Anda di halaman 1dari 6

Exergetic Analysis as a Tool for Optimization of

Distillation Processes
A. C. B. Arajo1, L. G. S. Vasconcelos2 and R. P. Brito2

1 BRASKEM/TRIKEM S/A Macei Brasil acbrandaoa@ig.com.br

2 Universidade Federal de Campina Grande Centro de Cincias e Tecnologia


Departamento de Engenharia Qumica brito@deq.ufpb.br

This work presents two forms of process thermodynamic analysis based on the concept
of exergy. Applications of those concepts are employed in compression, heat exchange
and separation processes, along with the computation of their irreversibility rates. A
case study based on a real process in a VCM Plant is presented. This study aims to
optimize the use of energy requirements in the part of the process that deals with the
purification of 1,2-dichloroethane (1,2-EDC) in a high purity distillation column.
Firstly, the steady state of the original configuration is simulated in a commercial
simulator (Aspen PlusTM). After that, the results of the simulation are validated with
data extracted from the actual distillation system. A proposed optimized heat pump
system is assembled and its associated irreversibilities and efficiencies are computed. A
measurement of process profitability is performed to complete the evaluation of the new
optimized configuration. It is verified that the more adequate optimized sub-
configuration in the exergetic viewpoint is that in which the temperature difference in
the reboiler tends to zero because, in this case, the irreversibility would reach its
minimum value.

1. Basic Concepts of Exergy and Performance Criteria


According to Kotas (1995) energy quality is an expression of the capacity to cause
changes. Therefore, it depends on its mode of storage that can be either ordered or
disordered. One important characteristic of the ordered case is that all forms of energy
transfer occur in a reversible way while in the later case of storage all processes
involved are irreversible and the Second Law is essential in a further analysis. In case
the mode of storage of energy is disordered it can be defined an energy quality standard
which can be defined in terms of the maximum amount of work involved in a given
process, taking the environment as the reference state. This standard is commonly called
exergy. If there is any gradient between the environment and a given process, it is
possible to produce work from this configuration. The exergy of a stream can be written
according to the following equation:

E = EK + EP + EPH + E0 (1)

The kinetic (EK) and potential (EP) exergy components as ordered forms of energy are
represented by kinetic and potential energy equations.

The physical exergy (EPH) is defined as the maximum amount of work obtainable when
a stream of matter is taking reversibly from its initial state at P and T to the environment
state at T0 and P0 by physical processes exclusively with the environment.
The chemical exergy (E0) is defined as the maximum amount of work obtainable when a
stream of matter is taking reversibly from its environment state to the dead state at T0
and P0 by a process involving solely mass transfer with the environment.

According to Cornelissen (1997) there are three forms to evaluate the thermodynamic
performance of a process. Henley and Seader (1981) and King (1971) defined other
forms to calculate the thermodynamic efficiency. In fact, what seems to be more
important is to stipulate unique criteria of performance in order to evaluate energetically
a given process in its original configuration and in its proposed improved
configurations. This work presents two types of thermodynamic evaluation of a study
case: one by Kotas (1995) and the other due to Henley and Seader (1981).

2. Industrial Case
This study case deals with the purification process of 1,2-EDC in a distillation column
currently operating in one of the VCM Plants of the Braskem Company in Brazil. This
substance is an intermediate product to the production of VCM which is used to
produce PVC. Figure 1 depicts the actual system used in Braskem Company to purify
1,2-EDC.

TOP
CONDENSER

FEED 1
FEED 2
FEED 3 REFLUX DESTILLATE

FEED 4
DISTILLATION
VAPOR COLUMN

REBOILER
BOTTOM
LIQUID

Steam at 4 kgf/cm abs

Steam at 9 kgf/cm abs

Figure 1 Actual configuration of the distillation system.

Aspen Plus was used in all simulations. Its internal routine called RateFrac was used
to calculate non-equilibrium stages. Wilson equation was chosen as the thermodynamic
model. The results for the simulation of the actual configuration are given in Table 1.

Table 1 Main results of the simulated column.


Actual Simulation
Unit Distillate Bottom Distillate Bottom
Temperature C 55.00 93.79 55.00 95.89
Composition
1,2 EDC % 99.618 94.291 99.606 96.312

Figure 2 shows the proposed configuration using a heat pump scheme. An adiabatic
compressor with an isentropic efficiency of 0.72 was considered. According to Van
Ness et. al. (2000) compressors well-designed have efficiencies between 0.70 and 0.80.
This inefficiency generates some degrees of superheating in its outgoing stream.
COM PRESSOR

CONDENSER

REFLUX

FEED DISTILLATE

DISTILLATION
COLUM N

BOTTOM
PRODUCT
REBOILER

Figure 2 Proposed configuration for vapor mechanical recompression.

The procedure used to simulate the optimized configuration consists on simulating five
sub-configurations in order to eliminate the use of steam in the reboiler. Four of these
sub-configurations are based on four levels of temperature difference in the reboiler: 1,
10, 20, and 30C, that is, for each one of these sub-configurations there is a related area
in the reboiler and a defined power to be supplied to the compressor. The fifth
configuration was defined by holding the original reboiler. In this case the temperature
difference in the reboiler was found to be 2.82C.

It can be seen from Figure 3 that the exchange area decreases toward a constant value
showing that large temperature differences would reduce slightly the exchange area and
at the same time would increase, at high rates, the power required by the compressor.

3500

3000

2500
Power (kW) or Area (m)

2000

1500

1000

500

0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32
Temperature Difference in the Reboiler (C)

Reboiler area Compressor power

Figure 3 Reboiler area and compressor power profiles.

3. Exergetic Analysis
Irreversibility rate and efficiencies for all equipment/processes involved in the system
were calculated based on the equations described in Arajo (2002). Efficiencies were
calculated both by Kotas (1995) and Henley and Seaders (1981) approaches in order to
compare the profile of their results. All thermodynamic properties were obtained from
simulations by using Aspen Plus.
Stream exergy calculation was performed based on the procedure proposed by
Hinderink et. al. (1996). As composition of the streams involved in the system is
practically due to 1,2-EDC (at least 94% w/w), there are some peculiar characteristics in
this study case: almost all chemical exergy is due to 1,2-EDC, exergy of mixture has a
very small value in comparison with other forms of calculated exergies, exergy changes
in the system are almost totally due to physical exergy.

Table 2 presents the thermodynamic evaluation for the original configuration according
to Kotas (1995). Irreversibility rate and efficiency calculated for the distillation column
were 242kW and 99.68%, respectively. There is a great possibility to optimize heat
exchange processes particularly in the condenser where it can be seen the greatest
exergetic loss and the lowest efficiency.

Table 2 Thermodynamic evaluation results of the original configuration.


Equipment/Process Irreversibility (kW) Efficiency (%)
Reboiler 887.00 71.76%
Condenser 1,731.97 17.59%
Column 242.45 99.68%
Global 2,861.42 2.16%

Figure 4 presents the overall irreversibility rate and efficiency for the optimized
configuration. The equipment/processes that have a great influence on this behavior are
the reboiler and condenser. It is clearly seen that the best sub-configurations are those in
which the temperature differences are smaller.

3500 8%

3000 7%

6%
2500
Irreversibility rate (kW)

5%
Efficiency (%)

2000

4%

1500
3%

1000
2%

500
1%

0 0%
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32
Temperature Difference in the Reboiler (C)

Irreversibility rate Efficiency

Figure 4 Overall irreversibility rate and efficiency of the optimized configuration.

Figure 5 shows a diagram in which is presented the irreversibility rate distribution for
the original and optimized configurations. The substantial reduction in the condenser
irreversibility rate at the original configuration in relation to that of the optimized sub-
configurations is due to the reduction of heat exchange in this equipment. That is
because all latent heat is exchanged in the reboiler. It can be infer that for temperature
differences more than 10C the reboiler irreversibility rate becomes larger than that of
the original configuration. For temperature differences lower than 10C the total
irreversibility of the optimized configuration is 50% lower than the original
configuration. At a temperature difference of approximately 30C, optimized and
original configurations present similar values of irreversibility rate because of the
reboiler influence.

A very important observation from Figure 6 is the similarity between the profiles of the
overall efficiency calculated by Kotas (1995) and Henley and Seaders (1981) method.
Although they have different numerical values, they provide the same response for the
study case.

It was observed that there was no large variation between the efficiencies of each
equipment/processes in the optimized and original configurations. This means that only
the calculation of efficiency does not provide a good hint about what can be
exergetically optimized in a given system.

3000

598.27 kW
2500

467.46 kW
Total Irreversibility (kW)

2000 1731.97 kW

352.28 kW
1501.78 kW
1500
1164.45 kW
261.76 kW 279.07 kW

1000 838.56 kW

561.56 kW 561.95 kW
887.00 kW

500 559.09 kW
469.31 kW
288.69 kW 305.71 kW 376.64 kW

242.45 kW 242.45 kW 242.45 kW 242.45 kW 242.45 kW 242.45 kW


0
O riginal 1C 2.82C 10C 20C 30C

Configuration

Column Compressor Reboiler Condenser

Figure 5 Irreversibility rate distribution of the studied configurations.

0.18% 8%

0.16% 7%

0.14%
6%
Efficiency by Seader (%)

Efficiency by Kotas (%)

0.12%
5%
0.10%
4%
0.08%

3%
0.06%

2%
0.04%

0.02% 1%

0.00% 0%
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32
Temperature Difference in the Reboiler (C)

Efficiency by Seader Efficiency by Kotas

Figure 6 Optimized configuration overall efficiency.

4. Conclusion
The exergetic method showed to be a very important tool in process thermodynamic
analysis, allowing the identification of local and global exergetic losses, and
efficiencies. The calculation of both these variables is of vital importance in order to
guarantee successfully a final accurate decision about process modifications. The use of
classical and well-known thermodynamic concepts permits an easy implementation of
the exergetic analysis mainly because of the large dissemination of more robust
computational softwares.

The procedures defined by Kotas (1995) and Henley and Seader (1981) to perform
thermodynamic analysis yield the same profiles and tendencies. An advantage of
Kotass (1995) procedure is to allow the computation of irreversibilities and efficiencies
of the various parts that comprise the process in a very explicit and objective way which
represents a great differential in the investigation of possible exergetic enhancements
within the process. It is important to note that the use of Kotass (1995) procedure is
more interesting when the three parts that form the exergy of a given stream are
calculated separately as proposed by Hinderink et. al. (1996).

In the study case it was verified that the more adequate optimized sub-configuration in
the exergetic viewpoint is that in which the temperature difference in the reboiler tends
to zero because in this case the irreversibility would reach its minimum value.

5. References
ARAJO, A. C. B., Master Thesis, Federal University of Campina Grande, Brazil, 2002.

CORNELISSEN, R. L., Thermodynamics and Sustainable Development The Use of Exergy


Analysis and the Reduction of Irreversibility, University of Twente, The Netherlands, 1997.

HENLEY, E. J. & SEADER, J. D., Equilibirum-Stage Separation Operations in Chemical


Engineering, John Wiley & Sons, New York, 1981.

HINDERINK, A. P.; KERKHOF, F. P. J. M.; LIE, A. B. K.;DE SWAAN ARONS, J. & VAN
DER KOOI, H. J., Exergy Analysis with a Flowsheeting Simulator I. Theory; Calculating
Exergies of Material Streams, Chemical Engineering Science, (51), 4693-4700, 1996.

KING, C. J., Separation Process, McGraw-Hill Book Company, 1971.

KOTAS, T. J., The Exergy Method of Thermal Plant Analysis, Krieger Publishing Company,
Florida, 1995.

SMITH, J. M.; VAN NESS, H. C. & ABBOTT, M. M., Introduo Termodinmica da


Engenharia Qumica, LTC Editora, 2000.

Anda mungkin juga menyukai