-0010153-wrl6026n authors: Freeman, Brubeck Lee; Cleall, Peter John; Jefferson, Anthony Duncan title: An indicator‐based problem reduction scheme for coupled reactive transport models date: 2019-08-27 journal: Int J Numer Methods Eng DOI: 10.1002/nme.6186 sha: b2bedd9406f6529cc4b36b446f8913d3df4c5e91 doc_id: 10153 cord_uid: wrl6026n A number of effective models have been developed for simulating chemical transport in porous media; however, when a reactive chemical problem comprises multiple species within a substantial domain for a long period of time, the computational cost can become prohibitively expensive. This issue is addressed here by proposing a new numerical procedure to reduce the number of transport equations to be solved. This new problem reduction scheme (PRS) uses a predictor‐corrector approach, which “predicts” the transport of a set of non‐indicator species using results from a set of indicator species before “correcting” the non‐indicator concentrations using a mass balance error measure. The full chemical transport model is described along with experimental validation. The PRS is then presented together with an investigation, based on a 16‐species reaction‐advection‐diffusion problem, which determines the range of applicability of different orders of the PRS. The results of a further study are presented, in which a set of PRS simulations is compared with those from full model predictions. The application of the scheme to the intermediate‐sized problems considered in the present study showed reductions of up to 82% in CPU time, with good levels of accuracy maintained. The prediction of chemical solute transport behavior in porous media is of great importance in a wide range of engineering applications. To this end, a significant number of numerical transport models have been developed. Typically, these are coupled models that consider advection and hydrodynamic dispersion, the latter of which comprises both mechanical dispersion and self-diffusion 1 of the chemical species coupled with heat flow and, often, the mechanical behavior of the medium. In addition to these flow and deformation processes, the solute can also be considered to be either reactive or nonreactive depending on the nature of the problem considered. The application of these models has varied considerably, with many studies concentrating on geochemical problems, such as modeling groundwater systems, 2-6 assessing the performance of engineered barriers, [7] [8] [9] [10] [11] or attenuation of mine water tailings. [12] [13] [14] The application of these models to cementitious materials has most often investigated the ingress of chloride ions [15] [16] [17] [18] or calcium leaching 19, 20 ; however, recently, these models have also been used for the investigation of self-healing concrete. 21, 22 One of the major disadvantages of these models is that they can become computationally expensive when the chemical system becomes relatively large. Cleall et al 23 suggest that the magnitude of a problem is governed by three main aspects, namely, the domain size, timescale, and the complexity of the analysis, which includes the number of variables, the degree of coupling between them, the number of processes considered, and the degree of nonlinearity of the system. The degree of nonlinearity depends on the underlying physical processes; for example, the system will be highly nonlinear if the reactions are sensitive to small changes in concentration or the permeability of the porous medium is sensitive to the degree of saturation. To deal with this issue, several techniques have been proposed to improve the computational efficiency of the associated solution process, including "operator splitting" techniques and reformulation of the coupled system. Operator splitting divides or "splits" a time step into a transport calculation and a reaction calculation, to be solved sequentially, with many models iterating between the two. This is an approximation of the time integral of the governing equation, 24 which therefore changes the numerical framework from a global implicit approach to a sequential iterative or a sequential non-iterative approach. Such calculation splitting methods have been found to reduce the computational cost 25 and have been used by many authors. [2] [3] [4] [8] [9] [10] [11] A disadvantage of the sequential non-iterative approach is its propensity to introduce splitting errors (for example, mass balance errors), whereas the sequential iterative approach tends to require prohibitively small time steps and a relatively large number of iterations to achieve convergence. 26, 27 Consequentially, these operator splitting approaches have been found to increase computational demand for certain "chemically difficult" cases (for example, cases with high kinetic reaction rates). 27, 28 A number of authors have taken an alternative approach, concentrating instead on improving the efficiency of global implicit approaches. Most studies have focused on reformulating the system by decoupling a number of the partial differential equations (PDEs) and eliminating certain local (spatially invariant) equations, including both ordinary differential equations and algebraic equations. 26, 27, [29] [30] [31] A common assumption used in such formulations is that all chemical species have the same diffusion coefficient. 26, 27 This has been justified by the fact that predicted responses are relatively insensitive to differences in diffusion constants because mechanical dispersion is normally the dominant transport mechanism. 27 An alternative view was expressed by Thomas et al, 10 who suggested that this assumption is not valid for some chemical systems. The presence of minerals, whose concentrations can reach zero, may also cause issues when such schemes are used, 26 although Kräutle and Knabner 26 have introduced a moving boundary condition (BC) and a complimentary function to alleviate these difficulties. The reduction in computational demand that can be achieved by the operator splitting and reformulation approaches described above is often limited by the chemistry of a problem; for example, only limited reductions are possible when a large number of kinetic reactions with high reaction rates are considered. To address this issue, an indicator-based multi-order problem reduction scheme (PRS) is proposed, which allows for greater reduction in problem size that is independent of the nature of reactions and which explicitly considers species-dependent diffusion coefficients. The scheme works by decoupling a number of the nonlinear PDEs from the global system of equations. In this new PRS, a small number of indicator species (≤ 3) are selected for full computation, and then, on a stepwise basis, the response of other species is inferred from these indicators using a predictor-corrector approach. The reduction in computational demand is achieved through the reduction in size of the nonlinear global system of equations and by the replacement of nonlinear PDEs with stepwise predictor-corrector computations for non-indicator species. Three different orders of the PRS are considered, each applicable to different chemical systems. The new PRS is presented along with a study to establish the range of applicability of each order considered. This is followed by a set of validation examples in which the reactive transport of chemical species through mortar and concrete specimens is simulated. The theoretical model is based on the approach of Gawin et al. 32 The porous medium is assumed to be composed of three main phases, namely, solid skeleton, liquid moisture, and gas, the latter of which consists of dry air and moisture vapor. It is also assumed that dissolved chemical species are present in the liquid moisture phase and that precipitation can occur. Following the approach of Chitez and Jefferson, 33 the combined gas pressure of dry air and moisture vapor is assumed to remain constant at atmospheric pressure but that the moisture vapor may diffuse through the gas phase. The primary variables considered here are the capillary pressure (P C ) and chemical concentration (c i ) of each species. The advantages of using capillary pressure as the primary moisture variable have been discussed by Gawin et al. 32 These include the fact that it is convenient for coupling moisture flow to mechanical behavior, due to the direct relationship that exists between the pressure and the components of stress, and that it provides a physically meaningful driving force for moisture flow. Macroscopic balance equations, as derived from the volume-averaging theorem and the hybrid mixture theory, 34 are presented in the following sections. Heat transport is included in the model, governed by the enthalpy balance equation; however, the problems considered here are isothermal, and so, this part of the model has not been used in the present study. Transport in the porous medium is described using mass balance equations for liquid moisture, moisture vapor, and chemical species. Defining the domain ∈ R d with boundary and noting that the time interval considered is given as t ∈ [t 0 , T], the mass balance of the liquid moisture and moisture vapor can be written as Summing these up gives the total mass balance for moisture, as follows: where n denotes the porosity, S w is the degree of liquid moisture saturation, w is the liquid density, S g is the degree of gas saturation, and v is the moisture vapor density. t denotes a time derivative, v w is the liquid moisture velocity, t (m v ) is the rate of moisture transfer from liquid to vapor, and J v is the moisture vapor diffusion described here by Fick's law. 32 That is, where g is the gas density and D mv is the moisture vapor diffusivity tensor given by 22 where A v = 1, B v = 1.667, and f v = 0.01. D v0 = 2.47 × 10 −5 m 2 /s is the moisture vapor diffusion coefficient in air, T r = 273 K is the reference temperature, and P g and P atm denote the gas pressure and atmospheric pressure, respectively. The flow of the liquid moisture through the medium can be described by Darcy's law. 32 That is, where K in is the intrinsic permeability tensor of the medium, w is the viscosity of the fluid, and K rw is the relative permeability of the moisture phase given by 8 where A w = 4 in the present work. The remaining constitutive relationships used in the moisture transport component of the model are given in the Appendix (see Table A1 ; see also the Nomenclature table and the notation on the Greek letters and subscripts/superscripts used in this paper). The mass balance of a dissolved chemical species, i, can be written as where the final term on the left-hand side is the source/sink (SS) term due to chemical reactions, S p is the degree of saturation of precipitated or sorbed material, p is the mass of the precipitated or sorbed material, and J i d is the dispersive flux given by the Poisson-Nernst-Planck equations, which can be written as Equation (9), 35 where the first term on the right-hand side represents mechanical dispersion and the remaining right-hand side terms account for diffusion. In the above equations, D mol i is the coefficient of molecular diffusion, z is the charge of an ion, is the dielectric permittivity, is the charge density, F is Faraday's constant, and is the electrical potential. D d i is the mechanical dispersion tensor, defined in Equation (14) . The second term in the bracket of Equation (9) represents dispersion due to the local electric field , which can be calculated from Equation (10) . It can be noted that the first term in Equation (10) is negligible, 35, 36 as is the charge density ; thus, Equation (10) reduces to the following charge neutrality condition 10, [35] [36] [37] : and since the pore solution is initially charge neutral, this condition is ensured through the "no electrical current" condition, 10, 36, 37 as follows: where is the diffusive flux of species i. Equation (12) can be used to eliminate the electrical potential gradient, giving the dispersive flux of species i as Hydrodynamic dispersion is the sum of molecular diffusion and mechanical dispersion. The definition of hydrodynamic dispersion of Bear and Bachmat 1 is adopted here, where the mechanical dispersion tensor is given as where L and T are the longitudinal and transverse dispersivities, respectively, and kj is the Kronecker delta ( kj = 1 if k = j or kj = 0 if k ≠ j). The chemical reactions considered here are all assumed to be nonequilibrium reactions, ie, transport of the ions may be faster than the rate of reactions such that binding or precipitation is not instantaneous. 38 In addition, it is assumed that the rates of the reaction can be computed using a Freundlich-type isotherm, as follows 35 : in which and are rate parameters and is a characteristic time that accounts for nonequilibrium behavior. It can be noted that molar concentrations are used to calculate the reaction rates throughout. In order to solve the system, both the initial conditions and BCs are required. The initial conditions considered here define the values of all variables at time t = t 0 throughout the domain and on the boundary, as follows: The BCs considered here are of the Cauchy type and the Dirichlet type. Those of the former type are a combination of imposed fluxes (which, in isolation, describe Neumann BCs) 32 and convective fluxes, which are a function of the difference in variables between the sample and the environment. Applied to the governing mass balance equations of moisture and of dissolved chemical species, the Cauchy-type BCs describe the rate of mass transfer from the environment to the sample and are given as where q w , q v , and q c are the prescribed fluxes of the moisture, moisture vapor, and chemical species, respectively, and c and c are the convective transfer coefficients of the moisture and chemical species, respectively. n is the unit normal vector to the boundary. The Dirichlet-type BCs fix the value of the variable on the boundary and are given by In the present study, governing Equations (3) and (8) are discretized using the finite element method. The resulting variational problem, after the application of the Gauss-Green divergence theorem, may be written as follows: find ∈ U, such that is the vector of primary variables, and the space for the trial functions is defined as In this study, the continuous Galerkin weighted-residual method 32 is employed such that the weight functions (W) are chosen to be equal to the shape functions (N), with the primary variables being the capillary pressure and the species' concentrations. The resulting system of discretized equations is as follows: where the superior dot denotes a time derivative and each variable (eg, P c ) is interpolated from the vector of nodal variables in the standard manner (ie, P c = ∑ N i P c i = NP c ). It is recognized that the continuous Galerkin method does not guarantee local mass conservation when applied to Equations (21) and (22) , and the system described by Equation (23) may be subject to spurious oscillations, 39 particularly for an advection-dominant case. To address this issue, a number of stabilization techniques may be used, including mass lumping, the streamline-upwind Petrov-Galerkin method, and enrichment of the Galerkin method. [39] [40] [41] Alternatively, a different method could be employed for spatial discretization, such as the discontinuous Galerkin method. 39 In the present work, however, no spurious oscillations were observed in the problems considered, and therefore, it did not prove necessary to employ any of the above stabilization techniques. The global matrices are given by where superscript "ne" is the number of elements and the element matrices are given as follows: The global system can be written in compact form as Applying an implicit Euler backward difference scheme 33 for time discretization leads to This set of nonlinear equations can then be solved using a standard Newton-Raphson procedure 33 based on a first-order Taylor series expansion of the mass/energy balance error, which leads to the following incremental-iterative update of the primary variable vector ( t+1 k+1 ): where is the approximation error, given here as Without loss of generality, bilinear quadrilateral elements were used throughout this study. Since convergence is not always guaranteed with the Newton-Raphson procedure, 39 the stability of a solution was checked using the following Courant-Friedrichs-Lewy condition, as suggested by Zhu et al 12 : The balance equations governing reactive transport in porous media (Equation (27)) are often highly coupled and nonlinear, and, as such, the computational demand associated with their solution can become prohibitively expensive, particularly when solutions are required for large domains and relatively long time periods. The present approach addresses this issue by decoupling a number of the nonlinear PDEs from the global system of equations. The proposed PRS is a predictor-corrector approach that employs one or more indicator species and then, in the "predictor" step, computes the transport of the other (non-indicator) species by interpolation. The "corrector" step then refines the predicted concentrations using error approximation. Reactions involving both indicator and non-indicator species are dealt with on a point-wise basis (at nodal points) at the end of each time step. Before describing the interpolations and detailed processes used in the PRS, the overall solution algorithm is presented to show where the PRS fits into the transient solution procedure. The aim of the PRS is to greatly reduce the computational cost of solving multispecies chemical transport problems by reducing the number of primary variables solved in the coupled system. In this scheme, a full solution is undertaken for the capillary pressure and a selected number of indicator species, and then, a predictor-corrector approach is used to compute the concentration of non-indicator species. The predictor first computes the concentration of non-indicator species from the concentration of the indicators, and then, a correction is applied to these values using error approximation derived from an appropriate balance equation. The concentrations of the indicator species at each time step are computed from the solution of the coupled equation system (Equation (27) and Algorithm 1). Once these indicator concentrations are known, the predictor step is applied to compute the transport of non-indicator species. To derive this predictor step, the governing mass balance equation for a conservative chemical species i is considered, which, neglecting the charge neutrality condition, is given as where D d i is a function of the liquid moisture velocity and the dispersivities (see Equation (14)). Noting that both D d i and v w are the same for all species, it can be seen from Equation (32) that, for a conservative chemical species, the difference between the rates of transport of chemical species depends on their diffusion coefficients, concentrations, and concentration gradients. In the present study, Lagrangian polynomial interpolation is used for diffusion, with the interpolated diffusion being used to weight the change in concentration of an indicator species over a time step, Δc ind , based on the relative difference between the known diffusion coefficients of the current non-indicator species D i and of the indicator species D ind and D jnd , as follows: in which g(D i , Δc ind ) is a diffusion interpolation function, which is incorporated into the final interpolation function below (Equation (34)), n ind is equal to the number of indicator species, subscripts "ind" and "jnd" denote specific indicator species, subscript i denotes non-indicator species, and D i and D ind represent the diffusion coefficient of any species and a specific indicator, respectively. It can be noted that, in the case of n ind = 1, Equation (34) Equation (33) does not account for the effect of the difference in concentration gradients between indicator and interpolated non-indicator species. This is addressed by introducing a concentration gradient function ( f c ), and adding terms to account for chemical reactions and the charge neutrality condition leads to Equation (34) for predicting the concentration of a non-indicator species at time t + Δt, as follows: where c i is the concentration of a chemical species, c gr refers to a concentration gradient, and superscript t denotes time. The penultimate term in Equation (34) represents the SS term due to chemical reactions, and the final term ( f) represents the diffusion due to the charge neutrality condition, which is elaborated in Section 4.3. In the scenario in which a reaction is "sufficiently fast" in comparison with the rates of the transport processes, this may be treated under the local equilibrium assumption. 38 As such, the change in concentration over a time step should be multiplied by a retardation factor, as detailed in the work of Fetter 42 , where R denotes the retardation factor for an equilibrium reaction and the SS term has been retained to account for any kinetic reactions). A number of options were considered for the concentration gradient function in Equation (34) ( f c ), including using the ratio of non-indicator-to-indicator gradients from the previous time step and using a weighted measure of concentrations from the sources. Both these options proved to be flawed and to have problem-dependent levels of accuracy. A more successful approach was to derive f c from a "source influence solution" (SIS), which is a linear steady-state solution of the diffusion problem for each chemical species that is carried out at the beginning of the solution procedure (thereby having a negligible effect on the overall solution time) and is given as noting that n, w , and D mol i are constants, for a uniformly saturated medium Equation (35) simplifies to the following Poisson equation: The outputs from the SISs are concentrations of each chemical species throughout the domain, which are subsequently employed in f c , as follows: The corrector refines the concentrations computed from the predictor step using the mass balance approximation error (Equation (30)) for each non-indicator species in turn. The concentration correction for each species i is then based on a first-order Taylor series expansion of a single species form of Equation (29), as follows: where c err i is the concentration correction vector for species i and M i is an approximate tangent coefficient matrix. Two forms of M i were considered as follows: The use of diagonal lumped matrices implies that the concentrations can be updated on a point-wise basis. The concentrations are refined as It was found that the two options gave similar results in terms of the accuracy of solutions, and therefore, the simpler option 2 was adopted for all subsequent computations. In this study, the corrector step was treated as non-iterative since it was found that sufficient accuracy could be obtained using a single corrector step, as illustrated in Sections 6 and 7 of this paper. In the present work, three different orders of the generalized reduction scheme are investigated, denoted PRS0, PRS1, and PRS2, which use one, two, and three indicator species, respectively. Recalling Equations (34) and (39), the predictor-corrector scheme is summarized as The initial conditions for the PRS can be defined in the same way as in the full model (ie, using Equation (16) for all species). The BCs for the PRSs are also defined in the same way as for the full model and should be of the same type for both "indicator" and "non-indicator" species, with the latter being calculated on a point-wise basis. Another key consideration is how the charge neutrality condition is satisfied in the PRSs, since the transport of non-indicator species is not calculated in the reduced system of governing equations. In the present approach, the diffusive flux due to the electric field is considered explicitly by using concentrations from the previous time step and moving these to the right-hand side of the governing equations in a similar manner to the way moisture flow under gravity is included in liquid transport computations. For non-indicator species, this can then be calculated on a point-wise basis and subtracted at the end of a PRS predictor step (Equation (40)). The diffusive flux due to the electric field is therefore given by The choice of indicator species is an important aspect of the PRS. For PRS1 and PRS2, species with the highest and lowest diffusion coefficients are always defined as indicators. This ensures that the computed responses of all non-indicator species are bounded by those of fully computed species. For PRS2, which requires a third indicator species, it was found, following an error analysis and sensitivity study (see Section 6.2), that the greatest accuracy was achieved by using a third indicator species with a mean diffusion coefficient. In cases where such a real species is not available, artificial species can be used. Artificial indicator species may also be used for other reasons; for example, as mentioned above, PRS1 and PRS2 use indicator species with the highest and lowest diffusion coefficients. However, the presence of any reactions associated with one of the bounding species, with reaction rates of the same order as (or higher than) those of the transport processes, may alter its rate of transport, such that it may no longer be the highest/lowest. In such situations, a nonreactive artificial indicator should be used in its place, thereby maintaining the solution bound. The criterion used to select the single indicator species required for PRS0 is given in Section 6.2 Another aspect of reactive transport problems, which is relevant to the choice of the indicator species, is that the transport of different species can take place over very different timescales. When this occurs due to a dominant (or "sufficiently fast" 38 ) reaction linked to a particular non-indicator species, this is dealt with by modeling the process as an equilibrium reaction. If this is due to other factors, then indicator species could be chosen to represent each different timescale. It is recognized that other factors may also affect the choice of indicators; however, the factors considered in the present work encompass a wide range of real problems. The number of indicator species, and associated order of the scheme, is also a key consideration when employing the PRS. The choice of the number of indicator species to use depends upon the problem under consideration and specifically upon the range of diffusion coefficients of the chemical species in the system. In general, the smaller the range, the lower the order of the scheme required, with PRS0 being applicable to problems in which the diffusion coefficients lie in a narrow range, as quantified in Section 6.2. An exception to this rule is problems with a significant degree of advection, for which PRS0 is not appropriate (unless the diffusion coefficients are equal for all species), since solutions for the non-indicator species transport are not bounded, which can lead to an overestimation of the advection. The specific selection criterion used to determine the indicators for each order of the PRS is given in Section 6.2, along with the expected accuracy for each degree of the scheme. Before investigating the behavior of the PRS, the validity of the full model is demonstrated. To this end, a non-steady-state diffusion problem reported by Baroghel-Bouny et al 35 is considered. This problem is based on experiments carried out by Francy 43 on cement discs of dimension 120(d) mm × 20(h) mm. In these tests, the left-hand side of the sample was exposed to a salt solution, whereas the remaining sides were sealed. The transport of Na + , OH − , K + , and Cl − ions was considered, and nonequilibrium chloride binding was also taken into account, using Equation (15) , in addition to the instantaneous formation of Friedel's salt. It is assumed that, since chloride ions sorb onto the solid mass, hydroxide ions are released to preserve charge neutrality. At the end of the test, measurements were taken of the free chloride and total chloride content (Tcc) of the sample at different locations. 35, 43 The problem setup can be seen in Figure 1 . The time period considered was 7 days, and, following a mesh and time-step convergence study, a time step of Δt = 3.6 s was selected, giving a total number of time steps of 168 000, along with a mesh of 20 bilinear quadrilateral elements with a maximum element size of 1 mm. To reflect experimental conditions, the sample was assumed to be initially saturated. The model parameters, BCs, and diffusion coefficients of the chemical species are given in Table 1 . It should be noted that, in this example, it was found that a tortuosity factor, D , was needed to correctly predict the chemical transport; this factor takes into account the tortuous pathways of the medium and is simply multiplied by the species diffusion coefficients. The concentration profile and the Tcc, as predicted by the full model, are compared with the corresponding experimental results in Figure 2 . The numerical results are considered sufficiently close to the experimental results to validate the full model. The CPU time for the simulation was 1382 seconds. The accuracy and range of applicability of the PRS with three different orders of the scheme (namely, PRS0, PRS1, and PRS2) are studied by considering a wick action test on a mortar sample in which the transport of 16 chemical species is simulated. The analysis undertaken in this study considered chemical reactions between the ions and the cement matrix, as well as advective and dispersive transport. It was decided to use an artificial set of chemical species for this study in order to have control over the range and spread of the diffusion coefficients considered. The reactions concerned the adsorption of the chemical ions onto the cement matrix, described by the nonequilibrium Freundlich isotherm (Equation (15)). The time period considered was 24 hours, and the initial concentration of each ion, as well as the sorbed chemical mass for each species in the sample, was assumed to be zero. The mortar sample was assumed to be initially saturated, prior to the left-hand side of the specimen being exposed to the chemical solution and the right-hand side being exposed to an environmental humidity of 60%, with all remaining sides being sealed, thereby ensuring one-dimensional transport. A nonuniform mesh of 25 bilinear quadrilateral elements was used along the length of the specimen, with a maximum element size of 4 mm. A time step of Δt = 36 s was used, giving a total number of time steps of 2400. As with the previous validation case, the mesh and time step were adopted following a convergence study. The diffusion coefficients for all species can be seen in Table 2 , whereas model parameters, including the boundary mass transfer coefficients for all chemical species, are given in Table 3 . To determine the accuracy of the reduction schemes, an analysis of the full problem was undertaken. The calculated profiles of chemical concentration, moisture content, and sorbed masses from this full analysis at time t = 24 h are given in Figure 3 . The responses are plotted in groups of species in order of increasing diffusion coefficient (D mol ). The relative difference between the four responses in each of the groups reflects the relative spread of diffusion coefficients across the group, with group 1 (species 1-4) representing a 6-fold increase in diffusion coefficient and group 4 having a relative increase of only 1.6. The moisture profile shows that some drying of the specimen has occurred over the 24 hour analysis period, but much of the sample remains saturated or near saturated. The sorbed mass profiles are very similar to the concentration profiles but with less penetration into the sample. The CPU time for the simulation was 312 seconds. In this section, the results from a set of analyses of the problem considered in Section 6.1, undertaken with the three orders of the PRS, are compared with the results from the full analysis described above. The indicator species chosen for each of the solutions are presented in Table 4 . An artificial species labeled "A" has been chosen for PRS0 and PRS2 in order to allow the use of an indicator with a diffusion coefficient corresponding to the mean value of the species diffusion coefficients. This problem is used to determine the range of applicability of each of the reduction schemes based on a maximum allowable error in any one chemical species at any one time. The tolerance measure considered is the mean absolute percentage error (MAPE), defined as An acceptable MAPE of 3% was selected for PRS0, and 1% was selected for both PRS1 and PRS2 to enable the definition of the range of applicability of each scheme. It is acknowledged that these tolerance values are a matter of judgement and that different problems will require different levels of accuracy. Here, 1% is considered to be a reasonable tolerance value for most practical purposes, and 3% is thought to be an acceptable error for the type of relatively coarse approximation associated with PRS0. For PRS0, it is possible to consider species with progressively larger or smaller diffusion coefficients until a profile exceeds the tolerance; however, this is not possible for PRS1 and PRS2 as they use the extremes of the diffusion coefficient range as indicator species. To overcome this issue, the range was successively varied on a trial-and-error basis until the maximum range that meets the tolerance was found. The resulting ranges, over which each PRS achieves the selected tolerance, are given in Table 5 (where superscripts u and l indicate upper and lower indicators, respectively). The intermediate indicator (D ind,m ) for PRS2 was chosen to be the mean diffusion coefficient, which was based on the results of a series of analyses aimed at finding the value of D ind,m that produced the most accurate solutions. To further explore the influence of changing the diffusion coefficient of an indicator species, a sensitivity analysis was undertaken on the value of the diffusion constant used for the intermediate indicator for PRS2. Using the problem considered in Section 6.1, it was found that changing this coefficient by 20% resulted in an increase in the MAPE of 0.13% and the mass balance error of 0.17%, which suggests that the scheme is relatively insensitive to the choice of the intermediate indicator. Profiles showing the two species with the largest departure from the results of the full analysis, within the range of applicability given in Table 5 , are presented for each PRS in Figure 4 . The relative percentage error plots for each PRS can be seen in Figure 5 , where the relative percentage error is given as RE = |(c prs i − c full i )∕c b i |. It can be seen from the profiles that PRS results are very close to those of the full model, particularly for PRS1 and PRS2. The profiles corresponding to the PRS 0 1 2 Diffusion coefficient 0.8D ind