You are on page 1of 26

CH05CH14-Miller ARI 19 May 2014 17:25

Carbon Capture Simulation


Initiative: A Case Study in
Multiscale Modeling and
ANNUAL
Further
REVIEWS
Click here for quick links to New Challenges
Annual Reviews content online,
including:
• Other articles in this volume David C. Miller,1 Madhava Syamlal,2
Annu. Rev. Chem. Biomol. Eng. 2014.5:301-323. Downloaded from www.annualreviews.org

• Top cited articles


• Top downloaded articles David S. Mebane,3 Curt Storlie,4
Access provided by University of Saskatchewan on 03/05/17. For personal use only.

• Our comprehensive search


Debangsu Bhattacharyya,5 Nikolaos V. Sahinidis,6
Deb Agarwal,7 Charles Tong,8 Stephen E. Zitney,2
Avik Sarkar,9 Xin Sun,9 Sankaran Sundaresan,10
Emily Ryan,11 Dave Engel,9 and Crystal Dale4
1
US Department of Energy, National Energy Technology Laboratory, Pittsburgh,
Pennsylvania 15236; email: david.miller@netl.doe.gov
2
US Department of Energy, National Energy Technology Laboratory, Morgantown,
West Virginia 26507; email: madhava.syamlal@netl.doe.gov, stephen.zitney@netl.doe.gov
3
Department of Mechanical and Aerospace Engineering and 5 Department of Chemical
Engineering, West Virginia University, Morgantown, West Virginia 26506;
email: david.mebane@mail.wvu.edu, Debangsu.Bhattacharyya@mail.wvu.edu
4
Los Alamos National Laboratory, Los Alamos, New Mexico 87545; email: storlie@lanl.gov,
cbdale@lanl.gov
6
Department of Chemical Engineering, Carnegie Mellon University, Pittsburgh,
Pennsylvania 15213; email: sahinidis@cmu.edu
7
Lawrence Berkeley National Laboratory, Berkeley, California 94720; email: daagarwal@lbl.gov
8
Lawrence Livermore National Laboratory, Livermore, California 94550;
email: tong10@llnl.gov
9
Fundamental and Computational Sciences Directorate, Pacific Northwest National
Laboratory, Richland, Washington 99352; email: avik.sarkar@pnnl.gov, xin.sun@pnnl.gov,
dave.engel@pnnl.gov
10
Department of Chemical and Biological Engineering, Princeton University, Princeton,
New Jersey 08544; email: sundar@princeton.edu
11
Department of Mechanical Engineering, Boston University, Boston, Massachusetts 02215;
email: ryanem@bu.edu

Annu. Rev. Chem. Biomol. Eng. 2014. 5:301–23 Keywords


First published online as a Review in Advance on optimization, uncertainty quantification, process synthesis, computational
April 10, 2014
fluid dynamics, process control, risk analysis
The Annual Review of Chemical and Biomolecular
Engineering is online at chembioeng.annualreviews.org Abstract
This article’s doi: Advanced multiscale modeling and simulation have the potential to dramat-
10.1146/annurev-chembioeng-060713-040321
ically reduce the time and cost to develop new carbon capture technologies.
Copyright  c 2014 by Annual Reviews. The Carbon Capture Simulation Initiative is a partnership among national
All rights reserved

301
CH05CH14-Miller ARI 19 May 2014 17:25

laboratories, industry, and universities that is developing, demonstrating, and deploying a suite of
such tools, including basic data submodels, steady-state and dynamic process models, process op-
timization and uncertainty quantification tools, an advanced dynamic process control framework,
high-resolution filtered computational-fluid-dynamics (CFD) submodels, validated high-fidelity
device-scale CFD models with quantified uncertainty, and a risk-analysis framework. These tools
and models enable basic data submodels, including thermodynamics and kinetics, to be used within
detailed process models to synthesize and optimize a process. The resulting process informs the
development of process control systems and more detailed simulations of potential equipment to
improve the design and reduce scale-up risk. Quantification and propagation of uncertainty across
scales is an essential part of these tools and models.
Annu. Rev. Chem. Biomol. Eng. 2014.5:301-323. Downloaded from www.annualreviews.org
Access provided by University of Saskatchewan on 03/05/17. For personal use only.

INTRODUCTION
The Carbon Capture Simulation Initiative (CCSI) is a partnership among national laboratories,
industry, and academic institutions that is developing, demonstrating, and deploying state-of-the-
art computational modeling and simulation tools to accelerate the commercialization of carbon
capture technologies from discovery to pilot scale, demonstration, and ultimately widespread de-
ployment to hundreds of power plants. The CCSI Toolset is a comprehensive suite of scientifically
validated multiscale models and computational tools to enable uncertainty quantification (UQ),
optimization, and risk analysis. The CCSI Toolset incorporates commercial and open-source
software currently used by industry with new software tools to fill technology gaps. The CCSI
Toolset (a) enables promising concepts to be more quickly identified through rapid computational
screening of processes and devices; (b) reduces the time to design and troubleshoot new devices
and processes by using optimization techniques to focus development on the best overall process
conditions and by using detailed equipment models to better understand and improve the internal
behavior of complex equipment; (c) quantifies the technical risk in taking technology from labo-
ratory scale to commercial scale by understanding the sources and effects of model and parameter
uncertainty; and (d ) will ultimately stabilize deployment costs more quickly by providing industry
with a more fundamental understanding of the entire capture process. The computational tools
and multiscale modeling techniques comprising the CCSI Toolset can be broadly applied for the
development of a wide variety of technologies well beyond carbon capture, including chemicals
production, petroleum refining, natural gas processing, and biofuel production.
At the inception of CCSI, a goal of the US Department of Energy (DOE) was to support
research and development, as well as to pilot carbon capture and storage (CCS) projects, so that
widespread deployment of CCS could begin in eight to ten years (1). CCS is a key element in the
strategy to reduce atmospheric emission of CO2 , the principal greenhouse gas linked to climate
change (2, 3). The fastest way to deploy carbon capture technology is to scale up existing technolo-
gies, such as amine scrubbing, to the capacity required for use in a power plant and to deploy the
technology to the hundreds of existing power plants; however, both the cost of existing technolo-
gies and the time required to develop new technologies were unacceptable. Estimates showed that
scaling up amine scrubbing could increase the cost of electricity by as much as 80% in new pul-
verized coal (PC) power plants, while reducing the power plant’s net efficiency by approximately
30% (4). Based on prior experience, alternative technologies that might be less expensive could
take up to 15 years of time to move from the laboratory to predeployment (5) and 20 to 30 years
for industrial-scale deployment because of the need for multiple pilot- and demonstration-scale
projects, each with its own design, construction, learning, and operation phases (6).

302 Miller et al.


CH05CH14-Miller ARI 19 May 2014 17:25

Advanced multiscale modeling and simulation have the potential to dramatically reduce this
development time. Computational science techniques have been used previously to reduce the cost
and time for developing energy technologies (7, 8). CCSI extends this concept by enabling the
development of complete and consistent multiscale models from basic data to devices to the entire
process. Integrating models at various scales prevents knowledge loss and ensures that models are
consistent across scales. It also allows uncertainty to be propagated across scales to ensure that
both the true uncertainty and the source of that uncertainty are known.
Information from science-based models with quantified uncertainty will also help inform de-
cision makers, who must make large capital investments to bring innovative energy technologies
to market (5). By validating the models at multiple scales, quantifying the uncertainty of model
predictions, and estimating the technical risk, companies can plan more effective demonstrations
that can accelerate development by several years, resulting in significant cost savings. For example,
Annu. Rev. Chem. Biomol. Eng. 2014.5:301-323. Downloaded from www.annualreviews.org

our rough estimates show that reaching a large-scale demonstration five years earlier would result
Access provided by University of Saskatchewan on 03/05/17. For personal use only.

in a cost savings on the order of $100 million. We estimate that avoiding rework with the help of
simulations in just one major device in a $600 million capture plant could help that plant reach its
design capacity six months earlier, resulting in cost savings of $30 million.
CCSI is led by the National Energy Technology Laboratory (NETL) and leverages the DOE
national laboratories’ core strengths in modeling and simulation, bringing together the best ca-
pabilities at NETL, Los Alamos National Laboratory, Lawrence Berkeley National Laboratory,
Lawrence Livermore National Laboratory, and Pacific Northwest National Laboratory. CCSI’s
industrial partners provide representation from the power generation industry, equipment manu-
facturers, technology providers, engineering and construction firms, and software vendors. CCSI’s
academic participants (Boston University, Carnegie Mellon University, Princeton University, the
University of Texas, and West Virginia University) bring unparalleled expertise in reactive multi-
phase flow, process synthesis and optimization, process control, and chemical reaction modeling
for energy processes.

CCSI APPROACH TO MULTISCALE MODELING


The CCSI Toolset consists of multiple computational tools and models organized around eight
technical areas: (a) basic data submodels, (b) steady-state and dynamic process models, (c) pro-
cess optimization and UQ, (d ) dynamic simulation and control, (e) high-resolution filtered
computational-fluid-dynamics (CFD) submodels, ( f ) validated high-fidelity CFD device models
with quantified uncertainty, ( g) risk analysis and decision making, and (h) crosscutting integration
tools. These categories correspond with many of the activities involved with the development
of a new chemical process, such as a carbon capture system. This includes the development of
submodels for basic data, such as thermodynamics and kinetics, and the use of those submodels
within process models to synthesize and optimize a complete process on which advanced control
strategies are then developed and evaluated. The optimized process design identifies the most
promising device configurations, which serve as the basis for more detailed, high-fidelity CFD
simulations of those devices. Results from these additional simulations are then used to refine
the overall process design. All the information is then pulled together to assess project risk and
determine whether or how to proceed. Unique within CCSI is the integration of UQ among the
simulation scales. To effectively design a new process with new technology, each of these activities
is important and must interact to achieve an efficient, cost-effective system.
Computational models are approximations of the actual physical system that result from using
an inexact set of governing equations, imprecise parameter values in the equations, finite-precision
computer arithmetic, nonzero algorithmic thresholds, and noisy observational data for calibration.

www.annualreviews.org • Carbon Capture Simulation Initiative 303


CH05CH14-Miller ARI 19 May 2014 17:25

As such, credible simulation capabilities must be equipped with the ability to quantify the confi-
dence in model predictions. The first step to quantify uncertainty associated with a model involves
identifying all relevant sources of uncertainty, for example, all the model parameters that are not
precisely known. In the second step, uncertainty information in the form of bounds and probability
distributions is prescribed for the identified sources. At this point, any available observational data
can be used to fine-tune the prescribed distributions, using, for example, Bayesian inference tech-
niques (9). Finally, samples are drawn from the uncertainty distributions, ensemble simulations
are launched using the sample values, and simulation results are analyzed to assess model output
uncertainties and parameter sensitivities. These steps constitute a systematic UQ methodology
that must be executed rigorously. In practice, the task of quantifying uncertainties can be further
complicated if observational (i.e., experimental) data are available at various levels (component,
subsystem, or full-system), or when there are nonnegligible discrepancies between model outputs
Annu. Rev. Chem. Biomol. Eng. 2014.5:301-323. Downloaded from www.annualreviews.org

and the associated experimental data set. In the former scenario, Bayesian inferences are applied
Access provided by University of Saskatchewan on 03/05/17. For personal use only.

in a hierarchical fashion to integrate data at all levels. In the latter case, more advanced Bayesian
inference methods are required to understand and correct the biases.
Because the ultimate goal of CCSI is to accelerate the commercial development of carbon
capture technologies, an industry advisory board has actively provided input to CCSI since its
inception in February 2011. Working closely with the industry partners ensures that the simulation
tools and models are relevant to real industry problems, can be readily used by industry, and
build on the capabilities of tools and modeling environments currently in use. An initial set of
components of the CCSI Toolset was released in September 2012 followed by an extended set in
October 2013. The final release is planned for January 2016 with interim releases scheduled in
2014 and 2015.

SOLID-SORBENT CARBON CAPTURE: A CASE STUDY


The capabilities of the multiscale CCSI Toolset are being demonstrated in the context of post-
combustion carbon capture technologies, which are broadly applicable to fossil energy power
production, especially PC power plants. PC power plants generate more than 40% of the elec-
tricity in the United States and emit approximately one-third of all CO2 from US sources (10). A
recent analysis suggests that roughly 325 coal-fired generating units, accounting for approximately
two-thirds (200 GW) of current US coal-based generating capacity, are suitable for retrofit with
postcombustion carbon capture systems (11).
The solid-sorbent-based postcombustion capture technology was selected to demonstrate the
capabilities of the CCSI Toolset. Although DOE/NETL is sponsoring several sorbent develop-
ment efforts, significant work remains to define and optimize the reactors and processes needed for
successful sorbent capture systems. Sorbents offer a potential advantage because they can reduce
the regeneration energy associated with CO2 capture and reduce parasitic load compared with
conventional, aqueous amine-based solvent-capture processes. Most of the work on sorbents has
been restricted to developing the sorbent itself (12); only very recently have studies considered the
design of the reactor system and integration with the power plant (13). Thus, at the start of the
initiative, solid-sorbent systems were just beginning the traditional process development cycle and
could most benefit from application of the CCSI Toolset. Working with industry partners, re-
searchers are using the Toolset to accelerate the scale-up from lab scale to 25-MWe demonstration
scale.
The second capture technology selected to demonstrate the Toolset is liquid solvents. Although
detailed device simulations for solid sorbents are much different than those used in liquid solvents,
the required computational capabilities and interfaces share many common features (e.g., ability

304 Miller et al.


CH05CH14-Miller ARI 19 May 2014 17:25

to handle multiphase devices). The extension of the Toolset to solvent-based capture systems
focuses on second- and third-generation solvents.

Basic Data and Uncertainty Quantification


The first step in using the CCSI Toolset is to develop the basic data submodels necessary to
represent the chemistry at the heart of the CO2 capture process. From a modeling and simula-
tion standpoint, this means anything from the quantum scale to the scale of continuum kinetics
and transport associated with chemical reactions. Such data can be generated computationally
or experimentally. Kim et al. (14) demonstrated the use of computationally generated material
properties for initial screening of materials by linking the CCSI Toolset with a materials database.
For the most promising materials, data are obtained experimentally from techniques such as
Annu. Rev. Chem. Biomol. Eng. 2014.5:301-323. Downloaded from www.annualreviews.org

thermogravimetric analysis, fixed-bed, or other bench-scale methods. The physical and chemical
Access provided by University of Saskatchewan on 03/05/17. For personal use only.

properties of the sorbent used throughout this study are based on experimental data collected for
a polyethylenimine-impregnated mesoporous silica sorbent developed at NETL (15).
As reflected in a recent review (16), a common strategy for considering quantum chemical and
empirical data in models of chemically reacting systems is to use quantum chemical calculations
to estimate key physical parameters, such as activation energies, in microkinetic and statistical
mechanical models of chemical processes. Often, experimental data, when available, or when
the complexity of quantum and atomistic calculations presents a hurdle, are employed to infer
parameter estimates or establish semiempirical approaches.
UQ in parameters, models, and scale-bridging methods is a relatively new area of study and is a
major emphasis of CCSI. The entire enterprise of multiscale modeling can be recast as an exercise
in uncertainty minimization using a Bayesian paradigm (9). The most important aspect of this
is the view that all quantities in a model, including model parameters and model output, can be
considered as random variables with associated probability distributions. The inherent uncertainty
pertaining to quantum calculations of thermodynamic or conventional, mean-field kinetic model
parameters may therefore be captured in the form of a probability distribution. A straightforward
way to construct such a distribution is to perform several quantum chemical calculations using a
variety of different approximations, chemical environments, and conformations for a given reaction
and to fit a standard-form distribution (e.g., Gaussian or inverse gamma) to the results.
Additional uncertainty comes in the form of the model itself, which inevitably contains ap-
proximations or omissions relative to the actual process chemistry. This uncertainty is accounted
for through a stochastic function called the model form discrepancy. The discrepancy function is
generally a Gaussian process (GP), the purpose of which is to add variability to the model such that
the validation (i.e., experimental) data can be reproduced (aside from observational error) using
a model-plus-discrepancy construct. The discrepancy function is calibrated to the experimental
data along with the model parameters through the Bayesian framework, yielding a new stochastic
tool for model predictions that takes both first-principles theory and bench-scale experimental
data into account.
Formally, the Kennedy–O’Hagan (KOH) calibration framework (17) is provided by the model
Zi = Y (ζi ; θ) + δ(ζi ) + ∈i , i = 1, . . . , n, 1.

where Zi is the ith experimental observation (e.g., rate of reaction), Y is the physical model (e.g.,
rate expression), θ is the model parameters (e.g., pre-exponential factor, activation energy), ζ i is
the experimental inputs (e.g., pressure, temperature), δ is the model form discrepancy, and ∈i is
the ith observation error (usually assumed to be normally distributed with mean zero and variance
σ 2 ). The unknown parameters θ , δ, and σ 2 are estimated in a Bayesian framework. That is, a prior

www.annualreviews.org • Carbon Capture Simulation Initiative 305


CH05CH14-Miller ARI 19 May 2014 17:25

distribution π (θ, δ, σ 2 ) is defined for the unknown parameters: Typically a GP is assumed for δ,
and expert knowledge and/or previous experience is used to determine a prior distribution for θ
and σ 2 . The posterior distribution of the parameters, given the data vector Z, π (θ, δ, σ 2 |Z), is
then approximated via Markov chain Monte Carlo sampling (18). This approach is well known
and widely used to calibrate models of physical systems.
Of paramount importance for multiscale modeling is the use of the discrepancy to provide
stochastic predictions at larger length scales. As such, it is critical that the discrepancy reflect
what is known about the mathematical form of the physical model. For equilibrium or steady-
state models, this is straightforward, but kinetic models present distinct challenges in this regard,
because the dynamic nature of a kinetic model must be reflected in the form of the discrepancy.
This requires that we move beyond the KOH approach to a dynamic discrepancy methodology.
The experimental observations Zi are time-dependent curves in this context and follow the model
Annu. Rev. Chem. Biomol. Eng. 2014.5:301-323. Downloaded from www.annualreviews.org

δxi
Access provided by University of Saskatchewan on 03/05/17. For personal use only.

(t) = f [xi (t), ζi (t); θ ] + δ[xi (t), ζi (t)], i = 1, . . . , n


δt
Zi (t) = g[xi (t)] + ∈i 2.
for the ith observed adsorbed-state xi with inputs ζ i (representing time-varying quantities, such
as temperature and gas composition), and in which g is a function that transforms model output
into the experimentally observable quantity (in this case, weight gain of the sorbent). Equation 2
imbues the discrepancy with a dynamic character that is missing from the KOH formalism. By
using a specially constructed GP for the discrepancy in Reference 4—known as Bayesian smooth-
ing spline–analysis of variance (BSS-ANOVA) (19)—the function δ can be decomposed into
an expansion of deterministic basis functions φ with stochastic coefficients β; in other words,

δ[x(t), ζ (t)] = β j φ j [x(t), ζ (t)]. The solution of the stochastic differential equation can then be
integrated seamlessly into the Markov chain Monte Carlo routine. The result of the calibration is a
joint posterior (i.e., consistent set of distributions) for model parameters, θ, stochastic coefficients,
β, and variance, σ 2 .
Because the uncertainty in the model form is captured in the parameter set, β, propagation of
uncertainty from the kinetic to the process scale becomes a straightforward exercise. Uncertainty
can then be minimized with the help of process-scale models. For example, propagation of kinetic
model predictions to a steady-state process model leads to a distribution of input trajectories, ζ .
Those input trajectories can be used to design a new set of experiments that will be closer to
the environment predicted at the process scale, thus minimizing extrapolation and the resulting
uncertainty. Figure 1 illustrates calibration and model upscaling of a hypothetical CO2 sorbent
and process model. As part of the calibration process, Figure 1a shows the bivariate marginal of
a posterior distribution for two physical parameters in a hypothetical kinetic model calibrated to
simulated experimental data where the blue dot is reality arising from a model pertaining to a
more complicated sorbent chemistry. When this calibrated kinetic model is incorporated into a
process model, the result is a set of profiles, such as the temperature profiles shown in Figure 1b,
in which the black line is reality.

Process Models
Industry uses commercial process simulators to aid in the design and analysis of various processes.
These packages include process models that accurately capture many aspects of vapor-liquid in-
teraction; however, process models for most solid-vapor interactions are not available. Thus, to
synthesize a process for a solid-sorbent-based carbon capture system consisting of a CO2 adsorber
and a sorbent regenerator, a suite of process models has been developed for equipment options

306 Miller et al.


CH05CH14-Miller ARI 19 May 2014 17:25

a b
5 350

4
340
θ1 (equilibrium)

Temperature (K)
3
330
2

Reality
320
1 Model realizations
Annu. Rev. Chem. Biomol. Eng. 2014.5:301-323. Downloaded from www.annualreviews.org

0 310
–120 –100 –80 –60 0 50 100 150
Access provided by University of Saskatchewan on 03/05/17. For personal use only.

θ2 (equilibrium) Time (s)

Figure 1
(a) Bivariate marginal of a posterior distribution for two physical parameters in a hypothetical kinetic model resulting from the
calibration to simulated (i.e., a known truth case) experimental data. The blue dot is “reality” arising from a model pertaining to a more
complicated sorbent chemistry. (b) A sampling from the posterior distribution of temperature profiles obtained upon forward
propagation of the posterior uncertainty of the calibrated kinetic model into a process model. The black line is “reality.” Data for the
figures courtesy of Sham Bhat, Los Alamos National Laboratory.

determined to be most suitable for large-scale capture. Models for both bubbling fluidized bed
(BFB) reactors and moving bed (MB) reactors were developed (20–24). BFB reactors are promis-
ing because they provide high mass- and heat-transfer rates owing to a large contact area (25), and
they are widely used in a range of industrial processes. MB reactors are promising because they
provide nearly countercurrent contacting that can improve working capacity of the sorbent as it
is cycled between the adsorber and regenerator. The models specifically consider thermal swing
processes and can function as either adsorbers or regenerators.
The models include both steady-state and dynamic modes to support process design as well
as process control. The models capture sufficient detail to provide accurate and predictive results
while remaining computationally tractable. They consist of a system of partial differential equations
(PDEs), which capture the hydrodynamic behavior, interactions of the gas and solids, heat- and
mass-transfer phenomena, and kinetics of the adsorption and desorption reactions.
The BFB reactor model is a 1D PDE-based, two-phase, nonisothermal model, incorporating
heat-transfer tubes to provide external heating and cooling capability (21, 22). It can be used to
simulate overflow-type configurations in which the solids leave the bed by flowing over a weir or
underflow-type configurations in which the solids leave from the bottom of the bed.
The MB reactor model is a 1D PDE-based, two-phase, nonisothermal model with heat-transfer
tubes (24). When used as a regenerator, an integral heat-recovery system heats the incoming
solid sorbent using steam that is produced by recovering heat from the hot sorbent leaving the
regenerator (23). The reaction kinetics, heat and mass transfer, and hydrodynamics are considered
in both the MB reactor and heat recovery system.
Because the high molecular weight of CO2 invalidates many compressor heuristics, process
models of both inline- and integral-gear multistage compressors, with interstage coolers and
knockout drums as appropriate, have been developed (20, 23) for simulating the compression of
CO2 from near-atmospheric pressure to a pipeline pressure of 2,200 psia. For applicability in
cases with wide variations in Mach numbers, nondimensional performance curves in terms of

www.annualreviews.org • Carbon Capture Simulation Initiative 307


CH05CH14-Miller ARI 19 May 2014 17:25

the impeller exit flow coefficient and polytropic head coefficient (26) have been generated using
vendor data. A model of a triethylene glycol absorber has also been developed and used to satisfy
the desired water content in the compressed CO2 for pipeline transport. The operating cost for
the integrated solid-sorbent process can be calculated easily by considering the power and utilities
requirements available from the process models.

Process Design and Optimization


When screening potential concepts, it is important to estimate their economics in the context
of a complete process that has been optimized for the particular characteristics of the materials
under evaluation. This process synthesis activity develops a set of interconnected equipment and
material and energy flows that achieves the desired specifications at minimal cost while meeting
Annu. Rev. Chem. Biomol. Eng. 2014.5:301-323. Downloaded from www.annualreviews.org

environmental, safety, and operating constraints (27). The technical and economic performance
Access provided by University of Saskatchewan on 03/05/17. For personal use only.

characteristics of a new technology are strongly dependent on the effectiveness of this process
synthesis activity. For many chemical processes, heuristic and evolutionary approaches to process
synthesis have been adequate because of their similarity to existing processes. However, large-
scale carbon capture processes are outside of current experience, and evolutionary approaches are
not suitable for accelerated technology development. Thus, the development of these processes
can especially benefit from a rigorous approach that uses a superstructure (i.e., a superflowsheet
that embeds structural and equipment alternatives as well as interconnections) (28) in conjunction
with advanced optimization tools (29) to identify the most cost-effective process configuration and
equipment interconnections.
Although conceptually simple, superstructure-based process-synthesis approaches are mathe-
matically complex and difficult to properly formulate and solve. They require models of the likely
process equipment that can predict performance over a wide range of conditions. Such models can
be created most easily using a process simulation tool that provides access to rigorous thermody-
namics packages. The downside of using these highly accurate models within a process simulator is
that their direct optimization is cumbersome and does not come with the guarantees and efficien-
cies of derivative-based optimization using algebraic models. This is because process simulators
act as a black box, determining outputs from a given set of inputs. Thus, to overcome this diffi-
culty, we have developed a tool to automatically convert process simulation models (such as those
discussed in Process Models, above) into surrogate algebraic models that have the appropriate
mathematical characteristics necessary for solving a superstructure formulation.
CCSI uses a two-stage, hybrid approach to process synthesis. In the first stage, the superstruc-
ture is solved with the help of surrogate models to identify the optimal process configuration based
on an estimated cost of electricity or cost of capture, which incorporates both capital and operating
costs. Then, rigorous process models are used in conjunction with simulation-based, derivative-
free optimization (DFO) algorithms to refine the process design (i.e., dimensions) and operating
conditions. This is a much smaller, more tractable optimization problem because it starts with a
known configuration and uses information from the superstructure solution as an initial solution.
It is important to reoptimize the system (i.e., dimensions and operating conditions) because the
superstructure solution is based on approximate surrogate models.
Estimating the cost of electricity or the cost of CO2 capture requires estimating capital and
operating expenses based on the actual process configuration, equipment, and operating expenses.
Because commercial entities often have their own proprietary data to estimate capital costs and have
different preferred methodologies, the CCSI approach is to use data from the open literature, such
as that found in Reference 30, and follow US DOE guidelines for estimating cost of electricity (31).
These are implemented in such a way that they can easily be modified to incorporate proprietary

308 Miller et al.


CH05CH14-Miller ARI 19 May 2014 17:25

correlations, approaches, and data. When linking process model data (such as diameter, heat-
transfer area, and pressure) to the capital cost-estimation models, users must specify appropriate
materials of construction to account for design temperature and pressure as well as potential
corrosivity of the materials. Corrosion, which is not an explicit part of the models, is also addressed
by the risk-analysis framework.
The majority of the operating costs of a carbon capture system result from using steam from
the PC plant’s steam cycle that would otherwise generate electricity. Thus, in order to properly
estimate the cost of electricity, it is necessary to have a model of the PC plant steam cycle or
empirical relationship that can estimate the parasitic power loss as steam is withdrawn (20).

Surrogates for optimization. The emergence of chemical process simulators over the past three
decades (32–34) has increased the range of chemical processes that can be accurately modeled,
Annu. Rev. Chem. Biomol. Eng. 2014.5:301-323. Downloaded from www.annualreviews.org

predicted, analyzed, and designed. As complexity of these simulators grows, so does the challenge
Access provided by University of Saskatchewan on 03/05/17. For personal use only.

associated with using them to optimize complex systems. Classical optimization techniques
require derivatives, which are difficult to calculate from proprietary simulators because equipment
and thermodynamic models appear to users as black boxes. Furthermore, as simulations grow
in complexity, they often become difficult to converge and may require a significant amount of
computing time to solve the underlying first-principles or empirical models. DFO algorithms
have been developed to cope with such challenges, but their success has been limited to relatively
small problems (35).
In the first stage of the CCSI approach, detailed process models developed in process simulators
(see the section on Process Models) are converted to surrogate algebraic models that can be used
for efficient optimization. This idea has been explored in many different forms in the literature
of experimental design, response surfaces, metamodeling, reduced-order modeling, and machine
learning (36). A new approach has been developed that provides algebraic models that are accu-
rate but simple and do not require an excessive number of simulations to compute. Metamodels
based on neural networks, kriging models, and the like are capable of accurately approximating
sophisticated, first-principles models but are fairly complex and do not offer accurate methods
for determining their derivatives. In contrast, our algebraic models are much simpler, thus lend-
ing themselves to optimization, especially when discrete decisions and complex flowsheets are
involved. An equally important feature of the proposed methodology is that it offers a systematic
approach to interrogate a surrogate model, identify parts of the domain where the model is inac-
curate, and add more simulation points to the training set to improve the model’s accuracy. The
complete methodology is detailed in Reference 37, and an overview is given in this section.
Rather than developing and simulating a complete flowsheet within a process simulator, the
major equipment items are simulated individually over a range of potential operating and de-
sign parameters. Whereas a large and complex process flowsheet may be difficult to converge, the
smaller models involve fewer variables and are far easier to converge, which drastically reduces the
computing time required to develop surrogate models (38, 39). The surrogate models thus derived
for the individual equipment items will later be combined within an overall optimization formu-
lation, ensuring that components act in concert and satisfy all relevant connectivity requirements.
The following steps are used to develop the surrogate model:

1. Run a small number of simulations over the domain of the independent (input) variables
based on Latin hypercube sampling or any preferred design of experiments technique.
2. Build a low-complexity surrogate by using an integer programming model to select a small
number of basis functions out of a very large number of possible functional forms (e.g.,
polynomial, multinomial, exponential, logarithmic, or expected bases from experience or

www.annualreviews.org • Carbon Capture Simulation Initiative 309


CH05CH14-Miller ARI 19 May 2014 17:25

physical phenomena, such as the log mean temperature difference) to add to the model. The
integer program makes this choice without enumerating all possible combinations. Contrary
to forward or backward regression techniques, the proposed approach captures synergies
between different potential functional components.
3. Use a DFO solver to sample the domain of the input variables and find points that maximize
the error between the current model and the simulator. If this error is small, stop; otherwise,
repeat steps 2 and 3.
In step 2, the integer program is solved by allowing increasingly more functional forms in the
model. The corrected Akaike information criterion (40) is computed for each number of terms,
and the progression is terminated once a local minimum of this criterion has been identified. This
criterion balances model complexity with fitting error, thus avoiding overfitting.
Annu. Rev. Chem. Biomol. Eng. 2014.5:301-323. Downloaded from www.annualreviews.org

ALAMO, the computational implementation of the above methodology, outperforms several


machine-learning techniques, including simple least-squares regression and the lasso method (37).
Access provided by University of Saskatchewan on 03/05/17. For personal use only.

In particular, it requires fewer simulations to obtain models of the same accuracy. Moreover, it
produces models that are significantly simpler than the other techniques. This simplicity of the
resulting surrogate models is highly advantageous when the models are used for optimization and
is key to solving the difficult mixed-integer problems that arise when integer variables are used to
determine an optimal combination of units in a process flowsheet.
Surrogate models for BFB adsorbers and regenerators with up to 15 input variables and
15 output variables have been generated. The input/output variables for these units modeled
flow conditions (concentrations, temperatures, and pressures of input/output streams) as well as
geometry of the units (length, diameter, number of heat-exchanger tubes). After an initial data set
of 150 sample points was obtained by simulating each unit, ALAMO performed adaptive sampling
that required an additional 150 simulations to derive surrogates that were deemed accurate.

Superstructure. The surrogate models of the adsorber and regenerator were coupled with first-
principles models of compressors and heat exchangers to develop a process superstructure model
for solid-sorbent-based CO2 capture that includes a number of adsorbers in countercurrent series
and a number of regenerators in countercurrent series along with required heat exchangers and
compressors. The superstructure model includes models for estimating capital and operating costs
and combining them to calculate the resulting cost of electricity. This superstructure provides
an optimization-based approach for process synthesis, determining the appropriate selection of
equipment among alternatives and balancing operating and capital costs.
The superstructure includes binary variables to determine the number and type of adsorbers
and regenerators to use (both the BFBs and MBs were considered as alternatives) and continuous
variables to model equipment geometry and flow conditions. The resulting mixed-integer nonlin-
ear optimization model was solved with GAMS/BARON (41) to obtain an optimal configuration
for the flowsheet along with corresponding operating conditions (20).

Simulation-based optimization (Turbine). Simulation-based optimization (42–46) is an


emerging field that integrates optimization with rigorous models of equipment or processes.
A simulation-based optimization framework has been developed, which consists of tools for in-
terfacing with a variety of process simulators, connecting simulations from different simulation
packages, linking with cost-estimation tools, storing and managing simulation data and results,
and performing DFO to optimize complicated process simulations, which are not capable of pro-
viding reliable derivatives (35, 47). The framework consists of a metaflowsheet that allows connec-
tions between different simulation modules, including recycle among modules. The framework is

310 Miller et al.


CH05CH14-Miller ARI 19 May 2014 17:25

capable of running parallel simulations of the metaflowsheet. Most DFO methods can be readily
adapted to parallel-objective function evaluation. Currently, two DFO methods are available in
the framework: CMA-ES (48) and SNOBFIT (49).
The simulation-based optimization approach is applied to a process configuration determined
using the superstructure-based process synthesis approach described in the previous section. Most
of the physical constraints on the size and behavior of the process are contained within the bounds
placed on the input variables for the optimization. In addition, the gas velocity within the BFB
adsorbers must remain within the bubbling or turbulent fluidization regime for the reactor model
to be applicable, and the minimum temperature approach in the solid heat exchanger is enforced.
To run the large number of simulations required for DFO and manage the resulting data, an
execution gateway named Turbine was developed (50). Turbine runs the simulations on a variety
of computing resources, including Amazon Cloud instances, clusters, and the local workstation
Annu. Rev. Chem. Biomol. Eng. 2014.5:301-323. Downloaded from www.annualreviews.org

where the optimization is taking place. It automatically detects and stops simulations that are failing
Access provided by University of Saskatchewan on 03/05/17. For personal use only.

to converge in a reasonable time and restarts simulations that do not run to completion. During
the optimization process, the gateway returns the results to the optimization framework, which
calculates the value of the objective function, such as cost of electricity. It then identifies new values
for the sets of variables according to the underlying DFO algorithm. This execution capability also
supports the application of UQ to the resulting process simulation. Turbine regularly manages
thousands of parallel simulations in support of optimization and UQ.
Algebraic surrogate models of the process models (see Process Models, above) incorporating
basic data submodels (see Basic Data and Uncertainty Quantification) were developed and used
within a superstructure for process synthesis. The resulting cost optimal process configuration was
further optimized by DFO to develop the CCSI process, which served as the basis for developing
and demonstrating the capabilities of the other components of the CCSI Toolset (20).

Uncertainty Quantification
When applying UQ to the CCSI process, the goal is to quantify the uncertainty in the CO2 capture
efficiency as well as uncertainties in the overall costs. The relevant sources of uncertainties include
the parameters for heat and mass transfer and the chemical kinetics submodels. Several model
assumptions were made to simplify the analysis, including the validity of the BFB model (21) and
the adequacy of the kinetics model (15). Uniform distributions were initially used for all parameters
with uncertainty bounds of +/−20%.
With 33 uncertain parameters at the process level, an initial screening was performed to identify
a smaller subset of parameters that accounted for the bulk of the output uncertainties. Initial anal-
ysis showed that the reaction parameters for the formation of carbamate are the most important
and that most nonreaction parameters had a negligible effect on the output uncertainty. As a re-
sult, a subset of 13 nonnegligible parameters was used for a subsequent study. Moreover, because
the reaction parameters were most significant, their uncertainty distributions were refined us-
ing Bayesian calibration techniques based on available experimental data. Comparison of the CO2
capture uncertainties with and without this calibration clearly demonstrates the benefit of incorpo-
rating experimental data when available, as shown in Figure 2. On the left, the large uncertainty in
overall CO2 capture is a result of overestimating the uncertainty in the carbamate reaction. When
using the calibrated reaction parameter uncertainty, the results show much less overall uncertainty.
Upon developing a better and smaller uncertainty related to the reaction parameters, the
screening process should be rerun to ensure that significant parameters were not inadvertently ex-
cluded. In this case, with the refined reaction parameter uncertainty, several fluidization parameters
(i.e., minimum fluidization velocity, bubble diameter, and bubble velocity) become nonnegligible.

www.annualreviews.org • Carbon Capture Simulation Initiative 311


CH05CH14-Miller ARI 19 May 2014 17:25

Assuming +/–20% of default input values After calibrating for reaction parameters
0.14 0.16
Mean = 72.36, standard deviation = 20.48 Mean = 88.48, standard deviation = 3.12
Uncertainty bounds = [0, 99.99] Uncertainty bounds = [75.81, 94.14]
0.14
0.12

0.12
0.1

0.1
Probabilities

Probabilities
0.08

0.08

0.06
Annu. Rev. Chem. Biomol. Eng. 2014.5:301-323. Downloaded from www.annualreviews.org

0.06
Access provided by University of Saskatchewan on 03/05/17. For personal use only.

0.04
0.04

0.02
0.02

0 0
0 50 100 0 50 100
ADS_Cap_CO2 ADS_Cap_CO2

Figure 2
Comparison between computed distributions of CO2 capture percentage.

Dynamics and Process Control


The design and operation of clean energy plants using carbon capture offer unique multiscale
modeling and control challenges (51). Effectively controlling CO2 capture rates in response to
plant disturbances and fluctuating electricity and carbon prices is required to achieve significant
reductions in the energy and efficiency penalties of capture (52, 53). The development of dynamic
modeling and control tools that will have a transformative impact on the deployment and use of
capture technologies requires new breakthroughs in rate-based (54) and distributed-parameter,
PDE-based dynamic modeling, dynamic model reduction, multiscale control (55, 56), and dynamic
UQ. CCSI is addressing these computational challenges and requirements for postcombustion
CO2 capture and compression processes, as well as their integration with PC power plants.
For maintaining process stability and CO2 capture rates under disturbances in flue gas flowrate
and composition, linear-model predictive control strategies were developed and found to be su-
perior to conventional proportional-integral-differential and feedback-augmented feed-forward
controller technologies (22, 23). In addition, surge detection and control algorithms have been
developed and implemented for the CO2 compression system. The surge-detection algorithm
uses dimensionless numbers for impeller exit flow coefficient and polytropic head coefficient. For
surge control, the gain-scheduling controller uses an adaptive λ-tracker control law (57, 58) that
drives the output to zero with a prescribed tolerance level.
Multiscale dynamic capture models are often too computationally expensive for use in real-time
applications, such as operator training and online process control. As a result, innovative methods
are required to reduce the complexity of PDE-based and rate-based dynamic models while pre-
serving input-output behavior. Dynamic reduced models (D-RMs) approximate the high-fidelity

312 Miller et al.


CH05CH14-Miller ARI 19 May 2014 17:25

capture models, thereby offering a trade-off between accuracy, range of applicability, and compu-
tational cost. CCSI researchers are developing a D-RM builder tool that automatically generates
fast D-RMs using precomputed results from repeated simulations of high-fidelity dynamic pro-
cess models (see Process Models, above) over a range of input and output variables. The D-RM
builder currently offers two data-driven black-box approaches: a nonlinear, autoregressive moving
average based on neural networks and the decoupled A-B net method (59) that combines a linear
state-space model with a nonlinear mapping from state-space to outputs using a neural network.
CCSI research is also focused on more complex state-space reduction methods, such as proper
orthogonal decomposition.
Future CCSI efforts in this area will be focused on the development of dynamic nonequilibrium
stage models with UQ capabilities, additional D-RM tools, and advanced process-control capabil-
ities for second-generation (e.g., high-viscosity) solvent-based CO2 capture systems. In addition,
Annu. Rev. Chem. Biomol. Eng. 2014.5:301-323. Downloaded from www.annualreviews.org

CCSI is also performing innovative research on dynamic UQ by considering uncertainties in the


Access provided by University of Saskatchewan on 03/05/17. For personal use only.

transients of states, measurements, and parameters.

Filtered Computational-Fluid-Dynamics Submodels for Gas-Particle Flows


Gas-particle flows in the BFB and MB reactors of the CCSI carbon capture system are inherently
chaotic, characterized by small-scale particle clusters spanning 10–100 times the particle diameter
(60). To obtain accurate CFD predictions, the grid size should be sufficiently small to resolve
these clusters. Device-scale simulations on such highly resolved grids are generally prohibitively
expensive.
To balance computational cost and accuracy, a multiscale CFD approach is proposed to simu-
late commercial-scale devices. New filtered constitutive relationships that model the influence of
fine clusters via subgrid corrections were developed (61–63), which were implemented in coarse-
grid simulations to make them accurate—similar to the subgrid closures used in large-eddy sim-
ulations of turbulent single-phase flows.
Results obtained from highly resolved simulations of model flow problems, each representing
possible flow conditions occurring in the full-scale device, are filtered (averaged) to construct
subgrid corrections (Figure 3). Closures were developed for the filtered gas-particle drag, solid-
phase stress and viscosity (61, 64, 65), heat-transfer rate (66), and reaction kinetics (67). The
filtered drag, heat-transfer, and reaction rates were found to be significantly lowered by clustering
inhomogeneities, sometimes by several orders of magnitude, which makes these filtered models
essential for obtaining accurate CFD predictions.
Previous work has focused on constructing filtered models for risers and fluidized beds without
immersed objects, whereas the CCSI capture devices contain arrays of cooling tubes too small in
diameter to explicitly resolve in device-scale simulations. Therefore, filtered drag models appli-
cable to flows with an immersed array of small cylinders are being developed (68). The presence
of the tube array is directly manifested as a drag force exerted on the suspension by the cylinder.
The closure for filtered cylinder-suspension drag has a familiar kinetic component proportional to
velocity squared (expected from flow past a single cylinder) and a contribution owing to buoyancy.
The presence of cylinders also affects the clustering behavior of sorbents: Gas-particle segre-
gation is enhanced around the cylinders in the denser flow regions, resulting in a further reduction
of the filtered gas-solid drag (Figure 3). The filtered models developed by Igci & Sundaresan (64)
were modified accordingly to account for the immersed cylinder array (68), where the cylinder
array is represented as an effective, stationary porous medium.
Verification and validation studies demonstrate that filtered models improve the predic-
tion of macroscopic quantities such as solids holdup and pressure variation at a fraction of the

www.annualreviews.org • Carbon Capture Simulation Initiative 313


CH05CH14-Miller ARI 19 May 2014 17:25

10
Filtered drag coefficient, no cylinders
Filtered drag coefficient with cylinders
8 Dcyl = 10 mm, acyl = 100 mm

gas–solid drag coefficient


Dcyl = 20 mm, acyl = 100 mm
Dimensionless filtered Dcyl = 31 mm, acyl = 100 mm
Dcyl = 41 mm, acyl = 100 mm
6
Dcyl = 31 mm, acyl = 75 mm
Dcyl = 31 mm, acyl = 150 mm 6

0.64
2 5 0.60
Annu. Rev. Chem. Biomol. Eng. 2014.5:301-323. Downloaded from www.annualreviews.org

4 0.50
3
Access provided by University of Saskatchewan on 03/05/17. For personal use only.

1 2 Particle 0.40
volume 0.30
0 fraction
0 0.1 0.2 0.3 0.4 0.5 0.6 0.20
Filtered solid volume fraction 0.10

0.00
1 2 3 4 5 6

Figure 3
Dimensionless gas-solid drag coefficient as a function of filtered solid volume fraction for varying cylinder diameters (Dcyl ) and spacing
(acyl ), shown with snapshots of the highly resolved computational-fluid-dynamics simulations used to construct the filtered closures.
The filtered drag models without and with immersed cylinder arrays are presented. The filtered drag coefficients for both cases are
considerably smaller than the corresponding unfiltered values.

computational cost of fine-grid simulations (69, 70). Refinement of the filtered models is ongoing,
and their predictive capabilities are continually being enhanced (65). Current work focuses on veri-
fication of the filtered corrections for immersed cylinder arrays, quantification of uncertainty asso-
ciated with the use of these filtered models, and the development of heat-transfer filtered models.

High-Fidelity Computational-Fluid-Dynamics Models with Hierarchical,


Uncertainty Quantification–Based Validation
The development of validated device-scale modeling capabilities is critical for reducing the time
and cost of scale-up and optimization of the solid-sorbent-based capture technology. Such models
enable device-scale performance for a given reactor configuration to be assessed using advanced
multiphase CFD simulations to obtain the spatially and temporally dependent device performance
predictions. Filtered constitutive relationships derived in the previous section have been imple-
mented in coarse-grid device-scale CFD simulations to account for the fine clusters and localized
flow inhomogeneities.
The results from these high-fidelity device-scale CFD simulations will enable system engineers
to derive accurate, reduced-order models for plant-scale synthesis and performance analyses; to

314 Miller et al.


CH05CH14-Miller ARI 19 May 2014 17:25

CCSI CFD validation hierarchy

25 MWe, 100 MWe,


650 MWe Demonstration-
solid-sorbent and full-scale
systems systems

1 MWe carbon Pilot-scale


capture system systems
Annu. Rev. Chem. Biomol. Eng. 2014.5:301-323. Downloaded from www.annualreviews.org

NETL carbon capture unit (C2U) Laboratory-scale


reacting unit subsystem
Access provided by University of Saskatchewan on 03/05/17. For personal use only.

(coupled
Bubbling bed adsorber Moving bed regenerator benchmark cases)

Laboratory-scale
Intermediate validation Intermediate validation
(adsorber without reactions and (regenerator without reactions
subsystem
heat transfer) and heat transfer) (decoupled
benchmark cases)

Upscaling Upscaling Upscaling Upscaling


(flow filtering) (reaction filtering) (energy filtering) (flow filtering)

Bubbling Reaction Heat Moving Unit


fluidized bed kinetic transfer fluidized bed problems
(adsorber) (regenerator)

Figure 4
The Carbon Capture Simulation Initiative (CCSI) computational-fluid-dynamics (CFD) validation
hierarchy, illustrating the various unit problems and the levels of validation that lead up to a quantitative
confidence on the predictions for the full-scale device systems (71). Abbreviation: NETL, National Energy
Technology Laboratory.

screen different device-scale designs; and to select a suitable final design. In addition, such device-
scale CFD models with adequate spatial resolution can also be used to optimize geometric and
operational parameters of the device or to use computational intuition to arrive at an entirely
novel device design.
To use the CFD-based device-scale predictions with quantitative confidence for devices that
are yet to be built, it is critical that the models are continually and methodically validated with
experimental data at various scales whenever possible, and that uncertainties at the various model
scales are characterized and quantified. A systematic, hierarchical approach is used to design
physical experiments (and/or use available measurement data sets) for comparison with predictive
computational models. This approach, which is depicted in Figure 4, divides the complex full-
scale solid-sorbent carbon capture system into several progressively simpler levels: pilot-scale
cases, laboratory-scale cases, and unit problems (71).
The lowest level of the hierarchy focuses on unit problems that are encountered in the adsorber
and regenerator and represent important physical phenomena of the system, such as fluid flow in

www.annualreviews.org • Carbon Capture Simulation Initiative 315


CH05CH14-Miller ARI 19 May 2014 17:25

a BFB. The second validation level addresses the effects of upscaling the physical models using
the filtered models in the section on submodels for gas-particle flows, above. The third and fourth
levels focus on laboratory-scale cases with decoupled physics (i.e., flow without reactions or heat
transfer) and coupled physics (i.e., flow with reactions and/or heat transfer). In these levels, the
CFD model predictions are compared with experimental data from the NETL Carbon Capture
Unit (72). The next level uses larger-scale pilot plants, such as a 1-MWe system, to investigate
the validity of the CFD models. Currently, there are no 1-MWe systems in operation. However,
one of CCSI’s industry partners is constructing a 1-MWe system that, when online, will be
used to validate a pilot-scale CFD simulation. The topmost level of the hierarchy is to deduce
quantitative confidence on the predicted quantities of interest for the larger device-scale systems
(25 MWe, 100 MWe, and 650 MWe). There are no plans to construct such a system within the
CCSI project life cycle. Therefore, the device-scale simulations will rely on the knowledge and
Annu. Rev. Chem. Biomol. Eng. 2014.5:301-323. Downloaded from www.annualreviews.org

expertise of the CCSI team as well as the quantitative prediction confidence gained through the
Access provided by University of Saskatchewan on 03/05/17. For personal use only.

smaller-scale CFD simulations and corresponding UQ performed on the lower validation tiers.
An example of a unit problem from the hierarchy is a BFB with immersed horizontal heat-
transfer tubes, which has been studied as part of the CCSI high-fidelity model development (73,
74). The system is based on experimental work (75) and was used to develop a UQ method for
CFD models. The research focused on the effects of model parameters on the accuracy of CFD
predictions compared with experimental data. In multiphase CFD modeling, several submodels
and model parameters must be defined, such as the drag model and coefficients of restitution. These
parameters can be difficult to measure and often are chosen based on previous studies or without
rationalization. Because there is no single correct choice for the aforementioned models and
quantities, there exists an associated uncertainty for each choice. Identification, quantification, and
reporting of these uncertainties are essential for using computational predictions with confidence.
Sensitivity analysis and Bayesian calibration methods (18, 76–78) were used to investigate the
uncertainty and variation in the CFD model parameters of the BFB and to develop a statistical
response surface model (emulator) that is capable of efficiently approximating the model system.
The emulator was used to complement CFD runs by allowing for additional model predictions
for sensitivity analysis and calibration. Initial analysis with this statistical framework revealed
fluidization problems with low gas velocities where no bubbles were recorded at the bottom of a
central probe in the bed. Sensitivity analysis via model emulation (74, 79, 80) showed that friction
angles for solid-solid interactions and drag models had significant effects on bubble frequency and
phase fraction predictions. Posterior distributions from calibration analysis identified low friction
angles for solid-solid interactions (approximately 25◦ ) and the Wen-Yu drag model (81) as the
optimal parameter settings. Additionally, the statistical emulator of the CFD model (along with a
model form discrepancy function), via out-of-sample cross-validation, was capable of capturing the
observed experimental data within 95% credible intervals for higher velocities where fluidization
is not an issue (73, 74). The information gained from this unit problem has been propagated up
the validation hierarchy to the filtered models and the laboratory-scale cases.

Risk Analysis
Risk analysis for better decision making is one of the critical objectives of CCSI, which seeks to use
information from science-based models with quantified uncertainty to inform decision makers who
are making large capital investments. These tools and capabilities facilitate the development of risk
models tailored for carbon capture technologies, quantify the uncertainty of model predictions,
and estimate the technical and financial risks associated with the system. These tools are being
integrated into a probabilistic risk-assessment methodological framework that encapsulates three

316 Miller et al.


CH05CH14-Miller ARI 19 May 2014 17:25

interconnected probabilistic modeling and simulation components: technology maturity, technical


risk, and financial risk.
A critical aspect of the risk-assessment framework is to identify and account for potential issues
that are not explicitly included in the multiscale models themselves. Such issues may include
material life, safety, and operability. Material life includes both the life of the capture material,
for which particle attrition and chemical degradation are potential concerns, as well as the life
of process equipment, for which erosion of internal components and corrosion are of concern.
Safety issues can arise from operating conditions, many of which can be addressed through dynamic
system modeling. Health risks may arise through degradation of capture materials or other by-
products of the capture process. Because many of these potential risks are difficult to identify early
in the development of a new technology, CCSI has developed a risk-assessment tool based on
technology maturity, which helps account for these types of risk.
Annu. Rev. Chem. Biomol. Eng. 2014.5:301-323. Downloaded from www.annualreviews.org
Access provided by University of Saskatchewan on 03/05/17. For personal use only.

Technology maturation. The technology readiness level (TRL)-based assessment component


identifies specific scientific and engineering targets required by each readiness level and applies
probabilistic estimation techniques to calculate the likelihood in technology maturity (82). An
expert elicitation system has been developed to assess carbon capture TRLs and use technical
maturity information to estimate the level of uncertainty owing to the lack of technology ma-
turity/knowledge for risk-analysis simulations. The TRL tools are also being developed to help
determine exit criteria from each TRL, including risk-factor analysis to identify the steps with
the greatest impact on risk (e.g., schedule, cost). This tool links to the stage-gate process typically
used in industry for managing technology and product innovation.

Technical risk model. The assessment of technical risk focuses on identifying and quantifying
risk contributors (both discrete and stochastic), performing scenario-based probabilistic risk anal-
ysis, and integrating with carbon capture process simulation and optimization toward achieving
greater reliability, maintainability, and availability (83). The technical risk model contains two
components: system mechanical risk and system operational risk. The system mechanical risk
model looks at the mechanical systems to estimate the (un)availability of each component within
each system. These results are then used within the system operational risk model to estimate
system performance.
Because carbon capture technology relies on both mechanical systems and phenomenolog-
ical (operational) behavior, the risk analysis must include probabilistic measures of off-normal
performance over time (84), as well as provide estimates of discrete failures (in time). In the current
stage of risk analysis, a 1-MWe hybrid solid-sorbent capture system design has been analyzed as
the basis for examining risk-based deviations to important decision and financial model parame-
ters. The important phenomena include the properties and behaviors of the sorbent flows and the
mechanical performance of the components containing those flows. The probabilistic measures
include both probabilistic risk-assessment and UQ estimates. The risk spectrum can be translated
into heuristic distributions for stochastic parameters in the decision framework. The technical risk
model can then be coupled with the financial risk model (both mechanical and operational results)
encapsulating economic risk factors and uncertainty to quantify financial returns measures.

Financial risk model. Retrofitting of carbon capture systems on existing power plants is likely
to incur additional costs owing to new construction, operational interruption, and increased op-
erations and maintenance, all of which must be taken into consideration to generate an accurate
assessment of financial risk. The financial-risk component estimates the long-term return on in-
vestment based on energy retail pricing, production cost, operating and power replacement cost,

www.annualreviews.org • Carbon Capture Simulation Initiative 317


CH05CH14-Miller ARI 19 May 2014 17:25

plant construction and retrofit expenses, and potential incentives. To provide a common basis for
comparison, the CCSI financial risk model converts annual revenue and expenditure estimates
into a common base year via a discount rate, which is summed to a final estimate of net present
value. The estimated net present values, presented as distributions, encapsulate the variability in fi-
nancial risk drivers integrated with uncertainty in technology maturity and technical performance
(85).

SUMMARY AND FUTURE CHALLENGES


Carbon capture, utilization, and storage—the capture of CO2 at a point source, such as coal-fired
power plants, followed by injection into oil reservoirs for enhanced oil recovery or into permanent
underground storage sites—is a key technology in the strategy to reduce CO2 emissions. To
Annu. Rev. Chem. Biomol. Eng. 2014.5:301-323. Downloaded from www.annualreviews.org

ready carbon capture technology components for demonstration in the next decade, the time for
Access provided by University of Saskatchewan on 03/05/17. For personal use only.

the commercial deployment of new energy technologies must be reduced. CCSI has addressed this
challenge by developing a Toolset of broadly applicable computational tools to take carbon capture
concepts from the laboratory to the power plant more quickly, at lower cost, and with reduced
risk. The Toolset (a) enables promising concepts to be more quickly identified through rapid
computational screening of processes and devices, (b) reduces the time to design and troubleshoot
new devices and processes by using optimization techniques to focus development on the best
overall process conditions and by using detailed equipment models to better understand and
improve the internal behavior of complex equipment, (c) quantifies the technical risk in taking
technology from laboratory scale to commercial scale by understanding the sources and effects
of model and parameter uncertainty, and (d ) will ultimately stabilize deployment costs more
quickly by enabling industry to have a more fundamental understanding of the entire capture
process.
The models in the Toolset span length- and timescales from particles and films to devices and
processes. Particle-scale CO2 capture reactions are modeled and a Bayesian calibration method-
ology is used to generate the posterior distribution of model parameters as well as a model-form
discrepancy function. The particle-scale models are needed as an input to device- and process-scale
models. The most economically optimal process topology is determined by using a superstruc-
ture optimization method. To make the superstructure optimization fast and robust, a method
to develop algebraic surrogate models from detailed steady-state process models has been devel-
oped. A steady-state simulation-based optimization method then refines the optimal conditions
for the CO2 capture and compression processes. The software tools necessary for determining
the uncertainty in the predictions of the process models (for example, the cost of electricity) and
for efficiently running the large number of process simulations required for UQ have been devel-
oped. A dynamic reduced model builder has been developed for generating fast, transient models
from higher-fidelity models, including PDE- and rate-based process models, for use in real-time
applications, such as operator training and online process control. Device-scale models based on
CFD are being developed for optimizing CO2 capture equipment and for providing device-scale
input to process-scale models (for example, parameters in 1D process models). Filtered models
have been developed to enable the effect of the numerous heat-transfer tubes in the CO2 ad-
sorbers to be included in the CFD models. A validation and UQ hierarchy has been developed
for systematically validating the CFD models. Risk-analysis and decision-making capabilities are
being developed to use information from the science-based models with quantified uncertainty to
inform decision makers who are making large capital investments.
CCSI developed a suite of models for sorbent-based capture processes and is currently focusing
on the development of models for solvent-based processes. The underlying framework of the

318 Miller et al.


CH05CH14-Miller ARI 19 May 2014 17:25

Toolset can be extended to model other processes and can be used to accelerate the development
of related technologies used for refining, chemicals production, and oil and natural gas production.

FUTURE ISSUES
1. Making the capabilities of the CCSI Toolset directly applicable to other types of carbon
capture systems will require developing additional rigorous models. The emphasis will be
on building capabilities not currently available in commercial software packages. Work
is currently focusing on advanced solvent-based capture systems.
2. As CCSI addresses other technologies, such as oxycombustion, chemical looping, and
gasification, the scope of the problem expands considerably, requiring the development
Annu. Rev. Chem. Biomol. Eng. 2014.5:301-323. Downloaded from www.annualreviews.org

of more efficient multiscale methods to enable solution of the entire integrated process.
Access provided by University of Saskatchewan on 03/05/17. For personal use only.

3. Ensuring industry realizes the benefits of the CCSI Toolset to accelerate the develop-
ment of carbon capture and related technology will require continued interaction with
the industry advisory board and commercial software vendors to enable the long-term
support and maintenance of the computational tools and models.
4. When generating the large amount of data at each scale to support multiscale UQ, it is
essential to use a data management system to track data provenance, ensure consistency
among models, and maximize the reuse of simulation data to efficiently use computational
resources. The development and integration of such a system is a major research thrust
as CCSI continues into its next phase.

DISCLOSURE STATEMENT
The authors are not aware of any affiliations, memberships, funding, or financial holdings that
might be perceived as affecting the objectivity of this review.
This presentation 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 respon-
sibility 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 by trade name, trademark, manu-
facturer, 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.

ACKNOWLEDGMENTS
This work was made possible by funding through the Office of Fossil Energy, US Department of
Energy. The authors acknowledge the technical contributions by the entire CCSI Team, especially
S. Bhat, T. Epperly, and M. Haranczyk, A. Lee, H. Kim, J. Morinelly, S. Modekurti, A. Cozad,
Z. Yuan, J. Eslick, J. Boverhof, J. Leek, B. Ng, J. Ou, J. Wendelberger, P. Mahapatra, J. Ma,
W. Lane, C. Montgomery, J. Dietiker, C. Lai, W. Pan, Z. Xu, E. Jones, and B. Edwards. We also
gratefully acknowledge M. Gray, J. Hoffmann, K. Resnick, J. Fisher, L. Shadle, J. Spenik, and
ADA-ES for providing experimental data.

www.annualreviews.org • Carbon Capture Simulation Initiative 319


CH05CH14-Miller ARI 19 May 2014 17:25

LITERATURE CITED
1. Chu S. 2009. Carbon capture and sequestration. Science 325:1599
2. Chu S, Majumdar A. 2012. Opportunities and challenges for a sustainable energy future. Nature 488:294–
303
3. Rubin ES. 2008. CO2 capture and transport. Elements 4:311–17
4. Ciferno JP, Fout TE, Jones AP, Murphy JT. 2009. Capturing carbon from existing coal-fired power
plants. Chem. Eng. Prog. April:33–41
5. Jenkins J, Mansur S. 2011. Bridging the Clean Energy Valleys of Death. Oakland, CA: Breakthrough Inst.
http://thebreakthrough.org/archive/bridging_the_clean_energy_vall
6. Haszeldine RS. 2009. Carbon capture and storage: How green can black be? Science 325:1647–52.
doi:10.1126/science.1172246
7. Hules KR, Yilmaz A. 2002. From bunker to stack: the cost-reduction and problem-solving benefits of computational
fluid dynamics for utility and industrial power generation. Presented at POWER-GEN Int., Orlando, FL
Annu. Rev. Chem. Biomol. Eng. 2014.5:301-323. Downloaded from www.annualreviews.org

8. Syamlal M, Guenther C, Cugini A, Ge W, Wang W, et al. 2011. Computational science: enabling


Access provided by University of Saskatchewan on 03/05/17. For personal use only.

technology development. CEP Magazine, January


9. Mebane DS, Bhat KS, Kress JD, Fauth DJ, Gray ML, et al. 2013. Bayesian calibration of thermodynamic
models for the uptake of CO2 in supported amine sorbents using ab initio priors. Phys. Chem. Chem. Phys.
15:4355–66
10. US Energy Inf. Adm. 2013. Annual Energy Outlook 2013 Early Release Overview. Report no.
DOE/EIA-0383ER, US Dep. Energy, Washington, DC. http://www.eia.gov/forecasts/aeo/er/pdf/
0383er%282013%29.pdf
11. Nichols C. 2010. Coal-fired power plants in the United States: examination of the cost of retrofitting with CO2
capture technology and the potential for improvements in efficiency. Report no. DOE/NETL-402/102309, US
Dep. Energy, Natl. Energy Technol. Lab., Pittsburgh, PA
12. Sjostrom S, Krutka H. 2010. Evaluation of solid sorbents as a retrofit technology for CO2 capture. Fuel
89:1298–306
13. Sjostrom S. 2010. Evaluation of solid sorbents as a retrofit technology for CO2 capture. Presented at 2010 NETL
CO2 Capture Technol. Meet., Pittsburgh, PA
14. Kim H, Haranczyk M, Epperly T, Abouelnasr M, Swisher JA, et al. 2012. Integrating the carbon capture
materials database with the process simulation tools of the carbon capture simulation initiative. Presented at
AIChE Annu. Meet., Pittsburgh, PA
15. Lee A, Mebane D, Fauth DJ, Miller DC. 2011. A model for the adsorption kinetics of CO2 on amine-impregnated
mesoporous sorbents in the presence of water. Presented at 28th Int. Pittsburgh Coal Conf., Pittsburgh, PA
16. Salciccioli M, Stamatakis M, Caratzoulas S, Vlachos DG. 2011. A review of multiscale modeling of metal-
catalyzed reactions: mechanism development for complexity and emergent behavior. Chem. Eng. Sci.
66:4319–55
17. Kennedy MC, O’Hagan A. 2001. Bayesian calibration of computer models. J. R. Stat. Soc. Ser. B Stat.
Methodol. 63:425–50
18. Higdon D, Kennedy MC, Cavendish JC, Cafeo JA, Ryne RD. 2004. Combining field data and computer
simulations for calibration and prediction. SIAM J. Sci. Comput. 26:448–66
19. Reich BJ, Storlie CB, Bondell HD. 2009. Variable selection in Bayesian smoothing spline ANOVA models:
application to deterministic computer codes. Technometrics 51:110–20
20. Miller DC, Sahinidis NV, Cozad A, Lee A, Kim H, et al. 2013. Computational tools for accelerating carbon
capture process development. Presented at 38th Int. Tech. Conf. Clean Coal Fuel Syst., Clearwater, FL
21. Lee A, Miller DC. 2013. A one-dimensional, (1-D) three-region model for a bubbling fluidised bed
adsorber. Ind. Eng. Chem. Res. 52:469–84
22. Modekurti S, Bhattacharyya D, Zitney SE. 2013. Dynamic modeling and control studies of a two-stage
bubbling fluidized bed adsorber-reactor for solid-sorbent CO2 capture. Ind. Eng. Chem. Res. 52:10250–60
23. Modekurti S, Bhattacharyya D, Zitney SE. 2012. Dynamic modeling and transient studies of a solid-sorbent
adsorber for CO2 capture. Presented at 29th Annu. Int. Pittsburgh Coal Conf., Pittsburgh, PA
24. Kim H, Miller DC. 2011. Development of a moving bed simulation model for carbon capture from fossil energy
systems. Presented at AIChE Annu. Meet., Minneapolis, MN

320 Miller et al.


CH05CH14-Miller ARI 19 May 2014 17:25

25. Kunii D, Levenspiel O. 1991. Fluidization Engineering. Boston: Butterworth-Heinemann


26. Lüdtke KH. 2004. Process Centrifugal Compressors: Basics, Function, Operation, Design, Application. Berlin:
Springer
27. Biegler LT, Grossmann IE, Westerberg AW. 1997. Systematic Methods of Chemical Process Design. Upper
Saddle River, NJ: Prentice Hall
28. Siirola JJ. 1996. Industrial applications of chemical process synthesis. In Advances in Chemical Engineering,
ed. JL Anderson, pp. 2–62. San Diego, CA: Academic
29. Grossmann IE. 1996. Mixed-integer optimization techniques for algorithmic process synthesis. In Ad-
vances in Chemical Engineering, ed. JL Anderson, pp. 172–239. San Diego, CA: Academic
30. Seider WD, Seader JD, Lewin DR, Widagdo S. 2008. Product and Process Design Principles: Synthesis,
Analysis and Design. New York: Wiley. 3rd ed.
31. Black JB, Haslbeck JL, Jones AP, Lundberg WL, Shah V. 2013. Cost and performance of PC and IGCC
plants for a range of carbon dioxide capture. Report no. DOE/NETL-2011/1498, US Dep. Energy,
Annu. Rev. Chem. Biomol. Eng. 2014.5:301-323. Downloaded from www.annualreviews.org

Natl. Energy Technol. Lab., Pittsburgh, PA. http://www.netl.doe.gov/File%20Library/Research/


Access provided by University of Saskatchewan on 03/05/17. For personal use only.

Energy%20Analysis/Publications/Gerdes-08022011.pdf
32. Seader JD, Seider WD, Pauls AC, Hughes RR. 1977. FLOWTRAN Simulation: An Introduction. Austin,
TX: CACHE
33. Evans LB, Boston JF, Britt HI, Gallier PW, Gupta PK, et al. 1979. Aspen: an advanced system for process
engineering. Comput. Chem. Eng. 3:319–27
34. Pantelides CC. 1988. Speedup—recent advances in process simulation. Comput. Chem. Eng. 12:745–55
35. Rios LM, Sahinidis NV. 2013. Derivative-free optimization: a review of algorithms and comparison of
software implementations. J. Glob. Optim. 56:1247–93
36. Wang GG, Shan S. 2007. Review of metamodeling techniques in support of engineering design optimiza-
tion. J. Mech. Des. 129:370–80
37. Cozad A, Sahinidis NV, Miller DC. 2014. Learning surrogate models for simulation-based optimization.
AIChE J. 60:2211–27
38. Caballero JA, Grossmann IE. 2008. An algorithm for the use of surrogate models in modular flowsheet
optimization. AIChE J. 54:2633–50
39. Henao CA, Maravelias CT. 2011. Surrogate-based superstructure optimization framework. AIChE J.
57:1216–32
40. Hurvich CM, Tsai CL. 1993. A corrected Akaike information criterion for vector autoregressive model
selection. J. Time Ser. Anal. 14:271–79
41. Tawarmalani M, Sahinidis NV. 2005. A polyhedral branch-and-cut approach to global optimization.
Math. Program. 103:225–49
42. Eslick JC, Miller DC. 2011. A multi-objective analysis for the retrofit of a pulverized coal power plant
with a CO2 capture and compression process. Comput. Chem. Eng. 35:1488–500
43. Deng G. 2007. Simulation-based optimization. PhD Thesis, Univ. Wis.-Madison
44. Mele FD, Guillén G, Espuña A, Puigjaner L. 2006. A simulation-based optimization framework for
parameter optimization of supply-chain networks. Ind. Eng. Chem. Res. 45:3133–48
45. Rajasree R, Moharir AS. 2000. Simulation based synthesis, design and optimization of pressure swing
adsorption (PSA) processes. Comput. Chem. Eng. 24:2493–505
46. Wan X, Pekny JF, Reklaitis GV. 2005. Simulation-based optimization with surrogate models—application
to supply chain management. Comput. Chem. Eng. 29:1317–28
47. Conn AR, Scheinberg K, Vicente LN. 2009. Introduction to Derivative-Free Optimization. Philadelphia:
Soc. Ind. Appl. Math.
48. Hansen N. 2006. The CMA evolution strategy: a comparing review. In Towards a New Evolutionary
Computation. Advances on Estimation of Distribution Algorithms (Studies in Fuzziness and Soft Computing), ed.
JA Lozano, P Larrañaga, I Inza, E Bengoetxea, pp. 75–102. Berlin: Springer
49. Huyer W, Neumaier A. 2008. SNOBFIT—stable noisy optimization by branch and fit. ACM Trans. Math.
Softw. 35:9:1–9:25
50. Boverhof J, Leek J, Eslick JC, Agarwal D. 2013. Turbine and Sinter: enabling management of paral-
lel process simulations on demand. CCSI Tech. Rep. Ser. https://www.acceleratecarboncapture.org/
drupal/sites/default/files/elfinder/CCSI_Directory/CCSI-TurbineSinter.pdf

www.annualreviews.org • Carbon Capture Simulation Initiative 321


CH05CH14-Miller ARI 19 May 2014 17:25

51. US Dep. Energy/Off. Foss. Energy. 2012. Clean Coal Research Program: 2012 Technology Readiness As-
sessment. Pathway for Readying the Next Generation of Affordable Clean Energy Technology—Carbon Capture,
Utilization, and Storage (CCUS). Washington, DC: US Dep. Energy
52. Lawal A, Wang M, Stephenson P, Koumpouras G, Yeung H. 2010. Dynamic modelling and analysis of
post-combustion CO2 chemical absorption process for coal-fired power plants. Fuel 89:2791–801
53. Lin Y, Pan T, Wong DS, Jang S, Chi Y, Yeh C. 2011. Plantwide control of CO2 capture by absorption
and stripping using monoethanolamine solution. Ind. Eng. Chem. Res. 50:1338–45
54. Ziaii S, Rochelle GT, Edgar TF. 2009. Dynamic modeling to minimize energy use for CO2 capture in
power plants by aqueous monoethanolamine. Ind. Eng. Chem. Res. 48:6105–11
55. Braatz RD, Alkire RC, Seebauer E, Rusli E, Gunawan R, et al. 2006. Perspectives on the design and
control of multiscale systems. J. Process Control 16:193–204
56. Nandong J, Samyudia Y, Tade MO. 2007. Control of multi-scale dynamics system. Presented at 16th IEEE
Int. Conf. Control Appl., Singapore
Annu. Rev. Chem. Biomol. Eng. 2014.5:301-323. Downloaded from www.annualreviews.org

57. Ilchman A. 1993. Non-Identifier-Based High-Gain Adaptive Control. New York: Springer
Access provided by University of Saskatchewan on 03/05/17. For personal use only.

58. Ilchman A, Ryan EP. 1994. Universal-tracking for nonlinearly-perturbed systems in the presence of noise.
Automatica 30:337–46
59. Sentoni GB, Guiver JP, Zhao H, Biegler LT. 1998. A state space approach to nonlinear process modeling:
identification and universality. AIChE J. 44:2229–39
60. Agrawal K, Loezos PN, Syamlal M, Sundaresan S. 2001. The role of meso-scale structures in rapid
gas-solid flows. J. Fluid Mech. 445:151–85
61. Igci Y, Andrews AT, Sundaresan S, Pannala S, O’Brien T. 2008. Filtered two-fluid models for fluidized
gas-particle suspensions. AIChE J. 54:1431–48
62. Parmentier JF, Simonin O, Delsart O. 2012. A functional subgrid drift velocity model for filtered drag
prediction in dense fluidized bed. AIChE J. 58:1084–98
63. Wang W, Lu B, Zhang N, Shi Z, Li J. 2010. A review of multiscale CFD for gas–solid CFB modeling.
Int. J. Multiph. Flow 36:109–18
64. Igci Y, Sundaresan S. 2011. Constitutive models for filtered two-fluid models of fluidized gas–particle
flows. Ind. Eng. Chem. Res. 50:13190–201
65. Milioli CC, Milioli FE, Holloway W, Agrawal K, Sundaresan S. 2013. Filtered two-fluid models of
fluidized gas-particle flows: new constitutive relations. AIChE J. 59:3265–75
66. Agrawal K, Holloway W, Milioli CC, Milioli FE, Sundaresan S. 2013. Filtered models for scalar transport
in gas-particle flows. Chem. Eng. Sci. 95:291–300
67. Holloway W, Sundaresan S. 2012. Filtered models for reacting gas-particle flows. Chem. Eng. Sci. 82:132–
43
68. Sarkar A, Sun X, Sundaresan S. 2013. Sub-grid drag models for horizontal cylinder arrays immersed in
gas-particle multiphase flows. Chem. Eng. Sci. 104:399–412
69. Igci Y, Pannala S, Benyahia S, Sundaresan S. 2011. Validation studies on filtered model equations for
gas-particle flows in risers. Ind. Eng. Chem. Res. 51:2094–103
70. Igci Y, Sundaresan S. 2011. Verification of filtered two-fluid models for gas-particle flows in risers. AIChE
J. 57:2691–707
71. Ryan EM, Montgomery C, Storlie C, Wendelberger J. 2012. CCSI validation and uncertainty quantification
hierarchy for CFD models. CCSI Tech. Rep. Ser. http://www.acceleratecarboncapture.org/drupal/sites/
default/files/elfinder/CCSI_Directory/Deliverables/Sept_30_2012/CCSI_V-UQ%20Hierarchy_
FINAL.pdf
72. Spenik J. 2011. Development of a circulating fluidized bed for flue gas carbon capture using solid sorbent. Presented
at NETL 2011 Workshop Multiph. Flow Sci., Pittsburgh, PA
73. Lane WA, Storlie C, Montgomery C, Ryan EM. 2013. Numerical modeling and uncertainty quantification
of a bubbling fluidized bed with immersed horizontal tubes. Powder Technol. 253:733–43
74. Storlie C, Lane WA, Ryan EM. 2013. Calibration of computational models with categorical parameters
and correlated outputs via Bayesian smoothing spline ANOVA. J. Am. Stat. Assoc. Submitted
75. Kim SW, Ahn JY, Kim SD, Lee DH. 2003. Heat transfer and bubble characteristics in a fluidized bed
with immersed horizontal tube bundle. Int. J. Heat Mass Transf. 46:399–409

322 Miller et al.


CH05CH14-Miller ARI 19 May 2014 17:25

76. Helton JC. 1997. Uncertainty and sensitivity analysis in the presence of stochastic and subjective uncer-
tainty. J. Stat. Comput. Simul. 57:3–76
77. Saltelli A, Chan K, Scott EM. 2000. Sensitivity Analysis. New York: Wiley
78. Oakley J, O’Hagan A. 2002. Bayesian inference for the uncertainty distribution of computer model outputs.
Biometrika 89:769–84
79. Storlie CB, Swiler LP, Helton JC, Sallaberry CJ. 2009. Implementation and evaluation of nonparametric
regression procedures for sensitivity analysis of computationally demanding models. Reliab. Eng. Syst. Saf.
94:1735–63
80. Konomi BA, Karagiannis G, Sarkar A, Sun X, Lin G. 2013. Bayesian treed multivariate Gaus-
sian process with adaptive design: application to a carbon capture unit. Technometrics. In press.
doi:10.1080/00401706.2013.879078
81. Wen CY, Yu YH. 1966. Mechanics of fluidization. Chem. Eng. Prog. Symp. Ser. 62:100–11
82. Engel D, Dalton A, Anderson K, Sivaramakrishnan C, Lansing C. 2012. Development of Technology Readiness
Annu. Rev. Chem. Biomol. Eng. 2014.5:301-323. Downloaded from www.annualreviews.org

Level (TRL) Metrics and Risk Measure. Richland, WA: Pac. Northwest Natl. Lab. http://www.osti.gov/
Access provided by University of Saskatchewan on 03/05/17. For personal use only.

bridge/product.biblio.jsp?osti_id=1067968
83. Dale C, Thompson J, Engel D, Dalton A, Jones E. 2013. Risk analysis and decision making. CCSI Tech.
Rep. Ser. www.acceleratecarboncapture.org
84. Zongxue X, Jinno K, Kawamura A, Takesaki S, Ito K. 1989. Performance risk analysis for Fukuoka water
supply system. Water Resour. Manag. 12:13–30
85. Engel D, Letellier B, Edwards B, LeClaire R, Jones E. 2012. New technical risk management development
for carbon capture process. Presented at 11th Annu. Conf. Carbon Capture Sequestration, Pittsburgh, PA

RELATED RESOURCES
Carbon Capture Simulation Initiative web site. https://www.acceleratecarboncapture.org
Carbon Capture Technology Program Plan, 2013. National Energy Technology Labo-
ratory. http://netl.doe.gov/technologies/coalpower/ewr/pubs/Program-Plan-Carbon-
Capture-2013.pdf

www.annualreviews.org • Carbon Capture Simulation Initiative 323


CH05-FrontMatter ARI 5 May 2014 12:43

Annual Review of
Chemical and
Biomolecular
Engineering

Volume 5, 2014 Contents


Annu. Rev. Chem. Biomol. Eng. 2014.5:301-323. Downloaded from www.annualreviews.org

Plans and Detours


Access provided by University of Saskatchewan on 03/05/17. For personal use only.

James Wei p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 1
Simulating the Flow of Entangled Polymers
Yuichi Masubuchi p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p11
Modeling Chemoresponsive Polymer Gels
Olga Kuksenok, Debabrata Deb, Pratyush Dayal, and Anna C. Balazs p p p p p p p p p p p p p p p p p p p35
Atmospheric Emissions and Air Quality Impacts from Natural Gas
Production and Use
David T. Allen p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p55
Manipulating Crystallization with Molecular Additives
Alexander G. Shtukenberg, Stephanie S. Lee, Bart Kahr,
and Michael D. Ward p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p77
Advances in Mixed-Integer Programming Methods for Chemical
Production Scheduling
Sara Velez and Christos T. Maravelias p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p97
Population Balance Modeling: Current Status and Future Prospects
Doraiswami Ramkrishna and Meenesh R. Singh p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 123
Energy Supply Chain Optimization of Hybrid Feedstock Processes:
A Review
Josephine A. Elia and Christodoulos A. Floudas p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 147
Dynamics of Colloidal Glasses and Gels
Yogesh M. Joshi p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 181
Rheology of Non-Brownian Suspensions
Morton M. Denn and Jeffrey F. Morris p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 203
Factors Affecting the Rheology and Processability of Highly
Filled Suspensions
Dilhan M. Kalyon and Seda Aktaş p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 229

vi
CH05-FrontMatter ARI 5 May 2014 12:43

Continuous-Flow Differential Mobility Analysis of Nanoparticles


and Biomolecules
Richard C. Flagan p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 255
From Stealthy Polymersomes and Filomicelles to “Self ”
Peptide-Nanoparticles for Cancer Therapy
Núria Sancho Oltra, Praful Nair, and Dennis E. Discher p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 281
Carbon Capture Simulation Initiative: A Case Study in Multiscale
Modeling and New Challenges
David C. Miller, Madhava Syamlal, David S. Mebane, Curt Storlie,
Debangsu Bhattacharyya, Nikolaos V. Sahinidis, Deb Agarwal, Charles Tong,
Annu. Rev. Chem. Biomol. Eng. 2014.5:301-323. Downloaded from www.annualreviews.org

Stephen E. Zitney, Avik Sarkar, Xin Sun, Sankaran Sundaresan, Emily Ryan,
Dave Engel, and Crystal Dale p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 301
Access provided by University of Saskatchewan on 03/05/17. For personal use only.

Downhole Fluid Analysis and Asphaltene Science for Petroleum


Reservoir Evaluation
Oliver C. Mullins, Andrew E. Pomerantz, Julian Y. Zuo, and Chengli Dong p p p p p p p p p p 325
Biocatalysts for Natural Product Biosynthesis
Nidhi Tibrewal and Yi Tang p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 347
Entangled Polymer Dynamics in Equilibrium and Flow Modeled
Through Slip Links
Jay D. Schieber and Marat Andreev p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 367
Progress and Challenges in Control of Chemical Processes
Jay H. Lee and Jong Min Lee p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 383
Force-Field Parameters from the SAFT-γ Equation of State for Use in
Coarse-Grained Molecular Simulations
Erich A. Müller and George Jackson p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 405
Electrochemical Energy Engineering: A New Frontier of Chemical
Engineering Innovation
Shuang Gu, Bingjun Xu, and Yushan Yan p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 429
A New Toolbox for Assessing Single Cells
Konstantinos Tsioris, Alexis J. Torres, Thomas B. Douce,
and J. Christopher Love p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 455
Advancing Adsorption and Membrane Separation Processes for the
Gigaton Carbon Capture Challenge
Jennifer Wilcox, Reza Haghpanah, Erik C. Rupp, Jiajun He, and Kyoungjin Lee p p p p p 479
Toward the Directed Self-Assembly of Engineered Tissues
Victor D. Varner and Celeste M. Nelson p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 507
Ionic Liquids in Pharmaceutical Applications
I.M. Marrucho, L.C. Branco, and L.P.N. Rebelo p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 527

Contents vii
CH05-FrontMatter ARI 5 May 2014 12:43

Perspectives on Sustainable Waste Management


Marco J. Castaldi p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 547
Experimental and Theoretical Methods in Kinetic Studies of
Heterogeneously Catalyzed Reactions
Marie-Françoise Reyniers and Guy B. Marin p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 563

Indexes

Cumulative Index of Contributing Authors, Volumes 1–5 p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 595


Cumulative Index of Article Titles, Volumes 1–5 p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 598
Annu. Rev. Chem. Biomol. Eng. 2014.5:301-323. Downloaded from www.annualreviews.org
Access provided by University of Saskatchewan on 03/05/17. For personal use only.

Errata

An online log of corrections to Annual Review of Chemical and Biomolecular Engineering


articles may be found at http://www.annualreviews.org/errata/chembioeng

viii Contents

You might also like