Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

HACA surface growth implementation

The HACA surface growth model utilized in this study is based on the HACA soot surface reaction
scheme developed by Appel et al. [1]. Table 1 displays the six surface reactions that are considered.
Except for the reaction of soot with OH radicals (S6) which is modeled based on free molecular regime
collision theory between OH and soot with a reaction probability of 0.13 [2] , the kinetics of the other
surface reactions are described using the concept of soot surface sites, which can either be saturated
(Csoot-H) or dehydrogenated (Csoot∙). In order to calculate reaction rates, concentrations of these two sites
are required for each soot section. The concentration of saturated sites for soot section i, [Csoot-H]i, is
calculated by:

χc As ,i N i
[Csoot − H ]i = soot − H
(1)
Av

where χ csoot − H is the number of sites per unit soot surface area and is set to a constant value of

2.3 ⋅1015 sites , As,i is the surface area of soot particles in section i, and Ni is the number of soot
cm 2
particles in section i. The concentration of dehydrogenated sites for soot section i, [Csoot∙]i, is calculated
by:

k1 H + k 2OH
[Csoot ⋅ ]i = [Csoot − H ]i ⋅ (2)
k−1 H 2 + k −2 H 2O + k4C2 H 2 + k5O2

where H, OH, H2, H2O, C2H2, and O2 are mole fractions. The fractional term represents the portion of
saturated sites that have been dehydrogenated and arises from a steady state approximation for [Csoot∙]i.

Table 1. HACA-based soot surface growth and oxidation reactions Appel et al. [1].

No. Reaction A [cm3 mol-1 s-1] b Ea [kcal/mol]


13
S1 Csoot-H+H  Csoot∙ +H2 4.2∙10 0 13
S2 Csoot-H+OH Csoot∙ +H2O 1∙1010 0.73 1.43
S3 Csoot∙ +H  Csoot-H 2∙1013 0 0
S4 Csoot∙ +C2H2 Csoot-H+H 8∙107 1.56 3.8
S5 Csoot∙ +O2 2CO + product 2.2∙1012 0 7.5
S6 Csoot-H+OH CO + product γOH=0.13

PAH condensation
The PAH condensation model utilized in this study is based on transition and continuum regime collision
theory between soot aggregates and pyrene (A4), with a prescribed collision efficiency, γ = 0.55 [3]. The
collision rate between pyrene and soot particles in the ith section, β , is given by [4]:
β = 4π Rabs ( Ds + DPAH ) f D (3)

where Rabs is the absorbing sphere cluster radius, Ds and DPAH are the diffusion coefficients for soot
particles in section i and pyrene respectively, and fD is the transition regime correction factor. The
diffusion coefficient D is given by:

k B ⋅ T ⋅ Cc ( Kn )
D= (4)
3π ⋅ µ ⋅ d m

where kB is the Boltzmann constant, T is the gas temperature, µ is the gas viscosity, dm is the mobility
diameter, and Cc(Kn) is the Cunningham slip correction factor as a function of the Knudsen number Kn
and is calculated as [5]

Cc ( Kn ) = 1 + 1.612 Kn (5)

The Knudsen number is defined as

2λmfp
Kn = (6)
dm

where λmfp is the mean free path of the gas, which is taken from the mean free path of air. The transition
regime correction factor fD is obtained from a simple approximation of flux-matching theory and is
calculated as

1 + KnD
fD = (7)
1 + 2 KnD (1 + KnD )

where KnD is the diffusion Knudsen number which characterizes the transition between the continuum
and free molecular diffusion as is defined as

λmfp ,12
KnD = (8)
Rabs

where λmfp,12 is the diffusion mean free path which is calculated as

Ds + DPAH
λmfp ,12 = (9)
2 k BT  1 1 
π ⋅  ms + mPAH 

where ms is the mass of the soot aggregates in section i, and mPAH is the mass of a pyrene molecule. The
mobility diameter of pyrene is taken to be the Lennard-Jones diameter. The mobility diameter of the
soot aggregate is calculated by
2rp n 0.43
p free molecular regime

dm =   Df −1 
0.7
(10)
2 R f   continuum regime
  2 

where rp is the primary particle radius, np is the number of primary particles in the aggregate, Df is the
fractal dimension, and Rf is the outer radius of the aggregate define by

R f = rp ( fn p )
1
Df
(11)

where f is the volume filling factor and is set to f = 1.43 [6]. Finally, the expression for the absorbing
cluster radius is the same as that suggested by Zurita-Gotor and Rosner [7].

The sectional aerosol dynamics model


The sectional aerosol dynamics model is based on the works of Park et al. [8] and Wen et al. [9] and is
described in detail by Zhang [3]. The model uses the classical sectional description of the particle
population balance equation based on the fixed pivot approach [10]. The particle size distribution (PSD)
is formulated on the basis of particle mass. The mass range of the fractal-like solid soot aggregates is
divided into 35 discrete sections. Each section has a representative mass that is prescribed and kept
unchanged. A geometric series is assumed for the representative mass of all sections due to
computational constraints. A sectional spacing factor fs, is used to denote the mass ratio of two
consecutive sections. In this work fs = 2.35. All the soot aggregates in a section are assumed to be
identical. Soot aggregates fall into individual sections according to their mass. A transport equation for
the number density of soot aggregates is constructed and solved in each section. The nucleation step
connects the gas-phase incipient species with the solid soot phase. The incipient soot particles are
assumed to be spherical and belong to the first section. By coagulation or surface growth, lower section
particles move to higher sections. On the other hand, higher section particles move to lower sections by
oxidation or fragmentation.

Consistent with experimental observations, each aggregate is assumed to comprise equally-sized


spherical primary particles, connected together by point touching. In the limit of containing only one
primary particle, an aggregate is the same as the primary particle. All aggregates are assumed to have
the same fractal dimension Df = 1.8 if they are larger than the primary spherule mass; whereas smaller
particles are assumed to be dense spheres (Df = 3). The fractal dimension, the primary particle diameter,
and the number of primary particles completely determine the structure of an aggregate. To
characterize soot aggregate structure, an additional quantity in each section being the number density
of primary particles is obtained by constructing and solving the transport equation for the number
density of primary particles in each section. By knowing the number densities of aggregates and primary
particles in a section, the average number of primary particles per aggregate in that section is known.
Since the mass of a single aggregate in the section under consideration is prescribed and known, the
volume of the aggregate, the diameter of primary particles forming the aggregate, and the total surface
area of the aggregates in that section are also known. Similarly, such information is known for all the
other sections. It is noted that all soot particles in the first section are spherical nascent soot particles or
monomers. Thus, aggregates and primary particles in the first section are essentially the same both
meaning spherical nascent monomers.

Taking into account flow convection, normal diffusion, thermophoresis, nucleation (nu), coagulation
(co), surface growth (sg), oxidation (ox), surface condensation (sc), and fragmentation (fr), the two-
dimensional transport equations for soot aggregate and primary soot particle populations in each
section are formulated and shown as follows:

∂N i ∂N 1 ∂  ∂N i  ∂  ∂N i  1 ∂
ρv
∂r
+ ρu i =
∂z 
r ∂r 
r ρ Di
∂r  + ∂z  ρ Di ∂z  − r ∂r ( r ρ N iVTs ,r )
  
∂  ∂N ∂N i ∂N i ∂N ∂N ∂N 

∂z
( ρ N iVTs , z ) + ρ  i
 ∂t
+
∂t
+
∂t
+ i + i + i 
∂t ox ∂t sc ∂t fr 
(12)
 nu co sg

( i = 1, 2,… , SN )
where Ni represents the number of the ith sectional particles per unit mass of the gaseous mixture (Nia
for aggregates and Nip for primary particles); SN is the total number of soot sections and equals to
SN = 35 in this work; Di is the diffusion coefficient of the ith section and is calculated in a similar way to
Eq. (4).

The thermophoretic velocities of soot aggregates VTs,r and VTs,z are calculated according to Talbot et al.
[11] as follows:

µ ∂T
VTs , xi = −0.55 ( xi = r , z ) (13)
ρT ∂xi

The nucleation term is based on the work or Apple et al. [1]and is calculated as follows:

∂N1a ∂N1p 2.2 4π k BT


= = ⋅ d PAH
2
⋅ Av2 ⋅ [ PAH ]2
∂t nu
∂t nu
ρ Cmass N C , PAH
(14)
∂N ia ∂N p
= i = 0, i = 2,..., SN
∂t nu
∂t nu

where Cmass is the mass of a carbon atom; NC,PAH is the number of carbon atoms in the incipient PAH
2 NC , PAH
species; dPAH is the diameter of the incipient PAH species and equals d PAH = d A , where dA
3
is the size of a single aromatic ring; Av is the Avogadro number and [PAH] denotes the mole
concentration of the incipient PAH species. The coagulation terms for aggregates and primary particles
in section i are calculated as:
∂N ia k ≤ j ≤i
 δ j ,k  SN

∂t
= ∑ 1 −
mi −1 ≤ m j + mk ≤ mi +1  2
ηβ ξ a
 j ,k j ,k j k
N N a
− N i ∑ βi , k ξi , k N k
a a

co  k =1
(15)
∂N i p k ≤ j ≤i
 δ j ,k  SN

∂t
= ∑ 1 −
mi −1 ≤ m j + mk ≤ mi +1  2
ηη β ξ a
 p j ,k j , k j k
N N a
− N i ∑ β i ,k ξ i ,k N k
p a

co  k =1

where mi is the representative mass of the ith section aggregate; δ is the Kronecker delta function; βj,k is
the collision kernel of two aggregates of the jth and the kth sections and is calculated similarly to Eq. (3);
ξj,k is the coagulation efficiency of two aggregates of the jth and the kth sections. The parameter η weights
the newly formed mass into two adjacent sections such that the number and the mass of aggregates are
conserved. The parameter η is calculated as:

 mi +1 − ( m j + mk )
 , mi ≤ m j + mk ≤ mi +1
 mi +1 − mi
η= (16)
 mi −1 − ( m j + mk )
 , mi −1 ≤ m j + mk ≤ mi
 mi −1 − mi

The factor ηp assigns primary particles to two adjacent sections so the primary particle size is conserved
and it is calculated as:

ηp =
mi
m j + mk
( n p, j + n p,k ) (17)

where np,j is the number of primary particles per aggregate of the jth section.

The surface growth and condensation terms are calculated as:

 I sg / si ,i
− , i =1
 m i + 1 − m i

∂N ia  I sg / si ,i −1 I sg / si ,i
= − , 2 ≤ i ≤ SN − 1
∂t sg / sc  mi − m i −1 m i +1 − m i
 I sg / si ,i −1
 , i = SN
 mi − mi −1
(18)
 I sg / si ,i
− n p ,i , i =1
 mi +1 − mi
∂N ip  I sg / si ,i −1 I sg / si ,i
= n p ,i −1 − n p ,i , 2 ≤ i ≤ SN − 1
∂t sg / sc  mi − mi −1 mi +1 − mi
 I sg / si ,i −1
 n p ,i −1 , i = SN
 mi − mi −1
where Isg/sc,i is the surface growth or surface condensation rate of soot aggregates of the ith section in the
unit of g/cc/sec, and is always non-negative.

The source terms due to surface oxidation causing a decrease in soot mass are calculated as follows:

 I ox ,i +1 I
− + ox ,i , i =1
 mi +1 − mi mi
∂N ia  I I ox ,i
= − ox ,i +1 + , 2 ≤ i ≤ SN − 1
∂t ox  m i +1 − m i m i − m i −1
 I ox ,i
 , i = SN
 mi − mi −1
(19)
 I ox ,i +1 I
− n p ,i +1 + ox,i n p ,i , i =1
 mi +1 − mi mi
∂N ip  I I ox ,i
= − ox ,i +1 n p ,i +1 + n p ,i , 2 ≤ i ≤ SN − 1
∂t ox  mi +1 − mi mi − mi −1
 I ox ,i
 n p ,i , i = SN
 mi − mi −1

where Iox,i is the oxidation rate of soot aggregates of the ith section in the unit of g/cc/sec, and is always
non-positive.

For fragmentation, 1:1 fragmentation pattern [3] was implemented. The source terms due to
fragmentation are calculated as follows:

Γi ,i +1Si +1 N ia+1 , i =1


∂N a 
i
= ( Γi ,i − 1) Si N ia + Γ i ,i +1Si +1 N ia+1 , 2 ≤ i ≤ SN − 1
∂t 
( Γi ,i − 1) Si N i ,
fr a
i = SN
 Γ i ,i +1Si +1 N ia+1n p ,i +1
 , i =1 (20)
 fs
 Γi ,i +1Si +1 N ia+1n p ,i +1
∂N ip 
= ( Γ i ,i − 1) Si N ia n p ,i + , 2 ≤ i ≤ SN − 1
∂t fr  f s
 Γ −1 S N an ,
( i ,i ) i i p ,i
i = SN


The breakage distribution functions Γ are calculated as:


fs − 2
Γ i ,i =
fs −1
(21)
f
Γ i ,i +1 = s
fs −1

Reference List
[1] J. Appel, H. Bockhorn, M. Frenklach, Kinetic modeling of soot formation with detailed chemistry and
physics: laminar premixed flames of C2 hydrocarbons, Combust. Flame. 121 (2000) 122-136.

[2] K. Neoh, J. Howard, A. Sarofim, Soot oxidation in flames, Particulate carbon formation during
combustion. (1981) 162-282.

[3] Q. Zhang, Detailed Modeling of Soot Formation/Oxidation in Laminar Coflow Diffusion Flames,
(2010), PhD Thesis, University of Toronto.

[4] S.N. Rogak, R.C. Flagan, Coagulation of aerosol agglomerates in the transition regime, J. Colloid
Interface Sci. 151 (1992) 203-224.

[5] C. Sorensen, G. Wang, Note on the correction for diffusion and drag in the slip regime, Aerosol
Science and Technology 33 (2000) 353-356.

[6] K. Naumann, COSIMA—a computer program simulating the dynamics of fractal aerosols, J. Aerosol
Sci. 34 (2003) 1371-1397.

[7] M. Zurita-Gotor, D. Rosner, Effective diameters for collisions of fractal-like aggregates:


Recommendations for improved aerosol coagulation frequency predictions, J. Colloid Interface Sci. 255
(2002) 10-26.

[8] S. Park, S. Rogak, W. Bushe, J. Wen, M. Thomson, An aerosol model to predict size and structure of
soot particles, Combustion Theory and Modelling. 9 (2005) 499-513.

[9] J. Wen, M. Thomson, M. Lightstone, S. Rogak, Detailed kinetic modeling of carbonaceous


nanoparticle inception and surface growth during the pyrolysis of C6H6 behind shock waves, Energy &
Fuels. 20 (2006) 547-559.

[10] S. Kumar, D. Ramkrishna, On the solution of population balance equations by discretization - I. A


fixed pivot technique, Chemical Engineering Science. 51 (1996) 1311-1332.

[11] L. Talbot, R. Cheng, R. Schefer, D. Willis, Thermophoresis of particles in a heated boundary layer, J.
Fluid Mech. 101 (1980) 737-758.

You might also like