Coal Oxycombustion Power Plant Optimization Using First Principles and Surrogate Boiler Models

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Available online at www.sciencedirect.

com

ScienceDirect
Energy Procedia 63 (2014) 352 – 361

GHGT-12

Coal oxycombustion power plant optimization using first principles


and surrogate boiler models
Alexander W. Dowlinga, John P. Easona, Jinliang Mab,c, David C. Millerc,
Lorenz. T Bieglera,*
a
Carnegie Mellon University, Department of Chemical Engineering, 5000 Forbes Avenue, Pittsburgh, PA 15203, USA
b
URS Corportation, 3610 Collins Ferry Road, Morgantown, WV 26507, USA
c
National Energy Technology Laboratory, 626 Cochrans Mill Road, Pittsburgh, PA 15236, USA

Abstract

Numeric optimization has successfully been applied in the chemical industry to improve process performance,
reduce emissions, create advanced control systems and intelligently explore process design alternatives. Regarding
power plants, optimization methods provide a systematic approach to balance trade-offs when designing efficient
and minimal cost carbon capture systems. Coal oxycombustion power plants are an ideal candidate for optimization,
given the complex trade-offs regarding flue gas recycle strategies, heat integration and design of cryogenic
separation systems.

Many oxycombustion studies consider sensitivity analysis instead of multi-variable optimization, which neglect
consequentially important interactions between subsystems. Furthermore, these studies rely on over- simplified
boiler models available in commercial process simulators. In this paper we present a hybrid 1D/3D boiler model that
balances accuracy and computational expense, making it well suited for optimization studies of an entire coal power
plant. This model is then incorporated into an equation-based optimization framework for power plants. Trust region
optimization methods are used to ensure convergence. Finally a proof-of-concept case study is presented, and future
work is discussed.

Keywords: coal oxycombustion; systems analysis; heat integration; optimization; computational fluid dynamics

* Corresponding author. Tel.: +1-412-268-2232; fax: +1-412-268-7139.


E-mail address: author@institute.xxx

1876-6102 © 2014 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/3.0/).
Peer-review under responsibility of the Organizing Committee of GHGT-12
doi:10.1016/j.egypro.2014.11.038
Alexander W. Dowling et al. / Energy Procedia 63 (2014) 352 – 361 353

© 2014
© 2013TheTheAuthors.
Authors. Published
Published by Elsevier
by Elsevier Ltd. is an open access article under the CC BY-NC-ND license
Ltd. This
Selection and peer-review under responsibility of GHGT.
(http://creativecommons.org/licenses/by-nc-nd/3.0/).
Peer-review under responsibility of the Organizing Committee of GHGT-12
1. Introduction

Oxycombustion is a promising technology for power generation from coal with CO2 capture. In contrast to CO2
flue gas scrubbing schemes, the main separation effort is before combustion due to O2 enrichment from air. This
allows coal to be combusted in a N2 lean environment, with two beneficial consequences: (1) reduced NOx
emissions and (2) a high purity of CO2 in the flue gas. This means much less flue gas processing is required to
purify CO2 for sequestration, unlike post-combustion technologies. See [1] for a review of oxycombustion
technology.

Many researchers have studied coal oxycombustion technologies in the past decade, ranging from combustion
experimentalists to CFD modelers. In the system analysis realm, researchers have focused on using exergy analysis,
heat integration, process simulation and other tools to both quantify the expected efficiency of oxycombustion
systems (for comparison with other carbon capture technologies) and identify pathways for efficiency improvement.
For example, Fu and Gundersen [2] used pinch analysis to estimate an overall system efficiency penalty of 10
percentage points due to the air separation unit and CO2 processing in an oxycombustion power plant. They also
used exergy analysis to improve the efficiency of air separation units tailored to the oxycombustion process [3, 4].
Recent work of Gundersen and coworkers [5, 6] has focused on applying heat integration methods to recover waste
heat from compressors and other parts of the oxycombustion process in the steam cycle.

Chemical process simulators are also popular tools for oxycombustion system analysis. For example, Hagi et al
[7] used a combination of Aspen Plus and exergy analysis to compare three flue gas recycle strategies. Similarly, Li
et al [7] explored sizing trade-offs between the air separation unit (ASU) and CO2 processing unit (CPU) in an
oxycombustion power plant. The advantages of operating an oxycombustion boiler slightly pressurized were
explored by Soundararajan et al [8] using simulations in Aspen Plus. Xiong et al [9] also used Aspen Plus to study
the sensitivity of an 800 MWe oxycombustion facility to several key design variables.

Although the findings of these studies are valuable, there are several limitations. Many of these studies make
drastically simplifying assumptions of various units in the oxycombustion process. For example, most Aspen Plus
oxycombustion modelers use simplified models of the air separation unit instead of modeling the more difficult
coupled double column system with accompanying multistream heat exchangers. As a result, these studies miss
important interactions between the ASU and other systems, such as heat integration opportunities between the ASU
and CPU. Zebian et al [10, 11] demonstrated the shortcomings of these sensitivity-based studies by applying multi-
variable numeric optimization to the oxycombustion process and found important interactions between key design
variables. However, their work was restricted to 17 variables and used older optimization technology (SQP in Aspen
Plus).

Another important limitation of most oxycombustion studies is the reliance on simplified reactor models in
process simulations, such as a Gibbs Reactor block in Aspen Plus. In actuality, the boiler in an oxycombustion
process is very complex. Over 90% of the heat transfer is radiative, which requires attention when determining the
amount of heat absorbed various parts of the boiler (e.g. walls vs superheaters). Furthermore, the combustion
properties of an oxy-fired boiler are drastically different than a similar boiler (same geometry, etc.) operating in an
air-fired mode. This is because the emissivity and fluid properties of CO2 and N2 are drastically different, causing
different thermal and flow profiles in oxy-fired vs air-fired boilers. Consequently the correlations typically used to
design air-fired boilers [12] most likely don’t apply to oxy-fired systems. Understanding combustion properties in
oxycombustion boilers has been the focus on many pilot-scale and computational experiments. See [13] for an
overview of combustion research and open questions related to coal oxy-firing. With this motivation, Edge et al [14,
15] coupled boiler models based on data from CFD studies with flowsheet simulators.
354 Alexander W. Dowling et al. / Energy Procedia 63 (2014) 352 – 361

In this paper, we propose an alternate methodology for directly combining detailed oxy-fired boiler models with
equation-based equipment models, and place an emphasis on optimization (not just simulation). Significant
advances have been made in the mathematical programming (optimization) community in the past two decades,
making the optimization algorithms included in commercial simulators (e.g. SQP in Aspen Plus) obsolete and
outdated. For example, both the CONOPT [16] and IPOPT [17] algorithms are both capable of efficiently solving
large sparse nonlinear optimization problems with 100,000+ variables and constraints. The goal of this work is to
simultaneously optimize an entire oxycombustion power plant, while considering detailed models for the air
separation unit, boiler, steam cycle, CO2 processing unit and compression trains. This approach will allow for many
of the previously studied trade-offs between subsystems, such as ASU and CPU sizing, heat integration with
adiabatic versus isotherm compression strategies and boiler operating pressure (gas side) to be rigorously analyzed
and balanced. The remainder of the paper is organized as follows. In Section 2, a hybrid 1D/3D zonal boiler model
is summarized. Next in Section 3, the equation-based steam cycle and heat integration model are presented. The
trust region framework used to link boiler and steam cycle (turbines, etc.) models is discussed in Section 4. A simple
air-fired case study is presented in Section 5. Ongoing work to extend the case study to an oxycombustion case (with
recycle) is discussed in Section 6. Models for the cryogenic separation systems (ASU and CPU) are presented in a
companion paper [18].

2. First Principles Boiler Model

In this study a hybrid 1D/3D zonal model developed by Ma [19] is used for the coal boiler. It requires less than
one minute to evaluate, providing a nice balance between speed and accuracy compared to over-simplified
equilibrium reactors (inaccurate and computational cheaper) and detailed CFD (high fidelity but expensive).
Furthermore, the low computational cost allows for more boiler design variables to be optimized, such as geometry,
compared with using expensive CFD simulations given a fixed computational budget. Similarly, when simulating an
oxy-fired flowsheet, multiple evaluations of the boiler may be required to converge the recycle loop, as the inlet
conditions to the boiler implicitly depends on its outlet.

2.1. Surrogate Model

Low computational costs are important because this boiler model, along with CFD models in general, does not
supply exact derivatives, which are required for efficient optimization. Instead of directly linking the boiler model to
the steam cycle model (Section 3), a surrogate model is used. The surrogate model is refit during the optimization
procedure, as described in Section 4. Moderately low computational costs for the boiler model keep the overall costs
of the optimization methodology reasonable. In contrast, Edge et al [14] construct a static surrogate boiler model
from data generated using 3D CFD simulations using nonlinear regression, and do not update the CFD model with
new data as part of the simulation methodology. As a result, there is a likely mismatch between their surrogate
model and CFD simulations at the final solution. The update procedure used in our methodology (see Section 4)
avoids this mismatch while maintaining mathematically provable convergence.

Linear surrogate models are fit for each outlet of the boiler model, as shown below

ߛ ൌ ‫ܣ‬ሺߜ െ ߜ଴ ሻ ൅ ߛ଴ (1)

where į is the vector of boiler inputs, Ȗ is the vector of boiler outputs and A contains the fitted coefficients. į0 and Ȗ0
are the inlet and outlet values around which the surrogate model is constructed (“center of the trust region”, see
Section 4). The boiler model inputs includes the following variables: primary and secondary air temperature (Tp, Ts),
flowrate (Fp, Fs) and composition (fp, fs); boiler height, width and other geometry measurements; and location of the
primary and secondary air inlets. The outputs include the following variables: heat lost to the wall (Qwall); heat lost
to the secondary superheater (Qsuper); flue gas temperature (Tf), flowrate (Ff) and composition (yf).
Alexander W. Dowling et al. / Energy Procedia 63 (2014) 352 – 361 355

2.2. Zonal Model Details and Validation

In the hybrid zonal mode, the boiler is discretized into N user specified vertical zones, as shown in Figure 2. Each
zone is modeled as a well-stirred reactor and the gas phase is assumed to be in chemical equilibrium. Reaction
kinetics are only considered for char oxidation and gasification. Uniform properties (velocity, temperature, etc.) are
assumed in each zone, except for radiation intensity, which is calculated using a discrete ordinate method on a 3D
mesh consisting of unstructured hexagonal cells (see Figure 2). Sample temperature profiles are shown in Figure 1.
As expected, the gas temperature is uniform in each zone, but the wall temperature varies in multiple dimensions.
This is because of the 3D radiative heat transfer calculations. The wall temperature profiles also includes radiant
superheaters at the top of the boiler. Additional reheaters, convective superheaters and the economizer are modeled
as (convective) heat exchangers (see Section 3). Additional details for this hybrid boiler model as discussed in [19].

Fig. 1. Zone and wall temperatures from hybrid boiler model

Fig. 2. 1D zones and 3D mesh for hybrid boiler model

Table 1. Comparison between hybrid boiler model and CFD simulations [19]
Air-fired Oxy-fired
Hybrid Model CFD Model Hybrid Model CFD Model
Horz. Nose Flue Gas Exit Temp. 1679 K 1674 K 1628 K 1656 K
Heat Loss to Wall (Qwall) 410.8 MW 436.0 MW 393 MW 403 MW
Heat Loss to Platen Superheater (Qsuper) 101.8 MW 102.8 MW 98.9 MW 109 MW
356 Alexander W. Dowling et al. / Energy Procedia 63 (2014) 352 – 361

3. Steam Cycle Equipment Models

The steam cycle portion of the power plant is modeled as a collection of connected equipment, including
mixers, turbines, the boiler, heat exhangers (reheaters, economizer, boiler feed water heater) and pumps, as
shown in Figure 3.

Leveraging modern large-scale optimization algorithms requires exact first (and preferably) second derivatives,
which are not available from most commercial simulators (especially in sequential-modular mode). However these
derivatives are automatically calculated in popular optimization modeling environments, such as GAMS [21]. Using
this environment and solvers requires algebraic models for the entire problem, which are presented below for the
steam cycle.

Fig. 3. Steam Cycle Flowsheet

3.1. Steam Table Interpolation

Steam enthalpies (H), entropies (S) and compressibility factors (Z) are calculated for streams in the steam cycle
flowsheet (see Figure 3, steam side) based on the IAPWS IF-97 standard using XSteam [22], a third party
MATLAB package. These values are a function of temperature (T), pressure (P) and phase. The latter is specified a
priori for each stream. Unfortunately, this package does not output derivative information, which are required by
optimization algorithms. Thus for each stream, XSteam is evaluated three times, at (ܶത, ܲത), (T1, P1) and (T2, P2),
where the latter two points are slight perturbations from the original point. A unique linear model is fit for each
property and stream, as shown below

‫ܪ‬௦ ሺܶ௦ ǡ ܲ௦ ሻ ൌ ܽ௦ ൅ ܾ௦ ൫ܶത௦ െ ܶ൯ ൅ ܿ௦ ሺܲത௦ െ ܲ௦ ሻ (2)


Alexander W. Dowling et al. / Energy Procedia 63 (2014) 352 – 361 357

where ܶഥ௦ and ܲഥ௦ are the initial temperature and pressure in stream s. Unlike the steam tables, this model is
differentiable. Similar models are also constructed for S and Z. As part of the trust region framework described in
Section 4, these linearized steam table models are updated throughout the optimization procedure by refitting at new
(ܶത௦ ,ܲത௦ ) points.

Phases of the streams in Figure 3 are specified a priori, and are constrained using

ܶ௩ ൒ ሺͳ ൅ ߳ሻܶ ௦௔௧ ሺܲ௩ ሻǡ‫ א ݒ׊‬ሼܸܽ‫ݏ݉ܽ݁ݎݐܵݎ݋݌‬ሽ (3)


ܶ௟ ൑ ሺͳ െ ߳ሻܶ ௦௔௧ ሺܲ௟ ሻǡ‫ א ݈׊‬ሼ‫ݏ݉ܽ݁ݎݐܵ݀݅ݑݍ݅ܮ‬ሽ (4)

where Tv and Tl are the temperatures of vapor and liquid streams, respectively, and Tsat is the saturation temperature
for those streams. One thousand Tsat data points were generated using XSteam and regressed, leading to the
following correlation:
ܶ ௦௔௧ ሺܲሻ ൌ ͲǤͷ͹Ͷ͹ͷͺͻ͵͵͹ʹͶ͹ʹͺ ൅ ͲǤͲͲͳʹͺͲʹʹͷͳͲͻ͵͸Ͳܲ ൅ ͲǤͲͻ͹ʹ͹Ͳͷͷ͵ͺ͵͹͹͸͸ Ž‘‰ଵ଴ ሺܲሻ
െ͵ǤͻͻͲ͸ͷͷ͵Ͷͳ͹ͺ͵ͻͺ͵ ൈ  ͳͲିସ ܲ Ž‘‰ଵ଴ ሺܲሻ ൅ ͲǤͲͳ͹ʹͶ͸ͷͶͲ͹ͳͶͶͳ͵ሾŽ‘‰ଵ଴ ሺܲሻሿଶ (5)
൅ͲǤͲͲʹͲͶͶ͵ͷ͹ʹͲʹͶʹʹሾŽ‘‰ଵ଴ ሺܲሻሿଷ

Although this expression is complex, it is accurate to less than 0.02% error from the triple point to critical point
of water. The nonlinearities do not appear to be problematic for CONOPT [16] (large-scale optimization algorithm).
߳ ൌ ͲǤͲͳ is used in (3) and (4) to restrict the vapor and liquid stream from operating near the saturation curve. This
helps avoid numeric issues with XSteam and accounts for mismatch between (5) and XSteam.

3.2. Steam Turbine Model

Each steam turbine section (shown in Figure 3) is modeled with a fixed isentropic efficiency, as shown below

‫ ܪ‬௢௨௧ ൌ ‫ ܪ‬௜௡ െ ߟ ௜௦ ȟ‫ ܪ‬௜௦ (6)

ܹ ௧௨௥௕௜௡௘ ൌ ‫ ܪ‬௜௡ െ ‫ ܪ‬௢௨௧ (7)

where Hin and Hout are the inlet and outlet enthalpy, respectively. ǻHis and Șis are the isentropic enthalpy and
isentropic efficiency, respectively, and Wturbine is the work produced in the turbine section. Unlike the turbine
hardware model [23] and extensions [24] that use an approximation for ǻHis, the isentropic enthalpy change is
calculated exactly by introducing a shadow stream for each turbine (ܶ෠ǡ ܲ෠ǡ ‫ܪ‬෡ ǡ ܵመ) and using following equations:

ܵ ௜௡ ൌ ܵመ (8)


ȟ‫ ܪ‬௜௦ ൌ ‫ ܪ‬௜௡ െ ‫ܪ‬ (9)

෡ are calculated using linear model (2). In the future, correlations from [24] for Șis will be added to the
where ܵመ and ‫ܪ‬
model.

3.3. Pump Model

Lower bounds for the work in pumps are calculated using

ܹ ௣௨௠௣ ൒ ‫ܨ‬ሺ‫ ܪ‬௢௨௧ െ ‫ ܪ‬௜௡ ሻȀߟ௣௨௠௣ (10)

ܹ ௣௨௠௣ ൒ ‫ ܸܨ‬௜௡ ሺܲ௢௨௧ െ ܲ௜௡ ሻȀߟ ௣௨௠௣ (11)


358 Alexander W. Dowling et al. / Energy Procedia 63 (2014) 352 – 361

where F is the flowrate through the pump, Șpump is the pump efficiency and V is specific volume (calculated from Z).
The objective function (maximize thermal efficiency, etc.) should drive Wpump down so that at least one of (10) or
(11) is active. Including both of these equations helps with convergence (see Section 4), as they rely on two different
thermodynamic property linearizations.

3.4. Remaining Equipment and Heat Integration Models

The remaining equipment models are extensions of flash calculations. See [25] for detailed equations. In
summary, valves and flash vessels (which are equivalent to mixers in this framework) are modeled using isenthalpic
flash calculations. The amount of pressure drop for each unit is an optimization variable. Similarly, each half side of
a heat exchanger (shown in Figure 3) is also modeled to allow for phase changes, where the heating or cooling load
is an optimization variable. 2% pressure drops are assumed in both the gas and steam side of all heat exchangers,
with the exception of the primary superheater (2.5%) and reheaters (8%). The latter are based on the
recommendations of Cotton [26]. The deaerator is currently modeled as a flash vessel/mixer.

Stand-alone Heat Exchanger Network Synthesis (HENS) methods seek a design with fixed stream data
(flowrates and temperatures) that minimizes a balance of capital (area and number of matches) and operating
(utility) costs [27]. For the steam cycle design problem, this approach is restrictive as we need to simultaneously
adjust flowsheet operating conditions (i.e. stream data) while performing heat integration by calculating minimum
utility demands. The Duran-Grossmann formulation [28] embeds the pinch equations (see [29]) into a nonlinear
flowsheet problem, allowing for simultaneous optimization of flowsheet variables and heat integration.

Our approach (see [25] for equations and details) is used to integrate heat exchanger halves for boiler feed water
heats, economizers, reheaters, air heaters and condensers in the steam cycle (Figure 3). As a result, exact pairing
between hot and cold sides of heat exchanger halves are not assumed a priori, allowing for other heat sources or
sinks, such as coolers after compression stages in the ASU and CPU, to be easily heat integrated with steam cycle.
In contrast to previous work [2, 5, 6], our approach couples heat integration with optimization, and operating
conditions in the steam cycle are allowed to change in response to the heat integration. This creates additional
degrees of freedom for optimization and leads to more tightly integrated (and efficient) power plant designs. See
[18] for a demonstration of this heat integration approach to optimize cryogenic separation systems for coal
oxycombustion. Furthermore, it is possible to use the concept of heat integration zones [25] to restrict the allowable
pairings between heat exchanger halves (such as restricting the hot and cold halves of the economizer to be paired
together) and maintain some practical constraints.

4. Trust Region Framework

The problem of obtaining derivatives for the steam thermodynamics and boiler model behavior is addressed
using concepts of surrogate models and trust region algorithms. Surrogate models are commonly used in the
chemical engineering literature to provide a simpler representation of a system that is more suitable for optimization
and analysis. However, inaccuracies in the surrogate model will usually cause the optimization to terminate at
suboptimal points. Trust region methods alleviate this difficulty, by managing trust region accuracy in a manner that
can guarantee convergence to the rigorous model optimum.

Instead of constructing a surrogate model for the whole design space, a sequence of surrogate models are
built that only capture local behavior along the optimization path. The local region in the design space, over which
the surrogate is constructed, is known as the trust region. The optimization is not allowed to extrapolate beyond the
trust region, so it is desirable to adaptively adjust the size ȟ௞ of the trust region at each iteration k. After solving the
optimization subproblem with surrogate k, the actual progress towards the optimum is checked. If the actual
progress is comparable to the predicted progress, the step is accepted and the trust region size may be increased. If
the progress is very small or the step was detrimental, the step is rejected and the trust region size is decreased.

By decreasing the trust region, it is easier to obtain an accurate surrogate. In fact, the requirement on the
accuracy of the surrogate model essentially states that as the model is built on a sequence of smaller domains, the
Alexander W. Dowling et al. / Energy Procedia 63 (2014) 352 – 361 359

error of the surrogate approaches zero. For details, see [30]. In this work, linear interpolation models, (1) and (2), are
used as the local surrogate model, which satisfy the required accuracy property assuming a smooth underlying
function and perform surprisingly well despite their simplicity.

Several comments are required on how the algorithm tests for progress towards the solution. In each
subproblem, the optimizer seeks to both reduce infeasibility (i.e., converge the flowsheet) and improve the objective
using the method developed in [31]. One option to judge the quality of the solution is to use a weighted combination
of objective function and feasibility measure known as a penalty function. However, the correct weighting of these
two terms is not known a priori. Filter methods were introduced to correct this. Borrowing concepts from multi-
objective optimization, this method is generally able to accept improvement in either the objective or feasibility. For
details on filter methods, see [32].

5. Case Study

We consider an air-fired case study to demonstrate the optimization methodology using the steam cycle
equipment models (Section 3) and the trust region framework (Section 4). A supercritical steam cycle with double
reheaters is considered, as shown in Figure 3. Currently, the boiler inlet and outlet gas stream conditions, along with
Qwall and Qsuper, are fixed at values from [33]. This demonstrates the framework is capable of synthesizing an optimal
steam cycle (include turbines, steam extraction, air heaters, steam reheaters and boiler feed water preheaters) for a
given boiler configuration. Furthermore, with these open models all of the heat exchangers in the steam cycle are
available for heat integration with other parts of the power plant. As future work, the hybrid boiler model (Section 2)
will be included in the optimization problem.

max Gross Thermal Efficiency


s.t. Fixed Boiler Inlets and Outlets
Steam Thermodynamics (2) - (5)
Steam Cycle Equipment Models (6) - (11)
Gas Side Equipment Models See [25]

Key results and assumptions for the optimized steam cycle are shown in Table 2. The composite curves, Figure 4,
show the results of heat integration. The horizontal gap between the ends of the top (red) and bottom (blue) curves
indicates external utility demands. The left gap (low temperature) is the cooling water demand and the right gap
(high temperature) is the heating demand. As expected, there is no heating requirement (due to a large cost specified
in the objective function). The external cooling requirement is large, also as expected, and is met with a cooling
tower. In the future, waste heat from compressors will also be available to the steam cycle for boiler feed water
heating.
360 Alexander W. Dowling et al. / Energy Procedia 63 (2014) 352 – 361

Fig. 4. Composite Curves from Heat Integration (ȟܶ௠௜௡ ൌ ͳͲԨ)

Table 2. Key Steam Cycle Optimization Results


Turbine Section Isentropic Efficiency (ߟ ௜௦ , fixed) 86%
Coal Feed Rate (fixed) 69.26 kg/s
Coal Feed Rate LHV (fixed) 1.546 GW
Heat from Boiler Wall (Qwall, fixed) 545.2 MW
super
Heat from Secondary Superheater (Q , fixed) 209.7 MW
Gross Thermal Efficiency (Optimized) 46.19%
Heat from Hot Flue Gas 835.7 MW
Heat into Air Heaters 232.3 MW
Total Pumping Work 12.4 MW
External Cooling Requirement 687.4 MW
Turbine Work 682.6 MW

6. Conclusions and Future Work

We present a methodology for embedding complex boiler models and steam thermodynamics into a flowsheet
for use with advanced optimization algorithms. A trust region framework is used to link the steam table
thermodynamic model with the equation-based steam cycle models. It is demonstrated in a proof-of concept case
study to optimize and the entire steam cycle for a fixed boiler. A computationally efficiency hybrid 1D/3D boiler
model is also presented.

As future work, we plan to extend the flowsheet to consider coal oxycombustion. This will enable the
exploration of many complex trade-offs, such as balancing and sizing the cryogenic separation systems, determining
the best flue gas recycle strategy and boiler firing configuration, and incorporating waste heat from compression into
the steam cycle. Model improvements are also planned, such as updated pressure drops, temperature bounds and
turbine efficiencies, along with alternative ways to incorporate the steam thermodynamics models into the
framework.

Acknowledgements

This work was made possible by funding through the Office of Fossil Energy (U.S. Department of
Energy) and RES contract DE-FE0004000. This paper was prepared as an account of work sponsored by an agency
of the United States Government. Neither the United States Government nor any agency thereof, nor any of their
employees, makes any warranty, express or implied, or assumes any legal liability or responsibility for the accuracy,
completeness, or usefulness of any information, apparatus, product, or process disclosed, or represents that its use
would not infringe privately owned rights. Reference herein to any specific commercial product, process, or service
Alexander W. Dowling et al. / Energy Procedia 63 (2014) 352 – 361 361

by trade name, trademark, manufacturer, or otherwise does not necessarily constitute or imply its endorsement,
recommendation, or favoring by the United States Government or any agency thereof. The views and opinions of
authors expressed herein do not necessarily state or reflect those of the United States Government or any agency
thereof.

References
[1] Scheffknecht , G., Al-Makhadmeh, L., Schnell, U., Maier, J., Oxy-fuel coal combustion A review of the current state-of-the-art, International
Journal of Greenhouse Gas Control 5 (2011) S16–S35.
[2] Fu, C., Gundersen, T., Heat Integration of an Oxy-Combustion Process for Coal- Fired Power Plants with CO2 Capture by Pinch Analysis,
Chemical Engineering Transactions 21 (2008) (2010) 181–186.
[3] Fu, C., Gundersen, T., Using exergy analysis to reduce power consumption in air separation units for oxy-combustion processes, Energy 44
(2012) 60–68.
[4] Fu, C., Gundersen, T.. Recuperative vapor recompression heat pumps in cryogenic air separation processes. Energy 2013;59:708–718.
[5] Fu, C., Anantharaman, R., Gundersen, T.. Optimal integration of compression heat with regenerative steam Rankine cycles. In: Mario R.
Eden, J.D.S., Towler, G.P., editors. Proceedings of the 8th International Conference on Foundations of Computer-Aided Process Design;
vol. 34 of Computer Aided Chemical Engineering. Elsevier; 2014, p. 519–524.
[6] Soundararajan, R., Anantharaman, R., Gundersen, T.. Design of Steam Cycles for Oxy-combustion Coal based Power Plants with Emphasis
on Heat Integration. Energy Procedia 2014;51(1876):119–126.
[7] Hagi, H., Nemer, M., Le Moullec, Y., Bouallou, C.. Assessment of the Flue Gas Recycle Strategies on Oxy- Coal Power Plants using an
Exergy-based Methodology. Chemical Engineering Transactions 2013;35:343–348.
[8] Soundararajan, R., Gundersen, T., Ditaranto, M.. Oxy-Combustion Coal Based Power Plants: Study of Operating Pressure, Oxygen Purity and
Downstream Purification Parameters. Chemical Engineering Transactions 2014;39:229–234.
[9] Xiong, J., Zhao, H., Chen, M., Zheng, C.. Simulation Study of an 800 MWe Oxy-combustion Pulverized-Coal- Fired Power Plant. Energy &
Fuels 2011;25:2405–2415.
[10] Zebian, H., Gazzino, M., Mitsos, A.. Multi-variable optimization of pressurized oxy-coal combustion. Energy 2012;38(1):37–57.
[11] Zebian, H., Rossi, N., Gazzino, M., Cumbo, D., Mitsos, A.. Optimal design and operation of pressurized oxy-coal combustion with a direct
contact separation column. Energy 2013;49:268–278.
[12] Economizers and Air Heaters. In: Steam: its generation and use; chap. 20; 41 ed. The Babcock & Wilcox Company; 2005, p.20.1–20.18.
[13] Chen, L., Yong, S.Z., Ghoniem, A.F.. Oxy-fuel combustion of pulverized coal: Characterization, fundamentals, stabilization and CFD
modeling. Progress in Energy and Combustion Science 2012;38(2):156–214.
[14] Edge, P., Heggs, P., Pourkashanian, M., Stephenson, P.,Williams, A.. A reduced order full plant model for oxyfuel combustion. Fuel
2012;101:234–243.
[15] Edge, P., Heggs, P., Pourkashanian, M., Stephenson, P.. Integrated fluid dynamics-process modelling of a coal-fired power plant with carbon
capture. Applied Thermal Engineering 2013;60(1-2):242–250.
[16] Drud, A.S.. Conopt – a large-scale grg code. ORSA Journal on Computing 1994;6(2):207–216.
[17] Andreas, W., Biegler, L.T.. On the implementation of an interior-point filter line-search algorithm for large-scale nonlinear programming.
Mathematical Programming 2006;106(1):25–57.
[18] Dowling, A.W., Balwani, C., Gao, Q., Biegler, L.T.. Equation-oriented optimization of cryogenic systems for coal oxycombustion power
generation. Energy Procedia (GHGT-12 Proceedings) 2014.
[19] Ma, J., Dowling, A., Eason, J., Biegler, L., Miller, D.. Development of First Principle Boiler Model and Its Reduced Order Model for the
Optimization of Oxy-combustion Power Generation System. In: 39th International Technical Conference on Clean Coal & Fuel Systems.
2014.
[20] Adams, B.R., David, K., Senior, C., Shim, H.S., Otten, B.V., Fry, A., et al. Characterizaton of oxy-combustion impacts in existing coal-fired
boilers. Tech. Rep.; NETL Technical Report NT0005288; 2013.
[21] GAMS Development Corporation, . General algebraic modeling system (gams) release 24.0.2. 2012.
[22] Homgren, M.. X steam, thermodynamic properties of water and steam. 2006. URL: http://www:mathworks:com/
matlabcentral/fileexchange/9817-x-steam--thermodynamic-properties-of-water-and-steam.
[23] Mavromatis, S., Kokossis, A.. Conceptual optimisation of utility networks for operational variations – I. targets and level optimisation.
Chemical Engineering Science 1998;53(8):1585–1608.
[24] Luo, X., Zhang, B., Chen, Y., Mo, S.. Modeling and optimization of a utility system containing multiple extractions steam turbines. Energy
2011;36(5):3501–3512.
[25] Dowling, A.W., Biegler, L.T.. A framework for efficient large scale equation-oriented flowsheet optimization. Computers & Chemical
Engineering 2014;72(2):3-20
[26] Cotton, K.C.. Evaluating and Improving Steam Turbine Performance. 2 ed.; Cotton Fact Inc; 1988.
[27] Biegler, L.T., Grossmann, I.E., Westerberg, A.W.. Systematic Methods of Chemical Process Design. Upper Saddle River, New Jersey, USA:
Prentice Hall PTR; 1997.
[28] Duran, M.A., Grossmann, I.E.. Simultaneous optimization and heat integration of chemical processes. AIChE Journal 1986;32(1):123–138.
[29] Linnho_, B.. Pinch analysis - a state-of-the-art overview. Trans IChemE 1993;71(A).
[30] Conn, A.R., Scheinberg, K., Vicente, L.N.. Introduction to derivative-free optimization. Philadelphia: SIAM; 2009.
[31] Omojokun, E.O.. Trust region algorithms for optimization with nonlinear equality and inequality constraints. Ph.D. thesis; University of
Colorado at Boulder; 1990.
[32] Agarwal, A., Biegler, L.T.. A trust-region framework for constrained optimization using reduced order modeling. Optimization and
Engineering 2013;14(1):3–35.
[33] Reference power plant model user manual. Tech. Rep.; Carbon Capture Simulation Initiative; 2013.

You might also like