J Cep 2021 108478

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Chemical Engineering & Processing: Process Intensification 166 (2021) 108478

Contents lists available at ScienceDirect

Chemical Engineering and Processing - Process


Intensification
journal homepage: www.elsevier.com/locate/cep

Mass transfer process intensification for SO2 absorption in a


commercial-scale wet flue gas desulfurization scrubber
Jiangyuan Qu a, b, Nana Qi a, b, *, Zhen Li c, Kai Zhang a, b, Pengcheng Wang c, Lifeng Li c
a
Beijing Key Laboratory of Emission Surveillance and Control for Thermal Power Generation, North China Electric Power University, Beijing 102206, China
b
Key Laboratory of Power Station Energy Transfer Conversion and System (North China Electric Power University), Ministry of Education, Beijing 102206, China
c
Shanxi Hepo Power Generation Company Limited, Yangquan 045011, Shanxi, China

A R T I C L E I N F O A B S T R A C T

Keywords: To intensify SO2 absorption process in wet flue gas desulfurization (WFGD) spray scrubber, a comprehensive CFD
Coal-fired power unit model coupled transfer process with chemical reactions is established by Eulerian‒Lagrangian method and
Flue gas verified in a commercial-scale device of 330 MW coal-fired power unit. The desulfurization performance is
Desulfurization
significantly affected by both the property of each droplet and the dispersion of droplet swarm, which are
Multiphase flow
Mass transfer
determined by the spray behaviors. Three dominant factors, including droplet diameter, injection direction and
Spray characteristics spray morphology, are explored to optimize the spray conditions for intensifying SO2 absorption process.
Differing from the ideal plug flow, the optimal diameter of droplet in non-ideal flow is obtained by considering
the synergistic effect of the uniformity of gas–slurry distribution and the dynamics of SO2 mass transfer.
Compared with the single-direction spray with downward or upward injection, the dual-direction spray reaches a
great balance between the competitive effects of the uniformity of gas‒slurry flow and the interfacial area for
mass transfer. The influence of spray morphology on mass transfer process is essentially the difference of
dispersion characteristics of droplets in the middle part of scrubber. A design scheme combining full-cone and
hollow-cone nozzles is suggested for improving the overall SO2 removal efficiency.

WFGD unit.
Spray scrubber is the heart of WFGD system where flue gas is
1. Introduction
essentially purified through chemical absorption into slurry droplets
[10]. Overall absorption performance is controlled by both the SO2
Coal combustion from power plants, industrial boilers and furnaces
transfer process from gas to slurry and the chemical reactions in solution
significantly contributes to undesired acidic gasses, trace heavy metals,
[11]. The existed results [5-7,12-15] identified that the hydration and
and particulate matters [1]. As one of the main air pollutants, sulfur
dissociation of sulfurous species could be treated as instantaneous re­
dioxide (SO2) causes a series of environmental problems, such as acid
actions. Accordingly, the species concentrations in solution are mainly
rain and formation of secondary aerosols [2]. The wet flue gas desul­
determined by thermodynamic equilibriums [7,15,16]. For the mass
furization (WFGD) using limestone is considered as the most effective
transfer process, Brogren and Karlsson [14] and Qin et al. [2] concluded
and reliable treatment of SO2 removal for industrial processes [3].
that the transport of SO2 into slurry droplet was largely controlled in
Among the several types of reactors for limestone scrubbing,
liquid side. In actual WFGD scrubber, the pH in slurry phase significantly
counter-current spray scrubber is most utilized in commercial-scale
decreases with the dissolution of sulfurous species. Frandsen et al. [17]
WFGD installations, which provides a large contacting area between
and Li et al. [18] optimized the conditions for acid–base reactions in
flue gas and slurry droplets [4,5]. The SO2 contained in flue gas is
WFGD system by using the additives to maintain a relatively stable pH
absorbed into slurry droplets, then dissolved sulfites are oxidized into
level in slurry phase based on the buffering effect. Besides, various
sulfates in the bottom tank and finally crystalized into gypsum extracted
promising technologies based on multi-field coupling method, such as
from the system [6,7]. With the increasingly stringent regulations for
magnetic field [19], electrified spray [20], and high-gravity rotating
SO2 emission [8,9], it is necessary to improve the desulfurization per­
technology [21], have been suggested to intensify the mass transfer
formance and flexibility under various load conditions for industrial

* Corresponding author at: Beijing Key Laboratory of Emission Surveillance and Control for Thermal Power Generation, North China Electric Power University,
Beijing 102206, China.
E-mail address: qinana2013@ncepu.edu.cn (N. Qi).

https://doi.org/10.1016/j.cep.2021.108478
Received 8 March 2021; Received in revised form 3 May 2021; Accepted 17 May 2021
Available online 27 May 2021
0255-2701/© 2021 Elsevier B.V. All rights reserved.
J. Qu et al. Chemical Engineering and Processing - Process Intensification 166 (2021) 108478

1
Nomenclature M molecular weight, kg kmol−
md mass of droplet, kg
AD projected area of droplet, m2 p static pressure, Pa
Ad surface area of droplet, m2 T temperature, K
CD drag coefficient u
→ velocity, m s− 1
CSO2 SO2 molar concentration in slurry, kmol m− 3 V cell volume, m3
cSO2 SO2 mass concentration in flue gas, mg Nm− 3 Y mass fraction of species
cd specific heat of droplet, kJ kg− 1 K− 1
D diffusion coefficient, m2 s− 1 Greek letters
dp droplet diameter, m ε turbulence energy dissipation rate, m2 s− 2

E enhancement factor η desulfurization efficiency, %


e internal energy per unit mass, kJ kg− 1 μ dynamic viscosity, kg m− 1 s− 1
→ ρ density, kg m− 3
F force acting on gas− slurry interface, N
σ d− g surface tension of droplet, N m− 1
→g acceleration of gravity, m s− 2 1 − 2
τ shear stress, kg m− s
H Henry’s constant, Pa m3 kmol− 1
h sensible enthalpy, kJ kg− 1 Subscripts

J diffusion flux, kg m− 2 s− 1 d droplet
K equilibrium constant g gas phase
Ktot overall mass transfer coefficient, kmol m− 2 s− 1 Pa− 1 i ith droplet
keff effective conductivity, W m− 1 K− 1 k kth species
kg mass transfer coefficient in gas side, kmol m− 2 s− 1 Pa− 1
l liquid phase
kl mass transfer coefficient in liquid side, m s− 1 0 initial value

process in wet scrubber. However, these developing technologies are the ideal counter-current mode of gas‒liquid (or slurry) system, the
still implemented in the special lab- and pilot-scale wet scrubbers. droplet with smaller diameter (dp) leads to higher volumetric mass
In a commercial-scale WFGD unit, the flue gas in a radial direction transfer coefficient resulting from larger interfacial area and higher
enters the scrubber whilst the bulk slurry is atomized into fine droplets liquid-side mass transfer coefficient [5,32,33]. For the non-ideal flow
commonly by pressure swirl nozzles [22,23]. It should be noted that the pattern, however, dp has two competitive effects on the uniformity of gas
flow pattern of flue gas and slurry droplets always deviates from the flow [30,34]. As expected, the increasing diameter of droplet decrease
ideal counter-current mode due to either the lateral inlet of raw gas or the interfacial area for interphase momentum transfer, while the droplet
the uneven dispersion of slurry, where the flow channeling significantly with large diameter can mitigate the gas channeling largely ascribed to
reduces the overall SO2 removal efficiency [10,16,24]. Therefore, the the high slip velocity between two phases. Focusing on the complicated
flow structure of two phases is one of key issues to intensify mass hydrodynamics of gas‒liquid flow in the scrubber, Wang and Dai [33]
transfer and thus to fulfill the requirement of reactions for industrial proposed that the proper sizes of droplets used in multilayer spray
desulfurization process. To improve the uniformity of gas distribution, scrubber could enlarge the liquid holdup in the premise of uniform gas
Montanes et al. [25] commented the synergistic rings mitigated the flow. However, the chemical absorption of SO2 should be paid more
channeling of flue gas around the wall, which could increase the consideration for deeply understanding the competitive mechanism
removal efficiency by about 4%. Chen et al. [26] found the deflectors between mass transfer coefficient and two-phase flow pattern depending
around the inlet scrubber efficiently avoided the localized high-speed on droplet diameter.
gas flow, which leaded to an increment for desulfurization efficiency The dispersion characteristics of droplet swarm is impacted by the
of 3% or so. For strengthening the interaction between gas and slurry spray conditions of nozzle, including ejection velocity, spray angle,
phases, great efforts have been devoted to modifying the flow pattern of discharge direction, and morphology of spray cone. Guo et al. [33]
two phases by the specific internals occupying the whole cross section of predicted the wet desulfurization process in a spray tower, which
scrubber, such as sieve tray [24], Venturi rod banks [27,28], flow concluded that increasing injection speed or spray angle resulted in
pattern controlling unit [8], and groove separator [29]. In this way, the better mixing degree of two phases and higher desulfurization effi­
slurry phase accumulates on the surface of internals, then SO2 absorp­ ciency. Feng et al. [31] and Sun et al. [35] explored that the spray di­
tion is locally intensified by the interaction between gas and slurry film rection influenced the momentum exchange behavior in gas‒droplet
or bubbles, rather than gas–droplet contacting mode, whereas the WFGD flow and obtained their optimal parameters to enhance the evaporation
scrubber requires high energy consumption and much maintenance cost. performance of droplets from a single nozzle. For the pressurized swirl
Anyway, slurry droplets are essentially the inherent internals of nozzle, the conical spray region is formed due to the action of centrifugal
scrubber to regulate the flow structure of two phases and to provide the force. Generally, the morphology of spray cone can be classified into
interfacial area for mass transfer [27,30]. SO2 absorption into droplets hollow cone and full cone [35,36]. Michalski [37] pointed out the
still makes considerable contributions for the whole desulfurization hollow-cone nozzle provided better dispersion characteristics than the
process. However, the removal efficiency is affected by both the prop­ full-cone pattern, while Sun et al. [23] thought that the full-cone nozzle
erty of each droplet and the dispersion behaviors of droplet swarm, generated uniform coverage in a pilot-scale spray device. Nevertheless,
which are largely determined by the operating conditions of nozzles to the best of our knowledge, only few investigations concerned to the
[23]. Accordingly, the influence of spray characteristics, such as droplet effect of spray direction and conical shape on the SO2 absorption per­
diameter and injection properties, on mass transfer process should be formances for commercial-scale WFGD scrubber with multi-nozzle
concerned to explore the cost-effective approach for improving SO2 arrangement.
absorption performance. A comprehensive computational fluid dynamics (CFD) model in
The property of droplet is one of the critical parameters to manipu­ Eulerian‒Lagrangian framework is established to describe the transport
late the transfer behavior between gas and slurry phases [13,27,31]. For process of gas‒slurry flow and chemical reactions in slurry droplets

2
J. Qu et al. Chemical Engineering and Processing - Process Intensification 166 (2021) 108478

inside WFGD scrubber, which is verified by the measured data from a Table 1
commercial-scale device with the capacity of 330 MW coal-fired power Chemical equilibriums and balance equations in the chemical reaction model.
unit. Based on the region identification in our previous work [38], the Name Equation No.
aim of this paper is to clarify the influence of three dominant factors,
Chemical KH2 O = [H ]∙[OH ] = exp(140.93− 13445.90/
+ −
(1)
including droplet diameter, injection direction, and spray morphology, equilibrium Td− 22.48lnTd)
on the overall desulfurization performance, which is utilized to explore KSO2,aq = [H+]∙[HSO−3 ]/[SO2,aq] = exp(− 1.96 + 637.40/ (2)
the optimal spray conditions for mass transfer process intensification in Td− 0.015Td)
WFGD scrubber. KHSO3 − = [H+]∙[SO2−
3 ]/[HSO3 ] = exp(− 21.27 +

(3)
1333.40/Td)
2. Model description KHSO4 − = [H+]∙[SO2−
4 ]/[HSO4 ] = exp(− 4.49)

(4)
KCO2,aq = [H+]∙[HCO−3 ]/[CO2,aq] = exp (5)
The CFD modeling concerns both the chemical reactions in aqueous (235.48− 12,092.10/Td− 36.78lnTd)
phase and the transport characteristics of SO2 from flue gas to slurry KHCO3 − = [H+]∙[CO2−
3 ]/[HCO3 ]= exp (6)

droplets within industrial WFGD scrubber. The reactions involved in (220.07− 12,431.70/Td− 35.48lnTd)

slurry bulk are assumed to be at equilibrium instantaneously. Accord­ Mass balance d (7)
([HSO3 − ] + [SO3 2− ] + [SO2,aq ]) = NS(IV)
dt
ingly, the kinetic-controlled steps are neglected, including oxidation of d (8)
([HSO4 − ] + [SO4 2− ]) = 0
sulfite species, dissolution of limestone, and precipitation of gypsum [7, dt
d (9)
13,22]. The transport process of SO2 is simulated by considering the dt
([HCO3 − ] + [CO3 2− ] + [CO2,aq ]) = 0
dynamics of SO2 mass transfer across the gas–slurry interface combined Charge balance 2[Ca2+ ] + 2[Mg2+ ] + [H+ ] = [Cl− ] + [HSO3 − ] + (10)
with the multiphase hydrodynamics, heat transfer, and H2O evaporation 2[SO3 2− ] + [HSO4 − ] + 2[SO4 2− ] + [HCO3 − ] + 2[CO3 2− ] +
in spray scrubber. [OH− ]
Auxiliary system KCaCO3 = [Ca2+ ]⋅[CO3 2− ] (11)
2[Ca2+ ] + 2[Mg2+ ] + [H+ ] = [Cl− ] + [HSO4 − ] + (12)
2.1. Chemical reactions in slurry droplet
2[SO4 2− ] + [HCO3 − ] + 2[CO3 2− ] + [OH− ]

To determine the dissolved constituents in aqueous phase, the re­


action model is established to describe the complex chemical in­ non-linear algebraic system can be referred to our previous work [38].
teractions within droplet. As illustrated in Fig. 1, the sulfite and bisulfite
ions are formed from the dissociation reactions after the hydration of 2.2. Governing equations for gas‒slurry flow
gaseous SO2, which increases the maximum dissolved concentrations of
sulfurous species compared with the physical transfer process [11]. The The interphase transfer process and aqueous phase chemistry are
compositions of sulfate and carbonate species vary with different equi­ simulated by Eulerian–Lagrangian method. In the Eulerian framework,
librium pH values in actual WFGD plant. In addition, some inert anions flue gas is treated as a compressible Newtonian fluid, composed of N2,
and cations, such as Mg2+ and Cl− , are also supposed to affect the O2, CO2, SO2, and H2O vapor. The motion of gas phase is described via
thermodynamic conditions in solution [10,24,39]. Therefore, totally 13 Reynolds-Averaged Navier–Stokes (RANS) equations closed by the
dissolved compounds are considered in slurry droplet in the proceeding Realizable k-ε turbulence model, which has been considered suitable for
of SO2 absorption, i.e., H+, OH− , SO2,aq, HSO−3 , SO2− 2−
3 , HSO4 , SO4 , CO2,

modeling the turbulence of gas flow in spray scrubber [4,24,31,34]. The
2− 2+ 2+
aq , HCO −
3 , CO 3 , Ca , Mg , and Cl −
. conservations of mass, momentum, energy, and species concentrations
The concentrations of species are taken from the acid–base equilib­ for the continuous phase are given as:
riums, mass balances, and electroneutrality for ions [7,11,15,20,22]. ( )
However, the initial conditions of slurry phase are related to the ther­ ∂ρg
+ ∇⋅ ρg → u g = Smass (13)
modynamic conditions in oxidation tank. Analogous to the approach by ∂t
Marocco and Inzoli [22], the chemical equilibriums for carbonate spe­ ( ) ( )
cies assisted with an auxiliary reaction system are employed to describe ∂
ρ→u g + ∇⋅ ρg →
u g→
ug = − ∇p + ∇⋅τ + ρg →

g + S mom (14)
the initial state of slurry in oxidation tank. The auxiliary system is ∂t g
comprised of the equilibrium for limestone dissolution and the charge ( ) ( ∑ →)
balance for dissolved ions. Table 1 lists the thermodynamic equations ∂( )
ρ e + ∇⋅ ρg e→
ug = − pg ∇⋅→
u g + ∇⋅ keff ∇T + τeff ⋅→
ug− hk J k
and pertinent equilibrium constants [5,40]. The closure scheme for this ∂t g
+ Sen
(15)
( )
∂ →
(ρYk ) + ∇⋅ ρg →
u g Yk = ∇⋅ J k + Sk,mass (16)
∂t

where Smass, Smom and Sen are the source terms accounting for the mass-,
momentum-, and energy-transfer between two phases, which are eval­
uated by averaging the contributions from droplets in the control vol­
ume [4,28,41]:
1∑ 1 ∑( )
Smass = − Si,mass = − Si,H2 O + Si,SO2 (17)
V i V i

→ 1 ∑→ 1∑
S mom = − Fi − Si,mass →
u d,i (18)
V i V i

Fig. 1. Schematic of gas–slurry interface and chemical equilibriums in


slurry phase.

3
J. Qu et al. Chemical Engineering and Processing - Process Intensification 166 (2021) 108478

1 ∑→ → 1∑ which is treated as a fitting parameter taken from Brogren and Karlsson


Sen = − F i u d,i − Si,en (19)
V i V i [14] and Warych and Szymanowski [15]. HSO2 is the Henry’s constant
obtained from the correlation of Maurer [40], kSO2 ,g is the mass-transfer
Slurry droplets are treated as the discrete phase. Each droplet is set as coefficient in gas side calculated by the modified Ranz-Marshall corre­
a spherical homogeneous mixture, which is mainly comprised of H2O
lation [44], and k0SO2 ,l is the liquid-side coefficient taken from the surface
and dissolved SO2 (SO2,aq). The motion of slurry phase is calculated by
stretch model [32]:
discrete phase model (DPM) in the Lagrangian framework. All the in­
dividual droplets are represented by a specified number of parcels. Each kSO2 ,g dp RT
Sh = = 2 + 0.6Red 1/2 Sc1/3 (26)
parcel contains several droplets with the same physical and chemical DSO2 ,g
quantities, including diameter, velocity, temperature, and composition
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
of solution [35,41,42]. Due to the small volume fraction of slurry phase √̅̅̅̅̅̅̅̅̅̅̅
8σ d− g
inside WFGD scrubber, the droplets are assumed to be noninteracting to
0
kSO2 ,l
= 0.88 DSO2 ,l (27)
3πmd
trade off the accuracy against computational load [22,37]. Obeying
Newton’s second law of motion, the flow characteristics for an isolated
droplet is tracked by the following equations: 3. CFD simulation for WFGD spray scrubber
( )
d→ud → ρg In this study, the aqueous chemistry and interphase transfer mech­
md = S i,mom + md →g 1− (20)
dt ρd anisms are coupled with the multiphase flow to optimize spray condi­
tions for the desulfurization process in a commercial-scale WFGD
md cd
dTd
= Si,en (21) scrubber. This section focuses on the operational parameters of scrubber
dt and the numerical strategy for CFD simulation.
dmd
= Si,mass (22) 3.1. Simulation configuration
dt

2.3. Closure model for gas‒slurry transfer behaviors The WFGD spray scrubber considered is in operation at a 330 MW
coal-fired power unit. The present work focuses on the effect of spray
For the fully specified source terms in Eqs. (17)–(19), the closure conditions on desulfurization performance mainly determined by the
models are required for the interphase transfer behaviors. In a WFGD contact condition between gas and droplets. Therefore, the oxidation
scrubber, the characteristics of gas–slurry momentum exchange and SO2 tank with slurry‒bubble flow pattern is ignored in the computational
mass transfer determine the overall desulfurization performance. Even domain, which has been widely accepted in the simulation for SO2 ab­
though the spray cooling has less impact on SO2 absorption process sorption process in spray scrubber [4,22,25-29]. Fig. 2 shows the con­
compared with the other transport processes [4,25], the consideration figurations of equipment and the schematic of grid generation. The
for interphase heat transfer combined with H2O evaporation can provide geometrical entity of scrubber is discretized using unstructured cells.
more detailed information about the evolution of droplet size [38]. The grid system is refined surrounding the pipelines and synergistic
The drag force plays a significant role in the momentum transfer rings. Table 2 summarises the boundary conditions for continuous and
between flue gas and droplets in spray scrubber [2,4,26,31]. The other dispersed phases. For the continuous phase, the raw gas enters the
terms contributing to the aerodynamic forces on droplet can be scrubber from the inlet specified by a mass flow rate, and the purified
neglected as suggested by Crowe et al. [41]. The drag force exerted on gas leaves from the pressure outlet. The mist eliminator with intricate
individual droplet is formulated as: geometrical structure is simplified as a porous zone, in which the pres­
⃒ ⃒( ) sure drop for flue gas is modelled through a specific inertial loss term.
1 ⃒
Si,mom = ρg CD ⃒⃒→ug− →

u d ⃒⃒ →ug− →
u d AD (23) For the dispersed phase, slurry droplets are released into the flow field
2 from totally 500 injection sources with the cone angle of 90◦ . The
droplets are excluded from further calculation until they encounter the
where CD is the drag coefficient evaluated by the Morsi and Alexander gas inlet, bottom, or mist eliminator.
correlation [43], which has been validated for the wide range of Rey­ Table 3 summarises the operational parameters under the design
nolds number (Red) in spray system [4,35]. condition and their corresponding assignments for boundaries. As to the
During the contact between gas and droplets, as aforementioned, the initial compositions of slurry phase, the sulphurous species are assumed
kinetic-controlled reactions can be neglected in slurry bulk, e.g., S(IV) to completely convert into S(VI) species and thus the concentrations of S
oxidation, CaCO3 dissolution, and CaSO4•2H2O crystallization. (IV) species are set as 0 [13,24]. The other initial physical and chemical
Accordingly, the transport of SO2 into droplet is controlled by the mass properties of slurry phase are the same as the measured data of oxidation
transfer resistance in both gas and liquid sides [2,11,14]. At the gas–­ tank. The size distribution of poly-dispersed spray is governed by Ros­
slurry interface, the phase equilibrium is assumed to be governed by in–Rammler function, which is divided into 10 groups with the discrete
Henry’s law [12]. The transfer rate of SO2 into a single droplet is diameters in the range of 0.5 mm to 5 mm. For the direction and
expressed as: morphology of injections, the original spray configurations in scrubber
( )
Si,SO2 = KSO2 ,tot PSO2 ,∞ − HSO2 CSO2 ,∞ Ad MSO2 (24) can be referred to our previous work [38]. The setups in different cases
for comparison will be illustrated in Section 4 to evaluate the effect of
where PSO2 ,∞ is the partial pressure of SO2 in gas bulk taken from Eq. spray conditions on the performance of scrubber.
(16), CSO2 ,∞ is the SO2 molar concentration in the slurry bulk obtained
from the reaction model in Section 2.1, and KSO2 ,tot is the global mass- 3.2. Numerical strategy
transfer coefficient defined as:
( )− 1 The desulfurization process is simulated in a commercial software, in
KSO2 ,tot =
1
+
HSO2
(25) which the models for chemical reactions and interphase transfer process
0
kSO2 ,g ESO2 kSO 2 ,l are implemented by the user defined subroutines. In the Eulerian
framework, the solution for staggered grid is obtained through pressure‒
where ESO2 is the enhancement factor for quantifying the ratio of mass velocity coupling method undertaken by segregated SIMPLEC algo­
flux into the droplet with and without reactions in aqueous phase [16], rithm. The second order upwind scheme is applied to spatially discretize

4
J. Qu et al. Chemical Engineering and Processing - Process Intensification 166 (2021) 108478

Fig. 2. Configuration of WFGD scrubber and schematic of grid generation.

and flue gas are conducted alternately until the variances of SO2 removal
Table 2
efficiency and pressure drop are lower than 0.1% and 20 Pa,
Boundary conditions for the WFGD spray scrubber.
respectively.
Geometry Boundary condition For the calculation by Eulerian–Lagrangian method, the numerical
Continuous phase (flue gas) uncertainties associate with grid resolution and parcel assumption [42].
Raw gas inlet Mass flow inlet Higher grid resolution generates lower numerical errors and tracking
Purified gas outlet Pressure outlet more parcels facilitates the accurate representation of actual droplet
bottom, pipelines and synergistic rings No-slip wall, adiabatic
Mist eliminator Porous zone
distribution, whereas greater computational loading is required. Ac­
Dispersed phase (slurry droplets) cording to the sensitivity analysis [38], the simulation with 2.4 × 106
Inlet, bottom and mist eliminator Escape tetrahedral cells is adopted, and the parcel number is assigned as 200 for
pipelines and synergistic rings Ideal reflection each nozzle under the slurry flow rate of 48 m3 h− 1.

4. Results and discussion


Table 3
Operating parameters and numerical settings under the design condition. In our previous work [38], three regions in full-scale WFGD scrubber
Parameter Value have been distinguished in terms of the gas–slurry flow patterns and
transfer characteristics, i.e., Gas Inlet Region (GIR), Dominant Absorp­
Continuous phase (flue gas)
Flow rate, Gin 1.25 × 106 Nm3 h− 1 on dry basis tion Region (DAR) and Slurry Dispersed Region (SDR). The desulfur­
Inlet temperature, Tg,in 393 K ization performance of scrubber is mainly dependent on the gas–slurry
Inlet SO2 concentration, cSO2 ,in 5000 mg Nm− 3 on dry basis transfer behaviors in DAR and SDR. This study tries to understand the
Viscosity, μg 1.96 × 10− 5 Pa s influence of droplet size, spray direction and morphology of spray cone
Operating pressure 101,325 Pa on the gas–slurry transfer process in the above three regions and to
Dispersed phase (slurry droplets)
Total flow rate, L 24,000 m3 h− 1
explore the optimal spray parameters for industrial desulfurization
Temperature of injection, Tl,in 323 K process.
Density, ρd 1140 kg m− 3
Solid content 15 wt.%
4.1. Model validation
Droplet diameter distribution Rosin–Rammler; de = 2.5 mm and n = 3.0
pH value 5.5
Concentrations of dissolved species, [Mg2+]0 = 0.17 kmol m− 3 Both the desulfurization efficiency of scrubber and the pressure drop
[X]0 [Cl− ]0 = 0.19 kmol m− 3 of flue gas are selected as the indicators to validate the reliability of CFD
[SO2−
4 ]0 = 0.15 kmol m
− 3
model for multiphase hydrodynamics combined with SO2 absorption
process. The measured data are obtained from the industrial WFGD
the convective nonlinear items in all transport equations. In the scrubber mentioned in Section 3.1, which are compared with predicted
Lagrangian framework, the length scale of 20 mm is specified to results in Table 4. Under the design condition shown in Table 3, the CFD
compute the time step for integrating the ordinary differential equations model evaluates the pressure drop with 5.19% deviation on the indus­
of droplets, i.e., Eqs. (20)–(22). trial data. With the SO2 inlet concentrations in the range of about 3800
Two-way coupling method is implemented to simulate the transfer mg Nm− 3 to 6400 mg Nm− 3, the deviations of simulated SO2 removal
process between flue gas and slurry droplets [31,41]. Firstly, the performances are lower than 1% compared with the measured data.
finite-difference equations of flue gas are individually solved in the Therefore, the CFD model in Section 2 can satisfactorily describe the
Eulerian framework without slurry droplets. Using this flow field, the desulfurization process in a commercial-scale WFGD scrubber.
physical properties of representative parcels are calculated in each time
step by integrating Eqs. (20)–(22) in Lagrangian framework. After the 4.2. Effect of droplet diameter
evaluation of source terms in Eqs. (17)–(19), the concentrations of dis­
solved species in slurry phase can be taken from the reaction model in Fig. 3 qualitatively illustrates the influence of droplet diameter (dp)
Section 2.1, meanwhile the gas flow field is solved again to update the on the desulfurization performance with non-ideal flow pattern. On the
interactions between two phases. Then the iterations for slurry droplets one hand, both interfacial area for momentum exchange (AD) and
volumetric mass transfer coefficient (KTotAd) increase with an increasing

5
J. Qu et al. Chemical Engineering and Processing - Process Intensification 166 (2021) 108478

Table 4
Comparison between simulated and measured data of desulfurization efficiency η and pressure drop ΔP (Gin = 1,246,212 Nm3 h− 1, L/G = 19.25 L Nm− 3).
3
Item η/% with different cSO2 ,in /mg Nm− ΔP/Pa
c1 = 3819.6 c2 = 4201.6 c3 = 4552.5 c4 = 5112.8 c5 = 5699.8 c6 = 6394.5

Simulation 98.89 98.66 98.49 98.40 98.10 97.76 1357


Measurement 98.26 97.80 97.68 97.61 97.42 96.86 1290
Error 0.63 0.86 0.81 0.79 0.68 0.90 67

Fig. 3. Desulfurization efficiency depending on uniformity of gas flow and


Fig. 4. Schematic of spray column with plug flow pattern.
dynamics of SO2 mass transfer with respect to droplet diameter.

displays the area with the slurry concentration of 0 kg m− 3 in GIR is


dp, which is positive for the uniformity of gas flow and the absorption
reduced by increasing dp from 2 mm to 4 mm, mainly because the
rate of SO2 [5,13]. On the other hand, the decrease of dp reduces the
droplets with stronger inertia detach from the path of raw gas. It can be
terminal velocity of droplet (uT) according to Eq. (28) derived from Eqs.
found that these cases present the similar evolution of two-phase hy­
(20) and (23) [5]. When droplets fall in the force-balanced state, the
drodynamic structure, where the gas flow in a lateral direction gradually
decline of uT decreases the slip velocity between two phases (|→
ug − →u d |)
turns to upward in DAR and thus the flow pattern progressively ap­
and thus reduces the source term related to momentum transfer in Eq.
proaches to the ideal counter-current mode in Fig. 5. For the slurry
(23), which is negative for mitigating the undesired high velocity of gas
droplets with diameter of 1 mm, a certain number of droplets are still
channelling [33,34,37]. Accordingly, there exists an optimal dp to ach­
entrained by flue gas channelling with extremely high velocity, even
ieve the best desulfurization performance with non-ideal flow pattern in
though their diameters are higher than the critical value of 0.63 mm
WFGD scrubber.
correlated by Morsi and Alexander [43]. The distribution of gas–slurry
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
( ) flow for the droplets with diameter of 1 mm is quite different from the
4gdp ρs
uT = − 1 (28) other three cases, where the two-phase flow pattern is locally trans­
3CD ρg
formed from the counter-current mode into co-current mode in DAR.
Accordingly, the diameter of droplet should be greater than 2 mm for
To clarify the competitive effects of dp on gas–slurry flow pattern and
maintaining a desired flow pattern in the present WFGD scrubber.
SO2 transfer characteristics, the performance of industrial WFGD
To further explore the influence of dp on the interphase momentum
scrubber is compared with its corresponding ideal plug flow operating as
transfer in three regions, the relative standard deviation of gas velocity
the counter-current mode with a uniform size distribution of droplets.
(RSDg) is employed to characterize the uniformity for flue gas. Fig. 7(a)
Fig. 4 shows the plug flow pattern in a spray column. Herein the uniform
exhibits the cross-sectional profiles of RSDg along the height of scrubber.
profile of gas velocity is specified for the inlet of flue gas at the bottom.
As mentioned in Fig. 3, the droplet with larger diameter (dp) generates
The slurry droplets are evenly injected from 4 planes with the same
higher slip velocity between two phases (|→ ug− → u d |) but lower interfa­
heights as the configurations of scrubber in Fig. 2. Following the cor­
relation of Morsi and Alexander [43], the critical diameter of droplet cial area for momentum exchange (AD), which has an opposite effect on
(dcrit) is evaluated as 0.63 mm, which matches with the entrainment the uniformity of gas flow [30,34]. Generally, the uniformity of gas flow
criterion in the gas flow at the superficial gas velocity of 3.04 m s− 1 and for dp = 1 mm reaches a stable state until the gas traverses SDR, rather
temperature of 323 K. The droplet diameters ranging from 1 mm to 4 than DAR in the other three cases, mainly due to the maldistribution of
mm are considered in this section. The other initial and boundary con­ gas–slurry flow as shown in Fig. 6. When dp increases from 1 mm to 2
ditions are the same as those listed in Tables 2 and 3. mm, RSDg decreases in the whole scrubber indicating that the summa­
Under the plug flow condition, the majority of droplets immediately tion of interphase drag force per unit volume of flue gas is largely
achieve the force-balanced state resulting in the stable slurry distribu­ controlled by |→ug− → u d |. When dp further increases from 2 mm to 4 mm
tion as shown in Fig. 5. In this stable state, the terminal velocity of the uniformity degree of flue gas is improved in both GIR and DAR.
droplet (uT) increases from 4.48 m s− 1 to 11.79 m s− 1 by rising dp from 1 However, the reduction of RSDg is less evident with increasing dp. This
mm to 4 mm according to Eq. (28). As expected, the increase of droplet reveals that AD becomes more significant in the uniformity of gas dis­
diameter contributes to a higher falling velocity of droplet in the force- tribution. As to the influence of dp on mass transfer behavior, Fig. 7(b)
balanced state, which decreases the slurry concentration. presents the SO2 concentration gradients with different droplet di­
In the industrial WFGD scrubber with non-ideal flow pattern, Fig. 6 ameters along the height of scrubber. With the diameter of 1 mm, a

6
J. Qu et al. Chemical Engineering and Processing - Process Intensification 166 (2021) 108478

Fig. 5. Distributions of slurry phase in plug flow with different droplet diameters (L/G = 19.25 L Nm− 3, ug,ave = 3.04 m s− 1).

Fig. 6. Distributions of slurry concentration and vector field of flue gas in non-ideal flow with different droplet diameters (L/G = 19.25 L Nm− 3, ug,ave = 3.04 m s− 1).

Furthermore, the transfer rate of SO2 increases with decreasing dp from


2 mm to 4 mm largely because smaller dp causes higher KTotAd in this
region.
With the superficial gas velocity of 3.04 m s− 1 and the ratio of liquid
to gas ratios (L/G) ranging from 16.04 L Nm− 3 to 22.46 L Nm− 3, Fig. 8
illustrates the desulfurization performances of scrubbers with different
flow patterns. In general, SO2 removal efficiency monotonically in­
creases in either plug flow or non-ideal flow pattern by decreasing dp

Fig. 7. Profiles of (a) uniformity degree for flue gas and (b) SO2 concentration
gradient along the height of scrubber with different droplet diameters (Gin =
1,246,212 Nm3 h− 1, L/G = 19.25 L Nm− 3, cSO2 ,in = 5000 mg Nm− 3).

lower absorption rate appears in DAR while a higher value in SDR


because the counter-current mode is restored in SDR, which is negative
for the overall contacting time of two phases in the whole scrubber.
Nevertheless, the profiles of SO2 concentration gradient for dp ranging
from 2 mm to 4 mm are similar to that under the design condition
Fig. 8. Effect of droplet diameter on desulfurization performances in plug flow
summarized in Table 3, which identifies the desulfurization process and non-ideal flow patterns (Gin = 1,246,212 Nm3 h− 1, cSO2 ,in = 5000
mainly occurs in DAR as mentioned in our previous work [38]. mg Nm− 3).

7
J. Qu et al. Chemical Engineering and Processing - Process Intensification 166 (2021) 108478

from 4 mm to 2 mm, though the deviation of η between two patterns gas and slurry in three regions. For the gas distribution in Fig. 10(a), the
increases because of the slightly deteriorated uniformity of flue gas in obvious difference of RSDg presents in DAR among three cases, where
Fig. 7(a). It demonstrates that the overall absorption performance of the uniformity degree of flue gas can be improved by decreasing the
actual scrubber mostly depends on the dynamics of mass transfer related ratio of upward slurry flow rate to overall flow rate (Lup/L). In contrast
to KTotAd for dp ranging from 4 mm to 2 mm. When dp further declines to RSDg evolution, the significant difference of slurry concentration in
from 2 mm to 1 mm, the desulfurization efficiency asymptotically ap­ Fig. 10(b) mainly exhibits in SDR, where the increase of Lup/L leads to
proaches to about 100% in the plug flow, which is consistent with higher slurry concentration related to the gas‒slurry contacting time.
Neveux et al. [5]. However, the corresponding value of η with non-ideal Fig. 11 provides the averaged residence time of droplets (τd) with respect
flow condition declines with reducing dp since the flow pattern of two to droplet diameter. Compared with Case A (Lup/L = 0), the residence
phases gradually deviates from the ideal counter-current mode. It can be time of slurry phase in Case B with Lup/L = 1.0 increases by 25.99% to
inferred that the two-phase flow structure is crucial to the overall 35.56% at the superficial gas velocity of 3.04 m s− 1 and the slurry flow
desulfurization performance of industrial scrubber operating with dp rate of 24,000 m3 h− 1.
from 2 mm to 1 mm. Taking the above two competitive effects into According to the above discussion on hydrodynamics, the influence
consideration, the optimal diameter of droplet is 2 mm for the selected of spray direction on SO2 transfer behavior is determined by the
conditions in the commercial WFGD scrubber, which can be obtained by competitive effects of uniformity of gas flow and concentration of slurry
designing nozzle structure and adjusting atomization pressure. phase. On the one hand, the downward injection effectively modifies the
gas flow due to the uniform distribution of slurry in DAR, which plays a
promotional role for SO2 absorption in a large cross section [10,26,27].
4.3. Effect of spray direction For this reason, the average SO2 concentration declines more rapidly in
DAR for Case A than the other two cases as shown in Fig. 12(a). Agreed
The spray direction of nozzle manipulates both the residence time with Eq. (24), on the other hand, the upward injection of Case B enlarges
distribution of droplets and the flow pattern of two phases [31,35]. For the gas–slurry interfacial area (Ad) accounted for higher slurry concen­
industrial desulfurization process, a lot of works [2,11,22,24‒30] tration in SDR, which is beneficial to increasing the SO2 transfer rate.
focused on the nozzles with downward injection in the spray scrubbers. Consequently, the scrubber in Case B with Lup/L = 1.0 has the best ab­
As listed in Table 5, three kinds of spray modes, i.e., downward direc­ sorption performance in SDR as shown in Fig. 12(b).
tion, upward direction and dual direction, are compared to investigate Fig. 13 displays pressure drops (ΔP) and desulfurization efficiencies
the influence of spray direction on SO2 absorption performance. (η) for three cases. In general, the simulated ΔP increases almost in a
Considering that the upward injections in Groups 1 and 2 can directly linear relationship with L/G agreed with the experiments of Chen et al.
impinge the pipelines or synergistic rings in Fig. 2, it is impractical to [8], and the predicted η asymptotically increases with the increase of
rigorously evaluate the difference of transfer behaviors between upward L/G confirmed by Qin et al. [2]. The comparison of ΔP suggests that
and downward sprays configured in these two groups. To ensure the increasing Lup/L requires more energy consumption for flue gas attrib­
comparability, the nozzles in Groups 1 and 2 are fixed as downward uted to higher interfacial area according to Eq. (23). As to transfer
injection mode. Different spray directions of nozzles in Group 3 are behavior of SO2, desulfurization performance is influenced by both
compared as below. gas–slurry dispersion in DAR and slurry concentration in SDR. The
Fig. 9 shows the distributions of slurry concentration with three improved flow rate of upward injections decreases the uniformity de­
spray directions under the operational conditions listed in Table 3. In gree of gas–slurry dispersion in DAR but increases the slurry concen­
Case A with downward spray mode, the slurry phase presents better tration in SDR. The downward spray mode in Case A causes a poor
dispersion performance below the spray layers compared with that in absorption performance especially at a low flow rate of slurry, while the
the other two cases. It should be noted that the upward injection in Case dual-direction spray in Case C results in obvious increment of η by
B is most likely to be intervened by the pipelines leading to the mal­ comparing to Case A. The main reason is that the increase of slurry
distribution of gas flow with higher velocity, however, this spray mode concentration in SDR related to the residence time of droplet effectively
could produce a higher slurry concentration above each spray layer and improve the overall desulfurization efficiency. However, with the
obtain better spatial distribution in the whole scrubber compared with further the increase of upward flow rate in Case B results in a slight
downward injection mode in Case A. decrease of η by comparing to Cases C especially at a large flow rate of
Fig. 10 presents the quantitative result about the distributions of flue

Table 5
Three cases for comparing the spray directions of nozzles.

Case Case A Case B Case C


Nozzle configuration Orifice type / spray direction

Single/down Single/down Single/down


Group 1
Single/down Single/down Single/down
Group 2
Single/down Single/up Dual/down & up
Group 3

Note: bold items indicate the differences among three cases.

8
J. Qu et al. Chemical Engineering and Processing - Process Intensification 166 (2021) 108478

Fig. 9. Distributions of slurry concentration in xz plane and gas axial velocity at z = 6.5 and 9.75 m with different spray directions (Gin = 1,246,212 Nm3 h− 1, L/G =
19.25 L Nm− 3).

Fig. 10. Profiles of (a) uniformity degree of flue gas and (b) slurry concen­ Fig. 12. Profiles of average SO2 concentration in (a) Dominant Absorption
tration along the height of scrubber with different spray directions (Gin = Region and (b) Slurry Dispersed Region with different spray directions (Gin =
1,246,212 Nm3 h− 1, L/G = 19.25 L Nm− 3). 1,246,212 Nm3 h− 1, L/G = 19.25 L Nm− 3, cSO2 ,in = 5000 mg Nm− 3).

4.4. Effect of spray morphology

As listed in Table 6, three kinds of nozzle arrangement schemes with


different conical shapes are considered to investigate the influence of
spray morphology on interphase transfer process. Taking the above Case
C using hollow-cone nozzles as the base case, Case D selects full-cone
nozzles, and Case E operates in a combination mode consisted of the
full-cone nozzles near the wall and the hollow-cone nozzles in the rest
area. It should be noted that hollow-cone spray is supposed to produce
finer droplet compared with full-cone nozzle [34,39]. In order to explore
the dependence of SO2 absorption performance on slurry dispersion
characteristics, the simulations for three cases in Table 6 are conducted
with the same initial size distribution of droplets to ensure the
comparability.
Figs. 14 exhibits the distribution of droplets in the cross section of z
= 9.0 m, which locates below the spray layers. The hollow-cone nozzles
produce better dispersion characteristics in the core area of scrubber
attributed to circumferential slurry distribution. However, the compar­
Fig. 11. Averaged residence time of droplets with different spray directions ison between Cases C and E elucidates that the full-cone morphology has
(Dashed line: fitting curve, ug,ave = 3.04 m s− 1, L = 24,000 m3 h− 1). better dispersion performance surrounding the wall because the full-
cone spray could generate the uniform coverage below the nozzle
slurry because the controlling factor for η gradually converts to the [23]. Moreover, there are no substantial changes in the distribution of
uniformity degree of slurry distribution in DAR. Overall, the slurry droplets with the proceeding of falling process, e.g., the plane of z
dual-direction spray mode in Cases C is more effective for different = 6.5 m in Fig. 15.
slurry flow rates, which is beneficial to the stable operation of WFGD Fig. 16 presents the uniformity degree of gas flow and slurry con­
scrubber to meet with the fluctuation of SO2 inlet concentration. centration in three regions. The profiles of RSD for gas and slurry are
more similar in SDR, while an obvious difference is found in DAR. In

9
J. Qu et al. Chemical Engineering and Processing - Process Intensification 166 (2021) 108478

Fig. 13. Effect of spray direction on (a) pressure drop and (b) desulfurization efficiency of scrubber (Dashed line: fitting curve, Gin = 1,246,212 Nm3 h− 1, cSO2 ,in =
5000 mg Nm− 3).

Table 6
Three cases for comparing the spray morphologies of nozzles.

Case Case C Case D Case E


Nozzle configuration Morphology of spray cone

Group 1 Hollow cone Full cone Hollow cone


Hollow cone Full cone Full cone

Group 2
Hollow cone Full cone Hollow cone
Group 3

Note: bold items indicate the differences with Case C.

Fig. 14. Distributions of droplets in the typical plane below spray layers (z = 9.0 m) with different morphologies of spray cone.

detail, the hollow-cone nozzles in Case C are prone to generate more with Case C due to the maldistribution of two phases is alleviated near
uniform dispersion of two phases in DAR compared with the full-cone the wall by full-cone nozzles. In fact, the influence of spray morphology
ones in Case D. As shown in Fig. 17, the gas flow with high velocity in on multiphase hydrodynamics is essentially the difference in dispersion
Case D could be mitigated in Case C owing to the better modification characteristics of slurry droplets in DAR. For the mass transfer process,
performance originating from more uniform droplet swarm. Moreover, the slurry distribution becomes the critical factor for determining SO2
the appropriate combination mode of nozzles in Case E could further transfer rate. As shown in Fig. 18, the better dispersion performance of
promote the uniformity degree of gas–slurry flow in DAR by comparing slurry droplet in Case D leads to the lower SO2 concentrations in both

10
J. Qu et al. Chemical Engineering and Processing - Process Intensification 166 (2021) 108478

Fig. 15. Distributions of droplets in the typical plane above gas inlet (z = 6.5 m) with different morphologies of spray cone.

performance of scrubber. However, the hollow-cone nozzles installed in


the central area of scrubber (Groups 1 and 3) results in higher desul­
furization efficiency than full-cone nozzles. According to above discus­
sions, the appropriate combination mode of nozzles in Case E is optimal
for the WFGD scrubber, which improves the removal efficiency by about
0.17% to 1.12% compared with the other two cases.

5. Conclusions

Based on the model validation by the measured data from a WFGD


scrubber of 330 MW coal-fired power unit, the influence of droplet

Fig. 16. Profiles of uniformity degree for (a) flue gas and (b) slurry concen­
tration along the height of scrubber with different morphologies of spray cone
(Gin = 1,246,212 Nm3 h− 1, L/G = 19.25 L Nm− 3).

DAR and SDR than other two cases.


Fig. 19 compares the pressure drops (ΔP) and desulfurization effi­
ciencies (η) of scrubber with different spray morphologies. The full-cone
nozzles in Case D causes lower desulfurization efficiency and energy
consumption in comparison to the hollow-cone nozzles in Case C, and
the gaps of η and ΔP between two cases declines with the increase of L/
G. The combination of full-cone and hollow-cone nozzles in Case E
presents the best SO2 absorption performance, and the increment of η
compared with Case C is more evident at a large slurry flow rate. It can Fig. 18. Profiles of average SO2 concentration in (a) Dominant Absorption
be inferred that full-cone spray mode is optimal for the nozzles sur­ Region and (b) Slurry Dispersed Region with different morphologies of spray
rounding the wall (Group 2) to improve the overall absorption cone (Gin = 1,246,212 Nm3 h− 1, L/G = 19.25 L Nm− 3, cSO2 ,in = 5000 mg Nm− 3).

Fig. 17. Distributions of gas axial velocity in the plane of z = 9.0 m with different morphologies of spray cone (ug,ave = 3.04 m s− 1).

11
J. Qu et al. Chemical Engineering and Processing - Process Intensification 166 (2021) 108478

Fig. 19. Effect of spray morphology on (a) pressure drop and (b) desulfurization efficiency of scrubber (Dashed line: fitting curve, Gin = 1,246,212 Nm3 h− 1, cSO2 ,in =
5000 mg Nm− 3).

diameter, spray injection, and spray morphology on desulfurization ef­ Declaration of Competing Interest
ficiency is investigated to intensify the mass transfer process for SO2
absorption in spray scrubber, which can be summarized as: The authors declare that they have no known competing financial
interests or personal relationships that could have appeared to influence
(1) In the industrial WFGD scrubber, increasing diameter of droplet the work reported in this paper
improves the uniformity of gas–slurry flow but decreases the
volumetric mass transfer coefficient. When the superficial gas Acknowledgments
velocity is 3.04 m s− 1, the overall desulfurization performance is
limited by the maldistribution gas–slurry flow with the droplet This work was supported by the National Natural Science Foundation
diameter ranging from 1 mm to 2 mm, whilst it is dominated by of China (51706070), and the Fundamental Research Funds for the
the dynamics of SO2 mass transfer for the diameters higher than Central Universities (2020MS008 and 2020MS078).
2 mm. The optimal diameter of droplet is 2 mm for the selected
condition in the commercial-scale WFGD scrubber by considering References
the synergy of two competitive effects.
(2) The spray direction of nozzle has a synergistic effect on the [1] M. Kang, J. Shin, T. Yu, J. Hwang, Simultaneous removal of gaseous NOx and SO2
by gas–phase oxidation with ozone and wet scrubbing with sodium hydroxide,
residence time of droplets and the flow pattern of two phases, and Chem. Eng. J. 381 (2020), 122601, https://doi.org/10.1016/j.cej.2019.122601.
further manipulates the overall SO2 absorption efficiency in [2] M. Qin, Y. Dong, L. Cui, J. Yao, C. Ma, Pilot-scale experiment and simulation
different regions of spray scrubber. Upward injection promotes optimization of dual-loop wet flue gas desulfurization spray scrubbers, Chem. Eng.
Res. Des. 148 (2019) 280–290, https://doi.org/10.1016/j.cherd.2019.06.011.
the uniformity of gas–slurry distribution in the middle part of [3] P. Córdoba, Status of flue gas desulphurization (FGD) systems from coal-fired
scrubber (Dominant Absorption Region, DAR), while downward power plants: overview of the physic-chemical control processes of wet limestone
injection produces higher slurry concentration in the upper part FGDs, Fuel 144 (2015) 274–286, https://doi.org/10.1016/j.fuel.2014.12.065.
[4] L. Marocco, F. Inzoli, Multiphase Euler–Lagrange CFD simulation applied to wet
of scrubber (Slurry Dispersed Region, DAR). Compared with the flue gas desulphurization technology, Int. J. Multiph. Flow 35 (2) (2009) 185–194,
single-direction injection, the scrubber arraying dual-direction https://doi.org/10.1016/j.ijmultiphaseflow.2008.09.005.
spray has better desulfurization performance under various [5] T. Neveux, Y.L. Moullec, Wet industrial flue gas desulfurization unit: model
development and validation on industrial data, Ind. Eng. Chem. Res. 50 (2011)
slurry flow rates, which enable to meet with the fluctuation of
7579–7592, https://doi.org/10.1021/ie102239q.
SO2 inlet concentration. [6] L.E. Kallinikos, E.I. Farsari, D.N. Spartinos, N.G. Papayannakos, Simulation of the
(3) Spray morphology influences the dispersion of slurry droplets in operation of an industrial wet flue gas desulfurization system, Fuel Process.
DAR, which essentially determines the overall absorption per­ Technol. 91 (12) (2010) 1794–1802, https://doi.org/10.1016/j.
fuproc.2010.07.020.
formance of scrubber. The full-cone nozzles surrounding the wall [7] J. Zhu, S. Ye, J. Bai, Z. Wu, Z. Liu, Y. Yang, A concise algorithm for calculating
could mitigate the local maldistribution of gas flow ascribed to absorption height in spray tower for wet limestone-gypsum flue gas
the uniform slurry distribution below the nozzle. However, the desulfurization, Fuel Process. Technol. 129 (2015) 15–23, https://doi.org/
10.1016/j.fuproc.2014.07.002.
hollow-cone injections in the core area of scrubber provide better [8] Z. Chen, H. Wang, J. Zhuo, C. You, Enhancement of mass transfer between flue gas
dispersion performance in large cross sections of industrial and slurry in wet flue gas desulfurization spray tower, Energy Fuel32 (1) (2018)
apparatus. The appropriate combination mode of hollow-cone 703–712. https://doi.org/10.1021/acs.energyfuels.7b03009.
[9] D. Flagiello, A. Erto, A. Lancia, F. Di Natale, Experimental and modelling analysis
and full-cone nozzles optimizes the gas‒slurry contacting condi­ of seawater scrubbers for sulphur dioxide removal from flue gas, Fuel 214 (2018)
tions, which leads to higher desulfurization efficiency compared 254–263, https://doi.org/10.1016/j.fuel.2017.10.098.
with the arrangement scheme with identical spray morphology. [10] H.G. Nygaard, S. Kiil, J.E. Johnsson, J.N. Jensen, J. Hansen, F. Fogh, K. Dam-
Johansen, Full-scale measurements of SO2 gas phase concentrations and slurry
compositions in a wet flue gas desulphurization spray absorber, Fuel 83 (9) (2004)
CRediT authorship contribution statement 1151–1164, https://doi.org/10.1016/j.fuel.2003.12.007.
[11] T. Bešenić, J. Baleta, K. Pachler, M. Vujanović, Numerical modelling of sulfur
dioxide absorption for spray scrubbing, Energy Convers. Manag. 217 (2020),
Jiangyuan Qu: Conceptualization, Data curtion, Software, Writing –
112762, https://doi.org/10.1016/j.enconman.2020.112762.
original draft. Nana Qi: Methodology, Project administration, Writing – [12] D. Eden, M. Luckas, A heat and mass transfer model for the simulation of the wet
review & editing. Zhen Li: Supervision, Resources, Project administra­ limestone flue gas scrubbing process, Chem. Eng. Technol. 21 (1998) 56–60,
tion. Kai Zhang: Supervision, Conceptualization, Writing – review & https://doi.org/10.1002/(SICI)1521-4125(199801)21:1<56::AID-CEAT56>3.0.
CO;2-9.
editing. Pengcheng Wang: Investigation, Resources. Lifeng Li: Inves­ [13] B. Dou, W. Pan, Q. Jin, W. Wang, Y. Li, Prediction of SO2 removal efficiency for wet
tigation, Resources. flue gas desulfurization, Energy Convers. Manag. 50 (2009) 2547–2553, https://
doi.org/10.1016/j.enconman.2009.06.012.

12
J. Qu et al. Chemical Engineering and Processing - Process Intensification 166 (2021) 108478

[14] C. Brogren, H.T. Karlsson, Modeling the absorption of SO2 in a spray scrubber using [29] L. Cui, J. Lu, M. Qin, Y. Dong, Simulation study on novel groove separator in a
the penetration theory, Chem. Eng. Sci. 52 (18) (1997) 3085–3099, https://doi. dual-loop wet flue gas desulfurization spray tower, Asia-Pac. J. Chem. Eng. 15 (4)
org/10.1016/S0009-2509(97)00126-7. (2020) e2442, https://doi.org/10.1002/apj.2442.
[15] J. Warych, M. Szymanowski, Model of the wet limestone flue gas desulfurization [30] Q. Zhang, S. Wang, P. Zhu, Z. Wang, G. Zhang, Full-scale simulation of flow field in
process for cost optimization, Ind. Eng. Chem. Res. 40 (12) (2001) 2597–2605, ammonia-based wet flue gas desulfurization double tower, J. Energy Inst. 91
https://doi.org/10.1021/ie0005708. (2018) 619–629, https://doi.org/10.1016/j.joei.2017.02.010.
[16] O. Levenspiel, Chemical Reaction Engineering, Wiley, New York, 1999. [31] S. Feng, L. Xiao, Z. Ge, L. Yang, X. Du, H. Wu, Parameter analysis of atomized
[17] J.B.W. Frandsen, S. Kiil, J.E. Johnsson, Optimization of a wet FGD pilot plant using droplets sprayed evaporation in flue gas flow, Int. J. Heat Mass Tran. 129 (2019)
fine limestone and organic acids, Chem. Eng. Sci. 56 (10) (2001) 3275–3287, 936–952, https://doi.org/10.1016/j.ijheatmasstransfer.2018.10.023.
https://doi.org/10.1016/S0009-2509(01)00010-0. [32] C.T. Hsu, S.M. Shih, Semiempirical equation for liquid-phase mass-transfer
[18] Z. Li, C. Xie, J. Lv, R. Zhai, Effect of calcium formate as an additive on coefficient for drops, AIChE J. 39 (6) (1993) 1090–1092, https://doi.org/10.1002/
desulfurization in power plants, J. Environ. Sci. 67 (5) (2018) 89–95, https://doi. aic.690390618.
org/10.1016/j.jes.2017.06.023. [33] X. Guo, L. Yin, L. Hu, J. Cao, H. Shen, J. Xu, Y. Hu, D. Chen, Numerical simulation
[19] S. Mori, M. Kumita, S. Shichida, Mass transfer enhancement during electrolysis of wet deacidification process of sludge incineration flue gas, Fuel 280 (2020),
with cylindrical electrodes by magnetic field exposure and its dependency on 118480, https://doi.org/10.1016/j.fuel.2020.118480.
electrode positions, Energy Convers. Manag. 43 (3) (2002) 383–397, https://doi. [34] P. Wang, G. Dai, Synergistic effect between spraying layers on the performance of
org/10.1016/S0196-8904(01)00108-X. the WFGD spray column, Asia-Pac. J. Chem. Eng. 13 (6) (2018) e2266, https://doi.
[20] F. Di Natale, C. Carotenuto, S. Caserta, M. Troiano, L. Manna, A. Lancia, org/10.1002/apj.2266.
Experimental evidences on the chemi-electro-hydrodynamic absorption of sulphur [35] Y. Sun, Z. Guan, H. Gurgenci, K. Hooman, X. Li, L. Xia, Investigation on the
dioxide in electrified water sprays, Chem. Eng. Res. Des. 146 (2019) 249–262, influence of injection direction on the spray cooling performance in natural draft
https://doi.org/10.1016/j.cherd.2019.04.006. dry cooling tower, Int. J. Heat Mass Transf. 110 (2017) 113–131, https://doi.org/
[21] T. Chen, Y. Chen, P. Chiang, Enhanced performance on simultaneous removal of 10.1016/j.ijheatmasstransfer.2017.02.069.
NOx-SO2-CO2 using a high-gravity rotating packed bed and alkaline wastes [36] T. Zhang, B. Dong, X. Chen, Z. Qiu, R. Jiang, W. Li, Spray characteristics of
towards green process intensification, Chem. Eng. J. 393 (2020), 124678, https:// pressure-swirl nozzles at different nozzle diameters, Appl. Therm. Eng. 121 (2017)
doi.org/10.1016/j.cej.2020.124678. 984–991, https://doi.org/10.1016/j.applthermaleng.2017.04.089.
[22] L. Marocco, Modeling of the fluid dynamics and SO2 absorption in a gas-liquid [37] J.A. Michalski, Aerodynamic characteristics of FGD spray towers, Chem. Eng.
reactor, Chem. Eng. J. 162 (1) (2010) 217–226, https://doi.org/10.1016/j. Technol. 20 (2) (1997) 108–117, https://doi.org/10.1002/ceat.270200208.
cej.2010.05.033. [38] J. Qu, N. Qi, K. Zhang, L. Li, P. Wang, Wet flue gas desulfurization performance of
[23] Y. Sun, A.M. Alkhedhair, Z. Guan, K. Hooman, Numerical and experimental study 330 MW coal-fired power unit based on CFD region identification of flow pattern
on the spray characteristics of full-cone pressure swirl atomizers, Energy160 (2018) and transfer process, Chinese J. Chem. Eng. 29 (2021) 13–26, https://doi.org/
678–692. https://doi.org/10.1016/j.energy.2018.07.060. 10.1016/j.cjche.2020.08.004.
[24] C. Tseng, C. Li, Eulerian–Eulerian numerical simulation for a flue gas [39] X. Wu, H. Zhao, Y. Zhang, C. Zheng, X. Gao, Measurement of slurry droplets in
desulfurization tower with perforated sieve trays, Int. J. Heat Mass Tran. 116 coal-fired flue gas after WFGD, Environ. Geochem. Health 37 (5) (2015) 915–929,
(2018) 329–345, https://doi.org/10.1016/j.ijheatmasstransfer.2017.09.024. https://doi.org/10.1007/s10653-014-9648-x.
[25] C. Montanes, S.A. Gomez, N. Fueyo, J.C. Ballesteros, P. Gomez-Yague, [40] G. Maurer, On the solubility of volatile weak electrolytes in aqueous solutions, ACS
Computational evaluation of wall rings in wet flue-gas desulfurization plants, Int. Symp. Ser. 133 (1980) 139–172, https://doi.org/10.1021/bk-1980-0133.ch007.
J. Energy Clean. Environ. 10 (2009) 15–36, https://doi.org/10.1615/ [41] C.T. Crowe, M.P. Sharma, D.E. Stock, The particle-source-in cell (PSI-CELL) model
InterJEnerCleanEnv.v10.i1-4.20. for gas-droplet flow, J. Fluids Eng. 99 (2) (1977) 325–332, https://doi.org/
[26] Z. Chen, H. Wang, J. Zhuo, C. You, Experimental and numerical study on effects of 10.1115/1.3448756.
deflectors on flow field distribution and desulfurization efficiency in spray towers, [42] S. Benyahia, J.E. Galvin, Estimation of numerical errors related to some basic
Fuel Process. Technol. 162 (2017) 1–12, https://doi.org/10.1016/j. assumptions in discrete particle methods, Ind. Eng. Chem. Res. 49 (21) (2010)
fuproc.2017.03.024. 10588–10605, https://doi.org/10.1021/ie100662z.
[27] P. Wang, L. Zhuang, G. Dai, Synergistic effect of droplet self-adjustment and rod [43] S.A. Morsi, A.J. Alexander, An investigation of particle trajectories in two-phase
bank internal on fluid distribution in a WFGD spray column, Chem. Eng. Sci. 162 flow systems, J. Fluid Mech. 55 (2) (1972) 193–208, https://doi.org/10.1017/
(2017) 227–244, https://doi.org/10.1016/j.ces.2016.12.062. S0022112072001806.
[28] P. Wang, G. Dai, Field synergy of the rod bank on the enhancement of mass transfer [44] P.A. Nelson, T.R. Galloway, Particle-to-fluid heat and mass transfer in dense
in a spray column, Ind. Eng. Chem. Res. 57 (2018) 12531–12542, https://doi.org/ systems of fine particles, Chem. Eng. Sci. 30 (1) (1975) 1–6, https://doi.org/
10.1021/acs.iecr.8b02656. 10.1016/0009-2509(75)85109-8.

13

You might also like