Download as pdf or txt
Download as pdf or txt
You are on page 1of 56

Journal Pre-proofs

Heat Transfer Calculation Methods in Three-Dimensional CFD Model for


Pulverized Coal-Fired Boilers

Heyang Wang, Chaoqun Zhang, Xin Liu

PII: S1359-4311(19)34550-8
DOI: https://doi.org/10.1016/j.applthermaleng.2019.114633
Reference: ATE 114633

To appear in: Applied Thermal Engineering

Received Date: 2 July 2019


Revised Date: 28 October 2019
Accepted Date: 3 November 2019

Please cite this article as: H. Wang, C. Zhang, X. Liu, Heat Transfer Calculation Methods in Three-Dimensional
CFD Model for Pulverized Coal-Fired Boilers, Applied Thermal Engineering (2019), doi: https://doi.org/10.1016/
j.applthermaleng.2019.114633

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version
will undergo additional copyediting, typesetting and review before it is published in its final form, but we are
providing this version to give early visibility of the article. Please note that, during the production process, errors
may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2019 Elsevier Ltd. All rights reserved.


Heat Transfer Calculation Methods in Three-Dimensional CFD Model for

Pulverized Coal-Fired Boilers

Heyang Wanga,b,*, Chaoqun Zhangc, Xin Liuc


a Key Laboratory of Efficient Utilization of Low and Medium Grade Energy (Tianjin

University), Ministry of Education, Tianjin 300350, China


b School of Mechanical Engineering, Tianjin University, Tianjin 300350, China
c Yantai Longyuan Power Technology Co. Ltd., Yantai, Shandong Province, 264006, China

Highlights:

 Developed efficient heat transfer models for practical boiler applications.

 Furnace wall heat transfer calculated in terms of a thermal boundary condition.

 Wall heat transfer parameters determined based on the thermal properties of ash.

 Predicted boiler heat transfer distributions in close agreement with the plant data.

*Corresponding author
E-mail address: heyang.wang@tju.edu.cn
Abstract
Computational fluid dynamics (CFD) has been used extensively to study the combustion

characteristics of coal-fired boilers. However, very limited of the boiler CFD studies were used

to quantitatively predict the heat transfer distributions in practical boiler applications. This

paper presents a systematic description on the heat transfer calculation methods in three-

dimensional CFD model for coal-fired boilers. The primary objective is to provide a set of heat

transfer submodels that can be easily employed to solve and analyze large scale practical boiler

problems. All types of boiler heating surfaces are considered with particular focus on the heat

transfer calculation of furnace water wall. It is implemented in the CFD model in terms of a

thermal boundary condition (B.C.) describing the energy balance between the boiler’s fire side

heat transfer and water side heat absorption. The physical significance of the associated B.C.

parameters and their accurate prescription are discussed in detail in this paper. It is found that

furnace wall heat transfer is largely affected by the wall ash deposit conditions such that the

heat transfer coefficient of furnace wall can be determined based on the thermal properties of

ash. Although ash thickness may vary dramatically over furnace wall and is extremely difficult

to predict, its thermal conductance was found to only vary within a small range for normally

operated coal-fired boilers. Establishing the connection between the thermal conductance of

wall ash deposits and the heat transfer calculation of furnace wall bypasses the tremendous

difficulty and uncertainty incurred in predicting the ash thickness, and thereby, greatly

simplifies the wall heat transfer calculation process. The heat transfer calculation methods

introduced in this study embody the key physics associated with boiler’s heat transfer process

while making a reasonable balance between the level of model complexity and their

applicability to practical boiler problems. In this paper they are illustrated and validated on a

330 MW tangentially-fired subcritical boiler. The results show that the predicted boiler heat

transfer distributions are in close agreement with the boiler’s operating data.
Keywords: Boiler; coal combustion; Computational Fluid Dynamics; heat transfer; furnace
wall; ash deposits
1. Introduction

With continuous advances in computer power, three-dimensional (3D) computational fluid

dynamics (CFD) has achieved significant progress in the simulation of coal-fired boilers during

the past decades [1-3]. The advances in CFD provide the basis for further insight into the

dynamics of multiphase reacting flow and allow us to look in more detail into any specific

process involved in the boiler combustion process that may not be available any other way. As

such, CFD is now becoming a powerful and efficient tool to predict the performance of boilers,

evaluate the design of new boilers and modification of existing boilers, optimize boiler

operations, and investigate and resolve specific boiler problems. For example, CFD has been

widely employed to evaluate the performance of various boiler modification and operation

optimization options to reduce NOx emissions, which is one of the major concerns of power

industry during recent years [4-13]. Some other studies focused on the prediction of char

burnout, which is one of the crucial factors affecting boiler combustion efficiency [9, 14-16].

It has also been applied to study many specific boiler problems, such as tube surface erosion

[17], coal switching [18], burner ignition [19], syngas reburning [20], and boiler slagging and

fouling [21, 22]. Nevertheless, so far few of these boiler CFD studies have presented a

systematic description on the calculation method of boiler heat transfer and been used to

quantitatively predict the boiler heat transfer distributions in practical boiler applications.

Boiler is, first of all, a heat exchanger. Heat transfer plays the central role in the design and

operation of boilers. A deep understanding and accurate prediction of boiler heat transfer are

essential to optimized system design and operation, improved thermal efficiency, and reduced

pollutant emissions. Thus, in this study we will present a systematic description on the heat

transfer calculation methods for different types of boiler heating surfaces, including radiative

heating surfaces, such as furnace water wall and division panel, and convective heating surfaces,

such as superheater and economizer. We aim to provide a set of easy-to-use heat transfer
submodels that can be readily incorporated into the 3D CFD model to solve and analyze large

scale practical boiler problems. Amongst all the components of boiler heating surfaces, the heat

transfer of furnace water wall plays the most important role because it also determines the

amount of available energy leaving the furnace to the boiler’s convection pass and the

subsequent heat transfer to the convective heating surfaces. Furnace wall heat transfer also

directly affects the overall furnace temperature level, which subsequently affects the coal

combustion and air pollutant formation processes. Furnace wall temperature distribution,

which is directly related with the wall heat flux distribution, is also a key factor associated with

many specific boiler problems, such as tube overheating and high temperature corrosion. Thus,

furnace wall heat transfer virtually affects every aspect of boiler operation. For this reason, its

calculation is going to be the primary focus of the present study. Many of the previous boiler

CFD studies used constant wall surface temperature as the thermal boundary condition (B.C.)

to implement the calculation of furnace wall heat transfer [3, 6, 11, 19]. However, as will be

seen later, furnace wall temperature distribution is highly non-uniform. It varies dramatically

with the wall heat flux distribution and is highly dependent on the local wall ash deposit

condition. In the lower furnace where flame temperature is high and radiation heat transfer is

intense, furnace wall surface temperature could be substantially higher than those at the upper

furnace. On the other hand, on the heavily ash deposited wall areas, since ash inhibits the heat

transfer through the water wall, the wall surface temperature may increase substantially. This

will lead to significantly increased self-radiation, and hence, decreased heat absorption of

furnace wall. Thus, furnace wall surface temperature should be solved as a crucial simulation

result rather than being given as a fixed boundary condition. In this paper a simple yet

physically straightforward method was developed in which the heat transfer calculation of

furnace water wall is implemented in terms of a thermal boundary condition (B.C.) describing

the energy balance between the boiler’s fire side heat transfer and water side heat absorption.
The physical significance of the associated B.C. parameters and accurate definition of their

values are discussed in detail in this paper. It is found that furnace wall heat transfer is largely

affected by the wall ash deposit status, and thus, the wall thermal B.C. parameters can be

determined based on the thermal properties of ash, particularly, its thermal conductance.

Although the thickness of ash deposits varies dramatically with boiler’s operating conditions,

its thermal conductance may only vary within a small range for normally operated coal-fired

boilers. Establishing the connection between the thermal conductance of ash deposits and the

heat transfer calculation of furnace wall is the most crucial finding of this paper since it

bypasses the tremendous difficulty and uncertainty incurred in the process of calculating the

thickness of wall ash deposits, and thereby, confines the prediction of furnace wall heat transfer

within a reasonably accurate range in a very simple way.

The heat transfer submodels developed in this paper embody the key physics involved in the

heat transfer process while making a reasonable balance between the level of model complexity

and their applicability to practical boiler problems. In particular, appropriate correspondence

between the model parameters and boiler’s key design and operating parameters are

incorporated into these models. With these heat transfer submodels the CFD model is now able

to accurately predict the heat transfer distributions in practical boiler applications. In this paper

they are illustrated and validated on a 330 MW tangentially-fired subcritical boiler and the

predicted heat transfer results are in close agreement with the boiler’s operating data.

2. Furnace geometry and operating conditions

The CFD models presented in this study is applied to a 330 MW tangentially-fired (T-fired)

boiler. Figure 1 shows the schematic of furnace geometry with the arrangement of coal and air

burners and boiler heating surfaces. Also shown in Fig. 1 includes the horizontal mesh

distribution across the burner region. The furnace has the dimensions of 14.0 m (W) × 14.0 m
(D) × 56.2 m (H) with the burners located at four corners. The firing system consists of five

levels of coal burners separated by auxiliary (Aux) air burners. Three levels of separated

overfire air (SOFA) burners are located 4.4 m above the top level of coal burners. The outlet

of the furnace geometry is placed at the economizer outlet in order to have all the convective

heating surfaces (superheaters, reheaters and economizers) included in the geometry. Inclusion

of all these heating surfaces in the model makes it possible to obtain a complete overview of

the boiler’s thermal performance. In the boiler’s convection pass, from upstream to

downstream, are reheaters (RH), secondary superheater (SSH), primary superheaters (PSH)

and economizer (ECON). These convective heating surfaces usually consist of a large number

of tube panels. It is not practical to represent detailed geometry of all the tubes in the furnace

geometry. Rather, they are modeled as porous media volumes that correspond to the physical

regions occupied by these heating surfaces. At the top of the furnace are division panel

superheater (DPSH) and platen superheater (PLSH). The tube panels of these heating surfaces

are modeled as interior double-side walls in the model, as shown in Fig. 1. The platen

superheater may also comprise a large number of tube panels, and treating each tube panel as

an interior wall may lead to increased difficulty in creating the furnace geometry and mesh.

However, it should not be modeled as a volume like the convective heating surfaces. Platen

superheater is a radiative type heating surface. Treating it as a volume cannot correctly model

the surface-flame radiation heat exchange and will significantly underestimate the heat transfer

of these heating surfaces [23].

The mesh is constructed with 3.54 million hexahedral structured cells. Most of the finer mesh

are located in the furnace burner region, and the mesh is further refined in the regions near the

burner inlets (as shown in Fig. 1) in order to better resolve the influence of fine burner design

features on the near-burner flow field.


Figure 1. Schematic of the boiler burner and heating surface arrangement

The boiler is designed to fire bituminous coal. The proximate and ultimate analyses data of the

coal (as-received basis) are shown in Table 1. The main operating conditions for the model

cases are listed in Table 2. As shown in Table 2, three cases are simulated in this study. Case

#1 is the validation case for demonstration and validation of the heat transfer models introduced

in this study. Case #2 is the burner tilt case where all the coal and air burners are tilt down by

20°. Case #3 corresponds to a coal switching case in which a lower heating value coal is used.

All these cases use the same total fuel heat input and furnace excess air ratio of 1.235. The

boiler water and steam parameters (corresponding to case #1) are shown in Table 3. Model

validation can be carried out by matching the gas side heat transfer prediction results with those

calculated from the boiler water and steam parameters.


Coal A Coal B
Carbon (ar) 63.24 % 52.47 %
Hydrogen (ar) 4.08 % 2.84 %
Oxygen (ar) 8.46 % 13.23 %
Nitrogen (ar) 0.75 % 0.34 %
Sulfur (ar) 0.81 % 0.41 %
Ash (ar) 15.16 % 11.41%
Moisture (ar) 7.50 % 19.30%
Volatile (ar) 28.63 % 25.64%
Heating Value (LHV) 24.58 MJ/kg 19.00 MJ/kg

Table 1. Coal proximate and ultimate analyses data, as-received (ar) base

Case #1 Case #2 Case #3


Coal Type Coal A Coal A Coal B
Total Coal Flow (t/h) 123.5 123.5 159.8
Excess Air Ratio 1.235 1.235 1.235
Total Air Flow (t/h) 1278.5 1278.5 1279.9
Total PA Flow Rate (%) 23.6 23.6 23.6
Total SA Flow Rate (%) 61.4 61.4 61.4
Total SOFA Flow Rate (%) 15.0 15.0 15.0
PA Temperature (%) 77.0 77.0 77.0
SA Temperature (ºC) 337.8 337.8 337.8
Burner Tilt Angle (Degree) 0 -20 0

Table 2. Main boiler operating conditions for model cases


Main Steam Flow 1065 t/h
Main Steam Temperature 543 ºC
Main Steam Pressure 18.54 MPa.g
Feed Water Temperature 261.0 ºC
Feed Water Pressure 20.30 MPa.g
Drum Pressure 19.91 MPa.g
Reheat Steam Flow 965 t/h
Reheat Steam Temperature (Inlet/Outlet) 345.8/543.0 ºC
Reheat Steam Pressure (Inlet/Outlet) 4.53/4.3 MPa.g

Table 3. Boiler feed water and steam parameters

3. Mathematical models

3.1 Overall models

ANSYS Fluent is used as the computational platform to implement the CFD models presented

in this study. The gas flow is described by the Favre-averaged conservation equations of mass,

momentum, energy and species mass fraction. Realizable k- turbulence model is used for

turbulence closure [23, 24]. The motion of discrete coal particles is described in Lagrangian

framework using a stochastic tracking method in which the instantaneous gas velocity is used

to compute the aerodynamic drag force on the dispersed particles. Coal combustion is modeled

following the sequential process of moisture evaporation, devolatilization, and char

combustion. As coal particle trajectories are tracked, the mass, momentum and energy

exchange between the gas phase and coal particles are accounted in the source terms of the gas

phase conservation equations using the Particle-Source-In-Cell (PSIC) method [1-3]. The

velocity, temperature and turbulence distributions at the furnace burner inlets are obtained from

separate burner flow models for each burner. The flow, temperature and turbulence

distributions obtained at the outlets of these burner models are then transformed to the

corresponding locations at the furnace burner inlets and prescribed as the inlet boundary
conditions [25]. This is a crucial step for the CFD model to be able to correctly predict the

boiler aerodynamic field and take into consideration the influences of different burner design

features and burner air flow distributions on the near-burner flow field. The initial velocity and

distributions of coal particles at the furnace coal burner inlets are also obtained in the same

way. The coal particle sizes are assumed to follow the Rosin-Rammler distribution [6, 12], and

are segmented into 10 discretized size bins (from 5.8 m to 230 m) with an average diameter

of 59 m and size distribution parameter of 1.0. A total number of 130,000 representative coal

particles are tracked in the furnace. A portion of the coal moisture is assumed to have been

evaporated into the primary air before entering the furnace. The evaporated moisture results in

a slight change of primary air composition which will be considered in the CFD model to ensure

overall mass balance.

3.2 Coal devolatilization

The devolatilization process of coal volatile matter is modeled using a first-order single

reaction rate model [25, 26]:

𝑑𝑉g
= 𝑘v(𝑉0 ― 𝑉g) (1)
𝑑𝑡

where 𝑉0 is the maximum yield of volatile matter from unit mass of raw coal, 𝑉g is the yield

of volatile matter at time t, and 𝑘v is the kinetic rate of volatile release (s-1) given in Arrhenius

form by

𝑘v = 𝐴vexp ― ( 𝐸v
𝑅𝑇p ) (2)

where Av and Ev are the pre-exponential factor and activation energy, respectively, Tp is the

particle temperature, and R is the universal gas constant. The values of V0, Av and Ev for both

coals are given in Table 4. They were obtained from separate FLASHCHAIN models in drop

tube furnace [25, 27]. It is noted that the values of V0 predicted by FLASHCHAIN are higher
than the proximate analysis data in Table 1. This is mainly due to the increased release of tar

under a much higher heating rate than the proximate analysis. Although coal volatile matter

contains many kinds of species, such as light gases (CH4, H2 etc) and tar, in this study the

volatile matter is represented by a single virtual substance CaHbOcSdNe where the values of a,

b, c, d, and e are determined by applying mass balance to each of the element of volatile

substance and assuming that the char is composed of pure carbon and ash [3, 19, 20, 28, 29].

Coal A Coal B
Coal devolatilization model
V0 (%) 43.70 36.03
Av (1/s) 9203 3447
Ev (kJ/mole) 27.67 16.96
Char surface combustion model
C1 5e-12 5e-12
C2 0.001 0.001
Ec (kJ/mole) 79.0 79.0

Table 4. The model parameters of coal devolatilization and char surface combustion

3.3 Char combustion

After volatile matter is completely released from coal, the residue char continues to react with

the surrounding gases. In conventional coal-fired furnaces, the partial pressures of CO2 and

H2O are quite low. Thus, in this study only the heterogeneous reaction of char with O2 is

considered:

1
Cchar + O2→CO (3)
2

The CO generated at the char surface is then transported to the bulk gas and further oxidized

to CO2. Assuming first order in O2 partial pressure 𝑝ox (atm), the char combustion rate can be

written as
𝑑𝑚p
= ―𝜋𝑑2p𝑝ox𝑘c (4)
𝑑𝑡

where 𝑚p is the particle mass, 𝑑p is the particle diameter, and 𝑘c is the overall char surface

reaction rate coefficient (kg/(sm2atm)). The overall char reaction rate is controlled by both

the rate of O2 diffusion from the bulk gas to the particle surface and the rate of char surface

reaction. It is modeled by the kinetic/diffusion-limited rate model [28, 30] in which the

diffusion rate coefficient

[(𝑇𝑝 + 𝑇∞)/2]0.75
𝐷0 = 𝐶1 (5)
𝑑𝑝

and the surface reaction rate coefficient

ℜ = 𝐶2exp( ― 𝐸c/𝑅𝑇𝑝) (6)

are weighted to yield the overall char surface reaction rate coefficient

𝐷0ℜ
𝑘c = (7)
𝐷0 + ℜ

In Eq. (5) 𝑇∞ denotes the surrounding gas temperature. In this model, the particle size 𝑑𝑝 is

assumed to remain constant while the density is allowed to decrease during char burnout. The

values of the rate constants used in the model, C1, C2 and Ec, are given in Table 4. Here, the

apparent activation energy Ec (79 kJ/mole) for coal A was obtained from the averaged

measurement values made in an entrained combustion reactor for high-volatile bituminous

coals [31]. The value of the pre-exponential constant C2 was then calibrated to match the

predicted flyash unburned carbon with the plant data. Coal B is used as a case study coal to

demonstrate that the heat transfer submodels developed in this study can naturally consider the

effect of coal heating value. There are no plant data for coal B to calibrate its char reaction rate

parameters. However, since it has very close proximate and ultimate analyses data with coal A

(dry-ash-free base), we assume that they have the same char reaction rate parameters.
3.4 Gas phase reaction

In this study, coal volatile matter is treated as a virtual substance CaHbOcSdNe and its

homogeneous reaction is modeled with a two-step reaction scheme [28, 29]:

CaHbOcSdNe + (𝑎
2
𝑏 𝑐
) 𝑏
+ 4 ― 2 + 𝑑 O2 → aCO + 2H2O + dSO2 + 2N2
𝑒
(8)

1
CO + 2O2 → CO2 (9)

The gas phase mixture is thus composed of the seven species, namely the volatile matter

(CaHbOcSdNe), oxygen (O2), water vapor (H2O), carbon dioxide (CO2), carbon monoxide (CO),

sulfur dioxide (SO2), and nitrogen (N2). Conservation equations for the mass fractions of all

these species but N2 are solved.

Homogeneous reaction involves the transport and chemical reaction of various gas species. For

engineering applications combustion is usually intense such that the reaction rate is overall

limited by the mixing rate between the fuels and oxidizers. For this reason, the eddy dissipation

model (EDM) [32] is used to model the homogeneous chemical reaction rate. In this model,

the mean reaction rate is controlled by the turbulent mixing between the fuels and oxidizers,

characterized by a turbulent mixing time scale, given by 𝑘 𝜀, where k is the turbulent kinetic

energy and  is the energy dissipation rate. The mean reaction rate of fuel is then given by:

𝑌ox 𝑌p
𝜀
𝑅f = 𝐴𝜌 min 𝑌f, ,𝐵
𝑘 (
𝑠 1+𝑠 ) (10)

where 𝑅f is the reaction rate of fuel (kg/(sm3)), 𝜌 is the density of gas mixture, 𝑌f is the fuel

mass fraction, 𝑌ox is the oxidizer mass fraction, 𝑌p is the total mass fraction of product species,

s is the stoichiometric coefficient of oxidizer, and A and B are empirical model constants.

3.5 Radiation

In coal-fired boiler, radiation is the dominant energy transport mechanism in the furnace

contributing more than 90% of total heat transfer to the furnace water wall [33]. Calculation of
radiative heat transfer is therefore a crucial element in accurate prediction of boiler heat transfer.

The accuracy of radiation calculation depends on a combination of the accuracy of the

calculation method and the radiation transport properties of the participating media. In this

study, the discrete ordinate method (DOM) is used to simulate the radiation heat transfer in

which the radiative transfer equation (RTE) is solved with an angular discretization of 3

divisions and 4 pixels in both polar and azimuthal directions. The emissivity of the gas mixture

depends on local variations in gas composition (mainly CO2 and H2O), temperature and

pressure. They are calculated using the cell-based weighted-sum-of-gray-gases model

(WSGGM). The calculation results of DOM for radiation yields the solution of incident heat

radiation to the furnace wall, which is the key quantity to the heat transfer calculation of boiler’s

radiative heating surfaces, as will be described in the next section.

4. Boiler heat transfer calculation methods

4.1 Furnace water wall

Radiation is the dominant heat transfer mechanism of furnace water wall. As illustrated in Fig.

2, strong incident thermal radiation from high temperature flame strikes the furnace wall. A

portion of the incident radiation is reflected back to the furnace. Furnace wall also emits

radiation energy to the surrounding. Then the net wall radiation heat absorption is the difference

between the incoming and outcoming radiation:

𝑞rad = 𝜀w(𝐺 ― 𝜎𝑇4w) (11)

where 𝑞rad is the net wall radiation heat flux, 𝜀w is the wall surface emissivity, G is the incident

radiation heat flux, 𝑇w is the wall surface temperature, 𝜎 is the Stefan-Boltzmann constant

(5.669 × 10 ―8 W/(m2K4)), and here we have assumed gray wall surface.


Figure 2. Schematic of wall surface radiation heat transfer

The heat absorption from the fire side of boiler is conducted through the furnace wall and then

absorbed by the water flow in the water tubes. This heat transfer process can be described by

the following heat balance equation:

𝑞rad = ℎext(𝑇w ― 𝑇ext) (12)

where 𝑇ext denotes the external heat sink temperature and ℎext the heat transfer coefficient.

Here, for the ease of analytical discussion we have neglected the contribution of fire side

convective heat transfer since it usually only contributes a very small portion of the overall

furnace wall heat transfer. Equation (12) describes the heat balance between the boiler fire side

heat transfer and water side heat absorption by which the thermal boundary condition (B.C.)

for the furnace water wall can be defined. Equation (12) can be implemented into Fluent by

means of a “convection” type wall thermal boundary condition. This type of wall thermal B.C.

requires three user-input parameters, namely ℎext, 𝑇ext and 𝜀w. Substituting Eq. (11) into Eq.

(12), we have

𝜀w(𝐺 ― 𝜎𝑇4w) = ℎext(𝑇w ― 𝑇ext) (13)

It can be seen from Eq. (13) that with the incident radiation G solved internally by the CFD

solver and ℎext, 𝑇ext and 𝜀w given as the input B.C. parameters, the only unknown in Eq. (13)

is the wall surface temperature, 𝑇w. Equation (13) along with the gas phase energy conservation
equation yields the solution of temperature distribution over the entire furnace domain. The

wall radiation heat flux, 𝑞rad, then can be calculated by substituting the wall surface

temperature, 𝑇w, back into Eq. (11). It is clear that, with the value of G calculated internally by

the CFD solver, accurate calculation of the wall surface temperature and heat flux distribution

is decided by the accuracy of the B.C. parameters, ℎext, 𝑇ext and 𝜀w. Accurate definition of

these parameters must depend on proper understanding of the physical significance of these

parameters and appropriately relating their values with the boiler design and operating

parameters.

For steam boilers, the water/steam flow acts as the external heat sink carrying away the heat

absorbed from the fire side of boiler. Thus, 𝑇ext apparently should adopt the temperature of the

water flow in the water tubes. However, the physical significance of the heat transfer

coefficient, ℎext, is not that straightforward. It is first to be noted that ℎext has different physical

significance from what is assumed in Fluent. Fluent assumes that the wall is of zero thickness.

This means that the tube wall has zero thermal resistance such that the temperature of the inner

surface of the wall tube, 𝑇in, coincides with that of the exterior surface, 𝑇w. In this case, it is

seen from Eq. (12) that ℎext should represent the convective heat transfer coefficient of the

water flow in the water tubes. However, the assumption of zero wall thickness is not applicable

to coal-fired boilers. In actual coal-fired boilers, the heating surfaces are, more or less, covered

by ash deposits. As illustrated in Fig. 3, the wall heat absorption from the fire side of boiler

will be conducted first through the ash deposit layer and then through the water tube wall before

being absorbed by the inside water flow.


Figure 3. Schematic of furnace wall heat transfer process

Using the circuit analogy to describe this heat transfer process, we have

𝑘ash 𝐴r𝑘tube
𝑞rad = (𝑇w ― 𝑇out) = (𝑇out ― 𝑇in) = 𝐴rℎs(𝑇in ― 𝑇ext) (14)
∆ash ∆tube

from which we can derive


―1

𝑞rad =
1
(+
∆tube
+
∆ash
𝐴rℎs 𝐴r𝑘tube 𝑘ash ) (𝑇w ― 𝑇ext) (15)

Comparing Eqs. (12) and (15), we have


―1

ℎext =( 1
+
∆tube
+
∆ash
𝐴rℎs 𝐴r𝑘tube 𝑘ash ) (16)

In Eq. (16), 𝑘ash and ∆ash are respectively the thermal conductivity and thickness of the ash

deposit layer, 𝑘tube and ∆tube are those of the water tube, ℎs is the convective heat transfer

coefficient of water flow in the water tubes, 𝐴r is the water side heat transfer area per unit plane

area, 𝑇w is now the exterior surface temperature of wall ash deposit layer, and 𝑇out and 𝑇in are

respectively the outer and inner surface temperatures of the water tube. Since furnace wall is

represented by plane surfaces in the model geometry, 𝐴r is the parameter to account for the

geometrical difference between a plane surface and actual water tube surfaces. In deriving Eq.
(16), for simplicity but without losing the essential physics, we have assumed planar tube wall

thermal conduction and clean inner surface of water tube. It is seen from Eq. (16) that ℎext

represents the (reciprocal of the) combined thermal resistance of the furnace wall heat transfer

process from the exterior surface of wall ash deposit layer to the water flow in the water tube.

If the furnace fire side heating surface is clean without ash deposits, Eq. (16) reduces to
―1

ℎext = (1
+
∆tube
𝐴rℎs 𝐴r𝑘tube ) (17)

Then, it is seen that the heat transfer coefficient of clean wall is entirely decided by the boiler

design and steam parameters. On the other hand, if the furnace wall is heavily covered with ash

deposits such that the thermal resistance of ash is much higher, then Eq. (16) reduces to

𝑘ash
ℎext = (18)
∆ash

Equation (18) states that under this condition the heat transfer coefficient, ℎext, is simply the

thermal conductance of wall ash deposits and becomes irrelevant to the boiler design and

water/steam parameters.

We now make an approximate estimation to the values of ℎext for typical coal-fired boilers. We

will use a furnace water wall made of Φ63.5 × 8 carbon steel tubes with center distance of

76.2 mm (Ar = 2). Using the typical thermal conductivity of carbon steel [33], the thermal

conductance of the water tube, 𝑘tube/∆tube, is estimated to be ~ 6 × 103 W/(m2K). The

convective heat transfer in the water tubes occurs by a complex combination of water heating

and boiling process, and the convective heat transfer coefficient, ℎs, cannot be easily estimated.

Here we used the correlation equation for nucleate boiling [33] to give an estimation (only on

the order of magnitude) of approximately 104 W/(m2K) for the subcritical boiler used in this

study. The thermal conductivity of wall ash deposit, 𝑘ash, varies widely from 0.05 to 2.4

W/(mK) depending on the temperature, composition, heating cycle and physical properties of
the deposit [33, 34]. The lower limit corresponds to friable particulate layer case at lower

temperature without sintering, while the upper limit to the completely molten case at higher

temperatures. The thermal conductance of ash deposit, 𝑘ash/∆ash, however, is less sensitive to

the boiler conditions. This is because as ash deposits grow in thickness, the thermal

conductivity of ash deposits, 𝑘ash, also increases due to the increase of wall surface temperature

and the resultant fusion of ash deposits. The net effect is that the thermal conductance, 𝑘ash/

∆ash, may only vary by a factor of four, from approximately 150 to 600 W/(m2K) for typical

coal-fired boilers, which is an order of magnitude smaller than the variations of thermal

conductivity [33, 34]. Using these numbers, Table 5 shows the estimation of ℎext under

different values of 𝑘ash/∆ash, ranging from 100 W/(m2K) for a heavily ash deposited case to

∞ for an absolutely clean wall case. The values of ℎs = 104 W/(m2K), 𝑘tube/∆tube = 6 × 103

W/(m2K) and 𝐴r = 2 are used in the estimation. To illustrate the influence of different water

wall design, the same set of estimation for 𝐴r = 1 (corresponding to a water wall with doubly

spaced tubes) is also given in Table 5. It is seen from Table 5 that under clean wall condition

(𝑘ash/∆ash→∞) the values of ℎext are on the order of 103~104 that is determined by the boiler

design (water wall) and water/steam parameters. However, even if the wall is only slightly

coated with ash deposits (e.g. 𝑘ash/∆ash~1000 W/(m2K)), the values of ℎext drop rapidly and

approach the value of 𝑘ash/∆ash. For typical boiler ash deposit conditions (𝑘ash/∆ash = 150 ~

600 W/(m2K)), it is seen that the values of ℎext are very close to the values of 𝑘ash/∆ash.

Moreover, comparing the values of ℎext for different values of 𝐴r it is seen that different furnace

wall design only makes considerable difference on the values of ℎext under clean wall

conditions, while with the growth of ash deposits this difference is rapidly diminished and ℎext

becomes more dependent on the wall ash deposit condition. This suggests that, under typical

boiler conditions, the values of the heat transfer coefficient, ℎext, are, in fact, largely determined
by the thermal conductance of wall ash deposits and only weakly dependent on the boiler

design parameters. Consequently, the typical range of thermal conductance of ash deposits,

from 150 to 600 W/(m2K), basically defines the range of hext for normally operated coal-fired

boilers.

𝑘ash/∆ash hext (Ar = 2.0) hext (Ar = 1.0)


100 98.7 97.4
200 194.8 189.9
300 288.5 277.8
400 379.7 361.4
500 468.8 441.2
600 555.6 517.2
800 722.9 659.3
1200 1034.5 909.1
∞ 7500.0 3750.0

Table 5. Values of hext (W/(m2K)) under different thermal conductance of ash deposit layer

In principle, accurate description of hext over furnace wall can be carried out by integrating the

3D CFD model with an advanced ash deposition model describing detailed coal ash

transformation, deposition and growth processes on wall surfaces. For example, Adamczyk et.

al. [35] simulated the ash growth process coupled with large-eddy simulation (LES) method to

resolve the flow field. The effects of both the properties of ash particles and the existing ash

layer on ash deposition were considered in this study. The thermal conductivity of ash deposits

was approximated using a packed-bed model [36]. This approach is clearly a very sophisticated

method in which detailed ash deposition and growth processes can be resolved by which the

impacts of different boiler design and operation options on ash layer growth can be analysed.

However, this is also going to make the problem extremely complicated and computationally

expensive, and hence, infeasible for the application of practical boiler problems. Moreover, it

is recognized that ash thickness in actual boilers in fact varies constantly due to the changes of
boiler operating conditions and the operation of sootblowers. In addition, the thermal

conductivity of ash may vary over a very large range and accurate determination of its value

needs to consider the physical properties, chemical composition, temperature, and thermal

history of the ash deposits [34, 36]. All these factors are very difficult to be accurately

incorporated into the CFD model. Consequently, determining the values of ℎext in terms of the

thickness ∆ash and thermal conductivity 𝑘ash of ash may incur dramatic complexity in the

calculation process and uncertainty in the results even though advanced ash deposition

submodels were used. The intent of this study is to develop CFD models that can be easily used

to solve large scale practical boiler problems in a timely-efficient manner. Thus, a balance

needs to be made between the level of model complexity and its applicability and efficiency.

Since for majority of the practical applications, only the total heat transfer of furnace wall is of

major concern, we can use a single value of hext to represent the averaged ash deposit status of

the entire furnace wall. In order to further define the value of hext, we first note that wall ash

deposit condition is a kind of boiler’s operating status. Thus, there should exist a

correspondence between ℎext and the boiler’s operating data based on which the value of ℎext

can be accurately determined. For an existing boiler the value of ℎext can be precisely

determined through a calibration calculation by comparing the predicted wall heat transfer

results with those estimated from the boiler’s water/steam parameters or by comparing the

predicted furnace exit gas temperatures (FEGT) with the measured values. In this way, the

influence of furnace wall ash deposit condition on heat transfer can be calibrated into the value

of ℎext so that the CFD model is precisely describing the boiler’s ash deposit status.

For the design calculation of new boilers, there are no such operating data available for the

calibration calculation. Nevertheless, we note that for normally operated boilers the severely

ash deposited areas are usually distributed at localized regions (e.g. near the burner inlets) and

majority of the wall surface can be considered relatively clean. For this reason, the value of
ℎext that is used to represent the averaged wall ash deposit status is expected to lie in the upper

part of its typical range, e.g. from 300 to 600 W/(m2K). With additional consideration on the

ash fusion properties of the design coal, the values of ℎext can be further refined to an even

narrower range. The values of ℎext determined this way have been extensively used in practical

applications, ranging from 60 MW subcritical to 1000 MW ultra-supercritical boilers, and most

of the predicted heat transfer results were in close agreement with the plant data.

Now the only wall B.C. parameter left to be determined is the wall surface emissivity, 𝜀w. Wall

ash deposits influence the heat transfer also by changing the emissivity of wall surface. The

emissivity of ash deposits depends on the temperature, ash particle size, chemical composition,

structure of the particulate layer, and whether the ash deposits are partially fused or molten [34,

36]. A friable particulate layer at lower temperature has lower emissivity because radiation is

scattered and reflected from individual particles and does not penetrate beyond a thin layer.

Below the sintering temperature the emissivity decreases when temperature increases, and then

increases sharply at higher temperatures as sintering and fusion of ash occur [34, 36]. The

emissivity of ash deposits at different wall locations, thus, may vary significantly due to

different wall temperature and ash deposit properties. Accurate description of the emissivity of

ash deposits over the furnace walls is not feasible due to the extremely complicated physical

and chemical processes involved in the ash transformation, deposition and growth processes.

Therefore, here we again assume that the heavily ash-deposited, sintered areas are distributed

only at localized regions and majority of the furnace wall are covered with friable or partially-

fused ash deposits. For pulverized coal-fired boilers, we can further assume that the sizes of

ash particles are smaller than the original coal particles. Figure 4 from Ref. [33, 34] shows the

influence of ash particle size and temperature on the surface emissivity of ash deposits. It is

seen that with these assumptions the emissivity of ash deposit typically fall within the range of

0.5 ~ 0.7 for small ash particles. In the current study, we will assume a constant 𝜀w = 0.6 to
represent the averaged influence of ash deposit on wall surface emissivity. This assumption,

however, should be examined case by case. For heavily slagged furnace walls with more areas

covered by molten ash deposit, a higher value of 𝜀w should be used.

Figure 4. Effect of coal ash size, structure and temperature on deposit emissivity (Reprinted

with permission from [34]).

4.2 Division panel and platen superheaters

Division panel superheater and platen superheater at the top of furnace are also radiative type

heating surfaces. Their heat transfer calculation can be handled in a similar way to the furnace

wall. However, since they are treated as interior double-side surfaces in the model, the

“convection” type wall thermal B.C. available in Fluent for exterior walls is not applicable to

these interior surfaces. We have known that the thermal boundary equation (12) is in fact used

to calculate the wall surface temperature, Tw. Thus, an alternative way to impose Eq. (12) as

the wall thermal B.C. is to directly assign the wall temperature calculated by Eq. (12) to these

interior surfaces by means of Fluent’s user-defined functions (UDFs). In the UDFs, the values

of 𝑞rad in Eq. (12) are obtained from the internal solution of Fluent, while the values of ℎext,

𝑇ext and 𝜀w are given directly as the input parameters. With the wall surface temperature given

this way, the wall radiation heat flux can be subsequently calculated by Eq. (11).
4.3 Convective heat exchangers

The convective heat exchangers in boiler’s convection pass usually consist of a large number

of tube panels. Because there exists dramatic scale discrepancy between the tubes of these heat

exchangers and boiler’s convection pass, it is not practical to represent the detailed tube

surfaces in the model geometry and resolve the flow and heat transfer processes across the

tubes. Simplifications, however, can be made based on the quantities of our primary interest

and the associated key physics. Since for convective heat exchangers the surface-flame

radiation heat exchange is no longer the dominant heat transfer mechanism, it is not necessary

to represent the tubes as surfaces in the model geometry. Alternatively, since for most practical

applications we are only concerned with the bulk heat transfer to each section of these heat

exchangers, they can be treated as volumes that correspond to the physical regions occupied

by these tube banks, and then the heat transfer can be handled as a bulk heat transfer process

occurring within these volumes. It is noted that as flue gas passes through these tube banks, the

large-scale swirling flows formed in the furnace will be largely inhibited and the gas flow

would become more “streamlined”. This effect can be taken into account by defining the tube

bank volumes as porous media with the inertial resistance in the transverse direction

(perpendicular to tube panels) much larger than those in the other directions. The values of the

inertial resistance can ideally be determined by matching the calculated pressure drop across

the tube banks in the CFD model with the measurement data. However, for applications

wherein no such data are available or pressure drop is not of a major concern we can simply

use reasonably-assumed values referred from similar studies, e.g. (50, 1, 1), where the larger

value is the inertial resistance in the transverse direction [12]. In this way, the impact of tube

panels on flow direction can be easily imposed without invoking extra complexity and work

load to the problem. The pressure drop across the tube banks may not be accurately calculated,
but it does not affect the calculation of overall heat transfer to these heat exchangers which is

the quantity of primary interest for most practical applications.

The heat transfer to the convective heat exchangers from the hot flue gas to the containing

steam in the tubes takes place in the sequence of gas side convective heat transfer, heat

conduction through the ash deposits and tubes, and convective heat transfer of steam flow. This

heat transfer process can be described by

𝑞tb = 𝐴cℎc(𝑇f ― 𝑇s) (19)

where 𝑞tb denotes the amount of heat transfer per unit volume, 𝐴c the tube outer surface area

per unit volume, 𝑇f the gas temperature, and 𝑇s the steam temperature, and

―1

ℎc = (
1 ∆ash ∆tube 1
+ + +
ℎf 𝑘ash 𝑘tube ℎs ) (20)

denotes the overall heat transfer coefficient wherein ℎf is the gas side convective heat transfer

coefficient. It is seen from Eq. (20) that ℎc represents the (reciprocal of the) total thermal

resistance of the heat transfer process from flue gas to steam. The thermal resistances of the

tube wall conduction and convective heat transfer of steam flow are usually much smaller, and

hence, can be neglected from Eq. (20) [36]. Then Eq. (20) reduces to

ℎc = 𝜓ℎf (21)

where we have used a thermal effectiveness factor, 𝜓, to represent the influence of ash deposits

on the overall heat transfer [37]. Now the heat transfer calculation problem is reduced to the

calculation of the gas side convective heat transfer coefficient, ℎf, and determination of

appropriate value for 𝜓.

The convective heat transfer coefficient ℎf can be estimated using the empirical correlation

equations for tube bank heat transfer problems. For example, the averaged heat transfer

coefficient hf for the tube banks in crossflow can be calculated by [38]


𝑁𝑢d = 𝐶1𝐶2Re𝑚
𝑑,maxPr
1/3
(22)

where 𝑁𝑢d = ℎf𝑑/𝑘 is the Nusselt number, Re𝑑,max = 𝜌𝑢max𝑑/𝜇 is the Reynolds number

calculated based on the maximum gas velocity 𝑢max in crossing the tube banks, and Pr = 𝜇𝑐𝑝/𝑘

is the Prantl number, d is the tube outer diameter, and 𝜌, 𝑘, 𝜇 and 𝑐𝑝 are respectively the density,

heat conductivity, dynamic viscosity and heat capacity of the flue gas. In Eq. (22), m, 𝐶1 and

𝐶2 are geometrical correction factors accounting for the effects of tube bank arrangement on

the overall heat transfer. Their values are dependent on the pitch ratios in longitudinal and

transverse directions, the number of tube rows, and tube bank alignment, such as those given

in [38]. The thermal effectiveness factor, 𝜓, accounts for the effects of surface ash deposits on

the heat transfer. For an existing boiler the values of the heat transfer coefficient ℎc can be

directly determined through a calibration calculation with the boiler’s operating data without

the need to prescribe the values for 𝜓. Usually the boiler’s control system provides detailed

steam parameters at the inlets and outlets of these heat exchangers. The values of the heat

transfer coefficient ℎc then can be fine-tuned to match the gas side heat transfer calculation

results with those estimated from the steam parameters. For new boilers without those

operating data available, appropriates values need to be assigned to the thermal effectiveness

factor 𝜓 in order to calculate the value of ℎc using Eq. (21). The recommended values of 𝜓 for

different applications are given by [37]. For pulverized coal fired boilers burning bituminous

coal, the value of 𝜓 generally can be prescribed as 0.65. The correlation equations for the tube

bank convective heat transfer coefficient ℎf and the recommended values for 𝜓 are widely

available in heat transfer text books and open literatures. These correlation equations can be

easily implemented into the CFD code by means of UDFs for each of the convective heat

exchangers. In these UDFs, the averaged inlet and outlet steam temperatures for each heat

exchanger is taken as the representative steam temperature, 𝑇s, used in Eq. (19). Since tube

banks are treated as porous media volumes in the model, these UDFs can be linked as the
energy source term for these volumes. As the flue gas flows through these volumes it loses heat

in the amount calculated by the UDFs. One of the key elements in implementing these UDFs

is that the design parameters of these heating exchangers, particularly, the heating surface area,

tube diameter and pitch ratios, should be incorporated into the formulations, e.g. into Eq. (22),

in order for the calculation to be able to account for the influence of different heat exchanger

designs on the overall heat transfer.

5. Results and discussion

5.1 Validation case

So far, we have systematically introduced the heat transfer calculation methods for all types of

boiler heating surfaces. These heat transfer submodels captured the essential physics involved

in the heat transfer process and incorporated the key boiler design and operating parameters

into the calculation while making a reasonable balance between the level of model complexity

and their applicability such that they can be easily used to solve and analyze practical boiler

problems. In this paper, they were applied to a 330 MW tangentially-fired boiler. Figure 5

shows the velocity distributions at the furnace burner and SOFA nozzle inlets. These inlet

velocity distributions incorporate the influences of burner design features and air flow

distributions and determine the aerodynamic field in the furnace. The furnace aerodynamics

strongly affects the dispersion of coal particles and their subsequent combustion processes, and

hence, also determines the overall temperature distribution in the furnace. Therefore, accurately

prescribing the burner nozzle inlet velocity distributions is a crucial step to accurately predict

the furnace flow field and temperature distributions, and thus, is also a key step for accurate

prediction of furnace heat transfer distribution.


Figure 5. Furnace burner and SOFA nozzle inlet velocity distributions

Figure 6 shows the velocity distributions on the furnace cross-corner and horizontal sectional

planes. An overall 3D furnace velocity distribution pattern can be constructed from these

sectional velocity distributions. It is seen that the air streams injected from the coal and air

burners at the furnace corners form a counter-clockwise swirling flow in the center of the

furnace. The air streams from the lower level burners have relatively deeper penetration than

those from the upper levels. This is because at lower level the overall strong swirling flow in

the furnace is not formed yet such that the air streams are less perturbed by the furnace bulk

flow. With more air flows injected into the furnace the overall swirling flow grows stronger

such that the directions of air streams from the upper level burners are deflected quickly near

the burner inlets. For this reason, the overall size of the swirling flow appears larger at higher

levels.
Figure 6. Furnace velocity distributions at different sectional planes

Figure 7 shows the temperature distributions on the furnace cross-corner and vertical middle

sectional planes (hext = 400 W/(m2K), w = 0.6). This figure presents the overall shape of the

flame in the furnace and the temperature changes over the entire boiler. The furnace

temperature is determined by the combined effects of coal combustion heat release and boiler

heat transfer. It is seen that at lower burner levels the temperature is relatively low since only

a portion of coal particles are injected into the furnace and these coal particles are still at the

early stage of their combustion process. With the progressing of coal combustion and more

coal particles injected into the furnace from higher level burners, combustion heat release

increases and this increase exceeds the amount of heat transfer to the furnace wall. As a result,

furnace gas temperature increases quickly from the bottom burner level to its max level in the

region near the top level burners. Downstream of this position, boiler heat transfer exceeds the

combustion heat release, and the temperature gradually reduces to ~ 989 ºC at the entrance of

boiler’s convection pass, and to ~ 357 ºC at the economizer outlet.


(ºC) Cross-corner plane Vertical middle plane

Figure 7. Furnace temperature distributions on cross-corner and vertical middle sectional

planes (hext = 400 W/(m2K), w = 0.6)

The temperature distribution in the furnace results from the combined effects of furnace

aerodynamics, coal combustion, and boiler heat transfer. It subsequently determines the

furnace wall radiation heat flux distribution which is obtained by solving the radiative transfer

equation (RTE) using the discrete ordinate method (DOM). Figures 8 and 9 show the

distributions of furnace wall heat flux and surface temperature, respectively. It is seen that they

exhibit very similar distribution patterns, and that the locations of the highest wall surface

temperature always coincide with those of highest wall heat flux. This trend can be clearly

illustrated by Eq. (12), which can be rewritten as

𝑇w = 𝑇ext + 𝑞rad/ℎext (23)

It is seen that the wall surface temperature 𝑇w is proportional to the wall heat flux, and inversely

proportional to the heat transfer coefficient ℎext. In the present case, we used a single ℎext to

represent the averaged influence of wall ash deposits on heat transfer and a constant water

saturation temperature 𝑇ext (~366 ºC). Thus, the wall surface temperature, Tw, should have the

same distribution pattern as that of wall surface heat flux.


(kW/m2) Left Front Right Rear

Figure 8. Furnace wall heat flux distribution (hext = 400 W/(m2K), w = 0.6)

(ºC) Left Front Right Rear

Figure 9. Furnace wall surface temperature distribution (hext = 400 W/(m2K), w = 0.6)

Further comparing the furnace wall heat flux distributions (Fig. 8) with the furnace gas

temperature distributions (Fig. 7), it is seen that the wall heat flux distributions along furnace

height closely follow the trend of furnace gas temperature distributions and that both of their

highest values are located near the top level coal burners. This trend can be verified in Fig. 10
that shows the profiles of the averaged wall heat flux and furnace gas temperature along furnace

height. It is seen that the gas temperature reaches its maximum value at the region between the

top level coal burners and bottom level SOFA nozzles, and that the wall heat flux distribution

maxes at exactly the same furnace height. This is because in the lower furnace where

temperature is high, wall surface heat absorption is predominantly by radiation that is

proportional to the fourth power of temperature. Therefore, the furnace wall heat flux

distribution is overall determined by the furnace gas temperature distribution such that their

peak values coincide at the same location.

Figure 10. Averaged furnace wall heat flux and gas temperature profiles along furnace height

(hext = 400 W/(m2K), w = 0.6)

Figure 11 shows the furnace wall heat flux distributions under different values of hext (300,

400, 600, 800, 1200 W/(m2K)). Here hext = 300 W/(m2K) represents a dirty (heavily ash-

deposited) wall case, 600 W/(m2K) a relatively clean wall case, and 1200 W/(m2K) an

extremely clean case. It is seen that all wall heat fluxes exhibit similar distribution pattern since

all cases have the same furnace aerodynamics and temperature distribution pattern. However,
the wall heat fluxes increase with the increase of hext especially in the lower furnace region

where temperature is high and radiation is intense. As we have discussed, for pulverized coal-

fired boilers the value of hext represents the thermal conductance of furnace wall ash deposits.

In this sense, it characterizes the cleanliness of furnace wall. Figure 11 clearly demonstrates

that wall cleanliness strongly influence the wall heat absorption, especially in the lower furnace

region where gas temperature is high.

(kW/m2) hext = 300 400 600 800 1200

Figure 11. Furnace wall heat flux distributions (right wall) under different values of hext (300,

400, 600, 800, 1200 W/(m2K)).

Figure 12 shows the furnace wall (fire side) surface temperature distributions at different values

of hext (300, 400, 600, 800, 1200 W/(m2K)). It is seen that with the decrease of hext, i.e. with

the growth of wall ash deposits, furnace wall surface temperature increases significantly.

Consequently, the self- radiation of furnace wall (~𝜎𝑇4w) increases, leading to reduced net heat

absorption of furnace wall, as shown in Fig. 11. For an extremely clean case with very little

wall ash deposits, the value of hext is large (~O(103)), as seen in Table 4. Then it is seen from

Fig. 12 and Eq. (23) that the wall surface temperature Tw would be approaching the water

saturation temperature, Text (~366 ºC). At this low temperature, the self-radiation of wall only
contributes a very small portion to the heat transfer of furnace wall. However, for heavily ash-

deposited case (very small hext), the wall surface temperature is going to increase to a level at

which the self-radiation (~𝜀w𝜎𝑇4w) balances the radiation absorption (𝜀w𝐺) of furnace wall such

that the net heat absorption of furnace wall approaches zero. The variations of wall surface

temperature with wall ash deposit status is the key mechanism by which wall ash deposits affect

the heat absorption of furnace wall.

(ºC) hext = 300 400 600 800 1200

Figure 12. Furnace wall temperature distributions (right wall) at different values of hext (300,

400, 600, 800, 1200 W/(m2K)).

This mechanism can be further corroborated by Fig. 13 from Ref. [39] that shows a visual

observation of the wall ash deposits in a pulverized coal boiler and the corresponding IR

photography of wall surface temperature. It is seen that wall surface temperature distribution

is highly non-uniform and is strongly dependent on the local wall ash deposit condition. The

surface temperature is above 700 ºC over the entire observed wall area, and on severely ash-

deposited areas it can be as high as 1400 ºC. This is significantly higher than the water

saturation temperature (~366 ºC). With wall surface temperature at this level, the self-radiation

of furnace wall can contribute a significant portion to the heat transfer of furnace wall. Figure
13 is a clear demonstration how ash deposits affect the heat transfer through their influence on

the exterior surface temperature of furnace wall. Apparently, the constant wall surface

temperature boundary condition employed by many of the previous boiler numerical studies

[3, 6, 11, 19] cannot correctly account this mechanism and will significantly overestimate the

heat absorption of furnace wall if the water saturation temperature (~366 ºC) was used as the

wall surface temperature.

Visual observation IR photography (ºC)

Figure 13. Visual observation and IR photography of furnace wall ash deposits [39]

Figure 14 shows the variations of furnace wall heat absorption with the heat transfer coefficient

hext. The gray area in this figure marks the range of heat transfer results confined by hext between

300 and 600 W/(m2K) and 𝜀w between 0.6 and 0.7. It is seen that as hext increases from smaller

values (e.g. 100 W/(m2K)) wall heat transfer quickly increases, but as hext increases above 800

W/(m2K) it becomes relatively insensitive to further increase of hext. This is because at larger

values of hext furnace wall is clean and wall surface temperature is low at which the self-

radiation of furnace wall is relatively insensitive to further changes of wall surface temperature.

Thus, further increasing hext poses little additional effect on the furnace wall heat transfer. As

indicated by Eq. (23), at very large values of hext, the wall temperature would be approaching

the water saturation temperature, Text. Then, it can be seen from Eq. (11) that the maximum
amount of furnace wall heat absorption is 𝜀w(𝐺 ― 𝜎𝑇4ext). This is the amount the curves in Fig.

14 would be approaching at very large values of hext.

Figure 14. Variations of wall heat absorption with hext.

Figure 15 shows the variations of furnace exit gas temperature (FEGT) with the heat transfer

coefficient hext. It is seen that FEGT decreases with the increase of hext due to the increased

wall heat absorption, and that as hext increases above 800 W/(m2K) the FEGT becomes

relatively insensitive to further increase of hext. This is consistent with the wall heat absorption

results shown in Fig. 14.


Figure 15. Variations of furnace exit gas temperature with hext.

Also shown in Figs. 14 and 15 include the results for 𝜀w = 0.7. The wall thermal B.C. parameter

𝜀w characterizes the emissivity of wall ash deposits. Its value depends on the temperature, ash

particle size, chemical composition, and structure of the particulate layer. It is extremely

difficult to give an accurate estimation on the distribution of 𝜀w over the furnace wall surface

in a practical boiler. As shown in Fig. 4, the values of 𝜀w usually fall within the range of 0.5 ~

0.7 for pulverized coal boilers. Experience shows that the values of 𝜀w between 0.6 and 0.7

generally give results in close agreement with the operating data for most of the practical

applications. Additionally, as we have discussed that the averaged hext usually lie in the range

of 300 ~ 600 W/(m2K) for normally operated pulverized coal boilers. Thus, the predicted

results of wall heat transfer and FEGT usually should fall within the range delimited by hext

between 300 and 600 W/(m2K) and 𝜀w between 0.6 and 0.7, as marked by the gray areas in

Figs. 14 and 15. It is seen that the predicted results of furnace wall heat transfer and FEGT are

already confined within a rather small range, and their variations within this range are due to

the variations in wall ash deposit status. With additional information, e.g., consideration of the

coal ash fusion properties or onsite observation of the boiler’s actual wall ash deposit status,
the predicted results can be further narrowed down to an even smaller range. We note that in

coal-fired boilers furnace wall heat transfer status is actually constantly changing with the

furnace operation status and wall ash deposit conditions. For example, if the boiler is running

at high loads for consecutive days, the wall ash deposits may gradually accumulate leading to

continuously reduced wall heat transfer and increased FEGT. On the other hand, the FEGT

may start from a relatively low level if the boiler, before raising to high loads, had been running

at low loads for a period of time during which a significant amount of wall ash deposits have

cracked and shed at lower temperatures. Operation of wall sootblowers will also have

significant impact on the wall ash deposit status, and hence, will affect furnace wall heat

transfer. For these reasons, the predicted results of wall heat transfer and FEGT within this

narrow range can be considered sufficiently accurate to solve and analyze most of the practical

boiler problems.

Figure 16 shows the variations of the self-radiation 𝑞w (~𝜀w𝜎𝑇4w) and absorbed incident

radiation 𝑞ab (~𝜀w𝐺) of furnace wall with the heat transfer coefficient hext. First, we note that

both 𝑞w and 𝑞ab increase with the decrease of hext (i.e., with the growth of wall ash deposits)

while the net heat transfer of furnace wall shows the opposite trend (as seen in Fig. 14). This

is because with the decrease of hext the gas temperature in the lower furnace increases as the

result of reduced heat absorption of furnace wall. Since the incident radiation G is proportional

to the fourth power of furnace gas temperature, increased gas temperature leads to significant

increase of wall incident radiation G and absorbed radiation (~𝜀w𝐺) as well. However, since

the wall surface temperature, 𝑇w, also increases with the decrease of hext, the self-radiation of

furnace wall increases dramatically at the same time. As seen in Fig. 16, the increase of the

self-radiation of furnace wall, 𝑞w, exceeds that of absorbed radiation, 𝑞ab. Consequently, the

net heat absorption of furnace wall decreases with the decrease of hext, i.e., with the growth of

wall ash deposits. Furthermore, it is seen from Figs. 14-16 that the influences of hext on the heat
transfer of furnace wall is only considerable at lower values of hext. As hext is above 800

W/(m2K) these heat transfer curves become insensitive to further increase of hext. As we have

discussed, this is because at higher hext the furnace wall temperature is reduced, and so radiation

becomes less sensitive to further increase of hext.

Figure 16. Variations of wall self-radiation and absorbed incident radiation with hext.

Figures 17 shows the heat flux and surface temperature distributions of division panel

superheater (DPSH). The heat flux and temperature distributions of the platen superheater

exhibit similar trend as the division panel, and hence, will not be shown here. It is seen that

the heat flux and temperature distributions on the DPSH surface are highly non-uniform. The

highest heat flux occurs at the lower region of DPSH that are in contact with the high

temperature gas flow and directly face the high temperature flame of lower furnace. The high

heat flux consequently leads to high surface temperature at the corresponding areas. In contrast,

at the upper region of DPSH, since the flue gas temperature there is lower and much of the

strong thermal radiation from the high temperature flame has been sheltered by the lower part

of DPSH, the heat flux, and consequently, the surface temperature, are substantially lower than

those at the lower region of DPSH. Moreover, it is noted that the surface temperature at the
high heat flux regions is much higher than that of the containing steam (~ 430 ºC) in the division

panel tubes. Many of the previous boiler CFD studies used the steam temperature in DPSH as

the wall surface temperature. Apparently, this is not going to correctly describe the surface

temperature distribution of DPSH, and will significantly underestimate the self-radiation, and

hence, overestimate the total heat absorption of the division panel superheater.

(kW/m2) (a) Heat flux distribution (ºC) (b) Surface temperature

Figure 17. Heat flux and surface temperature distributions of division panel superheater.

Figure 18 shows the predicted results of heat transfer to different parts of boiler heating surfaces

under different values of hext (300, 400, and 600 W/(m2K)) and their comparisons with the

operating data. The operating data in Fig. 18 was estimated from the boiler’s feed water,

superheat and reheat steam parameters, as given in Table 3. As we have discussed, the predicted

heat transfer results between hext = 300 and 600 W/(m2K) generally define the range of

variation of wall heat transfer for normally operated coal-fired boilers. hext = 400 W/(m2K) is

the value initially used in the calculation. This value was selected based on a combined

consideration of the boiler’s actual wall ash deposit status and previous experience of boiler

CFD studies. It is seen that using the heat transfer submodels introduced in this paper, the

calculated results are in close agreement with the operating data. It is recognized that the heat

transfer status of actual boiler system in fact changes constantly with the boiler’s operating
conditions and ash deposit status. Thus, exact agreement of the numerical results with a single

set of operating data is usually not necessary and agreement at this level can be considered

sufficiently accurate to solve and analyze practical boiler problems.

Figure 18. Comparison of boiler heat transfer results with operating data (𝜀w = 0.6)

Next we applied the heat transfer submodels introduced in this paper to investigate the effects

of different boiler operating conditions on boiler heat transfer distributions, such as burner tilt

angle adjustment and coal switching. Two additional cases are considered. In case #2 all the

coal and auxiliary air nozzles (except for the SOFA nozzles) are tilt down by 20º, and in case

#3 the coal is switched to another bituminous coal with lower heating value. All the other key

parameters are kept the same including total heat input, excess air ratio and burner air flow

distribution. Additionally, we have assumed that all three cases have similar furnace wall ash

deposit status. This means that all cases used the same wall thermal B.C. parameters (hext = 400

W/(m2K), w = 0.6).

Figure 19 shows the temperature distributions on the furnace cross-corner sectional plane for

all three cases. The impacts of burner tilt and coal switching on furnace temperature

distributions can be clearly seen from these figures. It is seen that burner tilt down (case #2)

drives more hot gas to the lower region of furnace. As a result, coal combustion heat release is
distributed over a wider region of furnace space. Consequently, the overall flame temperature

of case #2 appears lower than that of case #1. The temperature distribution of case #3 shows

that switching to lower heating value coal reduces the overall furnace gas temperature level

although total heat inputs are the same. Since radiation heat transfer is proportional to the fourth

power of gas temperature, the heat transfer distributions of furnace wall are going to be affected

by these different gas temperature distributions.

(ºC) Case #1 Case #2 Case #3

Figure 19. Furnace temperature distributions (hext = 400 W/(m2K), w = 0.6).

Figure 20 shows the furnace wall heat flux distributions for all three cases (right wall only). It

is seen that the wall heat flux distributions closely follow the trend of furnace gas temperature

distributions. Compared to case #1, the high heat flux regions of case #2 are distributed over a

larger furnace wall area. In particular, the heat flux in the region between the lower burners

and ash hopper is larger than that of case #1. For both cases the peak heat flux occurs in the

region between the top level coal burners and SOFA nozzles. But due to the reduced flame

temperature, the maximum heat flux of case #2 is smaller than that of case #1. Thus, burner tilt

down causes the heat flux to distribute more evenly on the furnace wall. Compared to case #1,

case #3 shows very similar wall heat flux distribution pattern since both cases have the same
burner settings. However, due to reduced flame temperature the wall heat flux of case #3 is

overall lower than that of case #1.

(kW/m2) Case #1 Case #2 Case #3

Figure 20. Furnace wall heat flux distributions (hext = 400 W/(m2K), w = 0.6)

Figure 21 shows the comparison of the boiler heat transfer results for all three cases. It is seen

that burner tilt down increases the total furnace wall heat transfer. Consequently, less thermal

energy is carried by flue gas to the downstream heating surfaces such that the heat transfer to

the superheaters and reheaters are reduced. Switching to a lower heating value coal shows the

opposite trend. The heat transfer to the economizer is relatively insensitive to the changes in

the wall heat transfer conditions. This is because economizer is located at the very downstream

of all boiler heating surfaces. Thus, most of the decreased (increased) heat transfer to the

furnace wall will be compensated by the increased (decreased) heat transfer to the superheaters

and reheaters upstream the economizer.


Figure 21. Comparison of boiler heat transfer results of different cases

6. Further discussion

This paper was primarily focused on the heat transfer calculation method of furnace water wall.

It was implemented into the CFD model through a thermal B.C. with the associated B.C.

parameters determined based on the thermal properties of wall ash deposits. However, it should

not be misunderstood that furnace wall heat transfer is entirely determined by the wall thermal

boundary condition. As we have discussed in this paper, the wall thermal B.C., in fact, only

defines the way to calculate the wall surface temperature, and thereby, to calculate the net heat

absorption of the furnace wall. Radiation heat transfer in coal-fired boilers is a highly complex

process, and is affected by a variety of interacting processes, such as coal properties, boiler’s

burner and heating surface arrangements, and furnace aerodynamics. For a given boiler design

and fuel type, furnace aerodynamics plays the predominant role in determining the furnace wall

heat transfer distribution since it largely determines the mixing of coal particles and the

subsequent ignition and combustion processes in the furnace, and consequently, determines the

furnace temperature distribution. Accurate calculation of furnace wall heat transfer, first of all,

needs accurate prediction of the furnace aerodynamics and proper incorporation of boiler
burner design features and air flow distributions into the calculation. This has been discussed

in section 3.1 of the present paper and the corresponding results are shown in Figs. 5 ~ 7.

Calculation of furnace wall heat transfer is also affected by the simulation of coal combustion

processes since it determines the rate and location of combustion heat release, and consequently,

affects the furnace temperature distribution. There are a variety of submodels to describe the

key elemental processes of coal combustion, such as coal devolatilization and char surface

combustion, with each having different level of model details and complexity. For the boiler

CFD model to serve as an efficient analytical tool to solve and analyze practical boiler problems,

these submodels should be able to capture the key physics involved in these processes, and at

the same time, need to be sufficiently simple and straightforward. The level of details that need

to be considered in the CFD model is firstly determined by the quantities to be obtained from

the CFD solution and the associated key physics. It is to be noted that use of sophisticated

model for an individual process does not ensure improved accuracy of the results since the

accuracy is also limited by the other submodels. Moreover, advanced models generally need

more model parameters that are usually difficult to be accurately determined in complex

practical systems. This further limits the accuracy of these advanced models. Therefore, for a

CFD model that is to be used to as an efficient analytical tool, all the submodels need to be

constructed in a balanced manner with each of them having the proper level of complexity.

Additionally, modern boiler systems are usually huge and contain many irregular and fine

structures with significantly distinct scales. What we can handle in a CFD model is usually a

simplified geometrical model with many details approximated or neglected from the real

system. How to treat these geometry details needs to consider both the key physics associated

with the problems of our interests and the limitation of our computational capability. This is

the reason why we can treat the convective heat exchangers as porous media volumes, but need

to treat the platen superheater as interior surfaces in the model geometry. Because of the scale
and complexity of practical boiler problems, the implemented models can only be seen as the

best possible approximation to meet our needs to solve and analyze specific boiler problems

within current computational capability. Aside from these model details, how accurately and

reliably a boiler CFD model can be used to solve and analyze practical problems also strongly

depends on if the CFD model is able to accurately describe the real system, i.e., if the boiler

design parameters, operating data, and coal properties can be appropriately translated into the

associated CFD model parameters, geometrical structures, and boundary conditions etc. This

is, in fact, one of the most critical steps for the CFD model to be used as a practical tool to

solve the real boiler problems and investigate the influence of different boiler design and

operation parameters on boiler performance.

7. Conclusions

The heat transfer calculation methods in three-dimensional CFD model for coal-fired utility

boilers were systematically introduced in this paper. Both the radiative and convective types of

heating surfaces were considered, including furnace water wall, division panel, superheater and

economizer. Particular attention was focused on the heat transfer calculation of furnace water

wall. It was implemented in the CFD model in terms of a thermal B.C. describing the energy

balance between the furnace fire side heat transfer and water side heat absorption. The physical

significance of the associated B.C. parameters, namely a heat transfer coefficient, hext, a wall

surface emissivity, 𝜺w, and an external heat sink temperature Text, and their accurate

prescription were discussed in detail in this paper. It was found that the heat transfer of furnace

wall is largely affected by the conditions of ash deposits formed on the furnace wall. The ash

deposits restrict the absorption of the incident radiation by forming an insulating layer on the

tube surface and by changing the radiation properties of wall surface. As such, the values of

hext and 𝜺w can be determined by establishing their connection with the thermal properties of
wall ash deposits. It is found that the heat transfer coefficient, hext, is largely determined by the

thermal conductance of wall ash deposit layer 𝒌ash/∆ash since ash constitutes most of the

thermal resistance of the heat transfer process, while the surface emissivity, 𝜺w, represents the

emissivity of the ash layer. These thermal properties of boiler ash deposits, however, have been

the most difficult operating variables to quantify in real boiler systems [34, 36]. In particular,

depending on the temperature, ash composition, heating cycle and physical properties of ash

deposits, the thermal conductivity of ash, 𝒌ash, may vary widely by a factor of 50 [33, 34], and

the ash thickness, ∆ash, is also extremely difficult to predict. The thermal conductance, 𝒌ash/

∆ash, however, is much less sensitive to the boiler conditions due to the concurrent increase of

𝒌ash and ∆ash with the growth of wall ash deposits. If a single value of hext is used to represent

the averaged thermal conductance of wall ash deposits, the range of variation of 𝒌ash/∆ash can

be narrowed down to a rather small range, from approximately 300 to 600 W/(m2K), for

normally operated coal-fired boilers. Establishing the connection between this particular

thermal property of wall ash deposits and the heat transfer coefficient of furnace wall is the

most crucial finding of this paper since it bypasses the tremendous difficulty and uncertainty

incurred in determining the physical and chemical properties of ash deposits, and thereby,

confines the prediction of furnace wall heat transfer within a reasonable range. It also makes

the heat transfer calculation a very simple and straightforward process while still capturing the

key physics involved in the furnace wall heat transfer process.

The heat transfer calculation of other types of boiler heating surfaces were also considered in

this paper, including division panel, superheaters, reheaters and economizers. The heat transfer

to the division panel is also radiative in nature, and hence, can be handled in a similar way as

the furnace wall. However, since they are treated as interior double-side surfaces in the model,

the type of boundary condition available in Fluent for exterior walls is not applicable to these

surfaces. Alternatively, the heat transfer of these heating surfaces can be calculated by the
Fluent’s UDFs. As for the convective heating surfaces, such as superheater and economizer,

the heating surfaces are treated as porous media volumes with large inertial resistance in the

transverse direction to handle the dramatic scale discrepancy between the tubes and boiler’s

convection pass and to impose the influence of tube panels on flow direction. The heat transfer

is then calculated as the energy sink of these porous volumes using the heat transfer correlation

equations by means of Fluent’s UDFs.

With the heat transfer submodels developed in this paper, the 3D CFD model is able to

accurately predict the heat transfer distribution in coal-fired boilers and incorporate the key

boiler design and operating parameters into the calculation in a very efficient way. Application

of the heat transfer models to a 330 MW T-fired subcritical boiler shows that the predicted

boiler heat transfer distributions are in close agreement with the boiler’s operating data. The

results show that with the decrease of hext (i.e. with the growth of wall ash deposits) furnace

wall surface temperature increases and this increase becomes greater at smaller values of hext

(< 300 W/(m2K)). Since thermal radiation is more sensitive to temperature changes at higher

temperature, this increase in wall surface temperature leads to dramatic increase of the self-

radiation, and consequently, decrease of the net heat absorption of furnace wall. At higher

values of hext (> 800 W/(m2K)), the wall surface temperature is relatively low at which

radiation is less sensitive to the temperature changes. Thus, the heat absorption of furnace wall

appears insensitive to further increase of hext. Variations of wall surface temperature with the

ash deposit status is the key mechanism by which the ash deposits affect the heat transfer of

furnace wall. Many of the previous boiler CFD studies used constant wall surface temperature

as the thermal boundary condition. Apparently, this will not correctly account this key

mechanism and will significantly overestimate the net heat absorption of furnace wall if water

saturation temperature was used as the wall surface temperature.

The heat transfer submodels introduced in this study embody the key physics involved in the
heat transfer processes of different types of boiler heating surfaces while making a reasonable

balance between the level of model complexity and their applicability to large scale practical

problems. With these heat transfer submodels, the 3D CFD model is now able to accurately

predict the heat transfer distribution in coal-fired utility boilers and incorporate the key boiler

design and operating parameters into the calculation. As such, it can analyze and solve a large

variety of practical problems, such as coal switching, gas conversion, and boiler overfire

system modification, wherein variations of boiler heat transfer distributions are of the major

concerns. Additionally, since these heat transfer submodels calculate the heat transfer using the

local flow and temperature information from the 3D CFD solution, they can predict the impacts

of flow and temperature non-uniformities on the heat flux and temperature distributions on

boiler heating surfaces. Therefore, these models can be further utilized to handle many related

boiler problems such as tube surface overheating and high temperature corrosion.

Acknowledgement

This work is supported by Tianjin University’s Double First-Rate Development Project (010-

0701020313). We would like to thank Yantai Longyuan Power Technology Co. Ltd. for

provision of boiler data and helpful comments.

References

[1] F.C. Lockwood, A.P. Salooja, S. S.A., A prediction method for coal-fired furnaces,

Combust Flame 38 (1980) 1-15.

[2] J.S. Truelov, The modelling of flow and combustion in swirled, pulverized-coal burners,

P Combust Inst 20 (1984) 523-530.

[3] R.K. Boyd, J.H. Kent, Three-dimensional furnace computer modelling, P Combust Inst 21

(1986) 265-274.
[4] A. Williams, M. Pourkashanian, P. Bysh, J. Norman, Modelling of coal combustion in

low-NOx p.f. flames, Fuel 73(7) (1994) 1006-1019.

[5] S.V. Belosevic, I.D. Tomanovic, N.D. Crnomarkovic, A.R. Milicevic, D.R. Tucakovic,

Modeling and Optimization of Processes for Clean and Efficient Pulverized Coal Combustion

in Utility Boilers, Therm Sci 20 (2016) S183-S196.

[6] L.I. Diez, C. Cortes, J. Pallares, Numerical investigation of NOx emissions from a

tangentially-fired utility boiler under conventional and overfire air operation, Fuel 87(7)

(2008) 1259-1269.

[7] T.L. Bris, F. Cadavid, S. Caillat, S. Pietrzyk, J. Blondin, B. Baudoin, Coal combustion

modelling of large power plant, for NOx abatement, Fuel 86 (2007) 2213-2220.

[8] R. Kurose, H. Makino, A. Suzuki, Numerical analysis of pulverized coal combustion

characteristics using advanced low-NOx burner, Fuel 83(6) (2004) 693-703.

[9] R.I. Backreedy, J.M. Jones, L. Ma, M. Pourkashanian, A. Williams, A. Arenillas, B.

Arias, J.J. Pis, F. Rubiera, Prediction of unburned carbon and NOx in a tangentially fired

power station using single coals and blends, Fuel 84(17) (2005) 2196-2203.

[10] M. Kuang, Z.Q. Li, S.T. Xu, Q.Y. Zhu, Improving Combustion Characteristics and NOx

Emissions of a Down-Fired 350 MWe Utility Boiler with Multiple Injection and Multiple

Staging, Environmental Science & Technology 45(8) (2011) 3803-3811.

[11] P. Tan, D.F. Tian, Q.Y. Fang, L. Ma, C. Zhang, G. Chen, L.J. Zhong, H.G. Zhang,

Effects of burner tilt angle on the combustion and NOx emission characteristics of a 700

MWe deep-air-staged tangentially pulverized-coal-fired boiler, Fuel 196 (2017) 314-324.

[12] C. Yin, S. Caillat, J.L. Harion, B. Baudoin, E. Perez, Investigation of the flow,

combustion, heat-transfer and emissions from a 609 MW utility tangentially fired pulverized-

coal boiler, Fuel 81(8) (2002) 997-1006.


[13] H. Zhou, G.Y. Mo, D.B. Si, K.F. Cen, Numerical Simulation of the NOx Emissions in a

1000 MW Tangentially Fired Pulverized-Coal Boiler: Influence of the Multi-group

Arrangement of the Separated over Fire Air, Energ Fuel 25(5) (2011) 2004-2012.

[14] J.Y. Chen, A.P. Mann, J.H. Kent, Computational modelling of pulverized fuel burnout in

tangentially fired furnaces, P Combust Inst 24 (1992) 1381-1389.

[15] J. Li, R. Jankowski, M. Kotecki, W.H. Yang, D. Szewczyk, A. Brzdekiewicz, W.

Blasiak, CFD Approach for Unburned Carbon Reduction in Pulverized Coal Boilers, Energ

Fuel 26(2) (2012) 926-937.

[16] J. Pallares, L. Arauzo, L.I. Diez, Numerical prediction of unburned carbon levels in large

pulverized coal utility boilers, Fuel 84(18) (2005) 2364-2371.

[17] M.B. Gandhi, R. Vuthaluru, H. Vuthaluru, D. French, K. Shah, CFD based prediction of

erosion rate in large scale wall-fired boiler, Appl Therm Eng 42 (2012) 90-100.

[18] N. Spitz, R. Saveliev, M. Perelman, E. Korytni, B. Chudnovsky, A. Talanker, E. Bar-

Ziv, Firing a sub-bituminous coal in pulverized coal boilers configured for bituminous coals,

Fuel 87(8-9) (2008) 1534-1542.

[19] T. Asotani, T. Yamashita, H. Tominaga, Y. Uesugi, Y. Itaya, S. Mori, Prediction of

ignition behavior in a tangentially fired pulverized coal boiler using CFD, Fuel 87(4-5)

(2008) 482-490.

[20] W.P. Adamczyk, S. Werle, A. Ryfa, Application of the computational method for

predicting NOx reduction within large scale coal-fired boiler, Appl Therm Eng 73(1) (2014)

343-350.

[21] R. Weber, M. Mancini, N. Schaffel-Mancini, T. Kupka, On predicting the ash behaviour

using Computational Fluid Dynamics, Fuel Process Technol 105 (2013) 113-128.
[22] X. Yang, D. Ingham, L. Ma, H. Zhou, M. Pourkashanian, Understanding the ash

deposition formation in Zhundong lignite combustion through dynamic CFD modelling

analysis, Fuel 194 (2017) 533-543.

[23] C. Schuhbauer, M. Angerer, H. Spliethoff, F. Kluger, H. Tschaffon, Coupled simulation

of a tangentially hard coal fired 700 degrees C boiler, Fuel 122 (2014) 149-163.

[24] T.H. Shih, W.W. Liou, A. Shabbir, Z. Yang, J. Zhu, A new k-e eddy viscosity model for

high Reynolds number turbulent flows, Computers & Fluids 24(3) (1995) 227-238.

[25] S. Park, J.A. Kim, C. Ryu, T. Chae, W. Yang, Y.J. Kim, H.Y. Park, H.C. Lim,

Combustion and heat transfer characteristics of oxy-coal combustion in a 100 MWe front-

wall-fired furnace, Fuel 106 (2013) 718-729.

[26] S. Badzioch, P.G.W. Hawksley, Kinetics of thermal decomposition of pulverized coal

particles, Industrial & Engineering Chemistry Process Design and Development 9 (1970)

521-530.

[27] S. Niksa, Predicting the devolatilization behavior of any coal from its ultimate analysis,

Combust Flame 100 (1995) 384-394.

[28] R.I. Backreedy, R. Habib, J.M. Jones, M. Pourkashanian, A. Williams, An extended coal

combustion model, Fuel 78(14) (1999) 1745-1754.

[29] A. Williams, R. Backreedy, R. Habib, J.M. Jones, M. Pourkashanian, Modelling coal

combustion: the current position, Fuel 81(5) (2002) 605-618.

[30] M.M. Baum, P.J. Street, Predicting the combustion behaviour of coal particles, Combust

Sci Technol 1 (1971) 231-243.

[31] R.E. Mitchell, R.H. Hurt, L.L. Baxter, D.R. Hardesty, Compilation of Sandia coal char

combustion data and kinetic analyses, milestone report. SAND 92-8208, 1992.

[32] B.F. Magnussen, B.H. Hjertager, On mathematical modelling of turbulent combustion

with special emphasis on soot formation and combustion, P Combust Inst 16 (1977) 719-729.
[33] Steam/its generation and use, 41 ed., The Babcock & Wilcox Company, Barberton,

Ohio, U.S.A., 2005.

[34] B.S.P. Wall T F, Zhang D K, Gupta R P, He X, The properties and thermal effects of ash

deposits in coal-fired furnaces, Progress in Energy and Combustion Science 19 (1993) 18.

[35] W.P. Adamczyk, B. Isaac, J. Parra-Alvarez, S.T. Smith, D. Harris, J.N. Thornock, M.M.

Zhou, P.J. Smith, R. Zmuda, Application of LES-CFD for predicting pulverized-coal working

conditions after installation of NOx control system, Energy 160 (2018) 693-709.

[36] A. Zbogar, F.J. Frandsen, P.A. Jensen, P. Glarborg, Heat transfer in ash deposits: A

modelling tool-box, Progress in Energy and Combustion Science 31(5-6) (2005) 371-421.

[37] D.F. Che, Boiler - theory, design and operation, Xi'an Jiaotong University Press, Xi'an,

China, 2008.

[38] F.P. Incropera, D.P. DeWitt, Fundamentals of heat and mass transfer, 4 ed., John Wiley

& Sons, New York, 1996.

[39] M.D. Tubio, R. Higginbotham, Slag control treatment at a southeastern utility, Water

Technologies & Solutions technical paper, 2017.


Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal
relationships that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be
considered as potential competing interests:

Heyang Wang

Chaoquan Zhang

Xin Liu

10/28/2019

You might also like