649
Views
0
CrossRef citations to date
0
Altmetric
Article

Linear Transport in Porous Media

, & ORCID Icon

Abstract

The linear transport theory is developed to describe the time dependence of the number density of tracer particles in porous media. The advection is taken into account. The transport equation is numerically solved by the analytical discrete ordinates method. For the inverse Laplace transform, the double-exponential formula is employed. In this paper, we consider the travel distance of tracer particles whereas the half-space geometry was assumed in our previous paper [Amagai et al. (Citation2020). Trans. Porous Media 132:311–331].

1. Introduction

The use of the transport equation for the flow in porous media was proposed by Williams (Citation1992a, Citation1992b, Citation1993a, Citation1993b). Recently it was experimentally shown that the concentration of tracer particles in column experiments obeys the transport equation (Amagai et al. Citation2020). In Amagai et al. (Citation2020), the advection u > 0 was taken into account. Then the spatial derivative term in the transport equation is given by (u+v0μ)/x instead of μ/x, where v0>0 is the inherent particle speed and μ[1,1] is the cosine of the polar angle. The transport equation with such a spatial derivative term has been explored in the context of the evaporation of rarefied gas (Loyalka, Siewert, and Thomas Citation1981; Scherer and Barichello Citation2009; Siewert and Thomas Citation1981, Citation1982).

In column experiments (Cortis et al. Citation2004), a column tube is filled with a granular material such as sand or glass beads, and water is poured from the top end of the column. Then tracer particles are injected to the water. They enter the column from the top and eventually exit from the bottom of the column. In Amagai et al. (Citation2020), the concentration of tracer particles was computed making use of the method of analytical discrete ordinates (ADO). Both the short-time growing behavior and long-time decay behavior of the breakthrough curve (the time dependence of the concentration) were well reproduced by the transport equation (Amagai et al. Citation2020).

In Amagai et al. (Citation2020), the solution to the transport equation for a semi-infinite medium was compared to the experimental data. In this paper, we give a formulation for the transport equation in the slab geometry taking into account the length of the column. We make use of the double-exponential formula for the inverse Laplace transform.

The rest of the paper is organized as follows. In Section 2, we formulate our transport problem in the slab geometry. In Section 3, a numerical scheme is developed using ADO. In Section 4, our numerical scheme of the inverse Laplace transform is described. Finally, concluding remarks are given in Section 6.

2. The transport equation

Let us consider the one-dimensional linear Boltzmann equation. The velocity v is given by v=u+v0μ. Let σa>0 and σs>0 be the absorption and scattering coefficients, respectively. Let L be the length of the column. We write our transport equation as follows. (2.1) {Pψ(x,μ,t)=σs211ψ(x,μ,t) dμ,0<x<L,1μ1,t>0,ψ(x,μ,0)=0,0<x<L,1μ1,ψ(0,μ,t)=n0δ(μ1),η<μ1,t>0,ψ(L,μ,t)=0,1μ<η,t>0,(2.1) where Pψ(x,μ,t)=[t+(u+v0μ)x+σa+σs]ψ(x,μ,t).

Here, δ(μ1) is the Dirac delta function, n0 is the initial particle number density, and η=uv0.

We call ψ(x,μ,t) the angular number density. The particle number density n(t) at x = L is given by n(t)=η1ψ(L,μ,t) dμ.

3. The laplace transform

Let us introduce the Laplace transform ψ̂(x,μ,p)=0eptψ(x,μ,t) dt,pC, and new variables μt=σa+σs+pv0C,μs=σsv0>0.

Then we have {(η+μ)xψ̂(x,μ,p)+μtψ̂(x,μ,p)=μs211ψ̂(x,μ,p) dμ,0<x<L,1μ1,ψ̂(0,μ,p)=n0pδ(μ1),η<μ1,ψ̂(L,μ,p)=0,1μ<η.

We note that the coefficient μt is a complex number and the boundary conditions are specified by intervals (η,1] and [1,η). We will obtain ψ̂(x,μ,p) in the above-mentioned equation using the analytical discrete ordinates method (ADO) (Barichello Citation2011; Barichello and Siewert Citation1999a, Citation1999b; Barichello, Garcia, and Siewert Citation2000; Barichello and Siewert Citation2001; Siewert and Wright Citation1999).

Remark 3.1.

By changing μ+ημ˜ and defining f̂(x,μ˜,p)=ψ̂(x,μ˜η,p), we can reformulate the equation as {μ˜xf̂(x,μ˜,p)+μtf̂(x,μ˜,p)=μs2η1η+1f̂(x,μ˜,p) dμ˜,0<x<L,η1μ˜η+1,f̂(0,μ˜,p)=n0pδ(μ˜η1),0<μ˜η+1,f̂(L,μ˜,p)=0,η1μ˜<0,

Such transformation is particularly useful for the evaporation problem, in which the integral on the right-hand side of the transport equation is taken from to (Scherer and Barichello Citation2009).

Let us write ψ̂(x,μ,p)=ψ̂b(x,μ,p)+ψ̂s(x,μ,p).

The ballistic term ψ̂b satisfies {(η+μ)xψ̂b(x,μ,p)+μtψ̂b(x,μ,p)=0,0<x<L,1μ1,ψ̂b(0,μ,p)=n0pδ(μ1),η<μ1,ψ̂b(L,μ,p)=0,1μ<η, and the scattering term ψ̂s obeys {[(η+μ)x+μt]ψ̂s=μs211ψ̂s(x,μ,p) dμ+q,0<x<L,1μ1,ψ̂s(0,μ,p)=0,η<μ1,ψ̂s(L,μ,p)=0,1μ<η, where q(x,μ,p)=μs211ψ̂b(x,μ,p) dμ=n0μs2pexμt/(η+1).

We note that ψ̂b(x,μ,p)=n0pexμt/(η+μ)δ(μ1).

Let us express ψ̂s(x,μ,p)=ψ̂s(x,μ) and q(x,μ,p)=q(x,μ) when there is no confusion. For the computation of ψ̂s, we discretize the integral by the Gauss-Legendre quadrature and obtain [(η+μi)x+μt]ψ̂s(x,μi)=μs2j=1Nwj[ψ̂s(x,μj)+ψ̂s(x,μj)]+q(x,μi), where μi,wi (i=1,2,,2N) are abscissas and weights, respectively. We have 0<μ1<<μN<1 and μN+i=μi (i=1,,N). Furthermore, we introduce Nη as the largest integer such that η<μNη.

Remark 3.2.

It is possible to assign different abscissas and weights for two intervals [1,η) and (η,1]. Since we assume η is small, we use one set of abscissas and weights for the interval [1,1] as described above.

The scattering part ψ̂s is obtained as ψ̂s(x,μi)=j=12N0LG(x,μi;x,μj)q(x,μj) dx, where the Green’s function defined for each p satisfies {[(η+μi)x+μt]G(x,μi;x,μj)=μs2k=12NwkG(x,μk;x,μj)+δ(xx)δij,G(0,μi;x,μj)=0,μi>μNη,G(L,μi;x,μj)=0,μi<μNη, where δij is the Kronecker delta.

Let us consider the following homogeneous equation. ((η+μi)x+μt)ψ̂(x,μi)=μs2j=12Nwjψ̂(x,μj).

We note that ψ̂ depends on p through μt. With separation of variables, we can write ψ̂ as ψ̂(x,μi)=ϕ(ν,μi)ex/ν, where ν is the separation constant. The function ϕ(ν,μi) satisfies the normalization condition, i=12Nwiϕ(ν,μi)=i=1Nwi(ϕ(ν,μi)+ϕ(ν,μi))=1.

We obtain ϕ(ν,μi)=μsν21μtνμiη, assuming ν(μi+η)/μt. If η = 0 and p is real, we can prove νμi/μt (Siewert and Wright Citation1999). The following orthogonality relation holds. i=12Nwi(μi+η)ϕ(ν,μi)ϕ(ν,μi)=N(ν)δνν, where N(ν)=i=12Nwi(μi+η)ϕ(ν,μi)2.

We can find 2 N eigenvalues ν=νn (n=1,2,,2N). The eigenvalues can be obtained as the reciprocal of eigenvalues of the following matrix (Amagai et al. Citation2020). (Ξ+1Ξ1)[μt(II)μs2(WWWW)], where I is the identity, Ξ±=diag(±μ1+η,±μ2+η,,±μN+η), and {W}ij=wj. The subroutine ZGEEV (LAPACK subroutine for a complex nonsymmetric matrix) was used to compute νn. Moreover, the free-space Green’s function G0 is obtained as G0(x,μi;x,μj)=±±Rνn>0wjN(νn)ϕ(νn,μj)ϕ(νn,μi)e(xx)/νn, where upper signs are chosen for x>x and lower signs are used for x<x.

Hence we can write G(x,μi;x,μj)=G0(x,μi;x,μj)+Rνn>0B1(νn)ϕ(νn,μi)ex/νn+Rνn<0B2(νn)ϕ(νn,μi)ex/νn, where coefficients B1(νn),B2(νn) are determined from boundary conditions. We obtain (3.1) Rνn>0B1(νn)ϕ(νn,μi1)+Rνn<0B2(νn)ϕ(νn,μi1)=y1(μi1),(3.1) (3.2) Rνn>0B1(νn)ϕ(νn,μi2)eL/νn+Rνn<0B2(νn)ϕ(νn,μi2)eL/νn=y2(μi2),(3.2) where 1i1Nη,Nη<i22N, y1(μi1)=Rνn<0wjN(νn)ϕ(νn,μj)ϕ(νn,μi1)ex/νn,y2(μi2)=Rνn>0wjN(νn)ϕ(νn,μj)ϕ(νn,μi2)e(Lx)/νn.

Let us multiply Equation(3.1) and Equation(3.2) by exp(xμt/(η+1)), integrate both sides of these equations over x, and take the sum with respect to j. We obtain (3.3) Rνn>0E1(νn)ϕ(νn,μi1)+Rνn<0E2(νn)ϕ(νn,μi1)eL/νn=z1(μi1),(3.3) (3.4) Rνn>0E1(νn)ϕ(νn,μi2)eL/νn+Rνn<0E2(νn)ϕ(νn,μi2)=z2(μi2),(3.4) where z1(μi1)=Rνn<0(η+1)νnN(νn)(η+1νnμt)[eL/νnLμt/(η+1)1]ϕ(νn,μi1),z2(μi2)=Rνn>0(η+1)νnN(νn)(η+1νnμt)[eLμt/(η+1)eL/νn]ϕ(νn,μi2), and {E1(νn)=j=12N0LB1(νn)eμtx/(η+1) dx,E2(νn)=eL/νnj=12N0LB2(νn)eμtx/(η+1) dx.

Thus E1(νn),E2(νn) are obtained from Equation(3.3) and Equation(3.4). To solve this linear system, the subroutine ZGESV (LAPACK subroutine for a complex system of linear equations) was used.

Hence we obtain ψ̂s(x,μi)=n0μs2pj=12N0LG(x,μi;x,μj)exμt/(η+1) dx=n0μs2pRνn>0(η+1)νnN(νn)(η+1νnμt)(eμtx/(η+1)ex/νn)ϕ(νn,μi)+n0μs2pRνn<0(η+1)νnN(νn)(η+1νnμt)exμt/(η+1)(1e(Lx)/νn(Lx)μt/(η+1))ϕ(νn,μi)+n0μs2p[Rνn>0E1(νn)ϕ(νn,μi)ex/νn+Rνn<0E2(νn)ϕ(νn,μi)e(Lx)/νn].

The Laplace transform of n(t) is obtained as n̂(p)=η1ψ̂(L,μ,p) dμ=η1ψ̂b(L,μ,p) dμ+η1ψ̂s(L,μ,p) dμ.

Therefore, n̂(p)=n0peLμt/(η+1)+n0μs2pRνn>0(η+1)νnN(νn)(η+1νnμt)(eLμt/(η+1)eL/νn)φ(νn)+n0μs2p[Rνn>0E1(νn)eL/νnφ(νn)+Rνn<0E2(νn)φ(νn)], where φ(ν)=i=1Nηwiϕ(ν,μi)=μsν2i=1Nηwiμtνμiη.

By the inverse Laplace transform, we have (3.5) n(t)=n02πiγiγ+ieptp{eLμt/(η+1)+μs2Rνn>0(η+1)νnN(νn)(η+1νnμt)(eLμt/(η+1)eL/νn)φ(νn)+μs2[Rνn>0E1(νn)eL/νnφ(νn)+Rνn<0E2(νn)φ(νn)]} dp,(3.5) where γ is taken to be greater than the largest real part of any singularity.

4. The inverse Laplace transform

Let us numerically evaluate the Bromwich integral in the inverse Laplace transform Equation(3.5). Although the trapezoidal rule was used in Amagai et al. (Citation2020), here we employed the double-exponential formula (Ooura and Mori Citation1991, Citation1999).

We note that In̂(γ+iω/t)=Rn̂(γ+i(ωπ2)/t) because n̂(γ+iωπ2t)=0eγti(ωπ2)n(t) dt=in̂(γ+iωt).

For t > 0, we have (4.1) n(t)=eγt2πteiωn̂(γ+iωt) dω=eγtπtcosω Rn̂(γ+iωt) dω=2eγtπt0cosω Rn̂(γ+iωt) dω.(4.1)

Let us introduce (4.2) ϕ(τ)=τ1eK sinhτ,K=6.(4.2)

Then the above integral can be written as n(t)=2eγtπtcos(Mϕ(τ)) Rn̂(γ+itMϕ(τ))Mddτϕ(τ) dτ, where ω=Mϕ(τ), M > 0. Let us define h=π/M. To numerically evaluate the above integral, the discretization by the trapezoidal rule is the suitable choice (Sugihara Citation1997; Trefethen and Weideman Citation2014). By the trapezoidal rule, we arrive at (4.3) n(t)2eγttk=kmaxkmaxcos(Mϕ(τ)) Rn̂(γ+iMtϕ(τ))ϕ(τ),(4.3) where kmax is an integer and τ=kh+π2M. We note that ϕ(τ)0 double exponentially as τ. We see that ϕ(τ)τ double exponentially as τ and cos(Mϕ(kh+π2M))cos(Mkh+π2)=0. The choice of ϕ(τ) in Equation(4.2) is not unique. The constant K = 6 was found by numerical experiments (Ooura and Mori Citation1991). In in Appendix A, we confirmed that any K6 gives the same numerical results. Although ϕ(τ) in Equation(4.2) sufficiently works, a more general expression ϕ(τ)=τ[1exp(2τα(1eτ)β(eτ1))]1 (β=1/4,α=β/1+Mln(1+M)/(4π)) has been proposed (Ooura and Mori Citation1999).

Table A1. The dependence of n(L,t)/n0 on K for N = 30, kmax=40, M = 40.

Table A2. The dependence of n(L,t)/n0 on N for K = 6, kmax=40, M = 40.

We set L=10 cm. We found γ=0.04 is suitable. Time was discretized as tj=jΔt (Δt=0.2 min,j=1,,250). Furthermore we set σa=108 min1,σs=5 min1,v0=5 cm/min, and u=1.5 cm/min. For the numerical calculation, parameters were chosen to be N = 30, kmax=40, and M=40. To confirm the validity of these values, in Appendix A compare n(L,t)/n0 for different parameters.The computation time was 90 sec on a laptop computer (MacBook Pro, 2.3 GHz Intel Core i5). The result is plotted in . In Amagai et al. (Citation2020), the particle number density was defined by (4.4) n(x,t)=11ψ(x,μ,t) dμ,(4.4) where ψ(x,μ,t) is the solution to Equation(2.1) when L. also shows n(L,t)/n0 for comparison.

Figure 1. The particle number density is plotted as a function of t. The blue curve is from (4.3) and the red curve is from (4.4).

Figure 1. The particle number density is plotted as a function of t. The blue curve is from (4.3) and the red curve is from (4.4).

Table A3. The dependence of n(L,t)/n0 on kmax for K = 6, N = 30, M = 40.

Table A4. The dependence of n(L,t)/n0 on M for K = 6, N = 30, kmax=40.

To confirm the convergence of the inverse Laplace transform, we have tried different N, kmax, and M.

Remark 4.1.

According to the error analysis in Ganapol (Citation2008), γt must be small. In Ganapol (Citation2008), it is suggested to move γ according to t, and the form γ=γ¯+α/t is proposed, where γ¯,α are positive constants.

5. Column experiment

A one-dimensional flow field is achieved in a column. We conducted a column experiment using adsorptive solute of zinc solution.

Before the tracer breakthrough curve was observed, ultra-pure water was injected by the peristaltic pump (MP-1000, Eyela) for 24 hours until the flow field became steady. The flow rate of the injected tracer solution was controlled by the peristaltic pump. Then the tracer solution was injected to displace the fresh water. The discharged solution was collected by the fraction collector (CHF161RA, Advantec) at regular intervals. Column experiments were performed at room temperature of 25–26 °C. The experiment was conducted until the discharge concentration became equal to the influent concentration. For the sake of the accuracy of measurements, we corrected the measurement time error from the tube volume and inlet volume of the column by subtracting the time lag due to the switchover from the breakthrough time.

The filling material is the standard sand (Tohoku silica sand No. 4, Kitanihon Sangyo). The median diameter of the sand is 750 μm. As a preparation, we eliminate the organic matter that may have been contained in the sand by soaking it in HNO3 solution. The zinc solution was 2 ppm. We set a filter on the top of the sand bed made with glass wool and put the same filter at the bottom of the column. A column of length 12.0 cm with internal diameter 3.1 cm was used. The bed height was L=10.7 cm and the section area was 7.54 cm2. It was confirmed by a blank test that the zinc was not absorbed on the surface of the column wall. The concentration was measured with the atomic absorption photometer (Z-2300, Hitachi High-Technologies), the compressor (SC820, Koki Holdings), and the Neo Cool Circulator (CF700, Yamato). The measured porosity was 0.289 and the flow rate was 11.25 cm3/h.

shows the zinc breakthrough curve. The measured values from the column experiment (black open circles) are compared with numerical values calculated from Equation(4.3) and Equation(4.4) using the linear transport equation in the slab geometry (blue dashed line) and half-space geometry (red line). In the left panel of , the vertical axis is the relative concentration C/C0 and the horizontal axis is time t. The semilog plot in the right panel of shows 1C/C0 as a function of t. When numerically obtaining C/C0, we determined the parameters by the Levenberg-Marquardt algorithm (Levenberg Citation1944; Marquardt Citation1963) using the FORTRAN library MINPACK (More, Garbow, and Hillstrom Citation1980). The fitted values are summarized in . The parameter values for the half space in were obtained in Amagai et al. (Citation2020). We note that C/C0 and n/n0 are related as C/C0=βn/n0 with constant β, which is determined by fitting. We see in that two curves from the transport equation are almost identical with different parameter values.

Figure 2. The breakthrough curve of C/C0 (Left) and 1C/C0 (Right) as a function of time. The experimental data (black open circles) are compared with numerical values computed from (4.3) and (4.4) using the linear transport equation in the slab geometry (blue dashed line) and half-space geometry (red line).

Figure 2. The breakthrough curve of C/C0 (Left) and 1−C/C0 (Right) as a function of time. The experimental data (black open circles) are compared with numerical values computed from (4.3) and (4.4) using the linear transport equation in the slab geometry (blue dashed line) and half-space geometry (red line).

Table 1. Fitted parameters.

6. Concluding remarks

Although solutions of the transport equation well described experimentally obtained breakthrough curves in Amagai et al. (Citation2020), the half space was assumed in the formulation. To take into account the length of the column, we gave a formulation in the slab geometry, which has both ends and tracer particles enter from one end and exit from the other end. It should be emphasized that indistinguishable breakthrough curves are obtained for different sets of fitted parameters. This means that it is important to impose the boundary conditions which correctly model the experimental setup.

For the numerical inversion of the Laplace transform, we could apply the double-exponential formula after expressing n(t) using cosω in Equation(4.1).

In this paper, isotropic scattering is assumed. The introduction of the anisotropy factor g0 is a future issue.

The most precise geometry for the column experiment is a cylinder in three dimensions. However, the essential nature of the transport is expected to be seen by the one-dimensional transport equation since the experimental setup is designed so that the flow is identical in horizontal directions.

Acknowledgment

The column experiment was conducted by Motoko Yamakawa.

Additional information

Funding

M.M. acknowledges support from Grant-in-Aid for Scientific Research (17K05572, 18K03438) of JSPS.

References

  • Amagai, K., M. Yamakawa, M. Machida, and Y. Hatano. 2020. The linear Boltzmann equation in column experiments of porous media. Transp. Porous Med. 132 (2):311–31.
  • Barichello, L. B. 2011. Explicit formulations for radiative transfer problems. In Thermal measurements and inverse techniques, ed. H. R. B. Orlande, O. Fudym, D. Maillet, and R. M. Cotta. Boca Raton, FL: CRS Press.
  • Barichello, L. B., and C. E. Siewert. 1999a. A discrete-ordinates solution for a non-grey model with complete frequency redistribution. J. Quant. Spec. Rad. Trans. 62 (6):665–75.
  • Barichello, L. B., and C. E. Siewert. 1999b. A discrete-ordinates solution for a polarization model with complete frequency redistribution. Astro. J. 513 (1):370–82.
  • Barichello, L. B., and C. E. Siewert. 2001. A new version of the discrete-ordinates method. Proceedings 2 International Conference on Computational Heat and Mass Transfer, Rio de Janeiro, 22–26.
  • Barichello, L. B., R. D. M. Garcia, and C. E. Siewert. 2000. Particular solutions for the discrete-ordinates method. J. Quant. Spec. Rad. Trans. 64 (3):219–26.
  • Cortis, A., Y. Chen, H. Scher, and B. Berkowitz. 2004. Quantitative characterization of pore-scale disorder effects on transport in “homogeneous” granular media. Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 70 (4 Pt 1):041108. doi:https://doi.org/10.1103/PhysRevE.70.041108
  • Ganapol, B. D. 2008. Analytical benchmarks for nuclear engineering applications case studies in neutron transport theory. Paris: Nuclear Energy Agency, OECD.
  • Levenberg, K. 1944. A method for the solution of certain non-linear problems in least squares. Quart. Appl. Math. 2 (2):164–8.
  • Loyalka, S. K., C. E. Siewert, and J. R. Thomas. Jr., 1981. An approximate solution concerning strong evaporation into a half space. Z. Angew. Math. Phys. 32 (6):745–7.
  • Marquardt, D. W. 1963. An algorithm for least-squares estimation of nonlinear parameters. SIAM J. Appl. Math. 11 (2):431–41.
  • More, J. J., B. S. Garbow, and K. E. Hillstrom. 1980. User guide for MINPACK-1. Argonne National Laboratory Report ANL-80-74.
  • Ooura, T., and M. Mori. 1991. The double exponential formula for oscillatory functions over the half infinite interval. J. Comp. Appl. Math 38 (1-3):353–60.
  • Ooura, T., and M. Mori. 1999. A robust double exponential formula for Fourier-type integrals. J. Comp. Appl. Math. 112 (1–2):229–41.
  • Scherer, C. S., and L. B. Barichello. 2009. Evaporation effects in rarefied gas flows. Proc. COBEM 2009 COB09:0684.
  • Siewert, C. E. 2000. A concise and accurate solution to Chandrasekhar’s basic problem in radiative transfer. J. Quant. Spec. Rad. Trans. 64 (2):109–30.
  • Siewert, C. E., and J. R. Thomas Jr. 1981. Strong evaporation into a half space. Z. Angew. Math. Phys. 32 (4):421–33.
  • Siewert, C. E., and J. R. Thomas Jr. 1982. Strong evaporation into a half space. II. The three dimensional BGK model. Z. Angew. Math. Phys. 33 (2):202–18.
  • Siewert, C. E., and S. J. Wright. 1999. Efficient eigenvalue calculations in radiative transfer. J. Quant. Spec. Rad. Trans. 62 (6):685–8.
  • Sugihara, M. 1997. Optimality of the double exponential formula – functional analysis approach. Num. Math. 75 (3):379–95.
  • Trefethen, L. N., and J. A. C. Weideman. 2014. The exponentially convergent trapezoidal rule. SIAM Rev. 56 (3):385–458.
  • Williams, M. M. R. 1992. A new model for describing the transport of radionuclides through fractured rock. Ann. Nucl. Energy 19 (10–12):791–824.
  • Williams, M. M. R. 1992. Stochastic problems in the transport of radioactive nuclides in fractured rock. Nucl. Sci. Eng. 112 (3):215–30.
  • Williams, M. M. R. 1993. A new model for describing the transport of radionuclides through fractured rock Part II: Numerical results. Ann. Nucl. Energy 20 (3):185–202.
  • Williams, M. M. R. 1993. Radionuclide transport in fractured rock a new model: Application and discussion. Ann. Nucl. Energy 20 (4):279–97.

Appendix A.

Parameters for numerical calculation