Open access

On the Analytical Formulations for Pollutant Dispersion Simulation in the Atmospheric Boundary Layer

Written By

Davidson Martins Moreira, Tiziano Tirabassi and Taciana Albuquerque

Submitted: 06 November 2014 Published: 21 October 2015

DOI: 10.5772/60023

From the Edited Volume

Current Air Quality Issues

Edited by Farhad Nejadkoorki

Chapter metrics overview

1,388 Chapter Downloads

View Full Metrics

1. Introduction

The management and preservation of air quality require information of the environmental status. Such information involves both cognitive and interpretative features. Networks and measurements for monitoring, together with emission inventory of sources, are of great significance for the building of the cognitive representation, but not the interpretative one. Air quality control demands interpretative tools that are able to extrapolate in space and time the values measured by analytical instrumentation at field sites, while environmental improvement can only be obtained by means of emissions reduction from a systematic planning by using tools (mathematical models of dispersion) adequate to link the causes (sources) with the effects (pollutants).

The pollutants’ transport and diffusion processes are complex and without mathematical models it is impossible to account for them. Such models therefore are an indispensable technical instrument of air quality management.

The theoretical approach to the atmospheric dispersion assumes different points of view. The diffusion in the K approach is proportional to the local gradient concentration from diffused material (at the fixed point in space). As a result, it is fundamentally Eulerian, once it assumes the motion of fluid within a fixed system of reference.

Lagrangian models differ from Eulerian ones in adopting a system of reference that follows atmospheric motions. This category contains all models that consider the pollutant cloud as discrete “elements” (puffs or computer particles). In Lagrangian approach, the dispersion is simulated through the motion of particles whose path allows the concentration field computation from the liberated material.

In this chapter, we consider the Eulerian approach.

Advertisement

2. K models

The Eulerian models takes into account the resolution of the mass conservation equation of pollutant chemical species [1]:

ct+uc=D2c+S.E1

In (1), u is the wind speed vector, D∇ 2c is the molecular diffusion, D is the molecular diffusion coefficient, S is the term source, is an operator, and ∇2 is the Laplacian.

To obtain the solution of equation (1), it is inevitable to know the wind field, something that is not possible because it is changing in space and time. As a consequence, wind is divided into two parts:

u¯, ensemble average

u, turbulent fluctuation

Therefore, the wind speed can be expressed as the sum of the two contributions:

u=u¯+uE2

with u¯'=0.

Similar hypothesis can be assumed for the concentration:

c=c¯+cE3

with c¯ = 0.

The new u and c are introduced into (1). Then, the following is obtained:

c¯t=u¯c¯cu¯+D2c¯+S¯.E4

This equation possesses new unknown variables. This feature leads to a number of unknowns greater than the number of equations. By solving this problem it is possible to parameterize some terms with known quantities.

The widely used approach is the parameterization of second order moments. Such an approach is called as K-theory or flux-gradient theory:

cu¯=Kc¯,E5

where K is the well-known eddy diffusivity.

A great variety of K-formulations exists [2]. The major part them take into account the similarity theory, and give different results for the same atmospheric stability.

Applying the following approximations:

  1. Tensor is diagonal

  2. c represents the concentration of a non-reactive pollutant (thusS = S)

  3. The molecular diffusion is insignificant

Then, equation (4) can be written as:

c¯t=u¯c¯+Kc¯+S.E6

Equation (6) can be solved analytically or numerically, with data for u, K, and S and the initial and boundary conditions forc. This equation can be resolved in two ways:

  1. With analytic methods, obtaining exact solutions

  2. With numerical methods, obtaining approximate solutions

In this chapter we take into consideration analytical solutions.

Advertisement

3. Analytical solutions

Analytical solutions of equations are very important to understand and describe physical phenomena. Moreover, they are fast, simple, and, generally, do not require complex meteorological inputs.

There are analytical solutions of two-dimensional advection-diffusion equation [3]:

uc¯(x,z)x=z(Kzc¯(x,z)z)+ S,E7

where u is longitudinal mean speed, c¯is the mean concentration, S is the source term, and Kz is the vertical eddy diffusivity. Furthermore, the longitudinal diffusion can be neglected because it is considered less in respect to the advection. Although, very recently a steady-state analytical solution for dispersion of contaminants in low winds by taking into account the longitudinal diffusion in the advection-diffusion equation was formulated [4].

The best-known solution is the so-called Gaussian solution, where both wind and turbulent diffusion coefficients are constant with height. So it is not a realistic solution of the transport and diffusion equation in the atmosphere. In the so-called Gaussian models, the solution is forced to represent real situations by means of empirical parameters, referred to as “sigmas.” The different versions of Gaussian models substantially differ in the techniques utilized to calculate the “sigmas” as a function of atmospheric stability and the downwind distance from the emission source. Gaussian models are fast, simple, and do not require complex meteorological inputs. For these reasons they are still widely used by all the environmental agencies over the world for regulatory applications. At this point, it is important to mention that there are models based on non-Gaussian analytical solutions.

[5] report a bi-dimensional solution considering cases where the wind speed and vertical eddy diffusivities follow power laws as a function of height:

u=u1(z/z1)αE8
Kz=K1(z/z1)β,E9

where z1 is the height when u1 and K1 are computed.

[6] report a bi-dimensional solution for linear profiles of the eddy diffusivity. [7] showed the bi-dimensional equation with u and K power functions of height with the exponents following the law of Schmidt.

[8] shows a solution considering u constant, but K as follows:

Kz=K0za(hz)b,E10

where Kzis a constant and a and b can be expressed as:

a³ 0 ; b = 0a = 0 ; b > 0 for 0£ha = 1 ; b > 0 for 0 z ha = 1 ; b = 0 for 0 h/2 ; a = 0 and b = 1 for h/2 h,

where h is the ABL height.

[9] proposed a solution with constant u and Kz as:

Kzzfor0zzsE11
Kz=Kz(z1)forzs<zh,E12

where zs is a predetermined height (generally, the height of the surface layer). This solution allows (as boundary conditions) a net flow of material toward the ground:

Kzc¯z=Vgc¯,E13

where Vg is the deposition velocity.

[10, 11] report bi-dimensional solutions considering elevated sources with wind speed and vertical eddy diffusivity power profiles, but for an unbound atmosphere. That is:

Kzc¯z=0at z = .E14

[12] shows a solution considering identical conditions, but for a vertically limited boundary layer. That is:

Kzc¯z=0at z = h .E15

The solutions [10, 11, 12] are used in the KAPPAG model [13, 14, 15].

[16] using the Monin-Obukhov similarity theory to diffusion, has derived a solution from continuous sources near the ground with u and K follow power profiles. SPM [17] is a model that uses this solution.

[18] report a solution, which was a particular case of [7, 8]. In the sequence, [19] extended the solution to the situation of a growing ABL. [20] extended the solution to the case of nonzero mean vertical wind profiles.

[21] presented a three-dimensional atmospheric dispersion-deposition model for a ground-level area source.

[22] extended the solution obtained in [12] with boundary conditions considering dry deposition to the ground.

In the work [23] were reported equations for point sources releases considering the first four moments of the vertical concentration distribution and the magnitude and downwind location of the maximum ground concentration.

[24] obtained a three-dimensional solution (with u and Kz constant) for low wind condition where the diffusion coefficients are function of down-wind distance from the source. [25] presented a two-dimensional solution with Kz constant and function of down-wind distance from the source, with a power low wind profile and for a finite boundary layer.

[26] obtained an analytical solution with dry deposition to the ground with any vertical function of wind and eddy coefficients but for a fixed vertical shape of contaminant concentrations.

c¯(x,z)=F(x)(1zh)2,E16

where F(x) is any function of x.

Using ADMM (Advection-Diffusion Multilayer Method) approach, [27] proposed a general solution for any wind and eddy coefficient profiles, but represented by a stepwise function in z. Recently, using the ADMM approach, [28] found a solution for two-dimensional steady-state solution considering eddy coefficient profile depending of x and z variables.

Finally, [29, 30] applying GILTT technique (Generalized Integral Laplace Transform Technique), found a general two-dimensional steady-state solution for any profiles of wind and eddy coefficient diffusions and a limited ABL.

3.1. GILTT approach solution

The General Integral Transform Technique (GITT) is a well-known hybrid method that solved a wide class of direct and inverse problems mainly in the area of Heat Transfer and Fluid Mechanics. There is a vast literature about this subject, including papers and books that become impossible to mention all of them. But, for instance, we mention the works of [31, 32, 33, 34], etc. In this chapter, we restrict our attention to the linear problem, because for nonlinear problem the linear result is iteratively used after the linearization of nonlinear transformed equation. Indeed, for the linear problems, the GITT transformed problem is analytically solved by the Laplace Transform technique. Consequently, the GITT solution becomes an analytical solution in the sense that no approximations are made along its derivation. It is important to notice that the GITT solution for nonlinear problems is semi-analytical because, in each iteration, the solution is analytical. This methodology is known as GILTT (Generalized Integral Laplace Transform Technique).

Bearing in mind that the novelty of the GILTT approach relies on the analytical solution of GITT linear transformed problem, in what follows, we restrict our analysis to the ensuing standard GITT time-dependent transformed problem:

EdY(t)dt+FY(t)=0t>0,E17

subject to the initial condition,

Y(0)=Y0.E18

Here, Y(t) is the unknown vector with N components, E and F are constant matrices of order (NxN), and Y0 is a known vector with N components. We begin our analysis recasting equation (17) like,

dY(t)dt+GY(t)=0t>0,E19

where G = E-1F and E-1 denotes the inverse of the matrix E. Now, applying the Laplace transform technique to equation (19), we obtain,

sY(s)¯+GY(s)¯=Y0,E20

where Y(s)¯ denotes the Laplace transform of the vector Y(t). Assuming that the matrix G is nondegenerate, we may write,

G=XDX1.E21

Here, D is the eigenvalue of the diagonal matrix, X is the matrix of the respective eigenfunctions, and X-1 its inverse.

Indeed, replacing equation (21) in equation (20) we get,

sY(s)¯+XDX1Y(s)¯=Y0,E22

or

(sI+XDX1)Y(s)¯=Y0,E23

where the matrix I is the (NxN) identity matrix. Recalling that X X-1 = X-1X = I, equation (23) reads like,

(sXX1+XDX1)Y(s)¯=Y0,E24

or

X(sI+D)X1Y(s)¯=Y0,E25

which has the well-known solution,

Y(s)¯=X(sI+D)1X1Y0.E26

Performing the Laplace transform inversion of equation (26), we have,

Y(t)=XL1{(sI+D)1}X1Y0.E27

Here, L-1 denotes the inverse Laplace transform operator. We must notice that the matrix (sI+D) has the form,

(sI+D)=(s+d1000s+d2000s+dN),E28

where di are the eigenvalues of the matrix G. From the diagonal structure of the matrix G, we can straightly write down its inverse like:

(sI+D)1=(1s+d10001s+d20001s+dN).E29

Performing the Laplace transform inversion of equation (29) by using the standard results of the Laplace transform theory we obtain:

L1{(sI+D)1}=E(t)=(ed1t000ed2t000edNt).E30

Finally, substituting the above ansatz in equation (27), we come out with the following solution for problem (17),

Y(t)=X(ed1t000ed2t000edNt)X1Y0.E31

To this point, it is relevant to underline that we are aware that problem (17) has a well-known solution. However, we must point out that the discussed solution is a robust algorithm, under computational point of view, to solve the problem with large N (N of order of 1, 500) and a small computational effort. Furthermore, this methodology can also be readily applied for the solution of equation (17) with boundary condition. This kind of a problem appears in the solution of discrete ordinates equation in a slab by the LTSN approach. For more information, see [35].

Next, we extend this methodology to the solution of more general linear ordinary differential equation having the entries of the matrices E and F varying with time. To reach this goal, let us consider next the problem:

E(t)dY(t)dt+F(t)Y(t)=0t>0,E32

subject to the initial condition,

Y(0)=Y0.E33

To solve this problem by the Laplace transform technique, we perform a stepwise approximation of the matrices E(t) and F(t). So far, bearing in mind the interest of finding a solution for the time ranging from zero to a prescribed time T, we split the interval (0, T) into sub-interval (tn-1, tn) for n=1:M, where M denotes the number of sub-intervals. For each sub-interval, we take the averaged values for the entries of matrix E(t) and F(t). It turns out that problem (28) simplifies to the recursive set of problems,

EmdYm(t)dt+FmYm(t)=0,E34

for t in the sub-interval (tm-1, tm) and m ranging from 1 to M. Here, Em and Fm are now constant matrices. Henceforth, the previously discussed solution can be applied in a straightforward manner. Indeed, the solution for problem (34) generically reads like,

Y(m)=X(m)(ed1(m)t000ed2(m)t000edN(m)t)X(m)1Y(m1),E35

for t in the interval (tm-1, tm) and the initial condition Y(m-1) given by the previously calculated solution at tm-1 for t in the interval (tm-2, tm-1).

Focusing our attention on the task of searching analytical solution for the GITT linear transformed equation, in the sequel, we report an analytical alternative approach to solve equation (19), skipping the stepwise approximation of matrices E(t) and F(t) entries. To hit this objective, we solve problem (33) by the decomposition method proposed by [36]. In order to apply the decomposition method, let us recast equation (34) as,

E¯dY(t)dt+F¯Y(t)=A(t)dY(t)dt+B(t)Y(t),E36

where the matrices A(t) and B(t) are, respectively, expressed as,

A(t)=E(t)+E¯E37

and

B(t)=F(t)+F¯.E38

Here, E¯ and F¯ are constant matrices properly chosen. Now, expanding the solution of equation (36) in the truncated series,

Y(t)=k=0Luk(t)E39

and replacing this ansatz in equation (36), we get,

E¯ddtk=0Luk(t)+F¯k=0Luk(t)=A(t)ddtk=0Luk(t)+B(t)k=0Luk(t).E40

From equation (40), we are in a position to construct the following recursive set of linear ordinary differential matrix equations with constant matrices,

E¯du0(t)dt+F¯u0(t)=0,E41
E¯du1(t)dt+F¯u1(t)=A(t)du0(t)dt+B(t)u0(t),E42
E¯du2(t)dt+F¯u2(t)=A(t)du1(t)dt+B(t)u1(t)E43

and so forth. Generically, we may write,

E¯dun(t)dt+F¯un(t)=A(t)dun1(t)dt+B(t)un1(t),E44

for n = 1:L. Given a closer look at the above equations, we readily realize that the RHS of the recursive set of equations from (42) to (44) are known terms. Therefore, to obtain the solution of these equations, we must solve the following problem with constant matrices and source, namely,

E¯dY(t)dt+F¯Y(t)=S(t)t>0,E45

which has the well-known solution,

Y(t)=Yh(t)Y0+Yh(t)S(t),E46

whereas the homogeneous solution Yh(t) is given by equation (35). Here, star denotes convolution. Therefore, the solution of the generic equation (44) has the form,

Y(t)=X(ed1t000ed2t000edNt)X1Y0++0t(X(t')(ed1t'000ed2t'000edNt')X(t')1S(tt'))dt'E47

and consequently the solution of equation (32) is well determined by equations (39) and (41) to (44).

Finally, regarding the task of solving second-order ordinary differential transformed equation, we must recall that this equation can be transformed into a set of first-order matrix differential equation by changing the variable. Indeed, performing the substitution of,

Y(t)=Z1(t)E48

and

dY(t)dt=Z2(t),E49

in the following transformed problem:

d2Y(t)dt+A(t)dY(t)dt+B(t)Y(t)=S(t),E50

we come out with the ensuing result:

ddt(Z1(t)Z2(t))+(01B(t)A(t))(Z1(t)Z2(t))=(0S(t)),E51

which is a set of first-order ordinary matrix differential equation that can be promptly solved by the previous discussed methods for nondegenerated matrix. Finally, to handle transformed problems with degenerated matrix, we proceed by performing the inversion of the symbolic matrix by the Schur decomposition approach. Furthermore, we make the Laplace transform inversion applying the Heaviside expansion formulation for eigenvalues with multiplicity larger than one. Hoping that we have completed the mathematical analysis regarding the task of solving analytically the GITT transformed problems, in the next section, we illustrate the GILTT aptness to simulate pollutant dispersion in the atmosphere by solving the two-dimensional steady-state advection-diffusion equation.

3.1.1. An example: Solution of the 2D advection-diffusion equation

To solve the problem (7) by the GILTT method we recast equation (7) as (S = 0):

u(z)c¯x=Kz(z)2c¯z2+(Kz(z)z)c¯zE52

with the following boundary conditions:

Kzc¯x=0at z = hE53
uc¯(x,z)=Qδ(zHs)at x = 0E54

and we construct the following associated Sturm-Liouville problem:

Ψi(z)+λi2Ψi(z)=0at 0 <z<hE55
Ψi(z)=0at z = 0,h ,    E56

which has the well-known solution:

Ψi(z)=cos(λiz),E57

where the eigenvalues λi are given by λi=iπh for i=0,1,2,3,.Furthermore, The eigenfunctions Ψi(z) satisfy the ensuing orthonormality condition:

1Nm1/2Nn1/2vΨm(z)Ψn(z)dv={0, m n 1, m = n,E58

where Nm is expressed by:

Nm=vΨm2(z)dv .E59

Having solved the Sturm-Liouville problem, we are in a position to construct the GILTT transform formula, which has the form:

c¯(x,z)=i=0ci¯(x)¯Ψi(z)Ni1/2.E60

Now, replacing the above ansatz in equation (52) we have,

ui=0ci¯(x)¯Ψi(z)Ni1/2=Kzi=0ci¯(x)¯Ψi(z)Ni1/2+(Kzz)i=0ci¯(x)¯Ψi(z)Ni1/2.E61

Here, we adopt the prime notation for the first derivative and double the prime notation for the second derivative. Multiplying equation (61) by Ψj(z)Nj1/2 and integrating from z equal zero to h, we read:

i=0ci¯(x) ¯ Ni1/2Nj1/20huΨiΨj dzi=0ci¯(x)¯λi2Ni1/2Nj1/20hKzΨiΨj dz++i=0ci¯(x)¯Ni1/2Nj1/20h(Kzz) Ψi Ψj dz=0E62

forj=0,1,2,,N. Rewriting equation (62) in matrix fashion, we obtain,

Y(x)+FY(x)=0,E63

where Y(x) is the column vector whose components are cj(x) and the matrix F is defined as F=B1E. Here, the entries of matrices B and E are, respectively, given by:

bi,j=0huΨiΨjdzE64

and

ei,j=0hKzzΨiΨjdzλi20hKzΨiΨjdz.E65

Transforming the boundary condition (54) by a similar procedure, we mean, multiplying the boundary condition (54) by Ψj(z)Nj1/2 and integrating from z = 0 to h, we obtain:

0hi=0ci¯(0) ¯uΨiΨj Ni1/2Nj1/2dz=0hQ δ (z-Hs)ΨjNj1/2dz.E66

Using the orthonormality property of the eigenfunctions Ψj(z), and the integration property of the generalized delta function, we get the following transformed boundary condition:

c0(0)=QΨ0(Hs)h  u(z)Ψ02(z)dzE67

for j = 0, and

ci(0)=QΨj(Hs)h2  0hu(z)Ψj2(z)dzE68

for j ranging from 1 to N. To this point, we are in a position to write down the solution of equation (39), subjected to the boundary condition given by equations (67) and (68), using the results discussed in the previous section. Indeed, the solution writes like,

c¯(x,z)=i=0ci¯(x)¯Ψi(z)Ni1/2.E69

Here, ci(x) are the N components of the vector Y(x) given by,

Y(x)=Z(ef1x000ef2x000efNx)G(x)Z1Y(0),E70

where fk, for k ranging from 1 to N, are the eigenvalues of the matrix of eigenfunctions and Z1 its inverse. For more details and recent developments, see [29, 30, 37, 38, 39, 40, 41].

Advertisement

4. Remarks

To deal with more realistic situations we need to change to numerical methods. However, it is helpful to check firstly possible analytical solutions in order to obtain a known framework and test solutions. The analytical solutions are efficient for many applications, for example: supply analysis of contaminant scenarios, leading sensitivity analyses to investigate the effects of parameters included in contaminant transport, extrapolation over large times and distances where numerical solutions may be impractical, as screening models for transport processes that cannot be solved exactly, and for validating numerical solutions.

Today, air pollution problems are not treated in the manner described in the present chapter. There are various air pollution situations that require the use of complex mesoscale models to properly describe the dispersion processes and properly represent the relevant chemistry and emission processes. Complex models, such as the CMAQ model (the Community Multiscale Air Quality model), have been designed to simulate air quality by including state of the art techniques for modeling multiple air quality issues. However, in complex models, increasingly more processes, such as sea breeze circulations, urban heat islands, and waves, are represented. Therefore, these models are often perceived as “closed” that cannot without difficulty or effort report the influence of individual processes on air quality. Finally, for many scientific applications, analytical solutions have utility in understanding air dispersion phenomena and some air chemistry phenomena, showing their usefulness in environmental management.

References

  1. 1. Arya, S. 1999. Air pollution meteorology and dispersion. Oxford University Press, New York, 310 pp.
  2. 2. Seinfeld, J.H., and Pandis, S.N. 1998. Atmospheric chemistry and physics. John Wiley & Sons, New York, 1326 pp.
  3. 3. Tirabassi, T. 2003. Operational advanced air pollution modeling. PAGEOPH 160(1–2), 5–16.
  4. 4. Moreira, D.M., Tirabassi, T., and Carvalho, J.C. 2005a. Plume dispersion simulation in low wind conditions in stable and convective boundary layers. Atmos. Environ. 39(20), 3643–3650.
  5. 5. Roberts, O.F.T. 1923. The theoretical scattering of smoke in a turbulent atmosphere. Proc. Roy. Soc. 104, 640–648.
  6. 6. Rounds, W. 1955. Solutions of the two-dimensional diffusion equation. Trans. Am. Geophys. Union 36, 395–405.
  7. 7. Smith, F.B. 1957a. The diffusion of smoke from a continuous elevated point source into a turbulent atmosphere. J. Fluid Mech. 2, 49–76.
  8. 8. Smith, F.B. 1957b. Convection-diffusion processes below a stable layer. Meteorological Research Committee, N. 1048 and 10739, London.
  9. 9. Scriven, R.A., and Fisher, B.A. 1975. The long range transport of airborne material and its removal by deposition and washout-II. The effect of turbulent diffusion. Atmos. Environ. 9, 59–69.
  10. 10. Yeh, G.T., and Huang, C.H. 1975. Three-dimensional air pollutant modelling in the lower atmosphere. Bound. Layer Meteor. 9, 381–390.
  11. 11. Berlyand, M.Y. 1975. Contemporary problems of atmospheric diffusion and pollution of the atmosphere. Translated version by NERC, USEPA, Raleigh, NC.
  12. 12. Demuth, C. 1978. A contribution to the analytical steady solution of the diffusion equation for line sources. Atmos. Environ. 12, 1255–1258.
  13. 13. Tagliazucca, M., Nanni, T., and Tirabassi, T. 1985. An analytical dispersion model for sources in the surface layer, Nuovo Cimento 8C, 771–781.
  14. 14. Tirabassi, T., Tagliazucca, M., and Zannetti, P. 1986. KAPPA-G, a non-Gaussian plume dispersion model: description and evaluation against tracer measurements. JAPCA 36, 592–596.
  15. 15. Tirabassi, T. 1989. Analytical air pollution advection and diffusion models. Water, Air Soil Poll. 47, 19–24.
  16. 16. Van Ulden, A.P. 1978. Simple estimates for vertical diffusion from sources near the ground. Atmos. Environ. 12, 2125–2129.
  17. 17. Tirabassi, T., and Rizza, U. 1995. A practical model for the dispersion of skewed puffs. J. Appl. Meteor. 34, 989–993.
  18. 18. Nieuwsadt, F.T.M. 1980. An analytical solution of the time-dependent, one-dimensional diffusion equation in the atmospheric boundary layer. Atmos. Environ. 14, 1361–1364.
  19. 19. Nieuwstadt, F.T.M., and de Haan, B.J. 1981. An analytical solution of one-dimensional diffusion equation in a nonstationary boundary layer with an application to inversion rise fumigation. Atmos. Environ. 15, 845–851.
  20. 20. Catalano G.D. 1982. An analytical solution to the turbulent diffusion equation with mean vertical wind. In: Proceedings of the 16th Southeastern Seminar on Thermal Sciences, Miami, 19–21 April 1982, 143–151.
  21. 21. Chrysikopoulos, C.V., Hildemann, L.M., and Roberts, P.V. 1992. A three-dimensional atmospheric dispersion-deposition model for emissions from a ground-level area source. Atmos. Environ. 26, 747–757.
  22. 22. Lin, J.S., and Hildemann, L.M. 1997. A generalised mathematical scheme to analytically solve the atmospheric diffusion equation with dry deposition. Atmos. Environ. 31, 59–71.
  23. 23. Brown, M.J., Arya, S.P., and Snyder, W. 1997. Plume descriptors from a non-gaussian concentration model. Atmos. Environ. 31, 183–189.
  24. 24. Sharan, M., and Yadav, A.K. 1998. Simulation of diffusion experiments under light wind, stable conditions by a variable K-theory model. Atmos. Environ. 32, 3481–3492.
  25. 25. Sharan, M., and Modani, M. 2006. A two-dimensional analytical model for the dispersion of air-pollutants in the atmosphere with a capping inversion. Atmos. Environ. 40, 3479–3489.
  26. 26. Essa, K.S.M., Etman, S.M., and Embaby, M. 2007. New analytical solution of the dispersion equation. Atmos. Res. 84, 337–344.
  27. 27. Vilhena, M.T., Rizza, U., Degrazia, G.A., Mangia, C., Moreira, D.M., and Tirabassi, T. 1998. An analytical air pollution model: development and evaluation. Contr. Atmos. Phys. 71, 315–320.
  28. 28. Moreira, D.M., Moraes, A.C., Goulart, A.G., and Albuquerque, T.T. 2014. A contribution to solve the atmospheric diffusion equation with eddy diffusivity depending on source distance. Atmos. Environ. 83, 254–259.
  29. 29. Wortmann, S., Vilhena, M.T., Moreira, D.M., and Buske, D. 2005. A new analytical approach to simulate the pollutant dispersion in the PBL. Atmos. Environ. 39, 2171–2178.
  30. 30. Moreira, D. M., Vilhena, M.T., Tirabassi, T., Buske, D., and Cotta, R.M. 2005b. Near source atmospheric pollutant dispersion using the new GILTT method. Atmos. Environ. 39, 6289–6294.
  31. 31. Cotta, R.M. 1993. Integral transforms in computational heat and fluid flow. CRC Press, Boca Raton.
  32. 32. Cotta, R.M., and Mikhaylov, M. 1997. Heat conduction lumped analysis, integral transforms, symbolic computation. John Wiley & Sons, Chichester.
  33. 33. Cheroto, S., Mikhailov, M.D., Kakaç, S., and Cotta, R.M. 1999. periodic laminar forced convection: solution via symbolic computation and integral transforms. Int. J. Therm. Sci. 38, 613–621.
  34. 34. Alves, L.S., Cotta, R.M., and Pontes, J. 2002. Stability analysis of natural convection in porous cavities through integral transforms. Int. J. Heat Mass Transfer 45, 1185–1195.
  35. 35. Segatto, C.F., and Vilhena, M.T. 1999. The state-of-the-art of the LTSN method, mathematics and computation, reactor physics and environmental analysis in nuclear applications – International Conference, Madrid, 2, 1618–1631.
  36. 36. Adomian, G. 1988. A review of the decomposition method in applied mathematics. J. Math. Anal. Appl. 1(135), 501–544.
  37. 37. Buske, D., Vilhena, M.T., Moreira, D.M., and Tirabassi, T. 2007a. Simulation of pollutant dispersion for low wind conditions in stable and convective planetary boundary layer. Atmos. Environ. 41, 5496–5501.
  38. 38. Buske, D., Vilhena, M.T., Moreira, D.M., and Tirabassi T. 2007b. An analytical solution of the advection-diffusion equation considering non-local turbulence closure. Environ. Fluid Mech. 7, 43–54.
  39. 39. Moreira, D.M., Vilhena, M.T., Buske, D., and Tirabassi, T. 2009. The state-of-art of the GLTT method to simulate pollutant dispersion in the atmosphere. Atmos. Res. 92, 1–17.
  40. 40. Tirabassi, T., Buske, D., Moreira, D.M., and Vilhena, M.T. 2008. A two-dimensional solution of the advection-diffusion equation with dry deposition to the ground. JAMC 47, 2096–2104.
  41. 41. Tirabassi, T., Tiesi, A., Buske, D., Vilhena, M.T., and Moreira, D.M. 2009. Some characteristics of a plume from a point source based on analytical solution of the two-dimensional advection-diffusion equation. Atmos. Environ. 43, 2223–2229.

Written By

Davidson Martins Moreira, Tiziano Tirabassi and Taciana Albuquerque

Submitted: 06 November 2014 Published: 21 October 2015