You are on page 1of 10

Applied Thermal Engineering 87 (2015) 175e184

Contents lists available at ScienceDirect

Applied Thermal Engineering


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

Research Paper

Influence of critical viscosity and its temperature on the slag behavior


on the wall of an entrained coal gasifier
Insoo Ye a, Changkook Ryu a, *, Ja Hyung Koo b
a
School of Mechanical Engineering, Sungkyunkwan University, Suwon 440-746, Republic of Korea
b
Coal Conversion System Development Team, Corporate R&D Institute, Doosan Heavy Industries & Construction, Yongin 448-795, Republic of Korea

h i g h l i g h t s

 Effects of critical viscosity temperature on slag behavior were numerically studied.


 Higher critical viscosity temperatures exponentially increased solid slag thickness.
 Higher critical viscosity gradually increased both liquid and solid thicknesses.
 Effect of flux supply was governed by the resultant critical viscosity temperature.
 Such trends were not changed by the wall location and angle, and gas temperature.

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

Article history: In entrained coal gasification, the properties of slag and its behavior on the wall are crucial in selecting
Received 22 December 2014 specific coal types and controlling the gasifier operation. The temperature of critical viscosity (Tcv) is the
Accepted 12 May 2015 key parameter that characterizes the slag flowability. This study evaluates the theoretical influences of
Available online 21 May 2015
Tcv, critical viscosity and flux addition on the slag behavior in a commercial coal gasifier using a nu-
merical model. The results showed that a lower slag Tcv value led to an exponential decrease in the solid
Keywords:
slag thickness (dS) owing to an increased heat flux to the wall and a reduced temperature difference
Coal gasification
between the slag and the coolant. In contrast, the influence of Tcv on the liquid slag thickness (dL) was
Critical viscosity temperature
Entrained gasifier
relatively small. When Tcv remained constant, a higher critical viscosity value decreased the velocity of
Flux the liquid slag close to the solid slag, leading to a gradual increase in both dS and dL. Addition of flux was
Slag effective in decreasing dS for coals with high Tcv values, while the impact of increased slag flow rates on dL
was minimal and could be disregarded. The relative magnitude of total slag thickness by changes in Tcv
and critical viscosity was almost the same for different ash:flux ratios, gas temperatures, and wall angles.
The influence of Tcv was dominant over that of critical viscosity for slags with critical viscosity values
above 10 Pa s.
© 2015 Elsevier Ltd. All rights reserved.

1. Introduction slag that becomes stagnant because of increased viscosities and


crystalline phase formation at low temperatures. The liquid slag
During entrained coal gasification, coal is converted to syngas layer has lower viscosities at high temperatures and flows down-
that is rich in H2 and CO, and the ash is molten into slag at ward by gravity. It is then discharged through the slag tap into the
the operating temperatures well above the melting point. Most of water-cooled pool. The two layers are distinguished by the slag
the slag is deposited on the wall of the gasifier to flow out and the viscosity and tapping temperature which are dependent mainly on
remaining fractions (fly slag) leave the gasifier with the syngas. the ash composition. The slag layer thickness must be maintained
The slag layer formed on the wall consists of two parts: the solid at an appropriate level to protect the refractory wall from the
slag in contact with the colder refractory and the liquid slag facing physical damage and corrosion by the hot syngas, while avoiding a
the hot syngas. The solid slag layer also includes a fraction of liquid blockage at the slag tap at the bottom of the gasifier [1]. The
behavior of fly slag that is contained in syngas is also important in
the design and operation of the downstream process, because
* Corresponding author. Tel.: þ82 31 299 4841. deposition of fly slag causes fouling/corrosion and, if severe,
E-mail address: cryu@me.skku.ac.kr (C. Ryu). blockage in the quenching and heat recovery equipment [2e5].

http://dx.doi.org/10.1016/j.applthermaleng.2015.05.027
1359-4311/© 2015 Elsevier Ltd. All rights reserved.
176 I. Ye et al. / Applied Thermal Engineering 87 (2015) 175e184

The ash/slag properties constitute important criteria in selecting study [29]. The model was used in a sensitivity study for key design/
candidate coals and determining the operating temperature for operation parameters and slag properties in a Prenflo gasifier [30].
entrained coal gasification [6e8]. Among them, the temperature of The results showed that the solid slag thickness was more sensitive
critical viscosity (Tcv) is considered as a key property that influences to changes in these parameters than the liquid slag thickness.
the behavior of slag on the gasifier wall. In general, Tcv refers to the Although Tcv is a crucial parameter for slag flowability and
temperature at which the viscosity of slag changes from that of a gasifier operation, its influence on the slag behaviors in the gasifier
Newtonian fluid to that of a plastic fluid upon cooling [9]. The rapid has rarely been investigated in details. In this study, the influence of
increase in viscosity at temperatures below Tcv occurs because of Tcv on the slag flow and heat transfer behaviors was analyzed using
the crystalline phase transformation in the compounds within the numerical modeling for simplified operating conditions in a com-
slag [10e12]. The critical viscosity varies depending on the slag mercial gasifier. Based on a slag viscosityetemperature correlation,
composition, but the value of 25 Pa s is considered as a typical value three parameters were varied: i) Tcv varied between 200 and
[1,13]. In gasification, Tcv represents the flowability of slag in rela- 200 K from the reference value, ii) critical viscosity varied between
tion to the syngas temperature. Slag with a low Tcv value may erode 2.5 and 100 Pa s, and iii) different ash:flux ratios. The results were
the refractory at the slag tap of the gasifier [13] or cause an analyzed for the thicknesses of solid and liquid slag layers, the
excessive deposition of fly slag downstream. In contrast, ash with a velocity and temperature profiles within the liquid slag, and the
high Tcv may cause a blockage at the slag tap by forming a thick slag heat transfer rate to the coolant.
layer. Increasing the gasifier temperature by supplying more O2 for
better slag flowability lowers the cold gas efficiency. In this case,
2. Numerical methods
CaO-based flux, such as limestone, can be fed together with coal to
lower the Tcv value and slag viscosity. A Tcv value below 1400  C has
2.1. Coal gasifier and test parameters
been proposed together with other ash property criteria for
entrained gasification [8,13]. It is also considered as the tempera-
Fig. 1 shows the entrained gasifier considered in this study,
ture at the interface of the solid and liquid slag layers on the gasifier
which was a Prenflo coal gasifier at Puertollano, Spain [14]. It
wall [14,15].
consisted of a cylindrical main body, a top cone leading to a syngas
Analyzing the slag flow properties and Tcv has been the main
quenching process and a bottom cone leading to a slag tap. The
topic of ash-related studies in the literature. The viscosity of slags in
bottom cone was inclined by 78 to gather the slag flowing
the Newtonian and non-Newtonian regions on cooling has been
downwards owing to gravity and to allow its discharge in the slag
measured for numerous coals [9,11,16,17]. Yun et al. [7] evaluated
tap.
the gasification and slag characteristics of nine candidate coals
Table 1 presents the reference condition prescribed for the
using a pilot-scale gasifier for integrated gasification combined
gasifier. The gas temperature (Tgas) was fixed at 1800 K along the
cycle (IGCC). To provide a guide for selection of coal types for
gasifier wall. The slag deposition temperature was assumed to be
gasification, Tcv has been correlated with ash fusion temperatures
50 K below Tgas, considering the possible effect of endothermic
or with chemical composition. Song et al. [18] experimentally
gasification. The coal ash consisted mainly of 57.4% SiO2 and 29.2%
assessed three existing models based on ash fusion temperatures
Al2O3, while CaO was the main constituent of the flux. Since the ash
for Chinese coals and proposed a new correlation for Tcv using the
had a low base/acid ratio resulting in a very high Tcv, the ash:flux
liquidus temperature calculated by FactSage. Owing to the large
ratio of 4:2 was considered as the reference condition to attain a
influence of inorganic constituents, many correlations have been
lower slag viscosity and Tcv. The slag deposition rate (mdep) for a
proposed with chemical composition of slag [19e21]. For example,
total of 6 kg/s was assumed to be evenly distributed per unit area of
Seggiani [19] analyzed data for 295 ash samples by linear regres-
the gasifier wall. Additionally, plausible distributions of Tgas and
sion and proposed a correlation for 49 parameters calculated based
mdep along the wall were considered, as presented in the
on slag composition. However, these correlations are subject to
Supplementary Information.
significant errors due to the complex nature of slag [8].
When the correlation of Kalmanovich and Frank [31] was
The slag flow behavior on the gasifier wall has been studied
adopted for the ash and flux blend ratio of 4:2, the viscosity for the
using cold flow experiments and, more frequently, mathematical
resultant slag composition was expressed to the following
modeling because of the extreme difficulty of investigation in
equation:
actual gasifiers. Wang et al. [22] carried out cold flow tests to
understand the slag deposition in the slag tap of a Shell gasifier.
m ¼ 6:37  1010 hðT  qÞexp½25049:5=ðT  qÞ (1)
Cohen and Reid [23] studied the trend in the slag thickness and
heat transfer for different slag viscosities and gas temperatures In Eq. (1), h and q were included in the original correlation in
using a very simple discretized model. Seggiani [14] proposed an order to vary the viscosity and Tcv values that are described later.
analytical model to calculate the slag flow and heat transfer, as part The reference condition had h ¼ 1 and q ¼ 0. According to Eq. (1),
of a dynamic process model for a Prenflo gasifier. With a linear the value of Tcv for the reference case at a viscosity of 25 Pa s was
temperature profile assumed in the liquid slag, this model derived 1465 K while that of the coal ash was 1852 K. The slag density,
algebraic functions for the key variables from the simplified gov- specific heat, thermal conductivity, and emissivity were deter-
erning equations. Seggiani's model was applied to gasifier studies mined using the respective correlations and values proposed by
of other researchers [24,25]. Ni et al. [26] used computational fluid Mills and Rhine [32,33].
dynamics (CFD) for the lower part (slag throat) of a gasifier with the For the wall condition, the refractory was assumed to have a
volume of fluid (VOF) model for the slag flow on the wall, and thickness of 16 mm and a thermal conductivity of 8 W/m K. The
analyzed the temperature, velocity and thickness of the slag layer. coolant tube had a thickness of 6.3 mm and a thermal conductivity
Yong et al. [15] proposed another analytical model for the slag flow of 43 W/m K. In regards to the coolant, the water/steam was
in which the temperature profile was assumed to be a cubic poly- assumed to have a temperature of 523 K with a convection coeffi-
nomial. This model was developed as the wall condition of oxy-coal cient of 1.0  104 W/m2 K.
combustors for CFD simulations [27,28]. In order to improve the Table 1 also lists the three parameters investigated in this study.
slag model without simplifying the temperature profile and slag Firstly, the slag viscosityetemperature correlation was varied to
properties, a new numerical model was proposed in our previous have Tcv values ranged between 1265 and 1665 K by changing q in
I. Ye et al. / Applied Thermal Engineering 87 (2015) 175e184 177

Fig. 1. Entrained coal gasifier considered in this study.

Table 1 assumed. In practice, the flux is not likely to be evenly distributed


Reference conditions of the gasifier and modeling cases for variations in Tcv, critical along the wall and mixed with coal ash to behave as a single phase,
viscosity, and ash:flux ratio.
especially at high flux feeding rates. Unlike the previous two
Parameter Values parameters, the changes in the ash:flux ratio accompanied an
Reference condition Gas temperature 1800 K increase in the slag deposition rate from 5 kg/s for the ratio of
Ash composition SiO2 57.4%, Al2O3 29.2%, 4:1e8 kg/s for 4:4.
Fe2O3 4.4% SiO3 3.1% CaO 2.7%, It is noteworthy that the numerical modeling on the slag
TiO2 1.3%, MgO 0.9%, K2O 0.7%,
behavior in this study was based on exemplary operating condi-
Na2O 0.3%
Flux composition CaO 85.67%, MgO 2.44% tions, without considering the actual heat and mass transfer
SiO2 6.80%, Al2O3 1.70%, interaction between the two regimes. For example, an increased
Fe2O3 0.51% Misc. 2.89% heat transfer rate to the wall by varying the aforementioned
Ash deposition rate 4 kg/s
parameters would lower the syngas temperature, which influences
Flux deposition rate 2 kg/s
Slag viscosity 6.37  1010 T exp(25049.5/T)
the coal conversion, gas flow, and slag formation/deposition char-
Test cases Tcv 1265, 1365, 1465 (reference), acteristics in the gasifier. In this study, the operating conditions for
1565, 1665 K the syngas were fixed constant in order to analyze the trends in the
Critical viscosity 100, 50, 25 (reference), 10, 2.5 Pa s slag thickness and heat transfer rate influenced by the selected
Ash:flux ratio 4:1, 4:2 (reference), 4:3, 4:4
parameters.

Eq. (1) between 200 and 200 K. The resultant correlations are 2.2. Numerical modeling for slag flow and heat transfer
illustrated in Fig. 2a. Secondly, the correlation was varied to have
critical viscosity values of 100, 50, 10, and 2.5 Pa s by multiplying The numerical model for the slag flow and heat transfer has
respective factors (h ¼ 0.1e4). In the measured data for 48 actual been proposed in our previous study [29]. The key equations of the
and synthetic ash samples, the critical viscosity ranged between model are summarized in Table 2. The mass, momentum, and en-
10.34 and 65.52 Pa s [18], but Yun et al. [7] reported some coals with ergy equations under the steady-state condition were expressed on
the critical viscosity approaching 100 Pa s in a pilot-scale gasifier. the cylindrical coordinates for a control volume i in a streamwise
The extreme value of 2.5 Pa s was included for comparison purpose. section j normal to the wall. In the solution procedure, the gov-
The value of Tcv was maintained constant at 1465 K, assuming that erning equations were discretized by the control volume method
the viscosity below this temperature abruptly increased. This for each section, and the solution marched from the top to the
assumption eliminated the need to model the low-viscosity region bottom of the gasifier. The slag deposition was treated as the
above Tcv, especially for the case with a critical viscosity of 2.5 Pa s. addition of a new control volume on the liquid slag surface.
Thirdly, the ash:flux ratio was varied from 4:2 to 4:1, 4:3, and 4:4. Therefore, the number of control volumes gradually increased from
Increasing the flux supply lowered the slag viscosity and Tcv, as the top to the bottom of the gasifier depending on the number of
shown in Fig. 2b, if the complete mixing of both materials was sections. In the iterative solution procedure for the equations that
178 I. Ye et al. / Applied Thermal Engineering 87 (2015) 175e184

Fig. 2. Viscosity-temperature curves for variations in (a) Tcv and (b) ash:flux ratio in the studied cases.

Table 2
Summary of the numerical model for the slag flow and heat transfer.

Equation For control volume i in streamwise section j

Mass conservation
mout ¼ min þ mdep (2)

Momentum conservation
   
dv dv
Mout ¼ Min þ Mdep þ 2pðr þ drÞdy$m   2prdy$m  þ rg sin a$dV
dr rþdr dr r
(3)
dv
at r ¼ ro : v ¼ 0; at r ¼ dL : ¼0
dr

Energy conservation
Hout ¼ Hin þ DHreact þ Qcond þ Hdep þ QGL (4)

k0 þ k1 Tcv  T1
at r ¼ ro : Qcond ð ¼ QLS Þ ¼ A0 (5)
2 r0 ln½r0 =ðr0  Dr1 =2Þ
 
4 4
at r ¼ rI : QGL ¼ AI εs Tgas  Tsurf (6)

Steady-state conduction in the wall


QLS ¼ QSR ¼ QRC (7)

where
Tcv  TR
QSR ¼ AR kS (8)
rR lnðrR =r0 Þ

and
QRC ¼ URC AC ðTR  TC Þ (9)

Liquid slag thickness


X
I
dL ¼ Dri ¼ rI  ro (10)
1

Solid slag thickness


dS ¼ rR  r0 (11)

Total slag enthalpy leaving the gasifier


X
I
Hout;total ¼ Hout;i (12)
1

Total heat transfer to the wall


X
J
QLS;total ¼ QLS;j (13)
1

Properties of slag [32,33]



Density ðkg=m3 Þ; r ¼ 2460 þ 18ðFeO%wt þ Fe2 O3 %wt þ MnO%wt (14)
X X . 
Specific heat ðJ=kg=KÞ; CL ¼ Xn CP;n ; CS ¼ Xn ðan þ bn T=1000  cn T 2 (15)
Thermal conductivity ðW=m=KÞ; k ¼ a$r$C where a ¼ 4:5  107 m2 =s (16)
Emissivity; ε ¼ 0:83 (17)
I. Ye et al. / Applied Thermal Engineering 87 (2015) 175e184 179

are coupled to each other, Eq. (2) was used to determine the liquid is noteworthy in Fig. 3 is that the temperature profile within the
slag thickness of each control volume, Eq. (3) the velocity, Eq. (4) liquid slag remained almost linear between Tcv to Tsurf within the
the temperature. Note that the temperature at the liquidesolid entire liquid slag layer, as indicated by the uniform horizontal
interface (i ¼ 0) was fixed at Tcv. The liquid slag thickness (dL) in a spacing between the contour lines.
section j was calculated by integration of the width of the control Fig. 4 compares dL for different values of Tcv within the range
volumes in the section. The heat transfer through the solid slag of 200 to 200 K with respect to the reference value. A lower Tcv
layer, and through the refractory to the coolant (water/steam) was value led to a thinner liquid slag throughout the gasifier wall. The
linked to the conduction of the liquid slag at the interface with the changes in dL at the slag tap for a difference in Tcv values of 400 K
solid slag (r ¼ ro). This relation of the steady-state conduction was were 6.4 mm (dL ¼ 14.0 mm at 1265 K and 20.4 mm at 1665 K). The
used to determine the solid slag thickness (dS) and the interface underlying mechanism that determines dL can be understood from
temperature of the refractory. the profiles of temperature, viscosity, and velocity within the liquid
At the end of calculation, the slag enthalpy (Hout) of the control slag at the slag tap, as shown in Fig. 5. The temperature profile
volumes at the bottom of the gasifier (slag tap) was added to (Fig. 5a) exhibited large differences between the studied cases,
calculate the total enthalpy of the liquid slag exiting the gasifier because Tcv was the interfacial temperature between the liquid and
(Hout, total). The heat transfer to the wall (QLS) on each section j was the solid slag layers. On the other end, however, the changes in Tsurf
also integrated along the wall to determine the total heat transfer were relatively small because Tgas was fixed at 1800 K. For slag with
rate (QLS,total). a higher Tcv, such a temperature profile resulted in much higher
The numerical model was developed into code using Microsoft viscosity towards the surface, leading to a slower slag flow and a
Excel Visual Basic for Applications (VBA). During the calculation for larger dL (Fig. 5b).
the gasifier, the wall was divided into 20 sections each for the top Fig. 6 shows that dS exhibits a similar trend with dL in the
cone, main body, and bottom cone. The grid sensitivity tests streamwise direction along the gasifier wall. However, dS increased
confirmed that the error was less than 1% for all key output rapidly for higher values of Tcv. When dL and dS were compared at
parameters. the slag tap (Fig. 7), the value of dS at Tcv ¼ 1665 K (158 mm) was ten
times larger than its value at 1265 K (15.4 mm). The reasons for
3. Results and discussion such large increases in dS elicited by Tcv are twofold. As mentioned
in Section 2.2, the steady-state conduction in Eq. (7) determined dS
3.1. Effect of shift in Tcv (note that Eq. (8) can be approximated as dS z (Tcv  TR)/QLS if r is
sufficiently large). For higher Tcv values, the temperature difference
Fig. 3 shows the temperature contour along the height of the (Tcv  TR) directly increased with Tcv changes because TR was almost
gasifier at the reference operating condition. With a uniform slag invariable as a result of dominant effect of TC. Moreover, the heat
deposition rate throughout the wall, the liquid slag quickly grew flux to the solid slag (qLS) was decreased, as indicated by the
thick on the top cone. This was because the slag velocity was low temperature gradient in Fig. 5a. In contrast, the increase in dL was
with the strong influence of the critical viscosity within the thin much smaller as shown in Fig. 4. It was because the velocity profile
layer. As more slag accumulated in the main body, the slag tem- was less sensitive to the change in the slag viscosity within the
perature near the surface became high enough to develop a low- liquid slag layer. With the viscosity at the liquidesolid slag interface
viscosity (i.e., high-velocity) region. Therefore, further increases fixed at 25 Pa s, the velocity profiles remained unchanged near the
in the amount of liquid slag contributed to a small increase in dL. As interface, as shown in Fig. 5b. In the authors’ previous study [30], dL
the slag entered the bottom cone, dL suddenly became very thick. has also been found less sensitive than dS to the changes in the slag
This occurred because of the abrupt change in the wall angle (78 ) properties and Tgas, whereas they are equally sensitive to the
and the corresponding decrease in the gravity force. One thing that changes in mdep and the bottom cone angle.

Fig. 3. Temperature contour within the liquid slag for the reference case. Fig. 4. Liquid slag thickness along the gasifier height for different values of Tcv.
180 I. Ye et al. / Applied Thermal Engineering 87 (2015) 175e184

Fig. 5. Profiles of (a) temperature, and (b) velocity and viscosity in the liquid slag layer for different values of Tcv.

Fig. 8 compares Hout, total and QLS,total for different values of Tcv.
QLS,total linearly decreased at higher Tcv values because of the
smaller temperature gradient within the liquid slag. In contrast, the
increase in Hout, total was relatively small despite the large tem-
perature differences between the studied cases (Fig. 5a). This was
because the mass flow (i.e., velocity) of the liquid slag near the
surface was large (Fig. 5b) at similarly high temperatures close to
Tgas. On the other hand, the influence of the large temperature
differences near the solid slag caused by different Tcv values was
diminished in Hout, total owing to the lower velocity.
The findings for the influence of Tcv are not dependent on the
distribution of Tgas and mdep, as presented in the Supplementary
Information. For plausible distributions of the parameters along
the gasifier wall (Fig. S1), the trends in the slag thicknesses, heat
transfer rate, and the slag enthalpy (Fig. S2) were found the same
with those in Figs. 7 and 8.

3.2. Effect of critical viscosity


Fig. 6. Solid slag thickness along the gasifier height for different values of Tcv.
For comparison with the influence of Tcv, the critical viscosity
was changed between 2.5 and 100 Pa s while Tcv was maintained
constant at 1465 K. Fig. 9 compares the profiles of temperature,

Fig. 8. Comparison of total wall heat transfer rate and slag enthalpy at the slag tap for
Fig. 7. Liquid and solid slag thicknesses at the slag tap for different values of Tcv. different values of Tcv.
I. Ye et al. / Applied Thermal Engineering 87 (2015) 175e184 181

viscosity, and velocity within the liquid slag at the slag tap. Sig-
nificant differences in the velocity profiles were observed at the
liquidesolid slag interface owing to the changes of the critical
viscosity (Fig. 9a). When the critical viscosity was lowered to
2.5 Pa s, a large velocity gradient (0.02 m/s/mm) was developed at
the interface, leading to a peak velocity of 0.196 m/s at the surface
in contact with the syngas. Since the mass flow rate was identical in
the studied cases, the higher slag velocity reduced dL to 8.1 mm
from 16.9 mm at the reference condition (25 Pa s). In contrast, the
slag with a critical viscosity of 100 Pa s had a velocity gradient at the
liquidesolid slag interface (0.002 m/s/mm) that was one tenth the
value attained for a critical viscosity of 2.5 Pa s. The peak velocity at
the surface was 0.064 m/s. As a result of the low velocity, dL
increased to 26.3 mm at the slag tap. Fig. 9b shows the temperature
profile within the liquid slag. Although Tsurf became slightly larger
at higher critical viscosities, its influence on the temperature
gradient was small. With Tcv fixed to a constant value on the other
end, the increases in dL accompanied smaller temperature
gradients.
Fig. 10. Liquid and solid slag thicknesses at the slag tap at the slag tap for different
Fig. 10 plots the variations in the values of dL and dS for different
critical viscosities.
critical viscosities. Because dS was inversely proportional to the
temperature gradient identified in Fig. 9b, it increased at higher
critical viscosity values (similarly to the trend of dL). However, both
dL and dS became less sensitive to the critical viscosity at values
above 25 Pa s.
Fig. 11 compares QLS,total and Hout, total for the studied cases. As
can be expected from the temperature gradient of the liquid slag
(Fig. 9b), QLS,total decreased at higher critical viscosities. In contrast,
the changes in Hout, total were very small for the same reasons as
those identified for Fig. 8.

3.3. Effect of ash:flux ratio

Fig. 12 shows the effects of the ash:flux ratio on the slag


thicknesses. The addition of flux not only lowered Tcv but also
increased the deposition rate (slag flow rate). However, the larger
slag flow rates did not increase dL. In fact, dL was decreased by
1.1 mm when the flux supply was doubled from 4:2 to 4:4. This
implies that the impact of a lower viscosity and lower Tcv values,
through the introduction of flux, was larger than that of the
increased mass flow rate. Fig. 13 indicates that the high-velocity
region near the liquid slag surface accommodated the additional Fig. 11. Comparison of total wall heat transfer rate and slag enthalpy at the slag tap for
increase in the mass flow rate. However, dS exponentially different critical viscosities.
decreased by the lowered Tcv values (Fig. 12), similarly to the

Fig. 9. Profiles of (a) velocity and viscosity, and (b) temperature in the liquid slag layer for different critical viscosities.
182 I. Ye et al. / Applied Thermal Engineering 87 (2015) 175e184

Fig. 12. Liquid and solid slag thicknesses at the slag tap for different ash:flux ratios. Fig. 14. Comparison of total wall heat transfer rate and slag enthalpy at the slag tap for
different ash:flux ratios.

elicited results of Fig. 7. Therefore, adding flux for coals with high
Tcv values is more effective in decreasing the slag thickness, while controlled by other parameters such as O2/coal ratio. The varia-
the impact of increased slag flow rates on slag thickness can be tions in these parameters influence the whole gasification char-
disregarded. acteristics including the syngas composition and slag deposition
Fig. 14 compares Hout, total at the slag tap and QLS,total at as well as the slag flow. Such complicated interactions between
different ash:flux ratios. Both Hout, total and QLS,total increased the slag layer and syngas regimes are not included in this study.
upon addition of flux. The observed trend in QLS,total was attrib- Further investigation is required using integrated process simu-
uted to the changes in Tcv, as previously explained in Section 3.1 lations or CFD studies for a comprehensive understanding of the
for Fig. 8. On the contrary, the trend in Hout, total was different impact of flux addition and Tcv of ash/slag for both the gas and
from that shown in Fig. 8. Despite the lower temperatures elicited wall regimes.
within the liquid slag with the addition of flux, the larger mass
flow rates directly increased Hout, total. Therefore, the heat loss
from the syngas could be considerably increased by the addition 3.4. Overall influence of Tcv and critical viscosity on the slag
of flux. For the increase of ash:flux ratio from 4:2 to 4:3, for thickness
example, the additional heat loss in Hout, total and QLS,total was
approximately 4.5 MWth. This was equivalent to a temperature The slag thicknesses change owing to a number of parameters
decrease of 55 K for the syngas (52.4 kg/s) leaving the gasifier at associated with the coal/slag properties and gasifier operation.
1803 K, when estimated using the operation data in Seggiani [14]. Among them, Tgas and the bottom cone angle were found to be
The endothermic calcination of flux, if it consists of lime, would the most important parameters [30]. To assess the influence of
further lower the syngas temperature. If Tgas decreases as a result these two parameters on the findings of this study, three addi-
of the larger heat loss, the slag behavior would be influenced in tional sets of test cases were also studied for i) a bottom cone
response. However, the syngas temperature can also be angle of 60 , ii) Tgas of 1750 K, and iii) Tgas of 1850 K. Fig. 15
compares the relative magnitude of the total slag thickness
(dS þ dL) at the slag tap normalized by the value for Tcv 1465 K
for each test set. The results also include those acquired for
different ash:flux ratios and at different locations (at the end of
reactor main body and at the slag tap). Regardless of these
parameters and locations, the total slag thickness exhibited
approximately the same trend against Tcv below 1565 K. For Tcv
above 1600 K, the deviations in the relative magnitude increased
with Tgas, but not with the values of the bottom cone angle. The
trend also suggests that it would be beneficial to maintain Tcv
below 1600 K, since the slag thickness rapidly increases above
this temperature.
Fig. 15b compares the effect of critical viscosity for the afore-
mentioned simulation cases. The trend in the relative magnitude of
the total slag thickness suggests that the critical viscosity of slag
does not have a significant influence on the slag thickness for values
above 10 Pa s. When the critical viscosity became extremely low (at
a value of 2.5 Pa s), the relative magnitude was largely varied by the
value of Tgas but not by the bottom cone angle. Comparing the two
graphs in Fig. 15, it can be concluded that determining the Tcv of slag
Fig. 13. Velocity and viscosity profiles of the liquid slag at the slag tap for different is more important on the slag behavior than the critical viscosity as
ash:flux ratios. long as it is above 10 Pa s.
I. Ye et al. / Applied Thermal Engineering 87 (2015) 175e184 183

Fig. 15. Relative magnitude of slag thicknesses for changes in (a) Tcv and (b) critical viscosity normalized by the values at the respective reference conditions (1465 K and 25 Pa s).

4. Conclusions Nomenclature

The trends in the slag flow and heat transfer characteristics were A area, m2
investigated by numerical modeling for the wall of a commercial C specific heat, J/kg/K
coal gasifier with different values of Tcv, critical viscosity, and g gravity, m/s2
ash:flux ratio. The key findings are as follows: H enthalpy, J/s
DH enthalpy of reaction, J/s
 A lower Tcv of slag exponentially decreased dS because of the k thermal conductivity, W/m/K
increased heat flux to the wall and the smaller temperature M momentum, kg m/s2
difference between the liquidesolid slag interface and the m mass flow rate, kg/s
coolant. However, the changes in Tcv did not noticeably in- Q heat transfer rate, W
fluence dL because the velocity profile near the liquidesolid r radius perpendicular to the wall, m
slag interface did not change at the same critical viscosity T temperature, K
(25 Pa s). U overall heat transfer coefficient, W/m2/K
 The slag with a higher critical viscosity values had a smaller V volume, m3
velocity gradient of the liquid slag near the solid slag layer for v streamwise velocity, m/s
the same Tcv. This led to a small increase in both dL and dS, but X mole fraction of slag compound
the impact on the total slag thickness was small above a critical y length parallel to the wall, m
viscosity value of 25 Pa s.
 Addition of flux was more effective for coals with high Tcv, since Greek
dS exponentially decreased at higher ash:flux ratios. The impact a angle from the horizontal plane,  thermal diffusivity, m2/s
of increased flux flow rate on dL can be disregarded because the d thickness of a slag layer, m
lowered viscosity accelerated the liquid slag flow. This accom- ε emissivity
panied an increase in both the wall heat transfer rate and the h multiplying factor for viscosity
slag enthalpy leaving the gasifier, which would lower the syngas q temperature shift for viscosity, K
temperature. m viscosity, Pa s
 The relative magnitude of the total slag thickness versus Tcv or r density, kg/m3
versus the critical viscosity was not significantly influenced by s StefaneBoltzmann constant, 5.67  108 W/m2/K4
the ash:flux ratio, gas temperature, or the angle of the bottom
cone. Overall, determining Tcv of slag was more important on the Subscript
slag behavior than the critical viscosity as long as the critical 0 liquidesolid slag interface
viscosity was above 10 Pa s. 1 outermost control volume facing the solid slag
C coolant (water/steam)
cond conduction
Acknowledgments cv critical viscosity
dep depositing slag
This work was supported by the New & Renewable Energy Core gas gas
Technology Program of the Korea Institute of Energy Technology GL from gas to liquid slag
Evaluation and Planning (KETEP) and Doosan Heavy Industries and I innermost control volume facing the gas
Construction with a grant from the Ministry of Trade, Industry & i index for a control volume within the liquid slag layer
Energy, Republic of Korea (2011951010001A). in inflow
184 I. Ye et al. / Applied Thermal Engineering 87 (2015) 175e184

J number of sections in the streamwise direction. [13] J.H. Patterson, H.J. Hurst, Ash and slag qualities of Australian bituminous coals
for use in slagging gasifiers, Fuel 79 (2000) 1671e1678.
j index for a section in the streamwise direction
[14] M. Seggiani, Modeling and simulation of time varying slag flow in a Prenflo
L liquid slag entrained-flow gasifier, Fuel 77 (1998) 1611e1621.
LS from liquid slag to solid slag [15] S.Z. Yong, M. Gazzino, A. Ghoniem, Modeling the slag layer in solid fuel
n each compound in slag gasification and combustion e formulation and sensitivity analysis, Fuel 92
(2012) 162e170.
out outflow to the section below [16] R.C. Corey, Measurement and Significance of the Flow Properties of Coal-ash
R refractory Slag, Bulletin 618, Bureau of Mines, United States Department of The Inte-
RC from refractory to coolant rior, 1964.
[17] A.Y. Ilyushechkin, S.S. Hla, D.G. Roberts, N.N. Kinaev, The effect of solids and
react reactions of residual carbon or the phase transformation phase compositions on viscosity behaviour and TCV of slags from Australian
S solid bituminous coals, J. Non-Cryst. Solids 357 (2011) 893e902.
SR from solid slag to refractory [18] W. Song, Y. Dong, Y. Wu, Z. Zhu, Prediction of temperature of critical viscosity
for coal ash slag, AIChE J. 57 (2011) 2921e2925.
surf liquid slag surface facing syngas [19] M. Seggiani, Empirical correlations of the ash fusion temperatures and tem-
total total value along the wall or in the liquid slag perature of critical viscosity for coal and biomass ashes, Fuel 78 (1999)
1121e1125.
[20] A. Lawrence, R. Kumar, K. Nandakumar, K. Narayanan, A Novel tool for
Appendix A. Supplementary data assessing slagging propensity of coals in PF boilers, Fuel 87 (2008) 946e950.
[21] A.R. McLennan, G.W. Bryant, C.W. Bailey, B.R. Stanmore, T.F. Wall, Index for
Supplementary data related to this article can be found at http:// iron-based slagging for pulverized coal firing in oxidizing and reducing con-
ditions, Energy Fuels 14 (2000) 349e354.
dx.doi.org/10.1016/j.applthermaleng.2015.05.027.
[22] J. Wang, H. Liu, Q. Liang, J. Xu, Experimental and numerical study on slag
deposition and growth at the slag tap hole region of Shell gasifier, Fuel Pro-
References cess. Technol. 106 (2013) 704e711.
[23] P. Cohen, W.T. Reid, The Flow of Coal-ash Slag on Furnace Walls, Technical
[1] C. Higman, M. Van der Burgt, Gasification, Gulf Professional Publishing, 2011. Paper 63, Bureau of Mines, United States Department of The Interior, 1944.
[2] G. Yu, J. Ni, Q. Liang, Q. Guo, Z. Zhou, Modeling of multiphase flow and heat [24] J. Kittel, F. Hannemann, F. Mehlhose, S. Heil, B. Meyer, Dynamic modeling of
transfer in radiant syngas cooler of an entrained-flow coal gasification, Ind. the heat transfer into the cooling screen of the SFGT-gasifier, in: 7th Modelica
Eng. Chem. Res. 48 (2009) 10094e10103. Conference, Como, Italy, September 2009.
[3] I.-S. Ye, S. Park, C. Ryu, S.K. Park, Flow and heat transfer characteristics in the [25] B. Li, A. Brink, M. Hupa, Simplified model for determining local heat flux
syngas quench system of a 300 MWe IGCC process, Appl. Therm. Eng. 58 boundary conditions for slagging wall, Energy Fuels 23 (2009) 3418e3422.
(2013) 11e21. [26] J. Ni, Z. Zhou, G. Yu, Q. Liang, F. Wang, Molten slag flow and phase trans-
[4] S. Park, I.-S. Ye, J. Oh, C. Ryu, J.H. Koo, Gas and particle flow characteristics in formation behaviors in a slagging entrained-flow coal gasifier, Ind. Eng. Chem.
the gas reversing chamber of a syngas cooler for a 300MWe IGCC process, Res. 49 (2010) 12302e12310.
Appl. Therm. Eng. 70 (2014) 388e396. [27] S.Z. Yong, A. Ghoniem, Modeling the slag layer in solid fuel gasification and
[5] J. Oh, I.-S. Ye, S. Park, C. Ryu, S.K. Park, Modeling and analysis of a syngas combustioneTwo-way coupling with CFD, Fuel 97 (2012) 457e466.
cooler with concentric evaporator channels in a coal gasification process, [28] L. Chen, A. Ghoniem, Development of a three-dimensional computational slag
Korean J. Chem. Eng. 31 (2014) 2136e2144. flow model for coal combustion and gasification, Fuel 113 (2013) 357e366.
[6] A.-G. Collot, Matching gasification technologies to coal properties, Int. J. Coal [29] I. Ye, C. Ryu, Numerical modelling of slag flow and heat transfer on the wall of
Geol. 65 (2006) 191e212. an entrained coal gasifier-Part I: model development and comparison with
[7] Y. Yun, Y.D. Yoo, S.W. Chung, Selection of IGCC candidate coals by pilot-scale analytical models, Fuel 150 (2015) 64e74.
gasifier operation, Fuel Process. Technol. 88 (2007) 107e116. [30] I. Ye, J. Oh, C. Ryu, Effects of design/operating parameters and physical
[8] P. Wang, M. Massoudi, Slag behavior in gasifiers. part I: Influence of coal properties on slag thickness and heat transfer during coal gasification, En-
properties and gasification conditions, Energies 6 (2013) 784e806. ergies 8 (2015) 3370e3385.
[9] S. Vargas, F.J. Frandsen, K. Dam-Johansen, Rheological properties of high- [31] D.P. Kalmanovitch, M. Frank, An effective model of viscosity for ash deposition
temperature melts of coal ashes and other silicates, Prog. Energ. Combust. phenomena, in: R.W. Bryers, K.S. Vorres (Eds.), Mineral Matter and Ash Depo-
Sci. 27 (2001) 237e429. sition from Coal, United Engineering Foundation, New York, 1990, pp. 89e102.
[10] J.W. Nowok, Viscosity and phase transformation in coal ash slags near and [32] K.C. Mills, J.M. Rhine, The measurement and estimation of the physical
below the temperature of critical viscosity, Energy Fuels 8 (1994) 1324e1336. properties of slag formed during coal gasification: 1.properties relevant to
[11] M.S. Oh, D.D. Brooker, E.F. de Paz, J.J. Brady, T.R. Decker, Effect of crystalline fluid flow, Fuel 68 (1989) 193e200.
phase formation on coal slag viscosity, Fuel Process. Technol. 44 (1995) [33] K.C. Mills, J.M. Rhine, The measurement and estimation of the physical
191e199. properties of slag formed during coal gasification: 2. Properties relevant to
[12] L. Kong, J. Bai, Z. Bai, Z. Guo, W. Li, Effects of CaCO3 on slag flow properties at heat transfer, Fuel 68 (1989) 904e910.
high temperatures, Fuel 109 (2013) 76e85.

You might also like