You are on page 1of 19

Energy Conversion & Management 41 (2000) 109±127

www.elsevier.com/locate/enconman

Thermodynamic analysis of a coke dry quenching unit


Marcelo Risso Errera*, Luiz Fernando Milanez
Departamento de Energia, Faculdade de Engenharia Mecanica, Universidade Estadual de Campinas (UNICAMP),
13081-970 Campinas, SP, Brazil
Received 12 November 1998; accepted 22 March 1999

Abstract

A thermodynamic analysis (®rst and second laws) was performed for a coke dry quenching (CDQ)
unit with all data obtained in the site and under the unit normal operating conditions. Expressions to
evaluate the coke speci®c heat, enthalpy, entropy and physical and chemical exergies, as well as
particulate material (dust) formation indices, and the loss of coke mass through chemical reactions, were
developed and compared with the literature. The analysis was extended to a coke wet quenching (CWQ)
unit, a process still largely used across the world. The procedure is fully described. The comparison
between the performances of the CDQ and CWQ processes supports that such analysis can be a useful
and accessible tool in the decision making process. # 1999 Elsevier Science Ltd. All rights reserved.

Keywords: Decision making; Exergy analysis; Thermodynamic analysis; Coke quenching systems

1. Introduction

The iron and steel making processes have provided many topics suitable for energetic
studies. The large amount of thermal processes, and their magnitudes, always justify analyses,
improvements and optimizations, mostly because even small percentile gains may represent a
considerable economy when absolute values are considered. This work may contribute to such
studies in the future.
This work is a sequence of a thermodynamic study [1±3] performed on a coke dry quenching
(CDQ) system, in a Brazilian steel making plant using actual operational data and an updated

* Corresponding author. Present address: Departamento de Engenharia MecaÃnica, Universidade Federal do


ParanaÂ-UFPR, CP 19011 Curitiba, PR 81531-990, Brazil. Tel.: +55(41)351-3488.
E-mail address: errera@demec.ufpr.br (M.R. Errera)

0196-8904/00/$ - see front matter # 1999 Elsevier Science Ltd. All rights reserved.
PII: S 0 1 9 6 - 8 9 0 4 ( 9 9 ) 0 0 0 9 0 - 4
110 M.R. Errera, L.F. Milanez / Energy Conversion & Management 41 (2000) 109±127

bibliography on such systems. The main conclusions drawn were based on an Exergy
(Availability) Balance [15]. Nevertheless, the authors preferred not to call this work a `Second
Law based analysis,' following the arguments of Giftopoulos [4]. Giftopoulos argues that any
consistent analysis takes into account the general laws of Classical Thermodynamics altogether.
To show how thermodynamic analyses can bring a broader perspective of a process, as well
as grounds for decision making, we used the case of the well known CDQ process. CDQ
systems are regenerative systems whose thermal function is to transfer energy from the
incandescent coke to water vapor. Several phenomena are involved in this process, including
combustion, turbulence, ¯ow through porous media, heat transfer and emissions to the
atmosphere.
The red hot coke from the coke ovens must be quenched to prevent its oxidization and ash
formation while it is conveyed to the blast furnace. This task is mainly accomplished by means
of two di€erent processes: a CDQ system or a coke wet quenching (CWQ) system. Recently,
Ref. [1] also discussed the exergetic and ecological advantages of CDQ over CWQ. The CDQ
system was also the object of a thermoeconomic study with environmental considerations [2],
where it was shown how environmental policies associated with penalties and economic losses
can in¯uence the decision making on systems of this kind. More informations and results can
be found in Ref. [3].
Thermodynamic studies of processes related to steel making can be often found in the
literature, as for example Refs. [5±8], among others. The process of dry quenching of coke
itself has also received some consideration in the literature [6,10±12] but not with the approach
and exposition presented in this work.
In particular, we recently discovered that a similar analysis was performed by Ref. [9]
(published in an Italian periodical) that, coincidentally, was based on the same CDQ plant by
the time the plant became operational in 1983. It is worth mentioning, however, that the
present work was completely independent of that work. The approach, methods and
conclusions presented here are not the same as shown in that ®rst work. In this paper, the
simpli®cations and hypotheses involved are also the subject of discussions in view of a general
treatment of similar thermal systems. The authors believe this procedure could be of help in
future analysis of other systems.
We expect to encourage engineers to make use of the complete thermodynamic analysis to
compare units and operational settings, as well as to the research and development of new
technologies.

2. The coke dry quenching (CDQ) system

To quench coke in the steel making industry means to cool it from approximately 1100 to
2008C in order to inhibit its combustion and gasi®cation. The CDQ technique was introduced
in Switzerland by the Sulzer brothers in the 1920s. The initial apparatus did not allow a
uniform steam generation, had operational safety problems, and it was economically less
ecient than the conventional wet quenching system. A few decades later, an improved
conception of the CDQ process with continuous operation was produced by the Giprokoks
Institute, in Russia. In 1965, the two ®rst units were installed in the Cherepovets Plant. Since
M.R. Errera, L.F. Milanez / Energy Conversion & Management 41 (2000) 109±127 111

Fig. 1. A schematic coke dry quenching system Ð CDQ.

then, Russian, German and Japanese plants have been using and improving this process.
Recent advances in the matter are described in Ref. [13].
A schematic illustration of the system studied can be seen in Fig. 1. A more detailed
description can be found in Refs. [3,9,12], or with manufacturers. The present study was
conducted for a quenching plant of ®ve units. The red hot coke leaving the coke oven is
conveyed to the top of the quenching fore-chamber where it is unloaded. In typical conditions,
the coke is at a temperature of 10508C. A lid closes the chamber until the next load. The
incandescent coke at the top of the chamber falls by gravity as the quenched coke is discharged
from the bottom at a temperature of 1808C. This cooling is the consequence of heat transfer
from the coke to the circulating gas that, in turn, is cooled in the recuperation boiler and
returns to the bottom of the quenching chamber, position (3b) in Fig. 1. Steam is generated at
3508C and 2.06 MPa (21 kgf/cm2). The coke considered in this study was composed of 85.1%
carbon, 0.64% sulfur and 14.26% of ashes (mass, dry basis). The circulating gas contained low
quantities of O2 and around 70% of N2, thus presenting a low oxidant behavior with respect
to the coke. The composition of the circulating gas varies considerably along the circuit shown
in Fig. 1, however, it was considered minimum due to the way the control volumes were
de®ned [14]. Therefore, it was possible to utilize the composition presented in Table 1 for the
circulating and exhaust gases at point (5)1 of Fig. 1 and the composition presented in Table 2
for the exhaust gas in point (8). Generation of particulates in suspension occurs at the top of

1
Fluxes in parentheses refer to Fig. 1.
112 M.R. Errera, L.F. Milanez / Energy Conversion & Management 41 (2000) 109±127

Table 1
Chemical composition of the the circulating and exhaust (5) gas (vol.)

Element CO H2 CO2 O2 N2 H2O(V)


Content (%) 14.65 3.98 7.55 0.36 71.9 1.56

the fore-chamber during loading of the incandescent coke, at the bottom of the chamber
during unloading and when dust is dragged by the cooling gas ¯ow. Points of emission to the
ambient are indicated in Fig. 1. All of the ¯uxes in Fig. 1 are described in Table 3.
The ¯ow regime of the cooling gas along its circuit (chamber-boiler-chamber) is turbulent
with the gas dragging solid particles along almost all the path. Inside the quenching chamber,
the ¯ow is through a porous medium (stacked coke) with chemical reactions. Just after the
internal chamber openings, point (4a), there is air injection that reacts with the carbon oxides,
hydrogen and with some hydrocarbons (see Table 2). The recuperation boiler is composed of
economizers, evaporators and superheaters, as in the majority of plants. In the circuit of the
cooling gas, there are also an inertial collector and two cyclones in parallel to collect
particulates. As mentioned above, assumptions and simpli®cations must be made in order to
allow a viable and yet suciently representative analysis. Therefore, some of the complex
phenomena listed were not considered. This is the main reason why the CDQ process is
represented in the way it is shown in Fig. 1. The right mindset in this kind of work consists of
knowing what to disregard or not to disregard.
It is worth mentioning that the operation points of the units studied are considerably below
the regime found to be optimum by Ref. [9], namely, steam generated at 5408C and 8 MPa.
Among the reasons why such optimum regime is not currently obtained is the fact that regime
was derived based upon design data, idealized conditions and with newly manufactured parts
in the units. On the other hand, the present work was performed at actual operational
conditions which provided more realistic and, thus, more useful results. All the issues
mentioned in the two preceding paragraphs are further discussed in the next sections.

3. Data acquisition and methodology

The analysis presented in this work consists of a study of an actual working plant, and
therefore, we made no adjustments on the process to suit our measurements. As a ®rst
estimate, a large number of samples (measurements) would have been necessary for statistical
treatment due to the diversity, location and magnitude of the measurements that should be
taken. Because of the enormous amount of work involved in this procedure, a particular
approach had to be devised to make this study feasible.

Table 2
Chemical composition of the exhaust gas (8) (vol.)

Element CO H2 CO2 O2 N2 H2O(v)


Content (%) 20.38 5.25 3.60 0.96 68.5 1.31
M.R. Errera, L.F. Milanez / Energy Conversion & Management 41 (2000) 109±127 113

Table 3
Description of the ¯uxes in the control volumes

No. Description of the ¯uxes in Fig. 1

1 Incandescent coke, m_ inc


2 Coke quenched, m_ ext
3 Circulating gas ( gin, ng, ngas)
4 Circulating gas ( gout, ngas)
5 Inferior bleeder or inferior relief, ninf
6 Fines retained in the inertial collector, m_ coll
7 Fines dragged by the circulating gas and later collected in the cyclone, m_ cycl
8 Superior (fore-chamber) bleeder, nsup
9 Fines collected in the loading coke dedusting system m_ char
10 Fines collected in the unloading coke dedusting system, m_ disc
11 Air intake for combustion of gases, nair
12 N2 for the dilution of the circulating gas, nN2
13 Electric power used in the fans, W _ CV
…a†
14 Heat ¯ux loss to the ambient, Q_ CV
15 Dermineralized feed-water for the boiler, m_ w
16 Steam generated, m_ v
17 Steam released to the atmosphere, m_ sat
…b†
18 Heat ¯ux to the ambient, Q_ CV
19 Fines collected in the cyclone, m_ cycl

Taking measurements in systems in situ, especially in industry, is not a simple task. There
are several factors, internal and external to the process, causing all sorts of variations in the
properties under observation. In this work, such factors are classi®ed as operational and
intrinsic.
Among the factors of the operational kind are the stops, the imposed production of the coke
and the quality of the resources in use (raw material, manpower). The intrinsic factors are the
turbulence in the gas ¯ow, its changes in composition and variations of the environment
conditions, as well as the uncertainties and accuracy that are usually found in the laboratory.
The operational factors are, generally, the predominant cause of the variations in the
measurements in those cases. To overcome this problem, the performance of the CDQ
installation was observed for several months, and a typical day of operation was selected. Once
this selection was made, all the measurements registered in that day were considered, and an
average hourly2 regime was determined.
The concept of the control volume is applied, and attention is focused on the ¯uxes of
energy and entropy crossing the control surface. Some of the variables that can be measured
with more accuracy and uniformity served as the basis for the analysis. The majority of the
hourly ¯uxes were determined by an arithmetic average of the readings. In the case of the
coke, the daily number of charges and discharges were counted since its ¯ow cannot be directly

2
A CDQ system does not operate steadily, but in a cyclic and transient regime. Nevertheless, the proposed treat-
ment was suciently representative for the informations needed.
114 M.R. Errera, L.F. Milanez / Energy Conversion & Management 41 (2000) 109±127

measured. A recent mass balance [3] indicates 22 tons of coke for each load and 1.675 tons for
each discharge. These ®gures will be used in this study. Thus, the ¯ow of the incandescent coke
(1) is 10.593 kg/s (38.133 ton/h), and the production of quenched coke (2) is 10.333 kg/s
(37.200 ton/h). The total of ®ne particulates, ®nes, generated was obtained through readings at
di€erent points of the installation. The small amount of coke that reacts was calculated
approximately because this is not possible to be measured. To obtain this value, the amount of
carbon liberated through the bleeders (points (5) and (8) and Tables 1 and 2) was evaluated,
and it was also assumed that the coke was the only source of the element carbon (85.1% C).
The mass balance is shown in Table 4, and it may be noticed that there exists an error of the
estimate of 0.77% when compared to the total loaded coke. That coke balance is
proportionally in accordance with the literature [12]. The uncertainties are not shown due to
their relatively high order of magnitude. To attempt to determine the large uncertainties would
not bring any relevant information to the analysis, but only show the impossibility of
performing such measurements. Therefore, the analysis was performed without considering the
uncertainties. More importantly, it will be shown that this does not jeopardize the results and
their implications.
In the following sections, the methods and considerations used in the thermodynamic
analysis itself are presented.

3.1. Delimitation of the control volumes

Because of the need of making estimates of some parameters, it was found appropriate to
divide the CDQ control volume, CDQ-CV, into two complementary control volumes:
Chamber-CV and Boiler-CV. By this procedure, it was possible to identify better how the
reading errors could a€ect the ®nal balance. The Chamber-CV consisted of a region physically
larger than the fore-chamber itself and the quenching chamber. Its boundaries reached into the
interior of the CDQ, to the limit where practically no chemical reactions in the circulating gas
were observed, and from where it ¯owed in a stable manner to the recuperation boiler (Fig.
2(a)). The Boiler-CV corresponded to the region between the internal boundary of the
Chamber-CV to the physical limits of the recuperation boiler and the cyclones (Fig. 2(b)).

Table 4
Mass balance in the CDQ-CV

Mass ¯ux kg/s kg/h

m_ inc 10,593 38,133


m_ ext ÿ10,333 ÿ37,200
m_ fines ÿ0.101 ÿ364
m_ r ÿ0.076 ÿ274
Total 0.082 295
M.R. Errera, L.F. Milanez / Energy Conversion & Management 41 (2000) 109±127 115

Fig. 2. Control volumes analyzed (a) Chamber-CV and (b) Boiler-CV; where CDQ-CV = Chamber-CV + Boiler-CV
(¯uxes described in Table 3).

3.2. Considerations on the balance of gases

The average chemical composition of the gases utilized in this work were determined in two
ways. The composition presented in Table 1 was obtained from readings in real time at a
location equivalent to point (3b). As for the composition of Table 2, it was the result of
samples in point (8) obtained a posteriori. It was observed that these measurements were
adequate, provided that the composition of the coke and the meteorological conditions of the
plant site did not vary very much. According to this procedure, the volumetric ¯ow of the
exhaust gas in point (8) was found to be V_ 8  126  10ÿ3 Nm3/s (455 Nm3/h). The exhaust
gas, point (5), and circulating gas, point (3b), ¯ows could not be measured and had to be
estimated. Using measurements related to the steam production and temperature and pressure
readings of 7508C and ÿ0.671 kPa in point (4b), and 1468C and 1.27 kPa just after the fan,
point (3b), the ¯ow of the circulating gas can be estimated as:
" #
m_ v …hv ÿ h w †
V_ …3b †  V_ g 
rg …hin ÿ hout †

ÿ 
V_ …3b †  18:7 Nm3 =s 67,400 Nm3 =h …1†

As found in Ref. [3], there is a certain proportionality between the circulating gas ¯ow, V_ g ,
and the exhaust gas ¯ow in point (5). This relation is approximately as follows:
116 M.R. Errera, L.F. Milanez / Energy Conversion & Management 41 (2000) 109±127

V_ …5 †  21:8  10ÿ3 V_ g …2†

Thus, V_ …5†  408  10ÿ3 Nm3/s (1470 Nm3/h).


These estimates are in accordance with the literature [14] and with the plant records, in
which the aim was to maintain a ratio of 1600 Nm3 of circulating gas per ton of coke: …V_ g †N ˆ
1600  10:593  10ÿ3 ˆ 17:0 Nm3/s, whereas V_ …3b† 118:7 Nm3/s (Eq. (1)).
There are two other gas streams across the boundaries of the control volumes, namely, air at
point (11) and nitrogen at point (12). Both are injected in the circulating gas ¯ow, mainly to
control the level of H2. The air enters the unit by natural draft at ambient conditions T0 and
P0, and the N2 is injected at T0 and P0 kPa. As there were no readings, the air ¯ux had to be
estimated through the balance of the element oxygen that leaves the system in points (5) and
(8), because the air was the only source of oxygen in the system. Using the compositions and
respective ¯ow rates at points (5) and (8), the air ¯ow was determined by
ÿ 
n_ air ˆ n_ O2 ‡ n_ N2 ˆ 9:814  10ÿ3 kmol=s 791 Nm3 =h …3†
where n_ O2 ˆ …335:7 ‡ 139:4†=2=32 ˆ 2:06  10ÿ3 kmol/s, and n_ N2 ˆ 3:76n_ O2 ˆ 7:75  10ÿ3
kmol/s. At last, the ¯ux of injected nitrogen in point (12) could be obtained from the balance
between the air, point (11), and the exhaust systems, points (5) and (8). Thus,
ÿ 
n_ …N122 † ˆ n_ …N52† ‡ n_ …N82† ÿ n_ …N112 † ˆ 9:192  103 kmol=s 741 Nm3 =h …4†

Recent measurements [3] indicated 638 Nm3/h as the air ¯ux in point (11), while our estimate
was 791 Nm3/h. The average monthly nitrogen consumption [3] per quenching unit is 420
Nm3/h, while it was estimated as 741 Nm3/h by Eq. (4), ¯ux (12). These agreements (in an
order of magnitude sense) indicate that the proposed model for the gases balance is suciently
adequate.

3.3. Thermodynamic properties of the coke

The thermodynamic behavior of the gases follows the ideal gas model. Particularly, the
thermo-mechanical and chemical exergies were treated as suggested by Ref. [15].
Throughout this study, the composition of the coke was held ®xed. For solid substances,

Fig. 3. Comparison of the value of the coke speci®c from Eq. (5) with the literature.
M.R. Errera, L.F. Milanez / Energy Conversion & Management 41 (2000) 109±127 117

once the speci®c heat is known, it is possible to evaluate almost all the properties needed.
Thus, an evaluation of the coke speci®c heat was obtained by linearly adjusting the
calorimetric measurements [3], namely:
c…T † ˆ a  T ‡ b …5†
where a ˆ 0:699  10ÿ3 (kJ/kg K2), b ˆ 0:588 (kJ/kg K) with T, temperature, in kelvin and
with a coecient of multiple determination, R 2, of 98.9%. Fig. 3 illustrates the agreement of
Eq. (5) with some results found in the literature.
The next step is the determination of the behavior of the speci®c enthalpy with temperature,
h(T ). As coke is incompressible, its enthalpy can be approximated by the relation
c…T † ˆ dh=dT. Eq. (5) was determined including all coke components, and Ref. [16] showed
that the reference enthalpies of the ash and moisture in the coke at T0 and P0 can be
considered zero. Thus, it was assumed href ˆ 0 at T0 ˆ 298:15 K. After the integration, the
expression for the enthalpy becomes:
h…T † ˆ a  T 2 ‡ b  T ‡ 237:5 kJ=kg …6†
It was also necessary to determine the behavior of the entropy with temperature. The
fundamental relations of thermodynamics, when approximated for incompressible substances,
can be written as: ds ˆ c…T †dT=T. The expression for the coke entropy can be given by
…T
0 1 c
s …T † ˆ s …T0 † ‡ dT …7†
Mcoque T0 T

After substitution of the speci®c heat c(T ) from Eq. (5), it becomes:
s…T † ÿ s0 …T0 † ˆ a  T ‡ bln T ÿ 3:56 kJ=kg K …8†
Now, the coke thermo-mechanical (or physical) exergy, (e.g., Ref. [15]), can be obtained,
namely:
eph ˆ …h ÿ h0 † ÿ T0 …s ÿ s0 † …9†

Substituting Eqs. (6) and (8) and arranging terms yields:


eph ˆ aT 2 ‡ bT ÿ T0 …aT ‡ bln T † ÿ 824 kJ=kg …10†
The simpli®cations used to determine Eqs. (5)±(10) certainly caused some imprecisions.
Nevertheless, we strongly believe they constitute a good approximation in the face of their
simplicity. The study continued with the determination of the coke chemical exergy. The
literature provides empirical relations and tabulated values (e.g., [14±18]) but not always
covering speci®c situations. For instance, in Ref. [17], there are presented expressions
developed by Szargut and Styrylska in 1964 proposing correlations between the lower heating
value (LHV) and the chemical exergy of hydrocarbons in its three phases, Ref. [18] suggests a
similar expression and Ref. [16] proposes an empirical relation for coal and tar. In view of the
limitations of those alternatives, it was decided to obtain our own approximation for the
speci®c coke used in this study and compare it with values found in the literature.
118 M.R. Errera, L.F. Milanez / Energy Conversion & Management 41 (2000) 109±127

In Refs. [16,17], the ashes chemical exergy was considered to be zero. Therefore, a useful
mass of coke (m 0 coke ) was considered, that is, the mass of the ashes was subtracted from the
original amount of coke. Hence, the coke chemical exergy can be approximated as:
m 0 coke 0
e0 ˆ e0 …11†
mcoke
In this study, the coke useful mass constituted a homogenous mixture of carbon and sulfur.
From the literature (e.g., [15]), the chemical exergy of an incompressible substance can be
estimated by:
X
e~ 0 ˆ c~e 0c ‡ s~e 0s ‡ T0  R  x i ln…x i † …12†
i

where c and s represent the molar fractions of carbon and sulfur, respectively, in the coke
useful mass and x i their molar concentrations.
Despite its simplicity, Eq. (12) is in good agreement with the literature (Table 5). The value
of Ref. [16] was estimated by Domalshi in 1978 for a breeze-type coke, and the value of Ref.
[18] was calculated by the expression recommended for black coal and tar. For Eq. (12), a coke
with 88.9% carbon and 0.9% sulfur was used, a more usual composition. In terms of
percentages, the deviations are of:

je…02 † ÿ e…01 † j je…02 † ÿ e…03 † j


ˆ 1:44 and ˆ 11:9%
e…02 † e…02 †

It was decided to use Eq. (12) because it provided an intermediate value, close to that obtained
according to Ref. [16], and also because it gives more information with regard to its origins.
Thus, the coke chemical exergy adopted in this work was e0 ˆ 29,274 kJ/kg for coke with
85.1% carbon and 0.64% sulfur, as used in this study.

4. Energy and exergy balance

The energy balance was based upon the classic approach of the ®rst law of thermodynamics
when applied to control volumes, namely H_ R ÿ H_ P ‡ Q_ CV ÿ W_ CV ˆ 0. The energy balance is
presented for each of the control volumes illustrated in Fig. 2, according to the assumptions
introduced in the previous section.

Table 5
Comparison of coke chemical exergy values

Source Coke chemical exergy (kJ/kg)

Singh et al. [16] (e…1†


0 ) 30,194
Equation (12) (e…2†
0 ) 30,636
Shieh and Fan [18] (e…3†
0 ) 34,054
M.R. Errera, L.F. Milanez / Energy Conversion & Management 41 (2000) 109±127 119

4.1. Chamber-CV

The energy ¯ows that entered the control volume were grouped in H_ R as follows:
     
0 0 0
H_ R ˆ n_ coke h f ‡ Dh 10508C,1 atm ‡ n_ r h f ‡ Dh 10508C,1 atm ‡ n_ N2 h f ‡ Dh T ,P
0 0

      
0 0 0
‡n_ O2 h f ‡ Dh O ‡3:76 h f ‡ Dh N ‡n_ gas h f ‡ Dh T,P,…3 † (13)
2 2
258C,1 atm

In the same manner, the ¯ows leaving the Chamber-CV were:


X      
0 0 0
_
HP ˆ n_   _   _  
i h f ‡ Dh …T,P†i ‡ n gas h f ‡ Dh T,P,…4 † ‡ n inf h f ‡ Dh T,P,…5 †
iˆ2,6,7,9,10
 
0
‡ n_ sup h f ‡ Dh T,P,…8 † (14)

Usually, the external walls of the whole CDQ systems are lined with refractories and thermally
insulated, and the external convective conditions do not vary considerably. Those observations
allowed us to use the same ®gure obtained in the previous analysis of the plant [3], namely
…a†
Q_ CV ˆ 219 kW.
It was not possible to obtain a direct reading of the power consumed by the fan, and
therefore, it had to be estimated. We then used the head of Dp ˆ 6563:9 Pa (670 mm H2O), as
used by the CDQ manufacturer at the design phase and the typical eciency for that type of
fan of Zf ˆ 0:95 [3]. The circulating gas ¯ow rate was provided by Eq. (1). Thus, W _ CV ˆ
_
…DpV g †=Zf ˆ 139 kW.
In Eq. 13, the coke contribution was split into the ®rst and second terms of the right-hand
side. The former accounts for the amount of coke that is not involved in chemical reaction,
and the latter for the amount that reacts with the circulating gas. Also, for the ®rst term, only
the enthalpy variation due to temperature changes was computed (sensible heat), whereas for
the other one, it was necessary to consider the enthalpy of formation of the coke. It has been
shown in the literature [16] that the contributions of ash and moisture in the enthalpy of
formation of the coke are negligible. In addition, we also considered a homogenous coke mass,
since the ash is mainly concentrated in an external layer. Thus, the amount of coke, nr,
involved in the chemical reaction consists mainly of graphite carbon which, in turn, has zero
enthalpy of formation. The coke streams were separated as n_ coke ˆ n_ inc ÿ n_ r , where nÇ inc and nÇ r
can be read from Table 4. A similar analysis was also applied for the coke streams in Eq. (14).
The terms accounting for the circulating gas in both Eqs. (13) and (14) were computed by
components according to Table 1 and from thermodynamic tables available in the literature
(e.g., [15,19]). The last two terms of Eq. (14) were computed considering the following
measurements: at point (5) P5 ˆ 1 atm and T5 ˆ 1468C and at point (8) P8 ˆ 1 atm and
T8 ˆ 4068C [3]. The respective ¯ow rates were discussed in Eq. (2) and in the paragraph before
Eq. (1).
The ¯uxes of nitrogen, point (11), and air, point (12), in Eq. (13) take place at ambient
conditions T0 and P0, and thus, the enthalpy contribution to the balance is zero.
120 M.R. Errera, L.F. Milanez / Energy Conversion & Management 41 (2000) 109±127

The energy balance in the Chamber-CV is brie¯y shown in Table 6. At the bottom is shown
an unbalance of ÿ590 kW, equivalent to 2.7% of the total energy that entered the control
volume.

4.2. Boiler-CV

The main energy interaction that takes place within this control volume is between the
circulating gas and the water (steam). There are no chemical reactions present in this CV as we
discussed in Section 2, when the control volumes limits were established. Therefore, the energy
balance for the Boiler-CV is expressed by:
m_ w hw ÿ m_ v hv ÿ m_ sat hsat ‡ m_ g …hin ÿ hout †
ÿ  …b †
‡ m_ coll ‡ m_ cycl hcoll ÿ m_ coll hcoll ÿ m_ cycl hcycl ‡ Q_ CV ˆ 0 (15)

where m_ sat ˆ m_ w ÿ m_ v .
The magnitude of the enthalpy variation undergone by the circulating gas corresponds to the
same as calculated for the Chamber-CV. The ¯ows of water (steam) and their respective
enthalpies are presented in Table 7. The net enthalpy variation is shown at the bottom line.
…b†
The heat loss to the environment, Q_ CV , in the case of the Boiler-CV also came from a previous
…b†
work [3], namely, Q_ CV ˆ 45 kW. The contribution of the coke ®nes to the energy balance is
due to two quantities: the amount held by the inertial collector and the amount held by the
cyclones. The amount of ®nes retained by the collectors does not undergo temperature change,
since we delineated the CV boundaries to do so, whereas the amount retained in the cyclones
leaves the CV at 1468C (Table 4). The total enthalpy variation on the coke particulates was
also estimated by Eq. (6). Thus,
ÿ  216:1…1109 ÿ 109 †
m_ cycl hcoll ÿ hcycl ˆ ˆ 60:1 …kW † …16†
3600
The resulting energy balance is shown in Table 8. In this case, there was an unbalance of 630
kW.
After we calculated the energy balances in the two control volumes, Chamber-CV and Boiler-

Table 6
Energy balance in the Chamber-CV

Energy/power (kW)

H_ R ÿ21,891
H_ P ÿ21,381
…a†
Heat loss to the ambient, Q_ CV ÿ219
Power consumed on the fans, W _ CV ÿ139
ÿ590
M.R. Errera, L.F. Milanez / Energy Conversion & Management 41 (2000) 109±127 121

Table 7
Energy ¯uxes associated to the water

Flow rate (kg/h) Temperature (8C) Speci®c enthalpy (kJ/kg) Enthalpy (kW)

Feed water 23,083 103 435.39 2791.7


Steam to atm. 982 218 2801.38 ÿ764.2
Steam 22,101 350 3132.00 ÿ19,227.9
Total ± ± ± ÿ17,200

CV, we were able to estimate the energy balance for the whole CDQ-CV by adding the
resulting unbalances of each one. This is summarized in Table 9.
We de®ned a relative error of balance based on the main energy ¯uxes that entered the
control volume, CDQ-CV, namely:
jDEj
eˆ  
…1 † …15 †
H_ ‡ H_inc water

ˆ 6:28%
This is considerably small due to the circumstances under which the data was obtained. The
major contribution to this ®gure is related to the coke via the approximations performed to
estimate its content of energy and the large magnitude of its streams. The secondary factor is
due to the approximations related to the circulating gas, even though being an internal ¯ow,
that was used to estimate the other energy ¯uxes.
There are a limited number of conclusions one can draw solely from an energy balance. In
the following paragraphs, we then introduce the concept of exergy in our analysis, so it
becomes complete. Almost all the information necessary for developing the exergy analysis was
already introduced in Section 3 and earlier in this section. Exergy, as a measure of available
work content, is given by e ˆ eth ‡ e0 where eth is the parcel related to thermophysical changes
and e0 to the chemical changes of each substance. The equation for the exergy balance on a
control volume in the literature is given by (e.g., [19]):

Table 8
Energy balance in the Boiler-CV

Energy/power (kW)

Water/steam ÿ17,200
Circulating gas 16,555
Heat loss to the ambient ÿ45
Particulates (cyclones) 60
Total ÿ630
122 M.R. Errera, L.F. Milanez / Energy Conversion & Management 41 (2000) 109±127

Table 9
Global energy balance on the CDQ-CV (kW)

Chamber-CV ÿ590
Boiler-CV ÿ630
CDQ-CV ÿ1220

p ÿ
X  Xq X
r
dX q
E_ W ˆ ÿ ‡ E_ l ‡ …me
_ †j ÿ _ †k ÿT0  S_ gen
…me …17†
dt lˆ1 jˆ1 kˆ1

where q is the total number of streams entering the control volume, r is the number of streams
leaving and p is the number of heat transfer interactions with the external ambient. To evaluate
the irreversilibities associated with the extinguishing process, we used the so-called Gouy±
Stodolla relation, namely, I_  T0  S_ gen
For this balance, it is necessary to establish reference states for the substances present in the
control volume. There are many choices for reference states, depending on the characteristics
of each case (see, for example, Ref. [23]). For this work, the consistent choice is
T0 ˆ T 0 ˆ 258C and P0 ˆ P 0 ˆ 1 st: atm ˆ 0:101325 MPa as the dead state (physical
reference). Also, for the chemical elements C, O, N and H, the reference substances are,
respectively, CO2, O2, N2 and water vapor at their equilibrium state in standard air. Finally,
we also considered the exergy streams due to the coke, separated according to whether the
streams underwent chemical reactions or not.
The exergy balance is summarized in Table 10 (more details are shown in Tables 11 and 12).
Using the exergy balance, the thermodynamic cost of quenching the incandescent coke
becomes clear. The ®gure at the bottom line of Table 10 can be used for comparison with
other operational settings or improvements in the same unit or with alternative processes.
Traditionally, the exergy balance is illustrated by means of the Grasmann diagram. This
diagram, Fig. 4, shows the magnitudes of the streams and the losses. An analogous diagram
for the energy balance is often preferred in the industry which, in turn, is implicitly taken into
account in the Grasmann diagram. For this reason, we only showed the magnitude of the
exergetic streams.
Another way to indicate the performance of the system under study is to consider the ratio
between the destroyed exergy and the main product, that is, the unit cost of exergy. In the
present case, the product is the quenched coke. Using the data from Tables 4 and 10, we ®nd
that for each kilogram of extinguished coke, 513 kJ of exergy is destroyed.
For the sake of illustration, we performed the same analysis in a CWQ, which is still very
common in today's plants. The CWQ process is quite basic,since it consists of only spraying

Table 10
Global exergy balance on the CDQ-CV (kW)

In-¯ows 16,277
Out-¯ows 10,972
Irreversibilities or destroyed exergy 5305
M.R. Errera, L.F. Milanez / Energy Conversion & Management 41 (2000) 109±127 123

Table 11
Exergy ¯uxes inward CDQ-CV

Flux kg/h T (K) Exergy (kJ/kg) Exergy ¯ux (kW)

Inc. cokea 37,859.14 1323.15 1,289.92 13,565.3


N2 (pure) 849.52 3,014.64 25.4
Fan ± ± ± 138.8
Air (amb.) 1023.51 298.15 0.00 0.00
Feed Water 22,083.00 34.69 222.4
Burnt cokea 273.85 1.32315 30,153.75 2325.0
Total of exergy (in) 16,277
a
37,859:14 ‡ 273:85138,133 kg/h.

Table 12
Exergy ¯uxes outward CDQ-CV

Flux kg/h T (K) Exergy (kJ/kg) Exergy ¯ux (kW)

Ext. coke 37,200 463.15 66.87 691.0


Cyclonesa 216.1 418.15 46.35 2.8
Collectora 61.27 1103.15 865.02 14.7
Chargea 26.23 1323.15 1289.92 9.4
Dischargea 60.07 453.15 66.87 1.1
Inf. bleeder 1879.82 419.15 3899.92 2036.4
Sup. bleeder 563.23 679.65 8603.33 1346.0
Heat to ambient ± ± ± 29.8
Steam 22,101 623.15 1073.27 6589.0
Steam to atm 982 491.15 922.86 251.7
Total of exergy (out) 10,972
a
Coke particles collected.

water over wagons loaded with incandescent coke, allowing the resultant vapor to ascend freely
to the atmosphere.
An exergy balance was also performed in situ on an actual CWQ plant with a few additional
assumptions [3]. For the comparison needed in this work, we set the production of quenched
coke to be the same as in the CDQ plant, i.e. 37,200 kg/h. The thermodynamic conditions for
the water are shown in Table 13, and the conditions for the coke are the same as those
discussed earlier. Further details of this analysis can be found in Ref. [3].
The exergetic balance is summarized in Table 14, where the bottom line indicates the
destroyed exergy (10,267 kW). This yields a unitary exergetic cost of 994 kJ/kg of dry coke. If
the boundaries of the control volume are extended from the surroundings of the CWQ unit up
to where the water vapor is fully in equilibrium with the atmosphere, the unitary cost is
increased to 1220 kJ/kg.3

3
For the sake of brevity, we decided not to show the Grasmann diagram for the CWQ.
124 M.R. Errera, L.F. Milanez / Energy Conversion & Management 41 (2000) 109±127

Fig. 4. Grasmann diagram for the CDQ-CV.

Certainly, one would expect the CDQ to be more ecient, and yet, the CWQ is still broadly
used. More ecient CWQ [22] and CDQ [21] plants in operation can be found, but that will
not change the trends presented in our work. The question of why not all the CWQ plants
were replaced by CDQ type plants in the steel making industry is addressed in Section 5 of this
paper.
Every analysis, per se, is not relevant unless it is part of a decision making process (a
problem solving matter, improvement or yet comparison between alternatives in hand). The
assurance that the resulting ®gures are reliable (and to what extent) is fundamental in reaching
a good decision. To evaluate how good were our assumptions, we compared our work with the
literature. For instance, in the work of Ref. [11], the energy output associated with the steam
per ton of incandescent coke was 1537 MJ, whereas in our work, it was 1826 MJ. It is in good
agreement when one considers the estimate error of 11% in the energy balance of Ref. [11] and
the 6.28% in ours. It is worth mentioning that those ®gures are related to the incandescent
coke ¯ux, meaning that all the errors and uncertainties are therein embedded. To strength our
previous statement, we also compared the energy net gain by the steam in the boiler. While in
Ref. [11] the net gain was 691,365 MJ/ton of coke, in the present work, it was 1634 MJ/ton of
coke. The results of Ref. [11] were also later reported in Ref. [12]. Our ®gures were also

Table 13
Water and steam readings in the CWQ unit

Pressure (atm) Temperature (8C) Entalpy (kJ/kg) Entropy (kJ/kg 8C)

Sprayed water 3 25 105.2 0.3653


Pool watera 1 100 419.04 1.3069
Steam to atm 1 100 2676.1 7.3549
a
Pool water = feed water + coke ®nes.
M.R. Errera, L.F. Milanez / Energy Conversion & Management 41 (2000) 109±127 125

Table 14
Exergy ¯uxes in the CWQ-CV

Flux kg/h T (K) Physical exergy (kJ/kg) Exergy ¯ux (kW)

Inc. coke 37,200 1323.15 1289.92 13,329.2


Ext. coke 37,200 423.15 49.11 ÿ507.5
Fines 227.65 373.15 24.33 ÿ1.5
Water (pool) 26,167.65 298.15 29.42 ÿ213.9
Sprayed water 43,612.76 298.15 0.20 2.4
Steam to atm 17,445.11 373.15 483.20 ÿ2341.5
Total 10,267.2

comparable with the work of Ref. [9] performed in the same unit. For instance, in our work,
the exergy contents of the produced steam corresponded to 60% of the total exergy output,
whereas in Ref. [9], it was 45%. Reference [9] does not show the error in the estimate.

5. Conclusion

In this section, we stress the importance of complete thermodynamic analysis, the


conclusions one can draw by accounting for the lost exergy or irreversibilities generated in a
system, the reasons why concepts such as the CWQ are still in use and the trends for the future
of CDQ plants.
We start with remarks on our approach. The comparison made at the end of Section 4
shows that our methodology was successful in capturing the main physical phenomena of the
CDQ process. What di€erentiates our work most from the others is that we pointed out the
main assumptions necessary to perform the balances (mass, energy and exergy) and their
consequences on the resulting analysis. We presented them in a clear way so one should be
able to reproduce them in similar analyses in di€erent systems.
The implications of the di€erence between the amount of lost (destroyed) exergy in the two
processes are of great importance. There is a broadly accepted trend in the scienti®c
community [20,24] that the exergy is one, or perhaps the one, measure of the ecological impact
associated with a given process. That was just covered in a meeting sponsored by NATO [20].
To some extent, exergy accounts for the expenditure of natural resources, for the degree that
emissions pollute the environment, for how processes are adequate for their purposes and how
well those processes are being operated. In the present work, the thermodynamic superiority of
the CDQ process over the CWQ process is veri®ed4 by the lower impact (almost half of the
destroyed exergy of the CWQ system) of the CDQ systems on the environment. It is worth
noting that this complies with the insight of plant engineers that the heat released from the hot
coke should be recovered. Furthermore, this analysis can be used as a communication tool
among engineers. There are the ones who advocate the CWQ process is economically superior

4
This ecological comparison has been widely accepted.
126 M.R. Errera, L.F. Milanez / Energy Conversion & Management 41 (2000) 109±127

to the CDQ process, and others claim the quality of the yield coke in the CWQ process is
better. That and the ending useful life of many plants are the main reasons why there are so
many CWQ units still in operation.
The CDQ concept has been used for decades in the steel making industry. It has been
gradually improved and has led to newer concepts as, for instance, a hybrid two stage system
that starts by partially cooling the coke by means of contact with an inert gas in a hermetic
chamber and then fully cooling it by pure water spray. Reference [25] claims it provides higher
eciency than the other conventional systems. It is worth mentioning that the bases for the
conclusions presented in Ref. [25] are mainly operational and economical. We believe a
thermodynamic perspective on this new system, as well as a comparison like the one we are
presenting, would be more than suitable. While some manufacturers are seeking newer
technology for cooling incandescent coke to comply with more strict environmental regulations
and competitive markets, at the same time, others strongly believe use of the conventional
preparation of coal for iron making will not last long [25].
Another current trend is that the whole conventional iron making process may be partially
replaced by electric arc furnaces and direct methods which eliminate all the coking processes.
Alternatives using natural gases and gasi®ed coal are being tested, and pilot plants being built.
Still, it is believed blast furnaces technology will be used for more than a decade, since the iron
and steel making industry evolves in decades cycles, and such plants are still being built across
the world [25].
One way or the other, one thing will never be changed and that is the need of decision
making and, thus, methods as the one used in this work, method that comprises all the
relevant aspects of the thermal processes, yields relevant information and still is not more
complex than any other conventional energy analysis.

Acknowledgements

The authors gratefully acknowledge the ®nancial support of the Conselho Nacional de
Desenvolvimento CientõÂ ®co e TecnoloÂgico, CNPq, of the Brazilian Government and from the
Companhia SideruÂrgica de TubaraÄo, CST.

References

[1] Errera MR, Milanez LF. Aspectos ecoloÂgicos da a naÂlise exergeÂtical. In: Proc. of the Fifth National Thermal
Sciences Meeting, ENCIT (ABCM). SaÄo Paulo, Brazil. 1994. p. 97±100.
[2] Errera MR, Milanez LF, Morimoto T. Thermoeconomic analysis with environmental considerations in coke
dry quenching system. In: Proc. of the International Conference ECOS'95, Istanbul, Turkey. 1995. p. 436±41.
[3] Errera MR. Environmental considerations in thermoeconomics. M.Sc. thesis, Universidade Estadual de
Campinas (UNICAMP), Campinas, Brazil, 1994 (in Portuguese).
[4] Giftopoulos EP. Energy Conver Mgmt 1997;38(15-17):1525±33.
[5] Donatelli JL, Nogueira LAH. J Braz Soc of Mechanical Sciences (RBCM) 1994;16:13.
[6] Bisio G. Energy savings in coke oven plants. In: Proc. 24th Intersociety Energy Conversion Engineering
Conference (IECEC), vol. 24, Washington, DC, USA. 1989. p. 1719±24.
[7] Bisio G. Energy: The Int Journal 1993;18:971.
M.R. Errera, L.F. Milanez / Energy Conversion & Management 41 (2000) 109±127 127

[8] Bisio G. Energy: The Int Journal 1996;21:147.


[9] Bisio G, Superina L. R Tecnica Italiana 1985;(2):77±85 (in Italian).
[10] Minasov AN, Kononenko VS, Dzyuba VYa. Koks i Khimyia 1989;7:21 (Coke and Chemistry 1989;7:28).
[11] Karcz A, Przybyla A, Warzecha A. The analysis of the process of dry quenching of coke. In: Proc. of the
Second International Cokemaking Congress, London, UK, vol. 1. 1992. p. 163±8.
[12] Ukhmylova GS. Koks i Khimyia 1993;8:21 (Coke and Chemistry 1993;8:43±6).
[13] Babanin VI, Zajdenberg MA. Koks i Khimyla (in Russian) 1995;12:22.
[14] Rod'kin SP, Korobeinikov AP, Ushakov EB, Chalykh GN, Zotkln VP, Kaplenko AA, Agarkov VI,
Skorobogatyi MA. Koks i Khimyia 1984;6:17 (Coke and Chemistry 1984;6:28).
[15] Kotas TJ. The exergy method of thermal power plants. Malabar, FL: Krieger, 1985.
[16] Singh SP, Weil SA, Babu SP. Energy: The Int Journal 1980;5:905.
[17] Rodriguez SJL. Calculation of available-energy quantities. In: Gaggioll R, editor. Thermodynamics: Second law
analysis. Washington, DC: ACS, 1983. p. 39±53.
[18] Shieh JH, Fan LT. Energy Resources 1982;6(1/2):1.
[19] Bejan A. Advanced engineering thermodynamics. New York: Wiley, 1988.
[20] Bejan A, Mamut E, editors. Thermodynamic optimization of complex energy systems (NATO Advanced Study
Institute, Romania). Dordrecht, The Netherlands: Kluwer, 1999.
[21] Stakheev SG, Sukhorukhov UI, Rod'kin SP. Koks i Khimyia 1992;10:17±21 (Coke and Chemistry 1992;10:25±
31).
[22] Minasov AN, Dzyuba VYa, Tarkanov AA. Koks i Khimyia 1991;11:47 (Coke and Chemistry 11:58±9).
[23] Gallo WLR, Milanez LF. Energy: The Int Journal 1990;15(2):113±21.
[24] Szargut J. Bulletin of Polish Academy of Sciences-Technical Series 1986;34:475±80.
[25] Ondrey G, Parkinson G, Moore S. Chemical Engineering 1995;102(3):37±41.
[26] Perry RH, Chilton CH, editors. Chemical engineers' handbook, 5th ed. New York: McGraw-Hill, 1973. p. 19.

You might also like