You are on page 1of 22

CH08CH16-Tsouris ARI 30 March 2017 13:55

V I E W Review in Advance first posted online


E on April 7, 2017. (Changes may
R

still occur before final publication

S
online and in print.)

C E
I N

N
A
D V A

Modular Chemical Process


Intensification: A Review
Yong-Ha Kim,1 Lydia K. Park,1 Sotira Yiacoumi,1
and Costas Tsouris1,2
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org
Access provided by Oak Ridge National Lab on 04/20/17. For personal use only.

1
Georgia Institute of Technology, Atlanta, Georgia 30332-0373
2
Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831-6181; email: tsourisc@ornl.gov

Annu. Rev. Chem. Biomol. Eng. 2017. Keywords


8:16.1–16.22
modular chemical process intensification, process intensification principles,
The Annual Review of Chemical and Biomolecular
Engineering is online at chembioeng.annualreviews.org chemical and manufacturing processes, intensified equipment, intensified
methods
https://doi.org/10.1146/annurev-chembioeng-
060816-101354
Abstract
Copyright  c 2017 by Annual Reviews.
All rights reserved Modular chemical process intensification can dramatically improve energy
and process efficiencies of chemical processes through enhanced mass and
heat transfer, application of external force fields, enhanced driving forces,
and combinations of different unit operations, such as reaction and sepa-
ration, in single-process equipment. These dramatic improvements lead to
several benefits such as compactness or small footprint, energy and cost sav-
ings, enhanced safety, less waste production, and higher product quality.
Because of these benefits, process intensification can play a major role in
industrial and manufacturing sectors, including chemical, pulp and paper,
energy, critical materials, and water treatment, among others. This article
provides an overview of process intensification, including definitions, princi-
ples, tools, and possible applications, with the objective to contribute to the
future development and potential applications of modular chemical process
intensification in industrial and manufacturing sectors. Drivers and barriers
contributing to the advancement of process intensification technologies are
discussed.

16.1

Changes may still occur before final publication online and in print
CH08CH16-Tsouris ARI 30 March 2017 13:55

INTRODUCTION
Energy consumption can significantly influence the operation of chemical processes and plants,
which are the core components of most industries, including chemical and petroleum, pulp and
paper, and energy industries, because fuel consumption, operational costs, and carbon dioxide
emissions strongly depend on the amount of energy consumed for a particular operation. The
largest energy-consuming industrial sector is generally the chemical and petroleum sector (1–3).
In 2006 and 2011, approximately 30% of the energy used in industry was attributed to the chem-
ical and petroleum industries (1, 3). In 2006, the chemical and petroleum industries consumed
approximately 10% of the global energy demand and produced approximately 7% of the global
carbon dioxide emissions (1), suggesting that energy savings in the industrial sector can be di-
rectly linked to reducing the world’s energy consumption and carbon dioxide emissions. Thus,
it is imperative to develop new processes and chemical plants that can operate with low energy
and minimal environmental impacts, as well as new technologies to retrofit energy-consuming
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org

conventional processes.
Access provided by Oak Ridge National Lab on 04/20/17. For personal use only.

Energy consumption is strongly influenced by the size of a chemical process. Large-scale


processes can be inefficient, consuming large amounts of energy for process operation. The size
of conventional processes is often large because their design is based on old paradigms of chemical
engineering, including “big is the best” (4). This paradigm can be supported by the two-thirds
power rule: C ∝ P2/3 , where C is the plant component cost and P is the production capacity (5).
However, because all products cannot be sold in markets and because the size of chemical plants
cannot increase infinitely, this paradigm should be replaced with “small is beautiful and can be
most efficient” (4, 5). To apply this new paradigm, necessary equipment and unit operations can
be appropriately integrated into a smaller space, interact properly with one another, and offer high
production capacity; i.e., processes can be intensified, modularized, and then combined.
An example of modular chemical process intensification applied to the chemical industry is the
reactive separation system of Eastman Chemical, which converts methanol into methyl acetate
(4–7). By combining reaction and separation, the new intensified process for methyl acetate pro-
duction consists of 3 pieces of equipment, instead of 28 pieces with the old system (Figure 1),
increasing compactness, decreasing energy consumption, and reducing manufacturing costs. Re-
active separation processes can typically reduce energy consumption and capital costs by as much
as 80% (8). Because its energy consumption is less than that of the large conventional system,
the intensified system has a lower impact on global warming, acidification, thermal pollution,
and fossil fuel depletion by a factor of 5, and on volatile organic carbon emissions by a factor of
10 (7). This simple example demonstrates that modular chemical process intensification can be
utilized to achieve substantial energy and cost savings in industrial process operations and reduce
environmental impacts.
Attempts have been made to gauge the effects of process intensification on global energy
consumption and cost savings (1–3). Table 1 shows the potential energy savings of the chemi-
cal and petrochemical industrial sector achievable by process intensification in various countries.
The potential worldwide energy savings achievable by process intensification relies mainly on the
potential savings in the United States, China, and Europe. In particular, the world’s energy sav-
ings potential, as well as cost-reducing opportunities that are proportional to the energy savings
potential (9), increases with time. These trends suggest that the role of modular chemical pro-
cess intensification in reducing energy consumption and manufacturing costs and in minimizing
environmental impacts will become more important in the future.
Modular chemical process intensification can substantially influence the operation of chemi-
cal processes and plants, and a high potential exists for lowering global energy consumption and

16.2 Kim et al.

Changes may still occur before final publication online and in print
CH08CH16-Tsouris ARI 30 March 2017 13:55

a Before b After
AA
M
C
MC
AA D
S Via process MC
intensification C ED

W RD
A
R
M RD
D
S W
E H
W
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org
Access provided by Oak Ridge National Lab on 04/20/17. For personal use only.

A: Azeo ED: Extractive distillation RD: Reactive distillation


AA: Acetic acid H: Heavies S: Solvent
W C: Catalyst M: Methanol W: Water
D: Distillation MC: Methyl acetate
E: Entrainer R: Reaction

Figure 1
The methyl acetate separative reactor process of Eastman Chemical (a) before and (b) after process
intensification. Figure adapted from Reference 4 with permission.

cost and minimizing environmental impacts. Since the late 1970s, much effort has been made to
define, develop, and apply process intensification. This review is focused on the current status
of process intensification in terms of definitions, generic principles, available tools, potential ap-
plications, drivers, and barriers, with the objective to contribute to the future development and
application of modular chemical process intensification in various industrial and manufacturing
sectors.

Table 1 Annual energy savings potential by process intensification technologies in the chemical
and petrochemical industrial sector (unit: EJ; 1 EJ = 1018 J) (1–3)
Year 2006 2010 2011
United States 0.15 0.21 0.20
OECD Europe 0.12a 0.17 0.21
China 0.15 0.23 0.25
Japan 0.08 0.10 0.08
Korea 0.05 0.10 0.10
India 0.05 0.06 0.06
Canada 0.04 0.03 0.03
Brazil 0.03 0.04 –
World 0.53 1.41 1.55

a
The estimation is based on data from Belgium, the Netherlands, Luxembourg, France, Germany, and Italy (1).

www.annualreviews.org • Modular Chemical Process 16.3

Changes may still occur before final publication online and in print
CH08CH16-Tsouris ARI 30 March 2017 13:55

Table 2 Various definitions of process intensification offered over the past four decades
Reference Definition
Ramshaw (10) “Devising exceedingly compact plant which reduces both the ‘main plant item’ and the installations costs.”
Stankiewicz & Moulijn “Any chemical engineering development that leads to a substantially smaller, cleaner, and more energy
(11) efficient technology.”
Tsouris & Porcelli (12) “Technologies that replace large, expensive, energy-intensive equipment or process with ones that are
smaller, less costly, more efficient or that combine multiple operations into fewer devices (or a single
apparatus).”
Portha et al. (13) “Holistic overall process based intensification (i.e., global process intensification) in contrast to the classical
approach of process intensification based on the use of techniques and methods for the drastic
improvement of the efficiency of a single unit or device.”
Baldea (14) “Any chemical engineering development that leads to substantially smaller, cleaner, safer and more energy
efficient technology or that combine[s] multiple operations into fewer devices (or a single apparatus).”
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org
Access provided by Oak Ridge National Lab on 04/20/17. For personal use only.

DEFINITION
Process intensification was first introduced by chemical engineers in the 1970s, and since then,
considerable efforts have been made to better define it to fit the needs of the chemical, manufac-
turing, pulp and paper, and energy industries, among others. Table 2 shows several definitions of
process intensification offered by different investigators. Definitions in the early 1980s focused on
reducing the size of chemical plants, which leads to substantial economic benefits as installation
and operational costs are decreased (10, 15). These definitions are still used in some cases (e.g., 16),
but they have been extended to include other benefits. For example, Stankiewicz & Moulijn (11)
introduced other beneficial aspects to the definition of process intensification, specifically energy
consumption, environmental impacts, and process efficiency. In addition to these benefits Tsouris
& Porcelli (12) and Reay (17) included multifunctionality and safety. In particular, the addition
of safety benefits was the result of considerable attention to several accidents involving traditional
chemical process plants (12, 18). Some of these definitions are mentioned in many process intensi-
fication studies (e.g., 8), and in other studies these definitions are occasionally combined (e.g., 14).
However, because these definitions focus on unit operations, single devices, and individual meth-
ods, recent attempts have been made to define process intensification at the macroscopic scale,
where intensified modular processes can interact with one another [e.g., at the plant scale (13,
18)]. Different areas of chemical engineering can contribute toward the goal of designing more
efficient chemical and manufacturing plants. Thus, process intensification has also been defined in
terms of chemical engineering principles, such as process systems engineering (19), which further
demonstrates that the definition is still evolving.
As Gerven & Stankiewicz (20) suggested, a universal definition of process intensification may
be difficult to obtain because it is a growing trend in chemical engineering (21, 22) and should
respond to the needs of stakeholders and markets [e.g., safety after severe plant accidents (12)], as
well as apply to various fields. So the definition will probably evolve continuously; however, it is
acceptable for process intensification to be clearly defined in such a way to fit current needs and
challenges until a universal definition is derived.

GENERIC PRINCIPLES
Gerven & Stankiewicz (20) proposed four generic principles that can be straightforward goals
of modular chemical process intensification. These principles are (a) maximize the effectiveness

16.4 Kim et al.

Changes may still occur before final publication online and in print
CH08CH16-Tsouris ARI 30 March 2017 13:55

of intra- and intermolecular events, (b) give each molecule the same processing experience,
(c) optimize the driving forces at every scale and maximize the specific surface area to which these
forces apply, and (d) maximize the synergistic effects from partial processes.
The first principle focuses on modifying the inherent kinetics of chemical reactions, which
imposes the intrinsic efficiency of chemical and manufacturing processes. An example of the
first principle includes the use of cyclodextrins to modify reaction products without significantly
changing the reaction yield (23, 24). For the reaction of indoxyl and isatin-5-sulfonate in aqueous
solution at pH 10, the normal reaction yields of indigo and indirubin-5 -sulfonate are 25% and
1.4%, respectively. When a β-cyclodextrin dimer is added to the system, however, the yields of
the reactions changed to <0.1% (detection limit) and 22%, respectively, because cyclodextrin
can increase the reactivity of isatin-5-sulfonate by selectively complexing sulfonate. This selective
complexation can increase the concentration of reactive isatin-5-sulfonate at equilibrium (23).
The second principle offers ideal conditions for chemical reactions. For a plug flow reactor,
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org

for example, ideal conditions to achieve the second principle include no mixing in the axial direc-
Access provided by Oak Ridge National Lab on 04/20/17. For personal use only.

tion, complete mixing in the radial direction, and constant flow velocity across any cross section
perpendicular to the flow.
The third principle is targeted to enhance rates of mass, heat, and/or momentum transfer.
Optimizing the driving forces (e.g., concentration gradient) and increasing surface areas can facil-
itate the transfer of mass, heat, and momentum. A typical example of the third principle includes
microreactors that can considerably enhance heat transfer.
The fourth principle emphasizes the importance of multifunctionality of process intensification.
Although individual processes are not interrelated, combining them may have a useful effect on the
operation of the integrated system. As shown in Figure 1, reactive separation is a good example
of the fourth principle.
These four principles should be applied to four important domains of process intensification:
spatial, thermodynamic, functional, and temporal domains, which for simplicity are called struc-
ture, energy, synergy, and time, respectively (20, 25). In the spatial domain, spatial structuring via
process intensification is important because intra- and intermolecular events can be controlled,
and thus, the effects of randomness on reaction kinetics can be minimized. In the thermodynamic
domain, appropriate use of alternative energy sources, as well as controlling energy transfer mech-
anisms, is important to facilitate heat and mass transfer. Process intensification in the functional
domain can focus on properly integrating functions and steps to maximize synergistic effects,
enhancing energy and process efficiencies. In the temporal domain, processes can be intensified
by either manipulating characteristic times of events or creating artificial and natural periodicity.
Applications of the generic principles to the four domains are available in the literature (20, 25,
26).

TOOLS
Typical tools for modular chemical process intensification are intensified equipment and methods
(11). Table 3 shows various types, drivers, and barriers to intensified equipment and methods.
Intensified equipment includes novel apparatuses, devices, or equipment that can largely enhance
rates of mass, heat, and/or momentum transfer. A general classification of intensified equipment
depends on the inclusion of chemical reactions. Intensified methods refer to innovative techniques
developed to intensify processes, including incorporating chemical reactions into unit operations,
combining two or more different processes, and adding various types of energy to processes. Most
process intensification studies have focused on developing and applying intensified equipment and
methods. Some examples are discussed below.

www.annualreviews.org • Modular Chemical Process 16.5

Changes may still occur before final publication online and in print
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org
Access provided by Oak Ridge National Lab on 04/20/17. For personal use only.

Table 3 Process intensification tools (11) and common drivers and barriers (27)
CH08CH16-Tsouris

Drivers Barriers

Engineering/
design
ARI

Less Higher concepts and


yield Better More Control/ Modeling initial

16.6
CO2
Energy emis- and con- quality Cost Less com- control and equipment New
Tools savings sion version product saving Safer waste pact Safety Cost systems scale-up design application

Reactor type Static mixer x x x x x x x x


reactor

Kim et al.
30 March 2017

Monolithic reactor x x x x x x x x

Microchannel x x x x x x
reactors
13:55

Nonreactor Static mixer x N/A N/A x x x x x


type
Plate heat x x x x x x

Intensified equipment
exchanger

Rotating packer x x x x x x x x
bed

Multifunctional Heat exchange x x x


reactors reactor

Reactive x x x x
distillation

Hybrid Membrane x x x x x x
separations distillation
Adsorptive N/A N/A x x x x
distillation

Alternative Centrifugal N/A x x x x x


energy liquid-liquid
sources contractor

Ultrasound x x x x x x x x x

Intensified methods
reactor

Microwave x N/A N/A x x


heating/
microwave
drying

Electric x x x x x x x

Changes may still occur before final publication online and in print
N/A N/A N/A N/A
field-enhanced
extraction

Plasma reactor x x x x x

Other Supercritical x x x
methods separation
CH08CH16-Tsouris ARI 30 March 2017 13:55

Intensified Equipment
Intensified equipment for process intensification is classified as nonreactor and reactor types, in-
cluding static mixers and static mixer reactors, respectively, which were included as core equipment
of approximately 2,000 US patents in 2003 (28). Static mixers normally refer to tubes, columns,
or pipelines equipped with stationary barriers that can split, stretch, reorder, and recombine fluids
without additional energy sources (29). Such stationary barriers are often called motionless in-
serts, elements, blades, or baffles. Because of these stationary inserts, miscible fluids are effectively
mixed, and small bubbles and droplets can be formed when immiscible phases of fluids are coming
into contact in a static mixer. Compared with typical mixing devices, such as mechanically agitated
vessels, static mixers of lower equipment cost, due to compactness and because additional energy
sources for mixing are unnecessary, can operate with less energy, in a smaller space, and with a
shorter residence time (28). Static mixer reactors are used for reactive systems, such as polymeriza-
tion reactions, because they provide effective radial mixing (30) and significantly enhance the rate
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org

of heat transfer. For the polymerization of poly(L-lactide), for example, better mixing conditions
Access provided by Oak Ridge National Lab on 04/20/17. For personal use only.

were demonstrated in a static mixer reactor than in a twin-screw extruder (31).


Microchannel reactors, with a sandwich-like structure and internal dimensions smaller than
1 mm, are another example of intensified equipment (32, 33). Possible effects of the extremely
small structural dimensions are dramatic improvement of heat transfer, production of a high ratio
of surface area to volume, generation of a large interfacial area between immiscible fluids, and
enhancement of mixing effects (34). Because of their small internal dimensions, microchannel
reactors can increase the rates of specific chemical reactions up to 1,000 times higher than those
in conventional systems (33).

Intensified Methods
Intensified methods for process intensification can be categorized as (a) multifunctional reactors,
(b) hybrid separation, and (c) alternative energy sources. Reactors intensified by adding functions
of conventional unit operations are called multifunctional reactors. Examples of multifunctional
reactors include heat exchanger reactors and reactive separation systems. Heat exchanger reactors
refer to reactors equipped with heat-exchanging operations mainly for thermal intensification
(35). For example, the specific area (m2 m−3 ), the global heat transfer coefficient (W m−2 K−1 ),
and the ratio of heat duty to reactor volume of compact multifunctional heat exchangers (kW
m−3 K−1 ) were shown to increase by 160, 8.75, and 1,400 times compared with those of conven-
tional technologies (e.g., batch reactor covered with a double jacket) (36). Intensified equipment,
such as static mixer reactors and microchannel reactors, can be combined with heat exchangers
to facilitate heat transfer during chemical reactions (35). Reactive separation couples chemical re-
actions with separation processes to avoid undesirable chemical reactions or phenomena, such as
catalyst/biocatalyst poisoning, or maintain a high driving force. A well-known example of reactive
separation is the methyl acetate reactive separation process of Eastman Chemical, which is based
on reactive distillation (Figure 1).
A continuous-jet hydrate reactor (CJHR) developed at the Oak Ridge National Laboratory
(ORNL) is another example of a multifunctional reactor (37–43). The CJHR combines a co-flow
injector/reactor to convert carbon dioxide (gas or liquid) or methane (gas) and water into gas
hydrates (Figure 2). The co-flow injector induces effective turbulent mixing and forms a high
surface area between two immiscible reactive fluids. For this system specifically, a high surface
area is desirable because the gas-hydrate product of the reaction is a solid phase formed at the
interface of the two fluids. As the reaction proceeds, gas hydrate acts as a barrier to mass transfer at

www.annualreviews.org • Modular Chemical Process 16.7

Changes may still occur before final publication online and in print
CH08CH16-Tsouris ARI 30 March 2017 13:55

Water
a CO2(l) c d
Water droplet
Hydrate-CO2(l)-
water composite

1 mm
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org
Access provided by Oak Ridge National Lab on 04/20/17. For personal use only.

Figure 2
Continuous-jet hydrate reactor (CJHR). (a) A schematic diagram of CJHR; (b) a photograph of the surface of
CO2 hydrate particles (37); (c) a photograph of CO2 hydrate stream produced by the CJHR reactor receiving
H2 O and liquid CO2 (38); (d) a photograph of dense 1-inch-diameter CO2 hydrate particles produced by the
CJHR reactor (42).

the interface; thus, a high conversion of gas into hydrate requires a high surface area and effective
removal of heat generated during the hydrate formation reaction (37, 39). The CJHR system is
a continuous-flow reactor that maximizes the interfacial surface area between reactants, i.e., CO2
and water or CH4 and water. Because gas hydrate crystals accumulate on the surface of droplets,
the production rates can increase by controlling some key parameters, such as the size of droplets
and fluid flow rates, that influence the surface area and mass transfer (37). To further increase the
surface area, it was also desirable to have the fluid with the highest volumetric flow rate as the
drop phase, so that the drop population and subsequently the surface area could be maximized.
Hybrid separation results from combining two or more unit operations. For example, typical
challenges of reactive distillation include the formation of an azeotropic mixture, which requires a
special separation process to break the azeotrope. Through selective vapor permeation, membranes
can be employed to break an azeotrope and therefore can be combined with reactive distillation.
Harvianto et al. (44) compared the economic benefits of various hybrid separation systems with
those of azeotrope distillation and found that the annual total costs of membrane distillation were
as much as 77% lower than those of azeotropic distillation.
The addition of external energy sources to unit operations can enhance the rates of mass, heat,
and momentum transfer. Typical energy sources added to intensify processes include high-gravity
fields, acoustic fields, electromagnetic radiation, and electric fields. Early process intensification
technologies focused on intensifying equipment and were developed with high-gravity fields, e.g.,
HiGee reactors (10, 45). Such technologies include rotating packed bed (45, 46) and spinning disc
reactors (47, 48). The mass transfer coefficients of the liquid side and gas side in the rotating packed
bed were 44 times and 9 times as high as those in conventional systems (45), thereby reducing the
volume by at least two orders of magnitude (46). Various operating and equipment parameters
of rotating packed bed reactors are available elsewhere (e.g., 46). Spinning disc reactors can also
significantly enhance mass and heat transfer rates (11, 47, 48) and offer much shorter residence
times than conventional tank reactors, such as a stirred batch reactor (48).
Another example of high-gravity field systems is the centrifugal contactor/reactor that inte-
grates liquid-liquid mixing, reaction, and phase separation within the same unit (49–52). The
centrifugal contactor/reactor developed at national laboratories under the US Department of

16.8 Kim et al.

Changes may still occur before final publication online and in print
CH08CH16-Tsouris ARI 30 March 2017 13:55

a Upper weir Splash plate


b Stationary state

More dense
phase exit

Less dense
phase exit
Upper
collector
Lower collector
More dense Lower weir
phase inlet

Less dense
Housing phase inlet
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org
Access provided by Oak Ridge National Lab on 04/20/17. For personal use only.

Continuous operation

Separation
zone

Mixing zone
Rotor

Bottom vane Rotor inlet

Figure 3
The centrifugal contactor/reactor developed for liquid-liquid systems. (a) A schematic diagram and a photograph of the centrifugal
contactor/reactor (49); (b) stationary operation, with organic and aqueous fluids colored red and blue, respectively, and continuous
operation (52).

Energy consists mainly of two zones: a mixing zone and a separation zone (Figure 3). Reactions
occur primarily in the mixing zone, where intensive shear mixing takes place. A high-gravity field
in the separation zone subsequently forces the separation of the two phases. Thus, the centrifugal
contactor/reactor is effective for mass transfer, chemical reaction, and phase separation. Biodiesel
production has taken advantage of these features (51). Owing to the high shear force and turbulent
flow created in the mixing zone, mass transfer effects were minimized and thus the biodiesel pro-
duction rate was assumed to be limited only by the reaction kinetics. The residence time of fluids in
a conventional centrifugal contactor/reactor is relatively short. For a slow reactive system, such as
biodiesel synthesis using esterification reactions, a modified centrifugal contactor was introduced
to control the residence time in the mixing zone. This centrifugal contactor/reactor can be used
for liquid-liquid systems at temperatures higher than 100◦ C and for gas-liquid systems.
Acoustic fields are often used for process intensification because sound wave energy can cause
acoustic cavitation that can augment mass transfer and generate radicals that can augment chemical
reaction rates (26, 53). In general, ultrasound reactors refer to reactors equipped with ultrasonic
transducers with a frequency range of 18 kHz and 1 MHz (26, 53–55). The use of ultrasound can

www.annualreviews.org • Modular Chemical Process 16.9

Changes may still occur before final publication online and in print
CH08CH16-Tsouris ARI 30 March 2017 13:55

a 1 kV 0.1 m s–1 b 2 kV 0.1 m s–1 c 3 kV 0.1 m s–1


Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org
Access provided by Oak Ridge National Lab on 04/20/17. For personal use only.

3 mm 3 mm 3 mm

Figure 4
Fluorescent particle trajectory at various values of applied potential between the capillary injector and a counter electrode in water:
(a) 1 kV, (b) 2 kV, and (c) 3 kV. The head and length of the arrows represent the direction and the relative velocity of the particles in
water. Figure adapted from Reference 70.

dramatically improve reaction rates and product yields (53–55). For example, when ultrasound
was used, the production yield of arylalkane oxidation was enhanced by a factor of 6.7. The effects
of ultrasound can be controlled by adjusting the power. When an ultrasonic transducer with a
frequency of 18 kHz and a power of 50 W was used to extract almond oil using supercritical CO2 ,
the product yield increased by 20% (54). When the power was increased to 110 W, however, the
reaction yield increased by approximately 90% (55).
Electric fields applied to multiphase systems can disrupt interfaces and facilitate surface inter-
actions to augment mass and heat transfer (56–59) and chemical reaction rates (60, 61). Typical
techniques that utilize electric fields include electrostatic spraying (59, 62–72), electrostatic pump-
ing (69, 70), electrostatic mixing (69), and electro-osmotic and electric field pulses (73). These
techniques are combined in the electrohydrodynamic micromixing reactor, developed at ORNL
for particle synthesis and water treatment (74–76). Compared with the energy supplied for con-
ventional mixing in multiphase systems, such as mechanical agitation, the energy introduced by
electric fields can be consumed more efficiently to facilitate mixing, heat and mass transfer, and
chemical reactions. This is because electric fields focus the energy on the interface, whereas in
conventional mixing a big fraction of the energy is dissipated in the bulk solution. Figure 4 shows
the trajectory of fluorescent particles in electrohydrodynamic flow fields produced by electrostatic
spraying in water at various applied potentials (70). The particles accelerate near the capillary tip
owing to the electrohydrodynamic flow generated by the electric field. As the applied potential
increases, the flow velocity near the tip of the electrified injector increases, indicating that mass
and heat transfer rates, as well as effective mixing, can be enhanced by appropriately adding an
electric field to a process. Weatherley et al. (71) reported that the extraction rates of penicillin G
into dichloromethane increased up to a factor of five when electrostatic spraying was used.
Electric fields can affect chemical reactions, such as electro-oxidation and hydrolysis (59–61,
72). Liu et al. (60) examined the effects of electric fields on oxidation of low-valence vanadium
in slag particles and found that the extraction yield of the vanadium could increase by as much
as 93.7% because the electric field modified the surface charge distribution of the particles, and

16.10 Kim et al.

Changes may still occur before final publication online and in print
CH08CH16-Tsouris ARI 30 March 2017 13:55

this modification facilitated the release of the vanadium contained in the particles. Hydrolysis of
natural oil can be enhanced by electric fields, and the enhancement depends on the field strength
(59, 72). The enzymatic hydrolysis rate of sunflower oil was augmented by applying an electric
field. Compared with the rates obtained in a mechanically agitated reactor, the specific reaction
rates (g m−2 h−1 ) obtained in an electrically enhanced enzyme reactor were up to nine times higher,
possibly because of better phase disengagement and stimulated enzyme activity on charged droplets
(72).
Electric fields may also be used to modify thermodynamic properties of mixtures. Blankenship
et al. (77, 78) and Tsouris et al. (79) examined the effects of electric fields on vapor-liquid equilibria
of water-alcohol mixtures. Some effects were reported but the magnitude of those effects was not
strong enough to justify the application of electric fields to modify thermodynamic properties.
Low-voltage electric fields between a solid electrode with high surface area and an aque-
ous solution can be employed to enhance sorption of ions by modifying the surface charge
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org

of the electrode, a phenomenon known as electrosorption. Electrosorption can be used to re-


Access provided by Oak Ridge National Lab on 04/20/17. For personal use only.

move ions from water of wide-ranging salinity (80–88). Nanostructured carbon materials of
high electrical conductivity and surface area have been developed to enhance this phenomenon
(89, 90).
Electric fields can be employed to control phase inversion in liquid-liquid systems. Tsouris
& Dong (91, 92) and Dong & Tsouris (93) applied an alternating-current field and flipped the
continuous and dispersed phases in a controllable manner. This phenomenon is applicable to
enhancing the rates of mass and heat transfer between immiscible liquids.
Magnetic fields can be used to capture ferromagnetic and paramagnetic particles from parti-
cle suspensions (94–96). High-gradient magnetic filtration (HGMF) was developed to separate
colloidal paramagnetic particles that are difficult to capture by other means (96–98).

APPLICATIONS
With intensified equipment and methods, process intensification can be applied to various indus-
trial sectors, including chemical manufacturing, biofuel production, power generation, petroleum
refining, oil and gas extraction, and water treatment and recycling (Figure 5).
In the chemical manufacturing industry, large amounts of energy are consumed to manufac-
ture acetone, ammonia, benzene, chlorine, ethanol, ethylene, ethylene dichloride, ethylene oxide,
methanol, nitrogen, oxygen, propylene, and sodium hydroxide, among others (9). In 2010, ap-
proximately 43% of the on-site energy used in the chemical manufacturing industry was attributed
to operating chemical manufacturing processes to produce the 13 chemicals identified above (9).
Using process intensification to produce these chemicals can save as much as 7.33 × 1018 J per
year, equating to production costs savings of up to $9.1 billion per year (9). Possible applications
of process intensification that lead to energy and cost savings include developing novel separa-
tion technologies, advancing heat transfer and recovery, improving reaction kinetics, integrating
process steps, developing innovative steam-reforming processes of natural gases, and developing
high-efficiency modular processes.
Modular chemical process intensification can be applied to the electric power industry, which
includes coal-fired power plants that can emit large amounts of carbon dioxide. This field could
be intensified by applications such as air separation, hydrogen generation, gasification, and carbon
capture and storage. For example, researchers have used process intensification tools to develop
a method to store carbon dioxide in the deep ocean (37–41). With the CJHR hydrate composite
particles containing water, liquid CO2 , and CO2 hydrate can be produced in mid-ocean depths
and then sink to the deep ocean (Figure 6). The key to successfully generating sinking CO2

www.annualreviews.org • Modular Chemical Process 16.11

Changes may still occur before final publication online and in print
CH08CH16-Tsouris ARI 30 March 2017 13:55

Chemical manufacturing
• Novel separations
• Advanced heat transfer and recovery
• Improved reaction kinetics
• Integrated process steps
Water treatment and recycling • Steam reforming of natural gas Petroleum refining
• Development of modular processes
• Anaerobic membrane digesters • Crude oil upgrading
• Desalination • Gas-to-liquid (fuels and
• Water and sewage treatment chemicals) conversion
• Metal recycling • Fractional distillation
• Electronics recycling • Steam reforming of natural gas
• Wastewater collection and treatment

Process
intensification
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org
Access provided by Oak Ridge National Lab on 04/20/17. For personal use only.

Power generation Oil and gas extraction


• Air separation • Wellhead gas separation
• Hydrogen generation • Helium recovery and refining
• Gasification • Hydro-fracturing water recycling
• Carbon capture and storage • Natural gas pipeline processing
Biofuel production
• Separation
• Improved reactions
• Distillation
• Gasification
• Integrated processes
• Production flexibility
• Biofuel upgrading

Figure 5
Possible applications of process intensification to various industrial sectors. Figure inspired by References 9 and 99.

hydrate composite particles in seawater is to maximize the conversion of liquid CO2 into CO2
hydrate, which is denser than seawater (100, 101). If the conversion rate is low, the density of the
produced particles does not increase sufficiently because of the presence of liquid CO2 ; therefore,
the particles float toward the water surface instead of sinking to the bottom of the ocean. To sink
particles generated at approximately 1 km, a minimum conversion rate of 25% to 29% is required
(39, 100). The CO2 hydrate conversion rate reached as high as 30% during a field investigation
in Monterey Bay, California, at ocean depths of 1.1 to 1.3 km (Figure 6).
Many countries face water scarcity issues because of severe droughts and water pollution.
Process intensification can be applied to various water and wastewater treatment technologies,
including anaerobic membrane digesters, wastewater and sewage treatment, and desalination, to
more efficiently and reliably supply drinking water, as well as more effectively manage the quality
of water sources. For example, in the water treatment industry, capacitive deionization (CDI)
is an alternative, energy-efficient desalination method based on electrosorption. CDI operates
at low voltage to avoid energy-intensive electrochemical reactions. Electrical energy stored in a
capacitor during deionization can be recycled during regeneration; thus, this process can be less
energy intensive than conventional processes, including reverse osmosis and thermal desalination
(83). Currently, the focus of CDI is to develop novel porous carbon materials and use ion-exchange
membranes. Novel carbon electrodes based on micelle formation can efficiently collect ions during

16.12 Kim et al.

Changes may still occur before final publication online and in print
CH08CH16-Tsouris ARI 30 March 2017 13:55

Test case number 1 2 3

Initial behavior Floating Sinking Neutral

Injection depth (km) 1.099 1.251 1.297

Particle velocity (m s–1) –0.06 0.02 0

CO2 hydrate
conversion (%) 21 30 26

Minimum conversion for


sinking composite (%) 28.6 25.9 25.5

Figure 6
Injection of carbon dioxide into the ocean using the continuous-jet hydrate reactor of the Oak Ridge
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org

National Laboratory. The image of hydrate composite particles in a strainer from dive V2501-03,
Access provided by Oak Ridge National Lab on 04/20/17. For personal use only.

illuminated by the laser Raman spectrometer, at a depth of 1,255 m in Monterey Canyon (credit  c 2004
Monterey Bay Aquarium Research Institute) (left) and the measurements and calculations (right) are from
Tsouris et al. (39).

charging. Tsouris et al. (89) reported that, compared with commercial carbon aerogel with a wide
pore size distribution, graphite electrodes coated with mesoporous carbon increased the ion storage
capacity per unit mass of the electrode by approximately a factor of 2. Ion-exchange membranes
can be combined with CDI to enhance the electrosorption rates of counterions and suppress the
electrosorption rates of co-ions, accelerating desorption kinetics. In a sodium-chloride solution
of 200 mg L−1 , for example, compared with CDI, the salt removal efficiency of membrane CDI
increased by at least a factor of 2 (102). New types of electrodes, such as flow electrodes, also
enhanced the performance of CDI and membrane CDI (103).
Biofuels produced with food and nonfood biomass resources are an alternative option for fossil
fuels, such as petroleum, but conventional biofuel production processes need to be intensified to
overcome remaining challenges. For example, biodiesel production faces the following challenges:
(a) The production rate of biodiesel is often limited by mass transfer (51, 104), (b) reaction time is
long (48, 104), and (c) commercial biodiesel production processes operate in batch reactors (104).
Intensified equipment and methods can be utilized to overcome these challenges. McFarlane et al.
(51) showed that the centrifugal contactor/reactor can eliminate mass transfer limitations during
transesterification of soybean oil. Compared with a typical batch reactor, an intensified spinning
disk reactor can reduce potentially by a factor of 40 the time to reach conversion equilibrium in
a mixture of canola oil, anhydrous methanol, and sodium hydroxide (48). Cases in which process
intensification tools are applied to continuous biodiesel synthesis are reviewed elsewhere (e.g.,
104). Biofuel production could be intensified by advancing novel separation methods, improving
reaction kinetics, enhancing gasification, integrating multiple processes, and developing upgraded
biofuels.
Process intensification can be applied to oil and gas extraction as well as petroleum refining.
For example, a conventional system to recover oil from oil contaminants can be substantially im-
proved by using alternative energy sources. Figure 7 shows continuous oil separation and recovery
processes, which are intensified using the bursting energy and buoyancy force of microbubbles,
and turbulent mixing induced by bubbly water (105). In the oil separation process, oil contami-
nants are washed with bubbly water and microbubbles are attached to the contaminants. When
the attached microbubbles are exposed to ambient air, the microbubbles burst spontaneously and
the bursting energy is dissipated, generating strong hydrodynamic forces (106, 107). According

www.annualreviews.org • Modular Chemical Process 16.13

Changes may still occur before final publication online and in print
CH08CH16-Tsouris ARI 30 March 2017 13:55

a
Polluted sands Rotary kiln
and gravels Conveyor
belt
Scum (oil)
outlet

(–) bubble
(+) bubble
Conveyor
belt

SIF (M) Treated water (recycle ratio: 15 %)

1st step: Separation process 2nd step: Water treatment process


Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org

b
Access provided by Oak Ridge National Lab on 04/20/17. For personal use only.

Raw sample Water (–) and (+) bubbles

Figure 7
Oil separation and recovery using microbubbles (105). (a) A continuous separation device utilizing
microbubbles; (b) images of sand and water, raw samples, and samples treated without and with charged
microbubbles. Abbreviation: SIF, superspeed impeller flotation.

to the energetics of bubble bursting, the magnitude of the hydrodynamic forces produced by a
0.6-cm-bubble burst is comparable to the fluid shear force of 95 N m−2 in laminar flow (106). The
bursting of smaller bubbles can produce stronger forces (107). Thus, when microbubbles burst,
strong hydrodynamic forces can impact oil contaminants near the bursting bubbles, leading the
oil to detach (105). In the water treatment process, bubbly water is introduced to induce turbulent
mixing and to generate microbubbles (108). During turbulent mixing, the detached oil can be
more efficiently collected by charged bubbles, via electrostatic interactions, and be removed via
flotation. The buoyancy force of the bubbles enables the detached oil to move quickly to the water
surface. Compared with tap water, pressurized bubbly water dramatically enhanced the efficiency
of oil separation and recovery (Figure 7).

16.14 Kim et al.

Changes may still occur before final publication online and in print
CH08CH16-Tsouris ARI 30 March 2017 13:55

DRIVERS AND BARRIERS

Drivers
Several key drivers, including energy consumption, cost savings, safety, environmental impacts,
and investments, can accelerate the development and application of modular chemical process
intensification (7, 8, 12, 17, 27, 109, 110).
Reducing energy consumption is an important driver of process intensification technologies.
Certain methods can reduce the energy consumed by chemical and manufacturing processes. As
shown in Figure 1, the amount of energy consumed to operate processes can be dramatically
reduced by decreasing their size and increasing their compactness. Compared with conventional
processes, reactive distillation and dividing wall column distillation can reduce energy consumption
from 20% to 80% and from 10% to 30%, respectively (8). Low-energy technologies or equipment
can also be used to reduce the amount of energy consumed by processes.
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org

Another strong advantage of process intensification technologies is cost savings, because lower
Access provided by Oak Ridge National Lab on 04/20/17. For personal use only.

costs mean higher business profits. Process costs typically consist of capital, energy, and raw
materials costs. Capital and energy cost savings can result from reducing the size of the process
and the amount of energy consumed, as well as increasing compactness. The capital costs of
reactive distillation, dividing wall column distillation, and reverse flow reactor can be 20–80%,
10–30%, and >20% cheaper than the capital costs of conventional processes, respectively (8). Raw
materials costs can decrease by reducing raw material consumption. Investigators have shown that
improving the chemistry that can enhance the intrinsic properties of chemical reactions such as
reaction yield, conversions, selectivity, and purity, as well as product quality (12, 20, 111), can
reduce the amount of raw materials being used.
Process safety depends on chemical inventory, flammability, explosiveness, toxicity, tempera-
ture, and pressure (110). Thus, safety is important for chemical and manufacturing processes that
handle hazardous, toxic, corrosive, flammable, and explosive chemicals. When chemical invento-
ries are complex and involve large amounts of dangerous chemicals, processes cannot be inherently
safe. Safety can be enhanced through small processes that require small chemical inventories and
consume small amounts of chemicals. Process intensification can exclude dangerous chemicals
from chemical inventories and thus can offer inherent safety (110).
Process intensification can lessen negative environmental impacts of chemical and manufactur-
ing processes. For example, by reducing energy consumption, intensified processes can emit less
carbon dioxide, which is suspected of causing global warming (17). Because process intensification
can reduce the amounts of needed feedstocks, intensified processes can produce less waste and
pollutants that can cause acidification (7).
Investments are the essential part of research and development (R&D). Since the late 1970s,
Europe has continuously made large R&D investments in process intensification (112, 113). The
focus of the investments has been on the development and application of process intensification in
chemical engineering. Recently, the Rapid Advancement in Process Intensification Deployment
(RAPID) Manufacturing Institute has been funded by the US Department of Energy. The RAPID
Manufacturing Institute is focused on developing and applying process intensification technologies
to manufacturing processes to lower costs, improve energy and resource efficiencies, and increase
overall productivity (114).
Energy consumption, cost savings, selectivity, and safety are important drivers in the chemical
manufacturing sector (8, 27). Cost savings and product quality and safety are vital to the food
manufacturing sector (27). Drivers of process intensification tools are given in Table 3.

www.annualreviews.org • Modular Chemical Process 16.15

Changes may still occur before final publication online and in print
CH08CH16-Tsouris ARI 30 March 2017 13:55

Barriers
Modular process intensification must overcome barriers including high risk of failure, lack of scale-
up knowledge, and safety (8, 12, 27, 109). Barriers to preventing the use of process intensification
tools and approaches are given in Table 3.
A major barrier to the development of high-efficiency modular chemical processing devices
is the high risk involved in its investment (12). Process intensification is focused on developing
novel and innovative equipment and methods that dramatically improve processes. Developing
such equipment and methods may require large monetary investments but can always include high
risk of failure, which increases the investment risk.
Safety can also be a barrier to modular chemical process intensification because of the potential
impact on the environment and human health (8, 109). Some intensified processes can include
hazardous chemicals in chemical inventories and thus are not inherently safe. New equipment and
methods incorporated into intensified modules may not always be tested for a sufficiently long
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org

period to assure safe operation. The chemical inventories of chemical manufacturing processes
Access provided by Oak Ridge National Lab on 04/20/17. For personal use only.

can be complex. Owing to a lack of information about intensified modules, a safety risk can be
created when they are used.
Intensified modules can also include novel processes and technologies that are not fully tested
for sufficiently long periods. Thus, the processes and technologies carry their own risk and un-
certainties even though high reliability is essential in the chemical manufacturing industry (8). In
particular, there is a lack of knowledge about the dynamic response and performance of intensi-
fied modules when many modules are connected to each other and operate simultaneously. For
example, Luyben & Hendershot (115) investigated the effects of modular process intensification
on the operation of a benzene nitration reactor converting 98% of benzene. Compared with a
single continuous stirred-tank reactor (CSTR) with a volume of 121.8 m3 , the initial operation of
two CSTRs in series with individual volumes of 13.9 m3 was more unstable. Thus, the reliability
of the single CSTR was higher than that of the two CSTRs for the first six hours.

FUTURE OF MODULAR CHEMICAL PROCESS INTENSIFICATION


Modular chemical process intensification is a growing trend in chemical engineering and can
substantially improve chemical and manufacturing processes. The dramatic improvements create
various benefits, including reduction of energy consumption, cost savings, safety enhancement,
and minimization of environmental impacts. These benefits are key drivers and could accelerate
future development and application of modular chemical process intensification to the chemical
industry, pharmaceutical industry, materials, pulp and paper, petroleum, water treatment and
desalination, and environmental management, among other fields. Thus, process intensification
is expected to include a wide range of academic and practical topics in the future. A universal
definition of process intensification may further enhance its range of applications.
Novelty, innovation, and creativity are the core features of modular chemical process inten-
sification. Successfully developing future process intensification technologies with these features
depends on a better understanding of the thermodynamic, kinetic, and transport phenomena oc-
curring at the molecular scale. Better knowledge of chemistry is essential. Future work on process
intensification should also include developing engineering approaches that control various phe-
nomena from the molecular to the process scale. An emerging engineering approach that can
benefit process intensification is advanced manufacturing, e.g., three-dimensional printing of de-
vices including process equipment for micromixing and droplet formation for liquid extraction
(116) and reactors for chemical synthesis and analysis (117). Advanced manufacturing may be a

16.16 Kim et al.

Changes may still occur before final publication online and in print
CH08CH16-Tsouris ARI 30 March 2017 13:55

key player in the development of modular chemical process intensification, and process intensifi-
cation can further accelerate the development of advanced manufacturing through applications in
process industries. Interdisciplinary research and close collaboration among universities, research
institutions, and industry will be needed. Educational modules of process intensification must be
introduced in the chemical engineering curriculum to train the future generations of researchers
and practitioners.
Successful implementation of modular chemical process intensification relies on overcoming
many barriers, high risk of failure, lack of scale-up knowledge, and safety. In particular, potential
problems caused by lack of scale-up knowledge should be eliminated to increase the reliability of
process intensification technologies. Mathematical modeling is expected to play a central role in
addressing scale-up, risk, and safety issues.

DISCLOSURE STATEMENT
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org
Access provided by Oak Ridge National Lab on 04/20/17. For personal use only.

The authors are not aware of any affiliations, memberships, funding, or financial holdings that
might be perceived as affecting the objectivity of this review.

ACKNOWLEDGMENTS
This manuscript has been authored by UT-Battelle, LLC under contract no. DE-AC05–
00OR22725 with the US Department of Energy. The United States Government and the pub-
lisher, by accepting the article for publication, acknowledge that the United States Government
retains a nonexclusive, paid-up, irrevocable, worldwide license to publish or reproduce the pub-
lished form of this manuscript, or allow others to do so, for United States Government pur-
poses. The Department of Energy will provide public access to these results of federally spon-
sored research in accordance with the DOE Public Access Plan (http://energy.gov/downloads/
doe-public-access-plan).

LITERATURE CITED
1. IEA (Int. Energy Agency). 2009. Energy Technology Transitions for Industry: Strategies for the Next Industrial
Revolution. Paris: IEA
2. IEA (Int. Energy Agency). 2013. Energy Technology Perspectives 2013. Paris: IEA
3. IEA (Int. Energy Agency). 2014. Energy Technology Perspectives 2014. Paris: IEA
4. Stankiewicz A, Moulijn JA. 2002. Process intensification. Ind. Eng. Chem. Res. 41:1920–24
5. Reay D, Ramshaw C, Harvey A. 2008. Process Intensification: Engineering for Efficiency, Sustainability and
Flexibility. New York: Butterworth-Heinemann
6. Siirola JJ. 1995. An industrial perspective on process synthesis. AIChE Symp. Ser. 91:222–33
7. Harmsen GJ, Korevaar G, Lemkowitz SM. 2004. Process intensification contributions to sustainable
development. In Re-engineering the Chemical Processing Plant: Process Intensification, ed. A Stankiewicz, JA
Moulijn, pp. 495–522. New York: Marcel Dekker
8. Harmsen J. 2010. Process intensification in the petrochemicals industry: drivers and hurdles for com-
mercial implementation. Chem. Eng. Process. 49:70–73
9. US Dep. Energy. 2015. Process intensification. Chapter 6: technology assessments. In Quadrennial Tech-
nology Review 2015, US Dep. Energy, pp. 1–29. Washington, DC: US Dep. Energy. https://energy.
gov/sites/prod/files/2015/11/f27/QTR2015-6J-Process-Intensification.pdf
10. Ramshaw C. 1983. Higee distillation - an example of process intensification. Chem. Eng. 389:13–14
11. Stankiewicz AI, Moulijn JA. 2000. Process intensification: transforming chemical engineering. Chem.
Eng. Prog. 96:22–34

www.annualreviews.org • Modular Chemical Process 16.17

Changes may still occur before final publication online and in print
CH08CH16-Tsouris ARI 30 March 2017 13:55

12. Tsouris C, Porcelli JV. 2003. Process intensification—Has its time finally come? Chem. Eng. Prog. 99:50–
55
13. Portha JF, Falk L, Commenge JM. 2014. Local and global process intensification. Chem. Eng. Process.
84:1–13
14. Baldea M. 2015. From process integration to process intensification. Comput. Chem. Eng. 81:104–14
15. Cross W, Ramshaw C. 1986. Process intensification: laminar flow heat transfer. Chem. Eng. Res. Des.
64:293–301
16. Jachuck RJ, Lee J, Kolokotsa D, Ramshaw C, Valachis P, Yanniotis S. 1997. Process intensification for
energy saving. Appl. Ther. Eng. 17:861–67
17. Reay D. 2008. The role of process intensification in cutting greenhouse gas emissions. Appl. Ther. Eng.
28:2011–19
18. Ponce-Ortega JM, Al-Thubaiti MM, El-Halwagi MM. 2012. Process intensification: new understanding
and systematic approach. Chem. Eng. Process. 53:63–75
19. Moulijn JA, Stankiewicz A, Grievink J, Górak A. 2008. Process intensification and process systems
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org

engineering: a friendly symbiosis. Comput. Chem. Eng. 32:3–11


Access provided by Oak Ridge National Lab on 04/20/17. For personal use only.

20. Gerven TV, Stankiewicz A. 2009. Structure, energy, synergy, time—the fundamentals of process inten-
sification. Ind. Eng. Chem. Res. 48:2465–74
21. Stankiewicz A. 2003. Reactive separations for process intensification: an industrial perspective. Chem.
Eng. Process. 42:137–44
22. Yildirim Ö, Kiss AA, Kenig EY. 2011. Dividing wall columns in chemical process industry: a review on
current activities. Sep. Purif. Technol. 80:403–17
23. Harper JB, Easton CJ, Lincoln SF. 2003. A cyclodextrin-based molecular reactor to template the for-
mation of indigoid dyes. Tetrahedron Lett. 44:5815–18
24. Barr L, Dumanski PG, Easton CJ, Harper JB, Lee K, et al. 2004. Cyclodextrin molecular reactors.
J. Incl. Phenom. Macrocycl. Chem. 50:19–24
25. Górak A, Stankiewicz A. 2011. Intensified reaction and separation systems. Annu. Rev. Chem. Biomol.
Eng. 2:431–51
26. Rivas DF, Kuhn S. 2016. Synergy of microfluidics and ultrasound. Top. Curr. Chem. 374:1–30
27. European Federation of Chemical Engineering. 2008. European roadmap for process intensification. Frank-
furt: DECHEMA. http://efce.info/efce_media/-p-531-EGOTEC-juaogjr2ecs2b1ljpebr3qmvs2.
pdf?rewrite_engine = id
28. Thakur RK, Vial C, Nigam KDP, Nauman EB, Djelveh G. 2003. Static mixers in the process industries—
a review. Chem. Eng. Res. Des. 81:787–826
29. Kandhai D, Vidal DJE, Hoekstra AG, Hoefsloot H, Iedema P, Sloot PMA. 1999. Lattice-Boltzmann
and finite element simulations of fluid flow in a SMRX static mixer reactor. Int. J. Numer. Methods Fluids
31:1019–33
30. Mutsakis M, Streiff A, Schneider G. 1986. Advances in static mixing technology. Chem. Eng. Prog.
82:42–48
31. Liu P, Wu J, Yang G, Shao H. 2017. Comparison of static mixing reaction and reactive extrusion
technique for ring-opening polymerization of L-lactide. Mater. Lett. 186:372–74
32. Jensen KF. 2001. Microreaction engineering—Is small better? Chem. Eng. Sci. 56:293–303
33. Lerou JJ, Tonkovich AL, Silva L, Perry S, McDaniel J. 2010. Microchannel reactor architecture enables
greener processes. Chem. Eng. Sci. 65:380–85
34. Pohar A, Plazl I. 2009. Process intensification through microreactor application. Chem. Biochem. Eng. Q.
23:537–44
35. Anxionnaz Z, Cabassud M, Gourdon C, Tochon P. 2008. Heat exchanger/reactors (HEX reactors):
concepts, technologies: state-of-the-art. Chem. Eng. Process. 47:2029–50
36. Ferrouillat S, Tochon P, Garnier C, Peerhossaini H. 2006. Intensification of heat-transfer and mixing in
multifunctional heat exchangers by artificially generated streamwise vorticity. Appl. Therm. Eng. 26:1820–
29
37. Lee S, Liang L, Riestenberg D, West OR, Tsouris C, Adams E. 2003. CO2 hydrate composite for ocean
carbon sequestration. Environ. Sci. Technol. 37:3701–8

16.18 Kim et al.

Changes may still occur before final publication online and in print
CH08CH16-Tsouris ARI 30 March 2017 13:55

38. West OR, Tsouris C, Lee S, McCallum SD, Liang L. 2003. Negatively buoyant CO2 -hydrate composite
for ocean carbon sequestration. AIChE J. 49:283–85
39. Tsouris C, Brewer P, Peltzer E, Walz P, Riestenberg D, et al. 2004. Hydrate composite particles for
ocean carbon sequestration: field verification. Environ. Sci. Technol. 38:2470–75
40. Riestenberg D, Chiu E, Gborigi M, Liang L, West OR, Tsouris C. 2004. Investigation of jet breakup
and droplet size distribution of liquid CO2 and water systems—implications for CO2 hydrate formation
for ocean carbon sequestration. Am. Mineral. 89:1240–46
41. Riestenberg DE, Tsouris C, Brewer PG, Peltzer ET, Walz P, et al. 2005. Field studies on the formation
of sinking CO2 particles for ocean carbon sequestration: effects of injector geometry on particle density
and dissolution rate and model simulation of plume behavior. Environ. Sci. Technol. 39:7287–93
42. Szymcek P, McCallum SD, Taboada-Serrano P, Tsouris C. 2008. A pilot-scale continuous-jet hydrate
reactor. Chem. Eng. J. 135:71–77
43. Taboada-Serrano P, Ulrich S, Szymcek P, McCallum SD, Phelps TJ, et al. 2009. A multi-phase, micro-
dispersion reactor for the continuous production of methane gas hydrate. Ind. Eng. Chem. Res. 48:6448–52
44. Harvianto GR, Ahmad F, Lee M. 2016. Vapor permeation–distillation hybrid processes for cost-effective
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org
Access provided by Oak Ridge National Lab on 04/20/17. For personal use only.

isopropanol dehydration: modeling, simulation and optimization. J. Membr. Sci. 497:108–19


45. Ramshaw C, Mallinson RH. 1981. Mass transfer process. US Patent No. 4283255
46. Rao DP, Bhowal A, Goswami PS. 2004. Process intensification in rotating packed beds (HIGEE): an
appraisal. Ind. Eng. Chem. Res. 43:1150–62
47. Aoune A, Ramshaw C. 1999. Process intensification: heat and mass transfer characteristics of liquid films
on rotating discs. Int. J. Heat Mass Transf. 42:2543–56
48. Qiu Z, Petera J, Weatherley LR. 2012. Biodiesel synthesis in an intensified spinning disk reactor. Chem.
Eng. J. 210:597–609
49. Birdwell JF Jr., McFarlane J, Hunt RD, Luo H, DePaoli DW, et al. 2006. Separation of ionic liquid
dispersions in centrifugal solvent extraction contactors. Sep. Sci. Technol. 41:2205–23
50. Birdwell JF Jr., McFarlane J, Jennings H, Tsouris C. 2012. Integrated reactor and centrifugal separator and
uses thereof. US Patent No. 8097219
51. McFarlane J, Tsouris C, Birdwell JF Jr., Schuh DL, Jennings HL, et al. 2010. Production of biodiesel at
the kinetic limit in a centrifugal reactor/separator. Ind. Eng. Chem. Res. 49:3160–69
52. Tsouris C, McFarlane J, Birdwell JF Jr., Jennings HL. 2008. Continuous production of biodiesel via an
intensified reactive/extractive process. Proc. Int. Solvent Extr. Conf. 2008 Tucson, AZ, Sep. 15–19
53. Thompson LH, Doraiswamy LK. 1999. Sonochemistry: science and engineering. Ind. Eng. Chem. Res.
38:1215–49
54. Riera E, Golas Y, Blanco A, Gallego JA, Blasco M, Mulet A. 2004. Mass transfer enhancement in
supercritical fluids extraction by means of power ultrasound. Ultrason. Sonochem. 11:241–44
55. Riera E, Blanco A, Garcı́a J, Benedito J, Mulet A, et al. 2010. High-power ultrasonic system for the
enhancement of mass transfer in supercritical CO2 extraction processes. Ultrasonics 50:306–9
56. Shin WT, Yiacoumi S, Tsouris C. 2004. Electric-field effects on interfaces: electrospray and electroco-
alescence. Curr. Opin. Colloid Interface Sci. 9:249–55
57. Tsouris C, Borole AP, Kaufman EN, DePaoli DW. 1999. An electrically driven gas-liquid-liquid con-
tactor for bioreactor and other applications. Ind. Eng. Chem. Res. 38:1877–83
58. Weatherley LR. 1993. Electrically enhanced mass transfer. Heat Recov. Syst. CHP 13:515–37
59. Weatherley LR, Rooney DW, Niekerk MV. 1997. Clean synthesis of fatty acids in an intensive lipase-
catalysed bioreactor. J. Chem. Technol. Biotechnol. 68:437–41
60. Liu Z, Nueraihemaiti A, Chen M, Du J, Fan X, Tao C. 2015. Hydrometallurgical leaching process
intensified by an electric field for converter vanadium slag. Hydrometallurgy 155:56–60
61. Liu Z, Li Y, Chen M, Nueraihemaiti A, Du J, et al. 2016. Enhanced leaching of vanadium slag in acidic
solution by electro-oxidation. Hydrometallurgy 159:1–5
62. DePaoli DW, Tsouris C, Feng JQ. 1998. Method of electrically producing dispersions of nonconductive fluids
into conductive fluids. US Patent No. 5762775
63. Feng JQ, DePaoli DW, Tsouris C, Scott TC. 1995. Spraying fine fluid particles in insulating fluid
systems by electrostatic polarization forces. J. Appl. Phys. 78:2860–62

www.annualreviews.org • Modular Chemical Process 16.19

Changes may still occur before final publication online and in print
CH08CH16-Tsouris ARI 30 March 2017 13:55

64. Kaufman EN, Harkins JB, Rodriguez M, Tsouris C, Selvaraj PT, Murphy SE. 1997. Development of
an electro-spray bioreactor for crude oil processing. Fuel Process. Technol. 52:127–44
65. Shin WT, Yiacoumi S, Tsouris C. 1997. Experiments on electrostatic dispersion of air in water. Ind.
Eng. Chem. Res. 36:3647–55
66. Tsouris C, DePaoli DW, Feng JQ, Basaran OA, Scott TC. 1994. Electrostatic spraying of nonconductive
fluids into conductive fluids. AIChE J. 40:1920–23
67. Tsouris C, DePaoli DW, Feng JQ, Scott TC. 1995. Experimental investigation of electrostatic dispersion
of nonconductive fluids into conductive fluids. Ind. Eng. Chem. Res. 34:1394–403
68. Tsouris C, Neal SH, Shah VM, Spurrier MA, Lee MK. 1997. Comparison of liquid-liquid dispersions
formed by a stirred tank and electrostatic spraying. Chem. Eng. Commun. 160:175–97
69. Tsouris C, Shin WT, Yiacoumi S. 1998. Pumping, spraying, and mixing of fluids by electric fields. Can.
J. Chem. Eng. 76:589–99
70. Tsouris C, Shin WT, Yiacoumi S, DePaoli DW. 2000. Electrohydrodynamic velocity and pumping
measurements in water and alcohols. J. Colloid Interface Sci. 229:335–45
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org

71. Weatherley LR, Campbell I, Kirton D, Slaughter JC. 1990. Electrically enhanced extraction of penicillin
Access provided by Oak Ridge National Lab on 04/20/17. For personal use only.

G into dichloromethane. J. Chem. Technol. Biotechnol. 48:427–38


72. Weatherley LR, Rooney D. 2008. Enzymatic catalysis and electrostatic process intensification for pro-
cessing of natural oils. Chem. Eng. J. 135:25–32
73. Bazhal M, Vorobiev E. 2000. Electrical treatment of apple cossettes for intensifying juice pressing.
J. Sci. Food Agric. 80:1668–74
74. DePaoli DW, Hu MZC, Tsouris C. 2001. Method for the production of ultrafine particles by electrohydrody-
namic micromixing. US Patent No. 6265025
75. Depaoli DW, Tsouris C, Hu MZC. 2003. EHD micromixing reactor for particle synthesis. Powder
Technol. 135:302–9
76. Tsouris C, Culbertson CT, DePaoli DW, Jacobson SC, De Almeida VF, Ramsey JM. 2003. Electrohy-
drodynamic mixing in microchannels. AIChE J. 49:2181–86
77. Blankenship KD, Shah VM, Tsouris C. 1999. Distillation under electric fields. Sep. Sci. Technol. 34:1393–
409
78. Blankenship KD, DePaoli DW, Hylton JO, Tsouris C. 1999. Effect of electrode configurations on phase
equilibria with applied electric fields. Sep. Purif. Technol. 15:283–94
79. Tsouris C, Blankenship KD, Dong J, DePaoli DW. 2001. Enhancement of distillation efficiency by
application of an electric field. Ind. Eng. Chem. Res. 40:3843–47
80. Hou CH, Liang C, Yiacoumi S, Dai S, Tsouris C. 2006. Electrosorption capacitance of nanostructured
carbon-based materials. J. Colloid Interface Sci. 302:54–61
81. Mayes RT, Tsouris C, Kiggans JO Jr., Mahurin SM, DePaoli DW, Dai S. 2010. Hierarchical ordered
mesoporous carbon from phloroglucinol-glyoxal and its application in capacitive deionization of brackish
water. J. Mater. Chem. 20:8674–78
82. Sharma K, Mayes R, Kiggans J, Yiacoumi S, Gabitto J, et al. 2013. Influence of temperature on the
electrosorption of ions from aqueous solutions using mesoporous carbon materials. Sep. Purif. Technol.
116:206–13
83. Sharma K, Kim YH, Gabitto J, Mayes RT, Yiacoumi S, et al. 2015. Transport of ions in mesoporous
carbon electrodes during capacitive deionization of high-salinity solutions. Langmuir 31:1038–47
84. Sharma K, Kim YH, Yiacoumi S, Gabitto J, Bilheux HZ, et al. 2016. Analysis and simulation of a blue
energy cycle. Renew. Energy 91:249–60
85. Yang KL, Ying TY, Yiacoumi S, Tsouris C, Vittoratos ES. 2001. Electrosorption of ions from aqueous
solutions by carbon aerogel: an electrical double-layer model. Langmuir 17:1961–69
86. Yang KL, Yiacoumi S, Tsouris C. 2003. Electrosorption capacitance of nanostructured carbon aerogel
obtained by cyclic voltammetry. J. Electroanal. Chem. 540:159–67
87. Ying TY, Yiacoumi S, Tsouris C. 2002. An electrochemical method for the formation of magnetite
particles. J. Dispers. Sci. Technol. 23:569–76
88. Wang X, Lee JS, Tsouris C, DePaoli DW, Dai S. 2010. Preparation of activated mesoporous carbons
for electrosorption of ions from aqueous solutions. J. Mater. Chem. 20:4602–8

16.20 Kim et al.

Changes may still occur before final publication online and in print
CH08CH16-Tsouris ARI 30 March 2017 13:55

89. Tsouris C, Mayes R, Kiggans J, Sharma K, Yiacoumi S, et al. 2011. Mesoporous carbon for capacitive
deionization of saline water. Environ. Sci. Technol. 45:10243–49
90. Tsouris C, Mayes RT, Kiggans JO, DePaoli DW, Bourcier W, Campbell R. 2014. Increasing ion sorption
and desorption rates of conductive electrodes. US Patent No. 8920622
91. Tsouris C, Dong J. 2000. Effects of electric fields on phase inversion of liquid–liquid dispersions. Chem.
Eng. Sci. 55:3571–74
92. Tsouris C, Dong J. 2002. Methods to control phase inversions and enhance mass transfer in liquid-liquid
dispersions. US Patent No. 6495617
93. Dong J, Tsouris C. 2001. Phase inversion of liquid–liquid dispersions under applied electric fields.
J. Dispers. Sci. Technol. 22:57–69
94. Tsouris C, Yiacoumi S, Scott T. 1995. Kinetics of heterogeneous magnetic flocculation using a bivariate
population-balance equation. Chem. Eng. Commun. 137:147–59
95. Tsouris C, Scott TC. 1995. Flocculation of paramagnetic particles in a magnetic field. J. Colloid Interface
Sci. 171:319–30
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org

96. Ying TY, Yiacoumi S, Tsouris C. 2000. High-gradient magnetically seeded filtration. Chem. Eng. Sci.
Access provided by Oak Ridge National Lab on 04/20/17. For personal use only.

55:1101–13
97. Tsouris C, Noonan J, Ying JY, Yiacoumi S. 2006. Surfactant effects on the mechanism of particle capture
in high-gradient magnetic filtration. Sep. Purif. Technol. 51:201–9
98. Taboada-Serrano P, Tsouris C, Contescu C, McFarlane J. 2013. Magnetic filtration process, magnetic
filtering material, and methods for forming magnetic filtering material. US Patent No. 8551617
99. Ozokwelu D. 2014. High efficiency modular chemical processes (HEMCP). Washington, DC: US Dep.
Energy. 12 pp. https://energy.gov/sites/prod/files/2014/10/f18/hemcp-topic-overview.pdf
100. Teng H, Yamasaki A, Shindo Y. 1996. Stability of the hydrate layer formed on the surface of a CO2
droplet in high-pressure, low-temperature water. Chem. Eng. Sci. 51:4979–86
101. Gabitto J, Riestenberg D, Lee S, Liang L, Tsouris C. 2005. Ocean disposal of CO2 : conditions for
producing sinking CO2 hydrate. J. Dispers. Sci. Technol. 25:703–12
102. Kim YJ, Choi JH. 2010. Enhanced desalination efficiency in capacitive deionization with an ion-selective
membrane. Sep. Purif. Technol. 71:70–75
103. Jeon SI, Park HR, Yeo JG, Yang S, Cho CH, et al. 2013. Desalination via a new membrane capacitive
deionization process utilizing flow-electrodes. Energ. Environ. Sci. 6:1471–75
104. Qiu Z, Zhao L, Weatherley L. 2010. Process intensification technologies in continuous biodiesel pro-
duction. Chem. Eng. Process. 49:323–30
105. Kim TI, Kim YH, Han M. 2012. Development of novel oil washing process using bubble potential
energy. Mar. Poll. Bull. 64:2325–32
106. Cherry RS, Hulle CT. 1992. Cell death in the thin films of bursting bubbles. Biotechnol. Prog. 8:11–18
107. Kunas KT, Papoutsakis ET. 1990. Damage mechanisms of suspended animal cells in agitated bioreactors
with and without bubble entrainment. Biotechnol. Bioeng. 36:476–83
108. Burns SE, Yiacoumi S, Tsouris C. 1997. Microbubble generation for environmental and industrial
separations. Sep. Purif. Technol. 11:221–32
109. Etchells JC. 2005. Process intensification: safety pros and cons. Process. Saf. Environ. Prot. 83:85–89
110. Hendershot DC. 2004. Process intensification for safety. In Re-engineering the Chemical Processing Plant:
Process Intensification, ed. A Stankiewicz, JA Moulijn, pp. 471–94. New York: Marcel Dekker
111. Belluomini GJ, Pendergast JG, Domke CH, Ussing BR. 2009. Performance of several ionic liquids for
the separation of 1-octene from n-octane. Ind. Eng. Chem. Res. 48:11168–74
112. Mercer AC. 1993. Process intensification—the UK programmes to encourage the development and use
of intensified heat exchange equipment and technology. Heat Recov. Syst. CHP 13:539–45
113. Efthimeros GA, Tsahalis DT. 2000. Intensified energy-saving technologies developed in EU-funded
research: a review. Appl. Ther. Eng. 20:1607–13
114. AIChE (Am. Inst. Chem. Eng.). 2016. U.S. Department of Energy taps AIChE to lead RAPID Modular
Process Intensification Institute. News Release, Dec. 9. http://www.aiche.org/about/press/releases/
12–20–2016/us-department-energy-taps-aiche-lead-rapid-modular-process-intensification-
institute

www.annualreviews.org • Modular Chemical Process 16.21

Changes may still occur before final publication online and in print
CH08CH16-Tsouris ARI 30 March 2017 13:55

115. Luyben WL, Hendershot DC. 2004. Dynamic disadvantages of intensification in inherently safer process
design. Ind. Eng. Chem. Res. 43:384–96
116. Shallan AI, Smejkal P, Corban M, Guijt RM, Breadmore MC. 2014. Cost-effective three-dimensional
printing of visibly transparent microchips within minutes. Anal. Chem. 86:3124–30
117. Symes MD, Kitson PJ, Yan J, Richmond CJ, Cooper GJ, et al. 2012. Integrated 3D-printed reactionware
for chemical synthesis and analysis. Nat. Chem. 4:349–54
Annu. Rev. Chem. Biomol. Eng. 2017.8. Downloaded from www.annualreviews.org
Access provided by Oak Ridge National Lab on 04/20/17. For personal use only.

16.22 Kim et al.

Changes may still occur before final publication online and in print

You might also like