You are on page 1of 10

Chemical Engineering Journal 279 (2015) 442–451

Contents lists available at ScienceDirect

Chemical Engineering Journal


journal homepage: www.elsevier.com/locate/cej

Effective radiation field model to scattering – Absorption applied


in heterogeneous photocatalytic reactors
Miguel Angel Mueses a, Fiderman Machuca-Martinez b,⇑, Aracely Hernández-Ramirez c, Gianluca Li Puma d
a
Photocatalysis and Solar Photoreactors Engineering, Department of Chemical Engineering, Universidad de Cartagena, A.A. 1382-Postal 195 Cartagena, Colombia
b
GAOX Research Group, School of Chemical Engineering, Universidad del Valle, A.A. 25360 Cali, Colombia
c
Universidad Autonóma de Nuevo León, Facultad de Ciencias Químicas, Cd. Universitaria, C.P. 66450 San Nicolás de los Garza, NL, Mexico
d
Environmental Nanocatalysis and Photoreaction Engineering, Department of Chemical Engineering, Loughborough University, Loughborough LE11 3TU, United Kingdom

h i g h l i g h t s g r a p h i c a l a b s t r a c t

! A new concept effective radiation


field model for solar heterogeneous
photoreactors.
! The model estimates the radiation
field for suspension and
polychromatic systems.
! Modified radiative transfer equation
for several reactors could be used
easily.
! The incident photons flow in the
photoreactor is an isotropic energy
cloud. Simulation of LVRPA in CPC solar reactor using ERFM to different catalyst concentration
! The global energy is independent of
the propagation angle and photon
frequency.

a r t i c l e i n f o a b s t r a c t

Article history: A new mathematical model for the calculation of the radiation field in heterogeneous photocatalytic reac-
Received 24 February 2015 tors using the new concept of ‘‘effective radiation field model’’ or ERFM is proposed. In this concept, the
Received in revised form 14 May 2015 incident radiation associated to the photons flow is an energy cloud. The generated space-phase and the
Accepted 15 May 2015
properties of the cloud are considered isotropic and independent of the propagation angle and photon
Available online 22 May 2015
frequency. The isotropic nature of the ERFM concept provides a simple estimation of the radiation field
of a catalyst in suspension (particles and fluid) for polychromatic radiation and the solar spectrum.
Keywords:
The ERFM is an alternative model for the calculus of the radiant energy distribution in heterogeneous
Radiation field
SFM
photocatalytic reactors as an extension of concept to the overall volumetric rate photon absorption –
LVRPA OVRPA. The local volumetric rate of photon absorption (LVRPA) predicted by the ERFM were compared
Solar photoreactors with the Six Flux Model (SFM) and the rigorous solution using Discrete Ordinate Method (DOM) for
Photocatalyst the radiative transfer equation (RTE).
The calculated LVRPA with the ERFM was found to be closer to the solution of the RTE-DOM. These
results were attributed to the performance of the phase function in both models.
! 2015 Elsevier B.V. All rights reserved.

1. Introduction
Abbreviations: ERFM, effective radiation field model; OVRPA, overall volumetric
rate photon absorption; SFM, Six Flux Absorption Scattering Model; RTE, radiative Heterogeneous photocatalysis is a platform technology, for
transfer equation; DOM, Discrete Ordinate Method. environmental, renewable energy and green synthesis applica-
⇑ Corresponding author. Tel./fax: +57 2 3312935. tions. A huge amount of literature exists about the fundamental
E-mail address: fiderman.machuca@correounivalle.edu.co aspects of photocatalysis [1,2] including the synthesis of
(F. Machuca-Martinez).

http://dx.doi.org/10.1016/j.cej.2015.05.056
1385-8947/! 2015 Elsevier B.V. All rights reserved.
M.A. Mueses et al. / Chemical Engineering Journal 279 (2015) 442–451 443

photocatalytic materials, the mechanisms of photocatalytic oxida- time [12,13,15,23]. Additionally, boundary conditions need to be
tion of contaminants [3], application of photocatalysis for environ- evaluated with accurate and specialized actinometric techniques
mental remediation [4], production of renewable energy [5] and [24].
for the green synthesis of chemicals [6,7]. The above limitations are a disadvantage in the simulation and
The spatial distribution of radiation field and the radiation scaling-up of photocatalytic reactors at the solar scale.
absorption are key aspects in heterogeneous photocatalytic pro- Furthermore, diurnal fluctuations of the solar energy do not permit
cesses. In the presence of a radiation field made of photons of to have a steady incident radiation flux and atmospheric effects
wavelength with energy higher than the band-gap of the semicon- and geographical conditions further increase the complexity of
ductor photocatalyst, electron–hole pairs can be generated, which the model solutions and actinometric treatment [26–31].
produce the reduction and oxidation reactions at the surface of the Regarding to solar photocatalytic reactors, the Six-Flux
photocatalyst. Highly reactive radical species such as hydroxyl, Scattering-Absorption Model (SFM) has been implemented to
peroxyl radicals and superoxide anion, [8–10] are often produced describe radiation field and rate of photon absorption in pollutants
which can oxidize or reduce organic and inorganic contaminants. degradation applications [25,32–33]. This method in photoreactors
There is consensus that the evaluation and optimization of the was first proposed as a modification of the Two-Flux Model (TFM)
radiation absorption in solar photoreactors is a very important [34,35], and it was validated by a comparison with Monte Carlo
step, in order to achieve better results on the application of the simulation of the radiation field in a flat heterogeneous photoreac-
photocatalytic processes [2,8,11]. tor [36].
The direct and indirect radiation with the photoreaction kinet- The original SFM describes photon scattering as a diffuse func-
ics is associated through of the quantum yield (the photoactivation tion (although this is not a limiting condition) and considers that
step of the semiconductor), and the distribution of the local volu- incident photons have the probability to disperse in trajectories
metric rate of photon absorption (LVRPA) [10–14]. Efforts in math- described by the six directions of the Cartesian coordinates (hence
ematical modeling are centered in the description of the LVRPA to the name) [25,26]. Its mathematical structure is of algebraic nat-
facilitate the prediction of the experimental data in the heteroge- ure; therefore, its implementation in reactors to different scales
neous photocatalytic degradation of contaminants in solution. and radiation sources is practical, with a low complexity in
The LVRPA is conventionally modeled from the contribution of numeric procedures and short computational times [36].
the specific radiation intensity (i.e., irradiance) Ik(x, X, t) integrated Nevertheless, the presence of a diffuse phase function can cause
over all propagation directions (Einstein"s#1) and the global volu- that the method to fail in correctly predict radiant field perfor-
metric coefficient of absorption kk (cm#1) [10]: mances generated by simulation by the solution of Mie theoretical
Z equation [10].
eak ¼ jk Ik ð~ ~ÞdXdk
x; X ð1Þ Other approaches in the descriptions of radiant fields in hetero-
X geneous photocatalytic reactors as stochastic models, which are
based on Monte Carlo simulation [37–39] and Computational
The value of Ik(x, X, t) is obtained by solving the radiative trans-
Fluid Dynamics (CFD) [21,40–43] based on the constitutive equa-
fer equation (RTE), which describes photon transport through an
tions of continuous-medium mechanics. Although these models
immobilized material or though a material dispersed in a fluid.
are as accurate as the DOM solution, they have not been currently
At steady state, the primary form of the RTE used to model the
implemented at a solar scale due to their high computational time
radiation fields in heterogeneous photocatalytic reactors is [15]:
and mathematical cost.
In this study, a new model for the evaluation of the radiation
d
Ik ð~ ~Þ ¼ #½jk ð~
x; X xÞ þ rk ð~
xÞ)Ik ð~ ~Þ þ 1 rk ð~
x; X xÞ
ds 4p field in heterogeneous photocatalytic reactors of different scales
Z is proposed. The model is based on the concept of effective radia-
~ÞdX
qðk0 ! k; X0 ! XÞIk ð~x; X ~0 ð2Þ
~ ¼4p tion [20], in which, the radiant field is quantified as isotropic
X
energy flux of photons, which is independent of the propagation
where Ik is the incident radiation at wavelength k, s is the general- direction and is particular suitable for application in solar powered
ized coordinate, jk is the absorption volumetric coefficient (m#1), rk photoreactors.
is the scattering volumetric coefficient (m#1), ~
x represents the coor-
dinate vector, and q (k0 ? k, X0 ? X) is the scattering phase func-
tion (the term in parentheses represents the frequency
re-directing in radiation k and propagation X [16,17]): 2. Methodology
Usually, the modeling of the radiation field in photocatalytic
reactors has been applied on systems using artificial radiation 2.1. Effective radiation model (ERFM) postulates
sources (UV lamps). The main modeling approach follows the rigor-
ous solution of the RTE (Eq. (2)) using the Discrete Ordinate Method The following postulates were made support on the
(DOM) [10,15]. This method has been applied to plane photoreactor effective-radiation concept [20]:
geometries [11], tubular photoreactors [10,13,18,19], and com- Postulate 1. In isotropic phase-space, which contains suspended
pound parabolic collectors (CPC) reactors [11,12,20–22] using particles in a fluid phase, the net global effective radiant power
monochromatic and polychromatic irradiation. that arrive to a surface a(v), identified by a vector of coordinate
The RTE solved by DOM method allow an accurate prediction of x, can be calculated summing up the energy E = hm (with h the
the radiation field in heterogeneous photocatalytic systems. This Planck constant) by each photon of wavelength k, frequency m
approach has been validated with simulation data based on the and propagation direction X. This net global energy will therefore
solution of Mie theoretical equation. Among different expressions be independent of the propagation direction and the radiation
of the scattering phase function, it has been found that the isotro- wavelength (Fig. 1).
pic phase function is one of the most appropriate to describe the Postulate 2. The net global effective energy associated to a
scattering properties of semiconductor photocatalyst [14,23]. phase-space of suspended solid particles is assumed to have an iso-
However, since the RTE is an integral–differential equation the tropic nature. Thus, the optical properties of suspension can be
application of the DOM for its solution requires a high degree of considered as independent global parameters of the geometric
numerical accuracy, which is reflected in a high computational coordinates [20,44].
444 M.A. Mueses et al. / Chemical Engineering Journal 279 (2015) 442–451

Fig. 1. Schematic diagram of the effective radiation concept model.

Postulate 3. Due to the isotropic nature of the system, the scat- propagation directions X, by a differential area ds, as follows:
tering probability is identical in each direction. Thus, the suspen-
sion is modeled exclusively using an isotropic phase function. ds ¼ dA cosðhÞ ð5Þ
Postulate 4. The net energy of particle absorption is quantified Z * Z
by the effective volumetric rate of energy absorption (EVREA), this dE ~ ÞdXdk
dX ¼ x; X
Ik ð~ ð6aÞ
parameter is equivalent to LVRPA in the traditional approach. X ds X

Z Z
d ~ÞdXdk
2.2. Mathematical formulation of effective radiation field model E* dX ¼ Ik ð~
x; X ð6bÞ
ds
(EFRM) X X

Then, nEff is defined as the global, net radiant energy of the pho-
The differential radiant energy dE⁄of a beam (J or Ein) in the ton flux with a wavelength dk through Eq. (7):
wavelength interval dk on any time t (system at steady-state), Z
which propagates in an unit solid angle dX, through a surface area nEff ¼ E* dX ð7Þ
X
dA (Fig. 2) results in a specific radiation intensity Ik(x, X, t) equals to:
Using the definition of Spectral Incident Radiation Gk(x) [15]:
! * " Z
~Þ ¼ dE
Ik ð~
x; X lim ð3Þ Gk ð~
x; tÞ ¼ Ik ð~ ~ÞdX
x; X ð8Þ
dA;dX;dt;dk!0 dAcosðhÞdXdk
X
At steady-state, the differential radiant energy equals to: Replacing (7), (8) in (6b), the expression for the net differential
* ~ÞdA cosðhÞdXdk incident energy on the surface ds of an arbitrary particle p in a
x; X
dE ¼ Ik ð~ ð4Þ
heterogeneous system of suspended solid particles is obtained.
As a result of postulate 2, the net radiant power is independent
of the direction of propagation of radiation, therefore, it is possible dnEff ð~
x; kÞ ¼ Gk ð~
x; tÞds ð9Þ
to express the radiation intensity as an integral function in all Integrating Eq. (9) over the external surface of an arbitrary par-
ticle considering that the particles are rigid non-porous spheres
with an uniform size and summing for all NV particles in the sus-
pension, it follows:
Z Nv Z
X Z
nEff
T;k ðx; kÞ ¼
~ dnEff ð~
x; kÞ ¼ x; tÞds ¼ NV
Gk;k ð~ x; tÞds
Gk;k ð~
s k¼1 s s

ð10Þ
Eq. (10) represents the total energy of the effective monochro-
matic spectral incident accumulated radiation in all semiconductor
particles in suspension, nEff
T , which is independent of the direction
of propagation of the incident radiation. According to Postulate 3
the net global effective radiant energy that is absorbed by the
semiconductor is:

dnEff
T;k ðx; kÞ
~
+ nEff Eff
N;k ¼ jk fN;k ðx; kÞ
~ ð11Þ
dt
where fEff
N;k is the EVREA for a monochromatic radiant energy of
wavelength k, jk is the volumetric absorption coefficient of the
Fig. 2. Schematic definition of specific radiation intensity Ik(x, X, t). semiconductor, that is the concentration and wavelength
M.A. Mueses et al. / Chemical Engineering Journal 279 (2015) 442–451 445

dependent, t is the reaction volume and fEff


N;k is the effective super-
For a plane geometry with constant radiation along the axial
ficial rate of energy absorption (ESREA per area unit). direction, a(v) is constant, v = z and the phase space is confined
For a polychromatic radiation, Eq. (11) is integrated over the between [0,d], with d the reactor thickness. Then, the correspond-
entire wavelengths range of the incident spectrum: ing radiative transfer equation is:
Z Z @ Eff
nEff nEff jk fEff f ðzÞ ¼ #½j þ r)fEff Eff
N ðzÞ þ r/N ðzÞ ð16Þ
N;k dk ¼ N;k ðx; kÞdk
N ¼
~ ð12Þ z N
k k

Applying the global, constant, optical parameter approach With the global boundary condition: fEff Eff
N;k = fN;0 for z = 0.

developed by Mueses et al. for solar radiation [20,44] Eq. (12) is For tubular, concentric or cylindrical reactor systems such as
finally expressed as follows: CPCs, where the transfer area of the net, effective energy flux is
not constant, v = r and the phase space is confined between [0,
nEff Eff
N ¼ jfN ðxÞ
~ ð13Þ R], with R the reactor radius. Thus, the corresponding radiative
transfer equation is:
where nEff
is the net effective volumetric rate of energy absorption
N
(W/m2), see Fig. 1, j is the global volumetric absorption coefficient 1 @ h Eff i
of the semiconductor, and fEff
rfN ðrÞ ¼ #½j þ r)fEff Eff
N ðrÞ þ r/N ðrÞ ð17Þ
N is the net, global effective radiation r r
per area unit, which is solved using a modification of the radiative
Boundary condition: fEff
N = fN,0
Eff
for r = 0.
transfer equation.

2.4. Estimation of the phase function of the global effective scattering


2.3. Modified radiative transfer equation
model, /Eff
N

We first consider an arbitrary, differential control volume in a


The description of the scattering properties of a catalyst suspen-
heterogeneous fluid system with suspended solid particles dt(v),
sion requires the knowledge of the scattering coefficient r and the
where v is the space-phase Cartesian position that involves the dif-
scattering function /Eff
N [10,11,36]. The isotropic phase function is
ferential control volume (see Fig. 3).
consistent with the concept of effective radiation field, and there-
The net global effective energy fEff
N (E"L
#2
) which is incident over
fore it was considered the most appropriate in this model [20]. A
the differential area da(v), under steady state, involves the follow-
mathematical formulation similar to the modified RTE (Eqs. (14)
ing balance for each component:
and (15)) around the phase space in Fig. 3 was considered:
EIN is the net radiant-energy flux entering the control volume
dt(v) through v; EABS is the energy flux absorbed by the semicon- /Eff Eff Eff Eff Eff
N ðvÞaðvÞjvþdv # /N ðvÞaðvÞjv þ ½rfN ðvÞ # j/N ðvÞ # r/N ðvÞ)dtðvÞ ¼ 0
ductor; ES,OUT is the energy lost by scattering effects; ES,IN is the ScatteringIN ScatteringOUT ScatteringS;IN ScatteringABS ScatteringS;OUT
energy gained from the surround by scattering effects; EOUT the
ð18Þ
net flux of radiant energy inside the control volume dt(v) by v + dv.
Mathematically: where, ScatteringIN is the net global effective energy by scattering
# $ entering the phase space; ScatteringOUT is the net global effective
fEff Eff Eff
N ðvÞaðvÞjv # fN ðvÞaðvÞjvþdv # ½j þ r) fN ðvÞdtðvÞ þ r/N
Eff
fEff
N ¼0 energy output; ScatteringS,IN is the gain by scattering related to
EIN EOUT EABS ES;OUT
ES;IN the net global effective energy fEf
N
f
not-absorbed; ScatteringABS is
ð14Þ the dispersed global energy lost by scattering effect. Analogous to
the modified RTE (Eq. (15)), an equation for radiation scattering is
where j and r are the global volumetric absorption and scattering
obtained, given by Eq. (19):
coefficients respectively, /Eff
N is the global phase function of effective
scattering, which is a function of fEff
N . 1 @ h Eff i
/N ðvÞaðvÞ ¼ ½j þ r)/Eff Eff
N ðvÞ # rfN ðvÞ ð19Þ
The new radiative transfer equation modified with the effective aðvÞ v
radiation approach considers a global net energy flux, which is
The expressions for plane (or axial) geometry and for cylindrical
independent of the direction of propagation. Considering a phase
systems are:
space in function of an arbitrary coordinate v, it is given by Eq.
(15): @ Eff
/ ðzÞ ¼ #½j þ r)/Eff Eff
N ðzÞ þ rfN ðzÞ ð20Þ
1 @h i z N
fEff
N ð vÞðvÞ ¼ #½j þ r )fEff
N ð vÞ þ r /Eff Eff
N ðfN Þ ð15Þ
aðvÞ v 1 @ h Eff i
r/N ðrÞ ¼ #½j þ r)/Eff Eff
N ðrÞ þ rfN ðrÞ ð21Þ
where a(v) is the perpendicular transfer area to energy flux (m ). 2 r r

3. Results and discussion

3.1. Model solution: Plane-geometry case

The global isotropic characteristics in the modified radiative


transfer equation make use of the characteristic photon path
length k0 proposed by Özisik (Özisik, 1973) [34,35,45], and defined
as the inverse of the volumetric coefficient of extinction b.
1 1
k0 ¼ ¼ ð22Þ
b jþr
Extending this definition to the effective radiation concept, and
applying it on to plane geometry for fEff Eff
N and /N , (Eqs. (16) and
Fig. 3. (a) Differential control volume, and (b) heterogeneous phase-space. (20)), then:
446 M.A. Mueses et al. / Chemical Engineering Journal 279 (2015) 442–451

Fig. 4. Radiant field simulation for EVREA dimensionless using ERFM. TiO2-P25 Degussa, r = 1297.75 m2/kg; j = 174.75 m2/kg.

Fig. 5. OVREA as a function of catalyst concentration for different optical thickness, sEff.

@ Eff 1h i
scattering (TFM ? SFM ? ERFM). Nevertheless, it is important to
fN ðzÞ ¼ xEff /Eff Eff
N ðzÞ # fN ðzÞ ð23Þ
z k0 note that TFM and SFM were obtained using a reflecting
diffuse-scattering phase function, while the ERFM has exclusively
@ Eff 1 h Eff i
isotropic characteristics. The system solution is given as an analo-
/N ðzÞ ¼ /N ðzÞ # xEff fEff
N ðzÞ ð24Þ
z k0 gous form to the one proposed by Brucato et al. [35,36] obtained
for TFM and SFM:
r
xEff ¼ ¼ rk0 ð25Þ fEff h i
jþr fEff
N;0 # z # z
e kEff 1 # Ce kEff
N ¼ ð27Þ
ð1 # CÞ
where xEff is the global scattering albedo for effective radiation. Eqs.
(23) and (24) are a coupled system with two unknowns, restricted
fEff h qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
# z
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
# z
i
by the boundary conditions at the plate ends: /Eff
N;0
1 # x2Eff Þe kEff # Cð1 þ 1 # x2Eff Þe kEff
N ¼ Eff
ð1 #
k ð1 # CÞ
B:C:1: fEff Eff
N ¼ fN;0 ; z¼0 ð26aÞ
ð28Þ
With
B:C:2: /Eff
N ¼ 0 ; z¼d ð26bÞ
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
At z = 0, the value for the net global energy corresponds to the 1 # x2Eff
1#
#2 d
value in the reactor wall exposed to the incident light. Scattering C¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi e kEff ð29Þ
1 þ 1 # xEff 2
radiation fluxes do not exist for z = d.
The equation system obtained for plane geometry has a mathe-
matical structure analogous to the Two-Flux Model (TFM) [35] and Eff x
kEff ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð30Þ
the Six-Flux Model (SFM) [36] balance equations, hence, it is postu- r 1 # x2Eff
lated that the ERFM is a multidimensional extension of radiation
M.A. Mueses et al. / Chemical Engineering Journal 279 (2015) 442–451 447

Fig. 6. Scattering function in a flat plate. TiO2- P-25, r = 1297.75 m2/kg; j = 174.75 m2/kg. (a) Low concentrations [TiO2] <0.1 g/L. (b) high concentrations [TiO2] >0.3 g/L).

Fig. 7. OVREA profiles for flat geometry. ERFM (––) and SFM (—).

Finally, the equation for the evaluation of EVREA using the con- (r = 1297.75 m2/kg; j = 174.75 m2/kg) were considered to simu-
cept of effective radiation is given by: late the ERFM for the evaluation of heterogeneous radiant fields
in photocatalytic reactors. It was validated for an hypothetic flat
fEff
N;0
h# qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi$ #z
# qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi$ z
i
plate photoreactor of infinite area and of thickness d, with constant
nEff
N ¼ xEff # 1 þ 1 # x2Eff ekEff þ C xEff # 1 # 1 # x2Eff ekEff
kEff xEff ð1 # CÞ incident light flux fEff
N;0 . The rate of radiant energy absorption was
ð31Þ
expressed independent of the incident light flux at the reactor
transparent window (fEff Eff
N /fN;0 ).
3.2. Simulation of radiant field Using ERFM of heterogeneous Fig. 4 shows the EVREA at different catalyst concentrations for
photocatalytic reactors the flat geometry. The main findings are:
At high catalyst concentrations (TiO2 > 1.0 g/L), the energy is
3.2.1. Radiation field simulation maximum near the reactor wall exposed to radiation (v/d < 0.1),
The optical properties of TiO2 P25 (Degussa-Evonik) photocata- but the energy abruptly decrease as the optical thickness increases
lyst reported by Satuf et al. [46] and Colina-Márquez et al. [25,26] [4,7].
448 M.A. Mueses et al. / Chemical Engineering Journal 279 (2015) 442–451

At the limit layer region (v/d < 0.1), the interaction of the pho- At low catalyst concentration, the scattering is more uniform
tons with the catalyst particles is high while the shielding effects (x/d > 0.3) across the reactor.
are insignificant. If, the scattering effects are highest then the cat- Fig. 5 shows the OVREA as a function of apparent optical thick-
alysts could have a better activation probability. ness, sEff is calculated from Eq. (32) [35]. The definition of Overall
If the optical thickness increases (at the same loading of the cat- Volumetric Rate of Energy Absorption (OVREA) was used by Eq.
alyst), the shielding effects overlap the scattering and the absorp- (33) [20,44].
tion of energy radically reduces [8,9,25,26].
sEff + d=kEff ð32Þ

Z !
1 nEff
N
OVREA + dd ð33Þ
d d fEff
N;0

If the optical thickness tends to zero (sEff > 0.001), the OVREA
reaches maximum values of energy availability, and it linearly
increases with loading catalyst. These results are according with
other works [9,25,26].
At low optical thicknesses (0.01–0.05), the OVREA tends to be
asymptotic with the concentration increase (14900 at 0.01; 6400
at 0.025 and 3500 at 0.05), and its values are not above the 10%
of the ones reached to low optic thicknesses and the same loading
catalyst.

3.3. Scattering effects simulation

Fig. 6 shows the scattering function /Eff


N as function of v/d at dif-
ferent loading catalyst. Radiation scattering increases linearly from
the back wall at low catalyst concentrations, below 0.1 g/L (Fig. 6a),
while it follows an exponential decrease function at higher catalyst
concentrations, above 0.3 g/L (Fig. 6b). This behavior is due to that
at higher catalyst loading the light shielding effect becomes signif-
icant. The performance found for this model is physically consis-
tent with literature reports [10–12,13,23,36].
The v/d coordinate can be used to distinguish two energetic
regions. A high energy region (near the reactor front wall) and an
area of low energy near the back wall. The amplitude of these areas
can be varied by changing the loading catalyst. The optimum for
this study is about 0.3 g/L.

3.4. Comparison of ERFM with SFM

Fig. 7 shows the dimensionless OVREA calculated with the


ERFM and SFM, at different catalyst loadings and at the same inci-
dent radiation flux. These data were obtained with using the same
coefficients: j = 174.75 m2/kg and b = 1472.50 m2/kg in the both
models.
The two models show numerical differences, the main differ-
ence appears in the correction of the scattering albedo x parame-
ter. This behavior can be explained by use of the different phase
function:
For the ERFM with isotropic phase function, the value for scat-
tering albedo is 0.8813, this parameter is the ratio between r and
j without any correction, so (x = r/j + r) (Eq. (25)). This value was
obtained with integrated values of absorption (j = 174.75 m2/kg)
and specific extinction (b = 1472.50 m2/kg) coefficients reported
by Colina-Marquez et al. and Satuf et al. [25,26,46].
However, for the SFM with diffuse-scattering phase function
[36], the value for scattering albedo must be corrected by the phase
function using the same values of j and b; however the scattering
albedo coefficient is calculated as relation between of probability
parameters a and b (x = b/a, where a and b are computed using
diffuse-scattering phase function) from SFM [26]. So a scattering
albedo corrected value of 0.7586 is obtained using the correction
by matching the SFM model results to the predictions from a
Monte Carlo model.
Fig. 8. Simulation of LVRPA in CPC solar reactor using ERFM to different catalyst Moreover, for low catalyst concentrations ([TiO2] <0.5 g/L) the
concentration: 0.1 g/L (up); 0.3 g/L (middle) and 0.5 g/L (down). OVREA do not present differences between the models (relative
M.A. Mueses et al. / Chemical Engineering Journal 279 (2015) 442–451 449

Fig. 9. Differential tubular photoreactor used by Zalazar et al. (2005). Adapted from [12].

Fig. 10. OVREA simulation in a differential tubular photocatalytic reactor. DOM (––) [12], ERFM (—) and SFM (" " ") Conditions: [TiO2-Aldrich] = 1.0 g/L; LR = dR = 0.052 m.
j = 411.32 m2/kg; b = 3897.4946 m2/kg.

errors less to 2% referred to SFM), then at this conditions, the dif- With
ference in the scattering phase function with the two models has
xP ¼ r cosðhÞ ð35Þ
slight influence on the OVREA profiles.
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
3.5. Extending the ERFM model to the cylindrical geometry d ¼ 2 R2 # x2P ð36Þ

Other significant advantage of this new modeling approach, is it r0p ¼ d=2 # r sinðhÞ ð37Þ
application on fluctuating radiation systems, such as solar photo-
catalytic reactors at different size. Among many reactor configura- If the CPC reactor is considered, the radiant energy re-directed
tions, the compound parabolic collector (CPC) is one of the most by the reflectors must be modeled by coupling the CPC geometry
efficient for application of heterogeneous photocatalysis at solar with the ray tracing technique [25,26]. So the distance d in the
scale [25,47–53]. reactor tube crossed by each reflected photon light beam is
The application of the RTE to this geometry requires a mathe- [25,26]:
matical change to cylindrical coordinates. Nevertheless, due to qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
the mathematical complexity to obtain an analytical solution for d¼ ðxiþ1 # xi Þ2 þ ðyiþ1 # yi Þ2 ð38Þ
these equations, the solution of a plane-geometry was extend to
The simulation using ERFM to calculate the LVRPA in a hetero-
the cylindrical geometry, using the methodology proposed by Li
geneous photocatalytic solar reactor is presented in the Fig. 8. The
Puma et al. [25,26,30–32,54]. In this approximation, the linear
performance of the new model is similar to the obtained for
solution of the RTE is extended to equations of bi-dimensional
Colina-Márquez et al. 2009, 2010 and Mueses et al. 2013 using
coordinate systems (r, h) [25,26]. The modified model is:
SFM with the same operational conditions [25,26].
&#
fEff
N;0
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi$ #rP ðr;hÞ
nEff
N ðr; hÞ ¼ xEff # 1 þ 1 # x2Eff e kEff
Eff
k xEff ð1 # CÞ 3.6. Validation the ERFM, SFM and RTE in a differential tubular reactor
# qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi$ rP ðr;hÞ '
þC xEff # 1 # 1 # x2Eff e kEff ð34Þ Finally, both the ERFM and SFM models were evaluated using
the cylindrical geometry approximation in a differential tubular
reactor (Fig. 9). The performance was compared with simulation
450 M.A. Mueses et al. / Chemical Engineering Journal 279 (2015) 442–451

data of the LVRPA provided by Zalazar and Cassano [12], this data [5] R. Vinu, G. Madras, Renewable energy via photocatalysis, Curr. Org. Chem. 17
(2013) 2538–2558.
was obtained by the rigorous solution of the RTE using the Discrete
[6] J. Wang, X. Wei, J. Shen, X. Lu, J. Xie, M. Chen, Photocatalytic selective
Ordinate Method – DOM and considering an isotropic scattering transformation of organics, Prog. Chem. 17 (2013) 2516–2537.
phase function. [10,12,15]. Experimental data are reported by [7] G. Palmisano, V. Augugliaro, M. Pagliaro, L. Palmisano, Photocatalysis: a
Zalazar and Cassano [12]. promising route for 21st century organic chemistry, Chem. Commun. 33
(2007) 3425–3437.
The obtained results (Fig. 10) show that SFM as well as ERFM [8] P. Valadés-Pelayo, F. Guayaquil, B. Serrano, H. de Lasa, Photocatalytic reactor
are far from the RTE solution in regions close to the lighted wall under different external irradiance conditions: validation of a fully predictive
of the photoreactor (0.1 < r/R > 0.9), but with errors less to 2.5%. radiation absorption model, Chem. Eng. Sci. 126 (2015) 42–54.
[9] P. Valades-Pelayo, J. Moreira, B. Serrano, H. De Lasa, Boundary Conditions and
However, OVREA values from simulations, when ERFM is used, phase functions in a Photo-CRECWater-II reactor radiation field, Chem. Eng.
they are low with respect to simulations with DOM. This is attrib- Sci. 107 (2014) 123–136.
uted to the phase function associated to both methods. According [10] A. Cassano, O. Alfano, Reaction engineering of suspended solid heterogeneous
photocatalytic reactors, Catal. Today 58 (2000) 167–197.
to Cassano et al. [10,12,46], the phase function that better repre- [11] O. Alfano, A. Cassano, Photoreactor modeling: applications to advanced
sents the radiant field distribution in this type of heterogeneous oxidation processes, Int. J. Chem. React. Eng. 6 (2008) 1–20.
systems is the isotropic phase function, which was validated by [12] C. Zalazar, R. Romero, C. Martin, A. Cassano, Photocatalytic intrinsic reaction
kinetics I: mineralization of dichloroacetic acid, Chem. Eng. Sci. 60 (2005)
comparing simulations obtained by solving the theoretical equa- 5240–5254.
tion of Mie function [10,46]. [13] G. Imoberdorf, A. Cassano, H. Irazoqui, H.O. Alfano, Optimal design and
Alfano et al. [11] found that the isotropic phase function pre- modeling of annular photocatalytic wall reactor, Catal. Today 129 (2007) 118–
126.
dicts, with discrepancies lower to 1%, in LVRPA simulations with
[14] R. Brandi, J. Citroni, O. Alfano, A. Cassano, Absolute quantum yields in
respect to theoretical solutions and quantum yields of the system. photocatalytic slurry reactor, Chem. Eng. Sci. 58 (2003) 979–985.
The latter results demonstrate that the proposed postulates are [15] J. Duderstadt, R. Martin, Transport Theory, Wiley, New York, 1979. pp. 21–422.
valid to formulate the effective radiation field model, ERFM. [16] L. Dombrovsky, The use of transport approximation and diffusion-based
models in radiative transfer calculations, Comput. Therm. Sci. 4 (4) (2012)
297–315.
[17] C. Bohren, D. Huffman, Absorption and Scattering of Light by Small Particles,
4. Conclusions Wiley, New York, 1983.
[18] J. Marugán, R. Van Grieken, A. Cassano, O. Alfano, Intrinsic kinetic modeling
with explicit radiation absorption effects of the photocatalytic oxidation of
A new model to estimate the radiation field in heterogeneous
cyanide with TiO2 and silica-supported TiO2 suspensions, Appl. Catal. B 85
photocatalytic reactors is presented. The model is focused on the (2008) 48–60.
concept of effective radiation, which considers that radiation dis- [19] J. Marugán, R. van Grieken, O. Alfano, A. Cassano, Comparison of empirical and
tribution into a heterogeneous phase-space can be described as a kinetic modeling of the photocatalytic oxidation of cyanide, Int. J. Chem. React.
Eng. 5 (A89) (2007) 1–9.
net global energy flux of incident photons, independent of propa- [20] F. Machuca-Martínez, J. Colina-Márquez, M. Mueses, Determination of
gation direction. This postulate permits the model to be considered quantum yield in a heterogeneous photocatalytic system using a fitting-
as an isotropic system with constant global optical properties, and parameters model, J. Adv. Oxid. Technol. 11 (2008) 42–48.
[21] C. Passalia, O. Alfano, R. Brandi, A methodology for modeling photocatalytic
with an isotropic phase function. From these postulates, modified reactors for indoor pollutions control using previously estimated kinetic
models of the radiative transfer equation and an equivalent for parameters, J. Hazard. Mater. 211–212 (2011) 357–365.
the radiation scattering function were formulated. [22] R. Brandi, G. Rintoul, O. Alfano, A. Cassano, Photocatalytic reactors reaction
kinetics in a flate plate solar simulation, Catal. Today 76 (2002) 161–175.
The balance equations of radiant energy obtained for a planar [23] L. Dombrovsky, D. Baillis, Thermal Radiation in Disperse Systems: An
system of infinite area and thickness d were topologically identical Engineering Approach, Begell House, New York, 2010.
to SFM and TFM equations; therefore, its solution complies with [24] C. Zalazar, M. Labas, C. Martin, R. Brandi, O. Alfano, A. Cassano, The extended
use of actinometry in the interpretation of photochemical reaction
the same mathematical structure. engineering data, Chem. Eng. J. 109 (2005) 67–81.
[25] J. Colina-Márquez, F. Machuca-Martínez, G. Li Puma, Radiation absorption and
optimization of solar photocatalytic reactors for environmental applications,
Acknowledgments Environ. Sci. Technol. 44 (2010) 5112–5120.
[26] J. Colina-Márquez, F. Machuca-Martínez, G. Li Puma, Photocatalytic
The authors thank Universidad de Cartagena-Colombia, mineralization of commercial herbicides in a pilot-scale solar CPC reactor:
photoreactor modeling and reaction kinetics constants independent of
Universidad del Valle-Colombia, Universidad Autónoma de Nuevo radiation field, Environ. Sci. Technol. 43 (2009) 8953–8960.
León – México and Lourgbourgh University – UK for support this [27] G. Li Puma, Modeling of thin-film slurry photocatalytic reactors affected by
paper. Mueses thanks Colciencias for financing his doctoral studies. radiation scattering, Environ. Sci. Technol. 37 (2003) 5783–5791.
[28] G. Li Puma, P. Lock-Yue, Modelling and design of thin-film slurry
Machuca-Martínez and Hernandez-Ramirez thank to photocatalytic reactors for water purification, Chem. Eng. Sci. 58 (2003)
CONACYT-México for support at the Programa de Estancias 2269–2281.
Sabáticas Nacionales, Estancias Sabáticas al Extranjero y [29] A. Gora, B. Toepfer, V. Puddu, G. Li Puma, Photocatalytic oxidation of herbicides
in single-component and multicomponent systems: Reaction kinetics analysis,
Estancias cortas para la consolidación de Grupos de Investigación, Appl. Catal. B 65 (2006) 1–10.
Grant 233199. Thanks to Professor Cassano and Dr. Cristina [30] G. Li Puma, J. Khor, A. Brucato, Modeling of an annular photocatalytic reactor
Zalazar by provide the simulation data of DOM-RTE. for water purification: oxidation of pesticides, Environ. Sci. Technol. 38 (2004)
3737–3745.
[31] G. Li Puma, A. Brucato, Dimensionless analysis of slurry photocatalytic reactors
References using a two-flux and six-flux radiation absorption-scattering models, Catal.
Today 122 (2007) 78–90.
[32] G. Li Puma, B. Toepfer, A. Gora, Photocatalytic oxidation of multicomponent
[1] M. Pelaez, N. Nolan, S. Pillai, M. Seery, P. Falaras, A. Kontos, P. Dunlop, J. systems of herbicides: scale-up of laboratory kinetics rate data to plant scale,
Hamilton, J. Byrne, K. O’Shea, M. Entezari, D. Dionysiou, A review on the visible Catal. Today 124 (2007) 24–132.
light active titanium dioxide photocatalysts for environmental applications, [33] B. Toepfer, G. Li Puma, A. Gora, Photocatalytic oxidation of multicomponent
Appl. Catal. B 125 (2012) 331–349. systems of herbicides: reaction kinetics analysis with explicit photon
[2] A. Hernández-Ramírez, I. Medina-Ramírez, Photocatalytic Semiconductors, absorption effects, Appl. Catal. B 68 (2006) 171–180.
Synthesis, Characterization, and Environmental Applications, Springer, [34] A. Brucato, L. Rizzuti, Simplified model of radiation fields in heterogeneous
Switzerland, (2015). doi: 10.1007/978-3-319-10999-2. photoreactors. 1. Case of zero reflectance, Ind. Eng. Chem. Res. 36 (1997)
[3] Y. Lan, Y. Lu, Z. Ren, Mini review on photocatalysis of titanium dioxide 4740–4747.
nanoparticles and their solar applications, Nano Energy 2 (2013) 1031–1045. [35] A. Brucato, L. Rizzuti, Simplified model of radiation fields in heterogeneous
[4] C. McCullagh, N. Skillen, M. Adams, P.K. Robertson, Photocatalytic reactors for photoreactors. 2. Limiting ‘‘Two-Flux’’ Model for the case of reflectance greater
environmental remediation: a review, J. Chem. Technol. Biotechnol. 86 (2011) than zero, Ind. Eng. Chem. Res. 36 (1997) 4748–4755.
1002–1017.
M.A. Mueses et al. / Chemical Engineering Journal 279 (2015) 442–451 451

[36] A. Brucato, A. Cassano, F. Grisafi, G. Montante, L. Rizzuti, G. Vella, Estimating [46] M. Satuf, R. Brandi, A. Cassano, O. Alfano, Experimental method to evaluate the
radiant fields in flat heterogeneous photoreactors by the six-flux model, AIChE optical properties of aqueous titanium dioxide suspensions, Ind. Eng. Chem.
J. 52 (2006) 3882–3890. Res. 44 (2005) 6643–6649.
[37] J. Farmer, J. Howell, Comparison of Monte Carlo strategies for radiative transfer [47] S. Malato, P. Fernández-Ibañez, M. Maldonado, J. Blanco, W. Gernjak,
in participating media, Adv. Heat Transfer 55 (1998) 333–429. Decontamination and disinfection of water by solar photocatalysis: recent
[38] T. Deutschmann, S. Beirle, U. Frieß, M. Grzegorski, C. Kern, L. Kritten, U. Platt, C. overview and trends, Catal. Today 147 (2009) 1–59.
Prados-Román, J. Puk ! ı!te, T. Wagner, B. Werner, K. Pfeilsticker, The Monte Carlo [48] S. Malato, J. Blanco, R. Maldonado, P. Fernández, D. Alarcón, M. Collares-
atmospheric transfer model McAtim: Introduction and validation of Jacobians Pereira, J. Farinha, J. Correia de Oliveira, Engineering of solar photocatalytic
and 3D features, J. Quant. Spectrosc. Radiat. Transfer 112 (2011) 1119–1137. collectors, Sol. Energy 77 (2004) 513–524.
[39] J. Petrasch, S. Haussener, W. Lipiński, Discrete vs continuum-scale simulation [49] S. Malato, J. Blanco, A. Campos, J. Cáceres, C. Guillard, J. Herrmann, A.
of radiative transfer in semitransparent two-phase media, J. Quant. Spectrosc. Fernández-Alba, Effect of operating parameters on the testing of a new
Radiat. Transfer 112 (9) (2011) 1450–1459. industrial titania catalyst at solar pilot plant scale, Appl. Catal. B 42 (2003)
[40] C. Passalia, O. Alfano, R. Brandi, Modelling and experimental verification of a 349–357.
corrugated plate photocatalytic reactor using computational fluid dynamics, [50] S. Malato, J. Blanco, A. Vidal, C. Richter, Photocatalysis with solar energy at a
Ind. Eng. Chem. Res. 50 (2011) 9077–9086. pilot-plant scale: an overview, Appl. Catal. B 37 (2002) 1–15.
[41] F. Trujillo, T. Safinski, A. Adesina, Oxidative photomineralization of [51] J. Colina-Márquez, D. Díaz, A. Rendón, A. López-Vásquez, F. Machuca-Martínez,
dichloroacetic acid in an externally-irradiated rectangular bubble tank Photocatalytic treatment of a dye polluted industrial effluent with a solar
reactor: computational fluid dynamics modeling and experimental pilot-scale CPC reactor, J. Adv. Oxid. Technol. 12 (2009) 93–99.
verification studies, Ind. Eng. Chem. Res. 49 (2010) 6722–6734. [52] E. Bandala, C. Estrada, Comparison of solar collection geometries for
[42] F. Jović, V. Kosar, V. Tomašić, Z. Gomzi, Non-ideal flow in an annular application to photocatalytic degradation of organic contaminants, J. Solar
photocatalytic reactor, Chem. Eng. Res. Des. 90 (2012) 1297–1306. Energy Eng. 129 (2007) 22–26.
[43] Z. Wang, J. Liu, Y. Dai, W. Dong, S. Zhang, J. Chen, CFD modeling of a UV-LED [53] E. Bandala, C. Arancibia-Bulnes, S. Orozco, C. Estrada, Solar photoreactors
photocatalytic odor abatement process in a continuous reactor, J. Hazard. comparison based on oxalic acid photocatalytic degradation, Sol. Energy 77
Mater. 215–216 (2012) 25–31. (2004) 503–512.
[44] M. Mueses, F. Machuca-Martinez, G. Li Puma, Effective quantum yield and [54] G. Li Puma, V. Puddu, H. Tsang, A. Gora, B. Toepfer, Photocatalytic oxidation of
reaction rate model for evaluation of photocatalytic degradation of water multicomponent mixtures of estrogens (estrone (E1), 17b-estradiol (E2), 17a-
contaminants in heterogenouos pilot-scale solar photoreactors, Chem. Eng. J. ethynylestradiol (EE2) and estriol (E3)) under UVA and UVC radiation: Photon
215–216 (2013) 937–947. absorption, quantum yields and rate constants independent of photon
[45] M. Özisik, Radiative Transfer and Interactions with Conduction and absorption, Appl. Catal. B 99 (2010) 388–397.
Convection, Wiley, New York, 1973.

You might also like