You are on page 1of 23

CH07CH17-Sadowski ARI 3 May 2016 12:4

ANNUAL
REVIEWS Further
Click here to view this article's
online features:
• Download figures as PPT slides
• Navigate linked references
• Download citations
• Explore related articles
Thermodynamics of
• Search keywords
Bioreactions
Annu. Rev. Chem. Biomol. Eng. 2016.7:395-414. Downloaded from www.annualreviews.org

Christoph Held and Gabriele Sadowski


Access provided by University of Nebraska - Lincoln on 06/10/16. For personal use only.

Laboratory of Thermodynamics, Department of Biochemical and Chemical Engineering,


Technische Universität Dortmund, 44227 Dortmund, Germany;
email: gabriele.sadowski@bci.tu-dortmund.de

Annu. Rev. Chem. Biomol. Eng. 2016. 7:395–414 Keywords


The Annual Review of Chemical and Biomolecular Gibbs energy of reaction, biothermodynamics, activity coefficients,
Engineering is online at chembioeng.annualreviews.org
solubility, reaction equilibrium
This article’s doi:
10.1146/annurev-chembioeng-080615-034704 Abstract
Copyright  c 2016 by Annual Reviews. Thermodynamic principles have been applied to enzyme-catalyzed reactions
All rights reserved
since the beginning of the 1930s in an attempt to understand metabolic
pathways. Currently, thermodynamics is also applied to the design and anal-
ysis of biotechnological processes. The key thermodynamic quantity is the
Gibbs energy of reaction, which must be negative for a reaction to occur
spontaneously. However, the application of thermodynamic feasibility stud-
ies sometimes yields positive Gibbs energies of reaction even for reactions
that are known to occur spontaneously, such as glycolysis. This article re-
views the application of thermodynamics in enzyme-catalyzed reactions. It
summarizes the basic thermodynamic relationships used for describing the
Gibbs energy of reaction and also refers to the nonuniform application of
these relationships in the literature. The review summarizes state-of-the-art
approaches that describe the influence of temperature, pH, electrolytes, sol-
vents, and concentrations of reacting agents on the Gibbs energy of reaction
and, therefore, on the feasibility and yield of biological reactions.

395
CH07CH17-Sadowski ARI 3 May 2016 12:4

1. INTRODUCTION
The application of thermodynamics to biological reactions began in the 1930s with the study
of the role of the energy carrier adenosine triphosphate (ATP) in metabolism (1, 2). Based on
the knowledge acquired, biochemical researchers, especially the groups led by Goldberg (3–8)
and Alberty (9–13), embarked on thermodynamic research into biological reactions and their
reacting agents. These researchers and others (14, 15) aimed to understand metabolism by applying
thermodynamic principles.
Bioreactions are catalyzed by enzymes (either isolated enzymes or via whole-cell catalysis) and
often require an aqueous reaction environment. Enzymes and their corresponding reactions can be
classified into six major classes, as recommended by the International Union of Pure and Applied
Chemistry (IUPAC) (16). All six reaction types have already been the focus of thermodynamic
investigations by, for example, Goldberg, Tewari, and their colleagues, who considered each of
Annu. Rev. Chem. Biomol. Eng. 2016.7:395-414. Downloaded from www.annualreviews.org
Access provided by University of Nebraska - Lincoln on 06/10/16. For personal use only.

the six major enzyme classes: oxidoreductases (17), transferases (18), hydrolases (19), lyases (20),
isomerases (21), and ligases (21).
Approximately 20 years ago, von Stockar & van der Wielen (22) applied thermodynamics to
the design and analysis of biotechnological processes, verifying its huge potential for designing
novel and sustainable processes for industrial biotechnology. This proof of concept established
a new awareness of how thermodynamic principles could be applied to biological processes, an
application that is known as biothermodynamics. Heijnen & Van Dijken (23) proposed a rela-
tionship between biomass yield and thermodynamic properties that has been applied to predict
biomass yield (24) and maintenance coefficients (25). Further applications of the use of biothermo-
dynamic tools include estimating maximum growth rates, analyzing bioprocess kinetics by means
of nonequilibrium thermodynamics, and predicting metabolic networks (26–28). A prominent
example for a metabolic pathway is the analysis of glycolysis, which is of particular interest in
biochemical research (29).
This review summarizes the basic properties of, and the relationships applied in, biothermody-
namics, as well as state-of-the-art approaches used to describe the influence of temperature, pH,
electrolytes, solvents, and the concentrations of the reaction agents on reaction equilibrium and,
thus, on the feasibility and yield of biological reactions. However, the influence of these system
properties on enzyme activity, stability, and kinetics is not the focus of this review.

2. GIBBS ENERGY OF REACTION


From the thermodynamic point of view, reactions are characterized by their driving force R g,
the Gibbs energy of reaction, which determines whether or not a reaction occurs spontaneously.
A spontaneous reaction occurs only when R g values are negative. These reactions are, therefore,
called thermodynamically feasible. R g is obtained from the sum of two terms: the standard Gibbs
energy of reaction, R g 0 , and a second term that contains the ratio of the thermodynamic activities
of the reacting agents:
 
ν
R g = R g 0 + RT ln ai i , 1.

where R and T are, respectively, the universal gas constant and the reaction temperature. The
a i values are the thermodynamic activities of the species i that participate in the reaction ac-
cording to their stoichiometric coefficient νi . The standard Gibbs energy of reaction, R g 0 (see
Section 3), depends only on temperature. The thermodynamic activities of the reacting agents
depend on temperature, concentration, and the activity coefficients of the reacting agents (see
Section 4).

396 Held · Sadowski


CH07CH17-Sadowski ARI 3 May 2016 12:4

pH 7.0
42 pH 7.5
pH 8.0

–ΔRg0 (kJ/mol) 39

36

33

30
Annu. Rev. Chem. Biomol. Eng. 2016.7:395-414. Downloaded from www.annualreviews.org
Access provided by University of Nebraska - Lincoln on 06/10/16. For personal use only.

27

20 24 28 32 36 40
T (°C)
Figure 1
Published standard Gibbs energies of reactions for ATP hydrolysis at different temperatures and pH values
(37–46).

By using Equation 1, tools from biochemistry, such as thermodynamic feasibility analysis, have
been applied to calculations for metabolic networks (26, 27, 30, 31). Mavrovouniotis (27) and
Vojinovic & von Stockar (32), as well as Maskow & von Stockar (33), analyzed glycolysis using
experimental nonequilibrium concentrations of the reacting agents and R g 0 data, which were
available from a previous study (34). They found that although there is no doubt that glycolysis
occurs, unreasonable concentrations of reacting agents had to be assumed to obtain a negative
value for R g. Although why this is so appears to be a mystery, it implies that the application of
thermodynamics to biological reactions is straightforward yet difficult, and requires careful and
thermodynamically correct analysis of the data.
Contradictory observations, such as positive R g data for obviously spontaneous reactions,
are usually assumed to be caused by experimental uncertainties (35). Additionally, however, using
the concentrations of reacting agents instead of their thermodynamic activities in Equation 1 and
applying available R g 0 data requires critical analysis (see Sections 3.3 and 6).
As an example, Figure 1 summarizes the R g 0 values for ATP hydrolysis found in the literature.
Obviously, there is a huge scatter in these R g 0 values. Data differ by up to 25%, even at a constant
temperature. As ATP hydrolysis is a key biochemical reaction and usually occurs simultaneously
with other reactions, the choice of its R g 0 value also influences the conclusions drawn for many
associated reactions (36).
Thus, to avoid confusion and even misleading conclusions, it is important to keep in mind that
published R g 0 values are usually not obtained at standard conditions and can, therefore, not be
directly used in Equation 1. Published R g 0 values have been obtained by different authors under
different conditions, which are sometimes not precisely detailed, and this explains the scatter of
data in Figure 1. Thus, R g 0 values must be carefully analyzed or recalculated, or both, before
use. This is discussed in Section 3.
In addition to inconsistencies in R g 0 values, the application of Equation 1 requires using
the thermodynamic activities of the reacting agents. These are obtained as the product of the
concentrations of the reacting agents and their activity coefficients (see Section 4). In summary,
R g and the equilibrium of a bioreaction are influenced by temperature, pH, the presence of

www.annualreviews.org • Thermodynamics of Bioreactions 397


CH07CH17-Sadowski ARI 3 May 2016 12:4

electrolytes and solvents, and the concentrations of all reacting agents. These properties must
be considered when performing thermodynamic feasibility studies of bioreactions. This review
discusses the basic ideas of, and existing approaches to, studying bioreactions.

3. STANDARD GIBBS ENERGY OF A REACTION, R g0

3.1. The Standard State


According to IUPAC, the index 0 is used to denote a standard state for the Gibbs energy of a
reaction, regardless of which standard state is meant. This means that different values for R g 0
can be obtained for the same reaction when using different standard states. Thus, the use of the
index 0 must always be accompanied by specification of the standard state that was actually used.
Annu. Rev. Chem. Biomol. Eng. 2016.7:395-414. Downloaded from www.annualreviews.org
Access provided by University of Nebraska - Lincoln on 06/10/16. For personal use only.

Two standard states are commonly used for the reacting agents: the standard-state pure com-
ponent and the standard-state hypothetical ideal solution. The first is usually applied to chemical
reactions and is rarely used for bioreactions (30, 31). As most biological reactions are performed
at low concentrations of reacting agents, their thermodynamic behavior is quite different from the
pure-component standard state. This usually makes the pure-component standard state unfeasible
in these cases. Therefore, the preferred standard state for the reacting agents in a bioreaction is a
solution of 1 mol reacting agent in 1 kg of water, assuming the interactions between the molecules
are the same as in an infinitely diluted solution (this is referred to as the hypothetical ideal solution)
(37–48).
The standard state can be chosen separately for each reacting agent and does not necessarily
need to be the same for all reacting agents of a reaction. This applies to reactions in which one
of the reacting agents is also used as solvent (for example, water) and is, therefore, present in
much higher concentrations than the other agents. The case in which a reacting agent acts as a
solvent is preferably described as the pure-component standard state, whereas the hypothetical
ideal solution is chosen as the standard state for the diluted reacting agents (48, 49).

3.2. Accessing R g0 via the Thermodynamic Properties of Formation


If the pure-component standard state is used for all reacting agents in a reaction, the standard
Gibbs energy of reaction, R g 0 , can be obtained as a linear combination of the standard Gibbs
energies of formation, F gi0 , of the reacting agents according to

R g 0 = νi · F gi0 , 2.

in which νi is the stoichiometric coefficient of the reacting agents in the reaction. The standard
Gibbs energy of formation of a species i, F gi0 , is the Gibbs energy change for a reaction that
exclusively forms 1 mol of the species from its elements. F gi0 refers to the pure species and, thus,
depends only on temperature.
F gi0 values are not directly accessible and are usually obtained using enthalpies of forma-
tion, F h i0 , which can be measured or determined from combustion enthalpies, and entropies of
formation, F s i0 , obtained from heat capacities measured at low temperatures or from quantum
chemistry (50, 51) according to (52)

F gi0 = F h i0 − T F s i0 . 3.

Enthalpies of formation of biological compounds have been extensively determined since the
beginning of the past century, mainly by heat-of-combustion measurements of pure compounds,

398 Held · Sadowski


CH07CH17-Sadowski ARI 3 May 2016 12:4

such as glucose (53–55), fructose (56), pyruvic acid (57), dihydroxyacetone (58), or glyceraldehyde
(59).
Quantum chemistry–based methods have already been applied to predict standard Gibbs en-
ergies of metabolic reactions and thermodynamic gas-phase properties of biological compounds
(for example, see References 60, 61).

3.3. Accessing R g0 via Equilibrium Reaction Analysis


Another possibility for accessing R g 0 is to perform reactions until thermodynamic equilibrium
is reached (R g = 0) and to determine R g 0 from the thermodynamic equilibrium constant K a :

R g 0 = −RT ln(K a ). 4.
Annu. Rev. Chem. Biomol. Eng. 2016.7:395-414. Downloaded from www.annualreviews.org

As  g depends on the choice of the standard state (see Section 3.1), K a must be chosen accord-
Access provided by University of Nebraska - Lincoln on 06/10/16. For personal use only.

R 0

ingly. If the hypothetical ideal solution is chosen as the standard state for all reacting agents, the
equilibrium constant K a also refers to that standard state. Thus, it is obtained by
 ∗ν  ν  ∗ν
Ka = ai i = mi i · γm,ii = K m · K γ ∗ . 5.

K a is defined as the product of the rational activities a i∗ (= mi · γ∗m,i ) of each reacting agent expo-
nentiated with its stoichiometric coefficient, and it can also be expressed as the product of K m and
K γ ∗ , which represent, respectively, the ratio of molalities, mi , and the ratio of rational activity
coefficients, γ∗m,i , being 1 at infinite dilution.
In this case, R g 0 is obtained, according to Equation 4, from Equation 6:
   
R g 0 = −RT ln K m − RT ln K γ∗ . 6.

Often the apparent Gibbs energy of reaction, R g 0,app , is used in the literature (34) instead of the
true standard value, R g 0 . In this case, apparent means that this property is directly determined
from measured equilibrium molalities, which are used to calculate K m . The apparent and the true
Gibbs energies of reactions are related by
 
R g 0,app = −RT ln (K m ) = R g 0 + RT ln K γ ∗ . 7.

In contrast to R g 0 , the apparent property, R g 0,app , depends on the activity coefficients of the
reacting agents and, therefore, also on the concentrations of reacting agents and on the reaction
medium (solvent, additives). This also explains the scatter of R g 0 data in Figure 1. Although
usually called the standard Gibbs energies of reaction, most published values obviously are R g 0,app
values determined under different conditions.
If the pure-component standard state is used for all reacting agents in a reaction, K a must be
determined from the mole fractions, xi , and the generic activity coefficients of the reacting species,
γi , which are 1 for the pure-component state. To avoid confusion with K a from Equation 5, some
researchers (30, 31) have renamed the equilibrium constant K th . Using xi · γi = mi · γm,i , mole
fractions can be replaced by molalities, which can then be used as concentration units:
 ν  ν  ν
K th = ai i = mi i · γm,i
i
= Km · Kγ . 8.

Unfortunately, many research groups do not use subscripts with the equilibrium constant K. In
these cases, the reader knows neither the considered standard state nor the concentration unit used
to obtain the K values; hence, these values should be used cautiously for further investigations.
As R g 0 depends only on temperature, according to Equation 4, the activity-based equilibrium
constants K a and K th also depend only on temperature. In contrast, both K m and K γ (or K γ ∗ )

www.annualreviews.org • Thermodynamics of Bioreactions 399


CH07CH17-Sadowski ARI 3 May 2016 12:4

depend on the reaction medium (solvent, cosolvents, additives, pH) and the concentration of
reacting agents. According to Equations 5 and 8, K a and K th values are accessible using measured

equilibrium molalities (→ K m ) and the activity coefficients of the reacting agents, γi,m or γi,m .
The latter depend on the concentrations of the reacting agents and the system composition and
must be estimated using a thermodynamic model (30, 31, 47, 48). For an infinite dilution of all
reacting agents, K γ ∗ becomes truly 1. In these cases, experimental K m values measured at different
molalities of reacting agents can be extrapolated to 0 molality of all reacting agents, leading to K a
(48). Often (for example, all references in 34, 62), however, K γ ∗ in Equation 5 is assumed to be
1 irrespective of the concentrations of the reacting agents and also in cases in which at least one
of the reacting agents (if not all) are not highly diluted. Consequently, this leads to inaccurate K a
values and, therefore, to inaccuracies in R g 0 . The same is true if K γ is neglected in Equation 8.
R g 0 , determined from K th via Equation 8, is also related to the Gibbs energies of formation,
Annu. Rev. Chem. Biomol. Eng. 2016.7:395-414. Downloaded from www.annualreviews.org

via Equation 2. Analogously, R g 0 , resulting from K th instead of K a in Equation 4, can also be


Access provided by University of Nebraska - Lincoln on 06/10/16. For personal use only.

obtained from the respective properties of the reacting species ∞ gi0 , which are the Gibbs energies
of species i in a 1 molal aqueous solution exhibiting the same interactions as in the infinitely diluted
state,

R g 0 = νi · ∞ gi0 . 9.

In turn, the species properties ∞ gi0 , which are hardly accessible directly, can be obtained from
Equation 9, once R g 0 is known for a set of reactions related to a number of species forming a
system of a certain number of equations and the same number of unknowns, ∞ gi0 . The obtained
∞ gi0 values are tabulated for many biological compounds, especially in the National Bureau of
Standards tables (63) or in Thermodynamics of Biochemical Reactions by Alberty (34). (It is worth
mentioning that the species properties ∞ gi0 are, of course, different from F gi0 . Nevertheless,
∞ gi0 is often also called the Gibbs energy of formation.) Based on these data, group-contribution
methods (64, 65) have been developed for calculating ∞ gi0 , which can now be used to determine
the standard Gibbs energies of metabolic reactions.

4. ACTIVITY COEFFICIENTS OF REACTING AGENTS


The application of Equation 5 and Equation 8 requires the activity coefficients for all reacting
agents. Generic activity coefficients can be obtained either by using the ratio of fugacity coefficients
from Equation 10 or by using the partial molar property of the excess Gibbs energy, GE /RT ,
according to Equation 11:
ϕi (T , p, xi )
γi = 10.
ϕ0i (T , p, xi = 1)
and

∂GE
RT ln(γi ) = . 11.
∂ni T , p,n j =i

The fugacity coefficient ϕi of component i in a mixture and the fugacity coefficient of the pure
component ϕ0i in Equation 10 can be calculated using an equation of state, such as Perturbed-
Chain Statistical Associating Fluid Theory (PC-SAFT) (66, 67). Moreover, GE models, such as
UNIQUAC Functional-Group Activity Coefficients (UNIFAC) (68, 69), Conductor like Screen-
ing Model for Realistic Solvents (COSMO-RS) (70, 71), and the nonrandom two-liquid model
(72), have already been used to calculate the activity coefficients of reacting agents in biological

400 Held · Sadowski


CH07CH17-Sadowski ARI 3 May 2016 12:4

reactions (31, 73, 74). The generic activity coefficient of a reacting agent in a solvent is also
accessible experimentally from solubility data of the reacting agent (see Section 5).
The so-called rational activity coefficient, γi∗ , is related to the generic activity coefficient by
the generic activity coefficient at infinite dilution, γi∞ = γi (xi → 0):
γi
γi∗ = . 12.
γi∞
γi∗ can thus be obtained based on Equation 12 by applying the models already mentioned for
calculating the generic activity coefficient.

5. SOLUBILITY OF REACTING AGENTS


The equilibrium of many bioreactions is located on the reactant side, whereby the reaction con-
Annu. Rev. Chem. Biomol. Eng. 2016.7:395-414. Downloaded from www.annualreviews.org
Access provided by University of Nebraska - Lincoln on 06/10/16. For personal use only.

version is equilibrium-limited. The yield of these reactions can be enhanced by increasing the
substrate concentration. However, this increase is limited by the solubility of the substrates in
the reaction media. Thus, the solubility of a reacting agent in the reaction medium is an impor-
tant property, especially as many reacting agents have very poor solubility in water [for example,
acetophenone (75) or oligopeptides (76)] or in organic solvents [for example, amino acids (77)].
Substrate solubility may also be influenced by additives present in the reaction medium (for
example, electrolytes, cosolvents) (78, 79). Ionic liquids have recently been proposed to serve
as solvents or cosolvents in reaction media (80–82). Dreyer & Kragl (75) have shown that the
solubility of acetophenone increases exponentially upon addition of the ionic liquid Ammoeng
110 (IoLiTec Ionic Liquids Technologies GmbH, Germany). Swatloski et al. (83) demonstrated
the huge influence of ionic liquids on the solubility of cellulose in water.
Solubility data are important not only for the obvious information that they offer but because
knowledge of the solubility of a reacting agent also provides access to the generic activity coef-
ficient of that reacting agent in a solvent. The solubility of a component i in mole fraction xi
can be calculated from the equilibrium condition between the liquid and the solid phase (52)
(Equation 13):


1 h SL T
xiL = · exp − 0i 1 − SL , 13.
γi RT T 0i
where h SL SL
0i and T 0i are, respectively, the enthalpy of melting and the melting temperature of the
pure component i. This relation assumes a pure solid phase and neglects the difference between
the solid and liquid heat capacities of component i. The melting properties are usually accessible
from calorimetric measurements (84) or group-contribution methods (85).
Because solubility is a function of solvent, temperature, pH, and additives, thermodynamic
models that are able to account for these influences become increasingly important. Numerous
thermodynamic models have been applied to describe the solubility of biomolecules by using
Equation 13, and these have been reviewed by Khoshkbarchi & Vera (86). In the past decade,
PC-SAFT has been applied to model the solubilities and activity coefficients of amino acids (87),
compatible solutes (88, 89), and sugars (90) in aqueous solutions, in multi-solvent systems (91,
92), and even in multi-solute systems (87, 88, 90, 93–97). Zeuner et al. (73) applied COSMO-RS
to model the solubility of methyl ferulate in different solvents and solvent mixtures. Other studies
have shown that thermodynamic models are, to a certain extent, able to predict the influences
(solvent, temperature, pH, and additives) on the solubility of biomolecules (51, 84, 98, 99, 100).
As an example, Figure 2 compares the capabilities of COSMO-RS and PC-SAFT to predict
the solubility of methyl ferulate in a mixture of ethyl acetate and an ionic liquid ([bmim][BF4 ] or
[omim][BF4 ]). In this case, prediction means that the experimental solubilities have not been used

www.annualreviews.org • Thermodynamics of Bioreactions 401


CH07CH17-Sadowski ARI 3 May 2016 12:4

COSMO-RS predictions
PC-SAFT predictions
0.4 Experimental data

0.3

xMFA(–)
0.2

0.1
Annu. Rev. Chem. Biomol. Eng. 2016.7:395-414. Downloaded from www.annualreviews.org
Access provided by University of Nebraska - Lincoln on 06/10/16. For personal use only.

0.0
[bmim] [BF4] [omim] [BF4]

Figure 2
Solubility of methyl ferulate (MFA) in an equimolar mixture of ethyl acetate and the ionic liquid
[bmim][BF4 ] or [omim][BF4 ] at 303.15 K and 1 bar. Experimental data from Reference 101; COSMO-RS
predictions taken from Reference 73. The pure-component PC-SAFT parameters were taken from the
literature (47, 67, 102, 103) without fitting any binary parameters to the solubility data shown in this figure.
Abbreviation: MFA, methyl ferulate.

to parameterize the two models. The results obtained with COSMO-RS are in fair agreement
with predictions of the solubility of methyl ferulate in the two solvent mixtures. In contrast, the
application of PC-SAFT allows for almost quantitative prediction of the solubility of methyl
ferulate in the two solvent mixtures. Thus, the solubility of reacting agents in different reaction
media is accessible by using thermodynamic models. These models are also able to account for
the effects of temperature, pH, and solvent on the solubility of reacting agents.

6. BIOREACTION EQUILIBRIA

6.1. Influence of Temperature


Equilibrium constants and, thus, R g 0 values depend on temperature. The temperature influence
on R g 0 can be obtained either from equilibrium constants determined at different temperatures
or by using the Van ‘t Hoff equation. The latter provides a relation between the standard enthalpy
of reaction R h 0 and the temperature dependence of the thermodynamic equilibrium constant
K a (or K th ) at isobaric conditions:

R h 0 ∂ln (K a )
= − . 14.
RT 2 ∂T p

R h 0 is usually directly determined by calorimetry (104). In particular, microfluid and chip


calorimetry are appropriate methods for measuring the R h 0 of bioreactions (105–108).
Van ‘t Hoff ’s equation can also be used to compare R h 0 from direct measurements (calorime-
try) and from indirect methods (via the temperature-dependent equilibrium constants K a or K th ).
This is useful because it allows for consistency tests of the data. Equation 14 can be conveniently
integrated, assuming that R h 0 is temperature independent. This is a reasonable assumption,
as bioreactions are typically considered within a small temperature range. The integration of
Equation 14 provides a linear relation between the logarithm of K a and the reciprocal absolute
temperature. The slope of this relation (−R h 0 /R) provides access to R h 0 . This procedure is

402 Held · Sadowski


CH07CH17-Sadowski ARI 3 May 2016 12:4

–1.50

–1.25

ln(Ka) (–)

–1.00

–0.75
3.20 3.25 3.30 3.35 3.40
Annu. Rev. Chem. Biomol. Eng. 2016.7:395-414. Downloaded from www.annualreviews.org

1/T 103 (1/K)


Access provided by University of Nebraska - Lincoln on 06/10/16. For personal use only.

Figure 3
Equilibrium constants K a versus the reciprocal absolute temperature of glucose-6-phosphate isomerization,
taken from Hoffmann et al. (48). The slope of the line equals −R h 0 /R, yielding the enthalpy of reaction,
which was determined to be 12.25 ± 0.3 kJ/mol.

demonstrated for the enzyme-catalyzed isomerization of glucose-6-phosphate (G6P) to fructose-


6-phosphate (F6P) in Figure 3. K a values were determined at three temperatures by Hoffmann
et al. (48). The R h 0 obtained using Van ‘t Hoff ’s equation (12.25 kJ/mol) was found to be in
excellent agreement with the R h 0 measured by calorimetry (12.05 kJ/mol). To conclude, the
R g 0 of bioreactions depends on temperature. This temperature dependence is accessible via
R h 0 data.

6.2. Influence of pH
The thermodynamic properties of a system may change dramatically with changes in pH. The main
reason for this is the possible dissociation of reacting agents resulting in a change in concentration
of neutral or ionized species upon changes in pH, even at a constant total concentration of a
reacting agent; this is called the pH-dependent species distribution of the reacting agent. The
more functional acidic or basic groups a reacting agent has, the higher the species diversity as
a function of pH. This dissociation behavior is usually described by dissociation constants and,
moreover, depends on temperature.
Dissociation constants of biomolecules can be found in the literature (109), allowing for the
prediction of species distribution (110). The concentration of a reacting agent in a biochemical
system is usually described as the sum of its species’ concentrations, leading to apparent K m
values (111). A change in a solution’s pH causes a change in the species contributing to R g 0 or
K m . Therefore, measured K m values apparently depend on pH (see, for example, Figure 1 for
ATP hydrolysis). As pH might strongly influence the equilibrium of enzyme-catalyzed reactions,
reaction media are usually buffered, or acids or bases are used, to keep pH constant during the
reaction.
Hoffmann et al. (47, 48) and Goldberg et al. (49) measured the influence of pH on different
bioreactions. They found that even small changes in pH strongly influenced the K m values of
methyl ferulate hydrolysis (47, 49). In that reaction, the substrate methyl ferulate is an ester,
whereas the product is an acid (ferulic acid). Hence, pH changes do not influence the species
distribution of the reactant, but they strongly influence the species distribution of the product. This
explains the strong change in apparent K m values occurring with pH in methyl ferulate hydrolysis.

www.annualreviews.org • Thermodynamics of Bioreactions 403


CH07CH17-Sadowski ARI 3 May 2016 12:4

In contrast, Hoffmann et al. (48) observed that apparent K m values for G6P isomerization did not
change with pH variation. They stated that this was caused by the similar dissociation constants
of the reactant (G6P) and the product (F6P), which led to similar changes in species distributions
upon pH change.
Because the influence of pH on K m values is caused mainly by the ratio of charged to neutral
species of a reacting agent, several research groups have developed equations that relate Gibbs
energies to dissociation constants to describe the effects of pH on reactions. Alberty (34) proposed
an approach that describes the change in ∞ gi0 (which he called Gibbs energies of formation;
see Section 3.3) upon pH variation based on dissociation constants of the reacting agents and the
solution’s pH. Similarly, Alberty (34) also developed a correlation for the pH dependence of the
apparent standard Gibbs energy of a reaction, R g 0,app (Equation 9), which neglects the activity
coefficients of the reacting agents (see Equation 7). He proposed the use of so-called binding
Annu. Rev. Chem. Biomol. Eng. 2016.7:395-414. Downloaded from www.annualreviews.org

polynomials, Pi , which are defined as the ratio of the total concentration, c itot , of a reacting agent
Access provided by University of Nebraska - Lincoln on 06/10/16. For personal use only.

species
i and the concentrations of its species formed by a dissociation reaction j (Pi = c itot /c i, j ).
Calculating this ratio requires the dissociation constants of the reacting agents (34). The final
expression for K m is then obtained as
  (Pi )νi
K m (pH) = K m (pH = 7) · . 15.
i
([H+ ])νi

To demonstrate the applicability of Equation 15, it is applied here to the methyl ferulate hydrolysis
experimentally investigated by Hoffmann et al. (47). Use of the experimental K m value of Hoffmann
et al. at pH 7 [K m (pH = 7) ≈ 62] in Equation 15 results in pH-dependent K m values, as illustrated
in Figure 4. As can be observed, the correlation in Equation 15 shows that the K m values of methyl
ferulate hydrolysis between pH 6 and pH 7 are in qualitative agreement with the experimental
data. This is a surprisingly good result considering that Alberty’s (34) correlation requires only
the dissociation constants of all reacting agents, the solution pH, and one experimental K m value.
Thus, K m values may strongly depend on pH, irrespective of the concentrations of the reacting
agents or of temperature. To calculate the influence of pH on K m , it is recommended that existing
correlations that use the dissociation constants of the reacting agents be applied.

70

60

50

40
Km (–)
30

20

10

0
6.0 6.2 6.4 6.6 6.8 7.0
pH
Figure 4
K m values of methyl ferulate hydrolysis at 298.15 K in phosphate buffer as a function of pH at
mmethyl ferulate, init = 6 mmol/kg. Circles are experimental data taken from Hoffmann et al. (47). The solid
black line represents an exponential regression of the experimental K m values. The green dashed line
represents the pH correlation according to Equation 15, using K m at pH 7 as an input.

404 Held · Sadowski


CH07CH17-Sadowski ARI 3 May 2016 12:4

PC-SAFT predictions
Equation 18 calculation
Experimental values
200

Km (–)
100
Annu. Rev. Chem. Biomol. Eng. 2016.7:395-414. Downloaded from www.annualreviews.org
Access provided by University of Nebraska - Lincoln on 06/10/16. For personal use only.

0
0.0 0.2 0.4 0.6 0.8 1.0
mNaCl(mol/kg)

Figure 5
K m values of methyl ferulate hydrolysis at 298.15 K and mmethyl ferulate, init = 6 mmol/kg as a function of
NaCl molality. Circles: experimental values (47); solid black line: prediction using electrolyte PC-SAFT
(47); green dashed line: calculation using Equation 18.

6.3. Influence of Electrolytes


Ions and charged molecules are known to influence reaction equilibria and phase equilibria (96).
On the one hand, bioreactions may contain charged reacting agents; on the other hand, the reaction
medium often contains charged components, for example, salts, cofactors, buffer, or cosubstrates.
The nonideality of a solution caused by the presence of charged species is often roughly accounted
for using the extended Debye–Hückel equation depending on the ionic strength of the solution:

 ∗  Azi2 I
log γi,m =− √ . 16.
1+B I
A and B represent temperature-dependent constants, and zi is the charge of a reacting agent
i. At 298.15 K, A and B are given as, respectively, 0.5093 L0.5 /mol0.5 and 1.6 L0.5 /mol0.5 (33).
Equation 17 was suggested to describe the influence of ionic strength on the apparent standard
Gibbs energy of a reaction (34):

 I
R g 0,app = R g 0 (I = 0) − 2.303RTA νi zi2 √ . 17.
1+B I
Using this expression in Equation 7 and assuming K γ ∗ is 1, K m is obtained as a function of ionic
strength by (34)

1.17582I 0.5 νi zi2
ln[K m (I )] = ln[K m (I = 0)] + . 18.
1 + BI 0.5
Equation 18 was applied to calculate the influence of NaCl on the K m of methyl ferulate hy-
drolysis (Figure 5). It becomes obvious that Equation 18 strongly overestimates the salt influence
on K m when compared with experimental data, even at fairly low concentrations of NaCl. This
is because the Debye–Hückel theory, and thus Equation 18, does not account for short-range
interactions among charged species, between charged species and neutral molecules (for example,
solvents), or between uncharged reacting agents.

www.annualreviews.org • Thermodynamics of Bioreactions 405


CH07CH17-Sadowski ARI 3 May 2016 12:4

This shortcoming can be overcome by applying a state-of-the-art electrolyte model to calculate


the activity coefficients of the reacting agents. In addition to electrolyte GE models, SAFT-based
equations of state have been combined with electrolyte theories (112) and have shown high accu-
racy in calculating activity coefficients for electrolyte systems. The big advantage of such models
over simple correlations, such as Equation 18, is that they describe both short-range and long-range
interactions among all components. Based on the K a of methyl ferulate hydrolysis, Hoffmann et al.
(47) predicted the NaCl influence on K m using ePC-SAFT (electrolyte PC-SAFT). This is also
illustrated in Figure 5. In contrast to Equation 18, the ePC-SAFT predictions show convinc-
ing results (47). Similarly good results have been obtained for the reaction equilibrium of G6P
isomerization in the presence of glutamate salts (48).
The influence on K m values is not pronounced when electrolytes are added, unless they change
the pH (see Section 6.2). Correlations based on Equation 18 cannot be recommended for cal-
Annu. Rev. Chem. Biomol. Eng. 2016.7:395-414. Downloaded from www.annualreviews.org

culating the influence of electrolytes on K m . Instead, either measurements or more sophisticated


Access provided by University of Nebraska - Lincoln on 06/10/16. For personal use only.

electrolyte models should be used for that purpose.

6.4. Influence of Reacting Agents’ Concentrations


Reacting agents are usually present in bioreaction media in a very diluted state. Under these
conditions, the rational activity coefficients of the reacting agents are roughly 1. Thus, setting K γ ∗
equal to 1 for the calculation of K a , according to Equation 5, might be considered an appropriate
and valid assumption at very low concentrations. Alberty (113) also assumed that R g 0 values are
quite similar to R g 0,app values for these cases.
However, studies from the past five years have shown that K m values might also strongly
depend on the concentrations of the reacting agents, even if the latter are only a few mil-
limole per kilogram of water (47, 48). Unfortunately, bioreaction equilibria are often investigated
with different concentrations of reacting agents. Even worse, the concentrations of the reacting
agents are not even specified, and this occurs even in textbooks and databases (for example, see
References 34, 114).
To investigate the impact of different concentrations of reacting agents on K m values of bioreac-
tions, some researchers analyzed enzyme-catalyzed reactions at different substrate concentrations.
Abbott et al. (115) studied the esterification of glycerol with lauric acid at different starting con-
ν
centrations of lauric acid. The observed mole fraction–based K x (= xi i ) values varied between
1.23 and 24.3 at equilibrium mole fractions of lauric acid between 0.02 and 0.05. Thus, even a
small change in the substrate concentration caused different K x values for this reaction. Hoffmann
et al. (47) have shown that for the hydrolysis of methyl ferulate, an increase in the equilibrium
concentration of methyl ferulate from 8 ·10−4 mmol/kg H2 O to 8 ·10−3 mmol/kg H2 O caused an
almost 50% decrease in K m , from 25.61 to 14.98.
Moreover, Hoffmann et al. (48) studied the influence of G6P molality on K m values of G6P
isomerization. The initial G6P molality was varied between 1 and 50 mmol/kg H2 O at pH 8.5.
Figure 6 shows the measured K m values obtained for varying G6P equilibrium molalities. K m
values increased up to 30% upon increasing the G6P molality from 0.7 to 33 mmol/kg H2 O.
Again, this illustrates that K m is not constant but instead strongly depends on the concentrations
of the reacting agents. Thus, it is highly recommended that molalities be reported together with
the K m values of biological reactions.
Figure 6 further illustrates the dependence on molality of K γ ∗ values, which were calculated
from the ratio of K a (determined from the extrapolation of K m values to 0 molality of all reacting
agents) to K m . At 0 molality of all reacting agents, K γ ∗ becomes exactly 1. However, as G6P
equilibrium molality increases, K γ ∗ decreases to values less than 0.8 at 33 mmol G6P/kg H2 O.

406 Held · Sadowski


CH07CH17-Sadowski ARI 3 May 2016 12:4

1.0

0.8

K(–) 0.6

0.4

0.2
0 5 10 15 20 25 30 35
Annu. Rev. Chem. Biomol. Eng. 2016.7:395-414. Downloaded from www.annualreviews.org

eq
mG6P(mmol/kg)
Access provided by University of Nebraska - Lincoln on 06/10/16. For personal use only.

Figure 6
K values for the isomerization of glucose-6-phosphate (G6P) as a function of G6P equilibrium molality at
pH 8.5 and T = 310.15 K. Experimental K m values (circles) and experimentally derived Kγ ∗ values (triangles)
were taken from Reference 48. K a , as derived from Equation 5, is represented as a black dashed line. The
green solid line represents electrolyte PC-SAFT-predicted K γ ∗ values, also taken from Reference 48.

Hoffmann et al. (48) applied ePC-SAFT to predict the K γ ∗ for that reaction. They achieved very
good agreement between the predicted and experimental K γ ∗ values, deviating by less than 2%
(also shown in Figure 6). This is considered an excellent result because ePC-SAFT was used solely
to predict the activity coefficients, and reaction-equilibrium data were not used to fit ePC-SAFT
model parameters.
In conclusion, it has been shown that K m values may strongly depend on concentration, even
at low concentrations of the reacting agents. This is caused by the non-one ratio of the activity
coefficients of reacting agents. Thus, neglecting the concentration dependence of K m by assuming
K γ ∗ to be 1 might result in incorrect K m values and the misinterpretation of data and, consequently,
in erroneous R g 0 values obtained from Equation 6.

6.5. Influence of Solvents


Biological reactions typically occur in aqueous reaction media, as most enzymes are active only
in water. Nevertheless, organic solvents are increasingly being used as media for bioreactions
because enzymes, especially lipases, have been developed that catalyze reactions that require only
small amounts of water [for example, see a study by Bornscheuer’s group (116)]. Lipases are
particularly worthwhile catalysts for esterification reactions, as the presence of a high amount of
water dramatically decreases product yield due to equilibrium limitations. Thus, esterifications
are usually performed in solvents other than water (e.g., in alcohols) (30, 31, 117).
The addition of a solvent to the reaction medium does not change K a or K th (Equation 5 and
Equation 8) and thus does not influence R g 0 values. However, solvents may have a tremendous
influence on the activity coefficients of reacting agents, thus also influencing their equilibrium
concentrations. Therefore, solvents can be used to manipulate K m values and reaction yields.
Research by Tewari et al. (118, 119), Fermeglia et al. (31), and Valivety et al. (120) has analyzed
the influence of solvents on the reaction equilibria of enzyme-catalyzed reactions to maximize
the yield of enzyme-catalyzed esterifications. Tewari & Bunk (121) studied the lipase-catalyzed
esterification of glycerol with octanoic acid at 303 K and atmospheric pressure. Compared with
the neat reaction, the addition of acetonitrile to the reaction medium was shown to increase the

www.annualreviews.org • Thermodynamics of Bioreactions 407


CH07CH17-Sadowski ARI 3 May 2016 12:4

1,600 n-Hexane
1,000
a n-Heptane
Cyclohexane
b
2,2,4-Trimethylpentane 800
1,200
Toluene

600
Kth(–) 800 Kth(–)
400

400
200
Annu. Rev. Chem. Biomol. Eng. 2016.7:395-414. Downloaded from www.annualreviews.org

0 0
Access provided by University of Nebraska - Lincoln on 06/10/16. For personal use only.

COSMO-RS UNIFAC PC-SAFT COSMO-RS UNIFAC PC-SAFT

Figure 7
K th values predicted for esterification reactions in different solvents using COSMO-RS, UNIFAC, and PC-SAFT. (a) Esterification of
(−)-menthol with lauric acid; (b) esterification of 1-dodecanol with lauric acid. Data for COSMO-RS and UNIFAC were taken from
Fermeglia et al. (31), and PC-SAFT parameters can be provided on request.

mole fraction–based K x by a factor of 32. Using toluene and benzene as solvents also increased
K x values, but to a much smaller extent than acetonitrile did.
Accounting for the activity coefficients of the reacting agents in the different solvents is the
key to predicting K a and, as a result, K m values and reaction yields in these solvents. Activity
coefficients can be obtained from thermodynamic models. For bioreactions, PC-SAFT, UNIFAC,
and COSMO-RS have primarily been applied (31, 47, 74). Hoffmann et al. (47) used PC-SAFT
to predict K γ ∗ values for enzyme-catalyzed hydrolysis. They showed that the application of PC-
SAFT is suitable for predicting the activity coefficients of the reacting agents and for modeling the
solvent’s influence on K m values. In other studies, COSMO-RS and UNIFAC (31, 74) have been
used to model the K th values of enzyme-catalyzed esterifications in different organic solvents. It
has been found that UNIFAC fails to predict the solvent’s influence on the activity coefficients
of the reacting agents, whereas COSMO-RS is considered to be an appropriate model for that
purpose.
To illustrate the suitability of thermodynamic models for predicting a solvent’s influence on
bioreactions, PC-SAFT, UNIFAC, and COSMO-RS were applied to predict the K γ values and,
from those, the K th of two esterification reactions in various organic solvents. Figure 7 compares
the K th values for the esterification of (−)-menthol with lauric acid (Figure 7a) and the esterifica-
tion of 1-dodecanol with lauric acid (Figure 7b). The K th values were obtained by using Equation
8 with experimental K m values from the literature (118, 119). As the equilibrium constant K th
should not depend on the solvent, the quality of the predictions can be judged by comparing the
K th values predicted for reactions in different solvents. It becomes obvious from Figure 7 that the
application of UNIFAC causes very high K th values compared with the predictions made using
COSMO-RS and PC-SAFT. Furthermore, the UNIFAC-predicted K th values in different sol-
vents deviated, on average, by 41% (Figure 7a) and 32% (Figure 7b), whereas substantially lower
deviations were found by using COSMO-RS (Figure 7a, 34%; Figure 7b, 17%) or PC-SAFT
(Figure 7a, 27%; Figure 7b, 9%). In conclusion, K m values strongly depend on the presence of
solvents in the reaction medium. State-of-the-art thermodynamic models allow for prediction of
the influence of the solvent on reaction equilibria and, from that, the prediction of the reaction
yield in various solvents.

408 Held · Sadowski


CH07CH17-Sadowski ARI 3 May 2016 12:4

7. CONCLUSIONS
The application of thermodynamic principles to biological reactions allows for the prediction of
yields and identification of reactions that may cause bottlenecks in metabolic networks. However,
applying these principles requires profound knowledge of thermodynamic properties, in particular
of the standard Gibbs energy of reaction, R g 0 , and the thermodynamic activities of the reacting
agents.
Although R g 0 depends only on temperature, R g 0 data from different literature sources usu-
ally show a huge scatter, even at a constant temperature. This is particularly worrying because these
data are used extensively for analyses of metabolic pathways. To a certain extent, these deviations
can be explained by uncertainties in the experimentally determined concentrations of the reacting
agents, which are usually small and hard to measure. However, this is only part of the problem.
Another reason for the huge scatter probably is that the activity coefficients of the reacting
Annu. Rev. Chem. Biomol. Eng. 2016.7:395-414. Downloaded from www.annualreviews.org
Access provided by University of Nebraska - Lincoln on 06/10/16. For personal use only.

agents are neglected when determining the R g 0 from equilibrium measurements of biological
reactions. The activity coefficients of reacting agents (as well as the K m values) depend on tem-
perature, on the concentrations of the reacting agents, on the presence of additives, and on the
solvent in which a reaction occurs. If this is not accounted for, these dependencies show up in the
R g 0 values.
The influence of temperature on biological reactions is usually quite small, as these reactions are
usually performed over a narrow temperature range. Furthermore, K m values strongly depend on
pH if the dissociation constants of the reactant(s) and product(s) differ significantly. Correlations
that use the dissociation constants of the reacting agents reliably predict the influence of pH on
Km.
In contrast, the influence of electrolytes on K m values is not very pronounced. Correlations for
calculating the influence of ionic strength on K m often overestimate this effect. Moreover, they
usually account only for ionic strength rather than for the nature of the ions actually present in
the reaction medium. Thus, it is recommended that either measurements or more sophisticated
electrolyte models be used to determine the influence of electrolytes on K m .
K m values strongly depend on the reacting agents’ concentrations and on the presence of
organic solvents in the reaction medium, due to the activity coefficients of the reacting agents.
State-of-the-art thermodynamic models now allow for estimates of these activity coefficients and,
from those, provide very good predictions of K m as a function of reacting agents’ concentrations
and solvent media (given reliable R g 0 values).
As for chemical reactions and phase equilibria, the application of thermodynamics to biological
reactions is straightforward. However, biological systems are usually more complex, as they contain
many different species, and the reacting agents often show a species distribution that varies as a
function of pH. This requires careful data analysis using state-of-the-art models to account for
a system’s nonidealities. If applied properly, thermodynamics is a powerful tool for assessing
biological reactions and evaluating the influence of a system’s properties on these reactions.

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.

ACKNOWLEDGMENT
The authors thank Peter Halling for inspiring discussions.

www.annualreviews.org • Thermodynamics of Bioreactions 409


CH07CH17-Sadowski ARI 3 May 2016 12:4

LITERATURE CITED
1. Krebs HA, Kornberg HL, Burton K. 1957. A survey of the energy transformations in living matter.
Ergeb. Physiol. 49:212–98
2. Lipmann F. 1941. Metabolic generation and utilization of phosphate bond energy. In Advances in Enzy-
mology and Related Areas of Molecular Biology, Vol. 1, ed. FF Nord, CH Werkman, pp. 99–162. Hoboken,
NJ: John Wiley & Sons
3. Gajewski E, Steckler DK, Goldberg RN. 1986. Thermodynamics of the hydrolysis of adenosine 5 -
triphosphate to adenosine 5 -diphosphate. J. Biol. Chem. 261:2733–37
4. Goldberg RN. 1975. Thermodynamics of hexokinase-catalyzed reactions. Biophys. Chem. 3:192–205
5. Tewari YB, Goldberg RN. 1984. Thermodynamics of the conversion of aqueous glucose to fructose.
J. Solut. Chem. 13:523–47
6. Tewari YB, Goldberg RN. 1991. Thermodynamics of hydrolysis of disaccharides: lactulose, α-D-
melibiose, palatinose, D-trehalose, D-turanose and 3-o-β-D-galactopyranosyl-D-arabinose. Biophys.
Annu. Rev. Chem. Biomol. Eng. 2016.7:395-414. Downloaded from www.annualreviews.org
Access provided by University of Nebraska - Lincoln on 06/10/16. For personal use only.

Chem. 40:59–67
7. Rekharsky MV, Goldberg RN, Schwarz FP, Tewari YB, Ross PD, et al. 1995. Thermodynamic and
nuclear magnetic resonance study of the interactions of α- and β-cyclodextrin with model substances:
phenethylamine, ephedrines, and related substances. J. Am. Chem. Soc. 117:8830–40
8. Goldberg RN, Kishore N, Lennen RM. 2002. Thermodynamic quantities for the ionization reactions
of buffers. J. Phys. Chem. Ref. Data 31:231–370
9. Alberty RA. 1968. Effect of pH and metal ion concentration on the equilibrium hydrolysis of adenosine
triphosphate to adenosine diphosphate. J. Biol. Chem. 243:1337–43
10. Alberty RA. 1969. Standard Gibbs free energy, enthalpy, and entropy changes as a function of pH and
pMg for several reactions involving adenosine phosphates. J. Biol. Chem. 244:3290–302
11. Alberty RA, Goldberg RN. 1992. Standard thermodynamic formation properties for the adenosine 5 -
triphosphate series. Biochemistry 31:10610–15
12. Alberty RA, Smith RM, Bock RM. 1951. The apparent ionization constants of the adenosinephosphates
and related compounds. J. Biol. Chem. 193:425–34
13. Alberty RA. 2002. Thermodynamics of systems of biochemical reactions. J. Theor. Biol. 215:491–501
14. von Stockar U. 2013. The role of thermodynamics in biochemical engineering. J. Non-Equilib. Thermo-
dyn. 38:225–40
15. Westerhoff HV. 2001. The silicon cell, not dead but live! Metab. Eng. 3:207–10
16. Webb EC, ed. 1992. Enzyme Nomenclature 1992: Recommendations of the Nomenclature Committee of the
International Union of Biochemistry and Molecular Biology on the Nomenclature and Classification of Enzymes.
San Diego, CA: Academic
17. Goldberg RN, Tewari YB, Bell D, Fazio K, Anderson E. 1993. Thermodynamics of enzyme-catalyzed
reactions: part 1. Oxidoreductases. J. Phys. Chem. Ref. Data 22:515–79
18. Goldberg RN, Tewari YB. 1994. Thermodynamics of enzyme-catalyzed reactions: part 2. Transferases.
J. Phys. Chem. Ref. Data 23:547–617
19. Goldberg RN, Tewari YB. 1994. Thermodynamics of enzyme-catalyzed reactions: part 3. Hydrolases.
J. Phys. Chem. Ref. Data 23:1035–103
20. Goldberg RN, Tewari YB. 1995. Thermodynamics of enzyme-catalyzed reactions: part 4. Lyases.
J. Phys. Chem. Ref. Data 24:1669–98
21. Goldberg RN, Tewari YB. 1995. Thermodynamics of enzyme-catalyzed reactions: part 5. Isomerases
and ligases. J. Phys. Chem. Ref. Data 24:1765–801
22. von Stockar U, van der Wielen LAM. 1997. Thermodynamics in biochemical engineering. J. Biotechnol.
59:25–37
23. Heijnen JJ, Van Dijken JP. 1992. In search of a thermodynamic description of biomass yields for the
chemotropic growth of microorganisms. Biotechnol. Bioeng. 39:833–58
24. von Stockar U, Liu JS. 1999. Does microbial life always feed on negative entropy? Thermodynamic
analysis of microbial growth. Biochim. Biophys. Acta 1412:191–211
25. Tijhuis L, Vanloosdrecht MCM, Heijnen JJ. 1993. A thermodynamically based correlation for main-
tenance Gibbs energy requirements in aerobic and anaerobic chemotropic growth. Biotechnol. Bioeng.
42:509–19

410 Held · Sadowski


CH07CH17-Sadowski ARI 3 May 2016 12:4

26. Pissarra PD, Nielsen J. 1997. Thermodynamics of metabolic pathways for penicillin production: analysis
of thermodynamic feasibility and free energy changes during fed-batch cultivation. Biotechnol. Prog.
13:156–65
27. Mavrovouniotis ML. 1993. Identification of localized and distributed bottlenecks in metabolic pathways.
Proc. Int. Conf. Intell. Syst. Mol. Biol. 1:275–83
28. Nielsen J. 1997. Metabolic control analysis of biochemical pathways based on a thermokinetic description
of reaction rates. Biochem. J. 321:133–38
29. Minakami S, Yoshikawa H. 1966. Studies on erythrocyte glycolysis. II. Free energy changes and rate
limiting steps in erythrocyte glycolysis. J. Biochem. 59:139–44
30. Braiuca P, Khaliullin I, Svedas V, Knapic L, Fermeglia M, et al. 2012. BESSICC, a COSMO-RS based
tool for in silico solvent screening of biocatalyzed reactions. Biotechnol. Bioeng. 109:1864–68
31. Fermeglia M, Braiuca P, Gardossi L, Pricl S, Halling PJ. 2006. In silico prediction of medium effects on
esterification equilibrium using the COSMO-RS method. Biotechnol. Prog. 22:1146–52
Annu. Rev. Chem. Biomol. Eng. 2016.7:395-414. Downloaded from www.annualreviews.org
Access provided by University of Nebraska - Lincoln on 06/10/16. For personal use only.

32. Vojinovic V, von Stockar U. 2009. Influence of uncertainties in pH, pMg, activity coefficients, metabolite
concentrations, and other factors on the analysis of the thermodynamic feasibility of metabolic pathways.
Biotechnol. Bioeng. 103:780–95
33. Maskow T, von Stockar U. 2005. How reliable are thermodynamic feasibility statements of biochemical
pathways? Biotechnol. Bioeng. 92:223–30
34. Alberty RA. 2003. Thermodynamics of Biochemical Reactions. Hoboken, NJ: Wiley-Intersci.
35. Minakami S, Yoshikawa H. 1965. Thermodynamic considerations on erythrocyte glycolysis. Biochem.
Biophys. Res. Commun. 18:345–49
36. Kjelstrup S, Rubi JM, Bedeaux D. 2005. Active transport: a kinetic description based on thermodynamic
grounds. J. Theor. Biol. 234:7–12
37. Bergman C, Kashiwaya Y, Veech RL. 2010. The effect of pH and free Mg2+ on ATP linked enzymes
and the calculation of Gibbs free energy of ATP hydrolysis. J. Phys. Chem. B 114:16137–46
38. Robbins EA, Boyer PD. 1957. Determination of the equilibrium of the hexokinase reaction and the free
energy of hydrolysis of adenosine triphosphate. J. Biol. Chem. 224:121–35
39. Alberty RA. 2003. Effect of temperature on the standard transformed thermodynamic properties of
biochemical reactions with emphasis on the Maxwell equations. J. Phys. Chem. B 107:3631–35
40. Panke O, Rumberg B. 1997. Energy and entropy balance of ATP synthesis. Biochim. Biophys. Acta
1322:183–94
41. Phillips RC, George P, Rutman RJ. 1969. Thermodynamic data for hydrolysis of adenosine triphosphate
as a function of pH, Mg2+ ion concentration, and ionic strength. J. Biol. Chem. 244:3330–42
42. Rosing J, Slater EC. 1972. Value of G degrees for the hydrolysis of ATP. Biochim. Biophys. Acta 267:275–
90
43. Guynn RW, Veech RL. 1973. Equilibrium constants of the adenosine triphosphate hydrolysis and
adenosine triphosphate-citrate lyase reactions. J. Biol. Chem. 248:6966–72
44. Ould-Moulaye CB, Dussap CG, Gros JB. 2002. A consistent set of formation properties of nucleic acid
compounds: nucleosides, nucleotides and nucleotide-phosphates in aqueous solution. Thermochim. Acta
387:1–15
45. Alberty RA. 2001. Effect of temperature on standard transformed Gibbs energies of formation of reactants
at specified pH and ionic strength and apparent equilibrium constants of biochemical reactions. J. Phys.
Chem. B 105:7865–70
46. Benzinger TH, Huebscher RG, Minard D, Kitzinger C. 1958. Human calorimetry by means of the
gradient principle. J. Appl. Physiol. 12:S1–24
47. Hoffmann P, Voges M, Held C, Sadowski G. 2013. The role of activity coefficients in bioreaction
equilibria: thermodynamics of methyl ferulate hydrolysis. Biophys. Chem. 173:21–30
48. Hoffmann P, Held C, Maskow T, Sadowski G. 2014. A thermodynamic investigation of the glucose-6-
phosphate isomerization. Biophys. Chem. 195:22–31
49. Goldberg RN, Lang BE, Selig MJ, Decker SR. 2011. A calorimetric and equilibrium investigation of
the reaction {methyl ferulate(aq) + H2 O(1) = methanol(aq) + ferulic acid(aq)}. J. Chem. Thermodyn.
43:235–39

www.annualreviews.org • Thermodynamics of Bioreactions 411


CH07CH17-Sadowski ARI 3 May 2016 12:4

50. Verevkin SP, Zaitsau DH, Emel’yanenko VN, Zhabina AA. 2015. Thermodynamic properties of glycerol:
experimental and theoretical study. Fluid Phase Equilib. 397:87–94
51. Verevkin SP, Emel’yanenko VN, Garist IV. 2015. Benchmark thermodynamic properties of alkanedi-
amines: experimental and theoretical study. J. Chem. Thermodyn. 87:34–42
52. Prausnitz JM, Lichtenthaler RN, Gomes de Azevedo E. 1999. Molecular Thermodynamics of Fluid-Phase
Equilibria. Upper Saddle River, NJ: Prentice Hall. 3rd ed.
53. Ponomarev VV, Migarskaya LB. 1960. Heats of combustion of some amino acids. Russ. J. Phys. Chem.
34:1182–83 (Engl. Transl.)
54. Oja V, Suuberg EM. 1999. Vapor pressures and enthalpies of sublimation of D-glucose, D-xylose, cel-
lobiose, and levoglucosan. J. Chem. Eng. Data 44:26–29
55. Boerio-Goates J. 1991. Heat-capacity measurements and thermodynamic functions of crystalline α-D-
glucose at temperatures from 10 K to 340 K. J. Chem. Thermodyn. 23:403–9
56. Clarke TH, Stegeman G. 1939. Heats of combustion of some mono- and disaccharides. J. Am. Chem.
Annu. Rev. Chem. Biomol. Eng. 2016.7:395-414. Downloaded from www.annualreviews.org

Soc. 61:1726–30
Access provided by University of Nebraska - Lincoln on 06/10/16. For personal use only.

57. Meyerhof O, Lohmann K. 1925. Synthesis of carbohydrate in muscle. Biochem. Z. 157:459–91


58. Kobel M. 1928. Heat of combustion and of solution of dihydroxyacetone. Biochem. Z. 203:159–63
59. Baer E, Flehmig HH. 1969. Refutation of alleged differences in the energy contents of optical isomers.
Can. J. Biochem. 47:79–83
60. Jinich A, Rappoport D, Dunn I, Sanchez-Lengeling B, Olivares-Amaya R, et al. 2014. Quantum chemical
approach to estimating the thermodynamics of metabolic reactions. Sci. Rep. 4:7022
61. Chen M, Lee JK. 2014. Computational studies of the gas-phase thermochemical properties of modified
nucleobases. J. Org. Chem. 79:11295–300
62. Villadsen J, Nielsen J, Lidén G, eds. 2011. Thermodynamics of Bioreactions. Boston: Springer
63. Wagman DD, Evans WH, Parker VB, Schumm RH, Halow I, et al. 1982. The NBS tables of chemical
thermodynamic properties: selected values for inorganic and C1 and C2 organic substances in SI units.
J. Phys. Chem. Ref. Data 11(Suppl. 2):21–392
64. Mavrovouniotis ML. 1990. Group contributions for estimating standard Gibbs energies of formation of
biochemical compounds in aqueous solution. Biotechnol. Bioeng. 36:1070–82
65. Mavrovouniotis ML. 1991. Estimation of standard Gibbs energy changes of biotransformations. J. Biol.
Chem. 266:14440–45
66. Gross J, Sadowski G. 2002. Application of the perturbed-chain SAFT equation of state to associating
systems. Ind. Eng. Chem. Res. 41:5510–15
67. Gross J, Sadowski G. 2001. Perturbed-chain SAFT: an equation of state based on a perturbation theory
for chain molecules. Ind. Eng. Chem. Res. 40:1244–60
68. Fredenslund A, Gmehling J, Michelsen ML, Rasmussen P, Prausnitz JM. 1977. Computerized design
of multicomponent distillation columns using UNIFAC group contribution method for calculation of
activity coefficients. Ind. Eng. Chem. Process Design Dev. 16:450–62
69. Fredenslund A, Jones RL, Prausnitz JM. 1975. Group-contribution estimation of activity-coefficients in
nonideal liquid-mixtures. AIChE J. 21:1086–99
70. Klamt A, Schuurmann G. 1993. Cosmo: a new approach to dielectric screening in solvents with explicit
expressions for the screening energy and its gradient. J. Chem. Soc. Perkin Trans. I 2:799–805
71. Klamt A. 1995. Conductor-like screening model for real solvents: a new approach to the quantitative
calculation of solvation phenomena. J. Phys. Chem. 99:2224–35
72. Renon H, Prausnitz JM. 1968. Local compositions in thermodynamic excess functions for liquid mixtures.
AIChE J. 14:135–44
73. Zeuner B, Kontogeorgis GM, Riisager A, Meyer AS. 2012. Thermodynamically based solvent design for
enzymatic saccharide acylation with hydroxycinnamic acids in non-conventional media. New Biotechnol.
29:255–70
74. Strompen S, Weiss M, Ingram T, Smirnova I, Groger H, et al. 2012. Kinetic investigation of a solvent-
free, chemoenzymatic reaction sequence towards enantioselective synthesis of a β-amino acid ester.
Biotechnol. Bioeng. 109:1479–89
75. Dreyer S, Kragl U. 2008. Ionic liquids for aqueous two-phase extraction and stabilization of enzymes.
Biotechnol. Bioeng. 99:1416–24

412 Held · Sadowski


CH07CH17-Sadowski ARI 3 May 2016 12:4

76. Luong TQ, Winter R. 2015. Combined pressure and cosolvent effects on enzyme activity. Phys. Chem.
Chem. Phys. 17:23273–78
77. Fuchs D, Fischer J, Tumakaka F, Sadowski G. 2006. Solubility of amino acids: influence of the pH value
and the addition of alcoholic cosolvents on aqueous solubility. Ind. Eng. Chem. Res. 45:6578–84
78. Cabral JMS, Aires-Barros MR, Pinheiro H, Prazeres DMF. 1997. Biotransformation in organic media
by enzymes and whole cells. J. Biotechnol. 59:133–43
79. Carrea G, Riva S. 2000. Properties and synthetic applications of enzymes in organic solvents. Angew.
Chem. Int. Ed. 39:2226–54
80. Kragl U, Eckstein M, Kaftzik N. 2002. Enzyme catalysis in ionic liquids. Curr. Opin. Biotechnol. 13:565–71
81. Schöfer SH, Kaftzik N, Wasserscheid P, Kragl U. 2001. Enzyme catalysis in ionic liquids: lipase catalysed
kinetic resolution of 1-phenylethanol with improved enantioselectivity. Chem. Commun. 5:425–26
82. Yang Z, Pan W. 2005. Ionic liquids: green solvents for nonaqueous biocatalysis. Enzym. Microb. Technol.
37:19–28
Annu. Rev. Chem. Biomol. Eng. 2016.7:395-414. Downloaded from www.annualreviews.org
Access provided by University of Nebraska - Lincoln on 06/10/16. For personal use only.

83. Swatloski RP, Spear SK, Holbrey JD, Rogers RD. 2002. Dissolution of cellose with ionic liquids. J. Am.
Chem. Soc. 124:4974–75
84. Emel’yanenko VN, Zaitsau DH, Shoifet E, Meurer F, Verevkin SP, et al. 2015. Benchmark thermo-
chemistry for biologically relevant adenine and cytosine: a combined experimental and theoretical study.
J. Phys. Chem. A 119:9680–91
85. Marrero J, Gani R. 2001. Group-contribution based estimation of pure component properties. Fluid
Phase Equilib. 183:183–208
86. Khoshkbarchi MK, Vera JH. 1996. Measurement of activity coefficients of amino acids in aqueous
electrolyte solutions: experimental data for the systems H2 O + NaCl + glycine and H2 O + NaCl +
DL-alanine at 25◦ C. Ind. Eng. Chem. Res. 35:2735–42
87. Held C, Cameretti LF, Sadowski G. 2011. Measuring and modeling activity coefficients in aqueous
amino-acid solutions. Ind. Eng. Chem. Res. 50:131–41
88. Held C, Neuhaus T, Sadowski G. 2010. Compatible solutes: thermodynamic properties and biological
impact of ectoines and prolines. Biophys. Chem. 152:28–39
89. Held C, Sadowski G. 2016. Compatible solutes: thermodynamic properties relevant for effective pro-
tection against osmotic stress. Fluid Phase Equilib. 407:224–35
90. Held C, Carneiro A, Macedo EA, Sadowski G. 2013. Modeling thermodynamic properties of aqueous
single-solute and multi-solute sugar solutions with PC-SAFT. AIChE J. 59:4794–805
91. Ferreira LA, Breil MP, Pinho SP, Macedo EA, Mollerup JM. 2009. Thermodynamic modeling of several
aqueous alkanol solutions containing amino acids with the perturbed-chain statistical associated fluid
theory equation of state. Ind. Eng. Chem. Res. 48:5498–505
92. Ruether F, Sadowski G. 2009. Modeling the solubility of pharmaceuticals in pure solvents and solvent
mixtures for drug process design. J. Pharm. Sci. 98:4205–15
93. Grosse Daldrup JB, Held C, Ruether F, Schembecker G, Sadowski G. 2009. Measurement and modeling
solubility of aqueous multisolute amino-acid solutions. Ind. Eng. Chem. Res. 49:1395–401
94. Prikhod’ko IV, Tumakaka F, Sadowski G. 2007. Application of the PC-SAFT equation of state to
modeling of solid–liquid equilibria in systems with organic components forming chemical compounds.
Russ. J. Appl. Chem. 80:542–48
95. Macedo EA, Peres AM. 2001. Thermodynamics of ternary mixtures containing sugars. SLE of D-fructose
in pure and mixed solvents. Comparison between modified UNIQUAC and modified UNIFAC. Ind.
Eng. Chem. Res. 40:4633–40
96. Held C, Reschke T, Müller R, Kunz W, Sadowski G. 2014. Measuring and modeling aqueous
electrolyte/amino-acid solutions with ePC-SAFT. J. Chem. Thermodyn. 68:1–12
97. Grosse Daldrup JB, Held C, Sadowski G, Schembecker G. 2011. Modeling pH and solubilities in aqueous
multisolute amino-acid solutions. Ind. Eng. Chem. Res. 50:3503–9
98. Emel’yanenko VN, Strutynska A, Verevkin SP. 2005. Enthalpies of formation and strain of chlorobenzoic
acids from thermochemical measurements and from ab initio calculations. J. Phys. Chem. A 109:4375–80
99. Gobble C, Chickos J, Verevkin SP. 2014. Vapor pressures and vaporization enthalpies of a series of
dialkyl phthalates by correlation gas chromatography. J. Chem. Eng. Data 59:1353–65

www.annualreviews.org • Thermodynamics of Bioreactions 413


CH07CH17-Sadowski ARI 3 May 2016 12:4

100. Verevkin SP, Sazonova AY, Emel’yanenko VN, Zaitsau DH, Varfolomeev MA, et al. 2015. Thermo-
chemistry of halogen-substituted methylbenzenes. J. Chem. Eng. Data 60:89–103
101. Panteli E, Voutsas E. 2010. Solubilities of cinnamic acid esters in binary mixtures of ionic liquids and
organic solvents. Fluid Phase Equilibr. 295:208–14
102. Ji XY, Held C, Sadowski G. 2012. Modeling imidazolium-based ionic liquids with ePC-SAFT. Fluid
Phase Equilibr. 335:64–73
103. Curras MR, Vijande J, Pineiro MM, Lugo L, Salgado J, Garcia J. 2011. Behavior of the environmen-
tally compatible absorbent 1-butyl-3-methylimidazolium tetrafluoroborate with 2,2,2-trifluoroethanol:
experimental densities at high pressures and modeling of PVT and phase equilibria behavior with PC-
SAFT EoS. Ind. Eng. Chem. Res. 50:4065–76
104. Beezer AE. 1980. Biological Microcalorimetry. London: Academic
105. Baier V, Fodisch R, Ihring A, Kessler E, Lerchner J, et al. 2005. Highly sensitive thermopile heat power
sensor for micro-fluid calorimetry of biochemical processes. Sens. Actuators A 123–24:354–59
Annu. Rev. Chem. Biomol. Eng. 2016.7:395-414. Downloaded from www.annualreviews.org

106. Lerchner J, Wolf A, Wolf G, Baier V, Kessler E, et al. 2006. A new micro-fluid chip calorimeter for
Access provided by University of Nebraska - Lincoln on 06/10/16. For personal use only.

biochemical applications. Thermochim. Acta 445:144–50


107. Lerchner J, Maskow T, Wolf G. 2008. Chip calorimetry and its use for biochemical and cell biological
investigations. Chem. Eng. Process. 47:991–99
108. Johannessen EA, Weaver JMR, Bourova L, Svoboda P, Cobbold PH, Cooper JM. 2002. Micromachined
nanocalorimetric sensor for ultra-low-volume cell-based assays. Anal. Chem. 74:2190–97
109. Sober HA. 1970. Handbook of Biochemistry: Selected Data for Molecular Biology. Cleveland, OH: Chem.
Rubber Co.
110. Reschke T, Naeem S, Sadowski G. 2012. Osmotic coefficients of aqueous weak electrolyte solutions:
influence of dissociation on data reduction and modeling. J. Phys. Chem. B 116:7479–91
111. Alberty RA, Cornish-Bowden A, Goldberg RN, Hammes GG, Tipton K, Westerhoff HV. 2011. Recom-
mendations for terminology and databases for biochemical thermodynamics. Biophys. Chem. 155:89–103
112. Tan SP, Adidharma H, Radosz M. 2008. Recent advances and applications of statistical associating fluid
theory. Ind. Eng. Chem. Res. 47:8063–82
113. Alberty RA. 1992. Equilibrium calculations on systems of biochemical reactions at specified pH and
pMg. Biophys. Chem. 42:117–31
114. Goldberg RN, Tewari YB, Bhat TN. 2004. Thermodynamics of enzyme-catalyzed reactions—a database
for quantitative biochemistry. Bioinformatics 20:2874–77
115. Abbott AP, Harris RC, Ryder KS, D’Agostino C, Gladden LF, Mantle MD. 2011. Glycerol eutectics as
sustainable solvent systems. Green Chem. 13:82–90
116. Ganske F, Bornscheuer UT. 2005. Optimization of lipase-catalyzed glucose fatty acid ester synthesis in
a two-phase system containing ionic liquids and t-BuOH. J. Mol. Catal. B 36:40–42
117. Riechert O, Husham M, Sadowski G, Zeiner T. 2015. Solvent effects on esterification equilibria. AIChE
J. 61:3000–11
118. Tewari YB. 1998. Thermodynamics of the lipase-catalyzed esterification of 1-dodecanoic acid and 1-
dodecanol in organic solvents. J. Chem. Eng. Data 43:750–55
119. Tewari YB, Schantz MM, Vanderah DJ. 1999. Thermodynamics of the lipase-catalyzed esterification of
1-dodecanoic acid with (−)-menthol in organic solvents. J. Chem. Eng. Data 44:641–47
120. Valivety RH, Johnston GA, Suckling CJ, Halling PJ. 1991. Solvent effects on biocatalysis in organic-
systems: equilibrium position and rates of lipase catalyzed esterification. Biotechnol. Bioeng. 38:1137–43
121. Tewari YB, Bunk DM. 2001. Thermodynamics of the lipase-catalyzed esterification of glycerol and
n-octanoic acid in organic solvents and in the neat reaction mixture. J. Mol. Catal. B 15:135–45

414 Held · Sadowski


CH07-FrontMatter ARI 14 May 2016 8:27

Annual Review of
Chemical and
Biomolecular
Engineering

Contents Volume 7, 2016


Annu. Rev. Chem. Biomol. Eng. 2016.7:395-414. Downloaded from www.annualreviews.org
Access provided by University of Nebraska - Lincoln on 06/10/16. For personal use only.

Neutron Scattering from Polymers: Five Decades of Developing


Possibilities
J.S. Higgins p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 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
Membrane Desalination: Where Are We, and What Can We Learn
from Fundamentals?
Joseph Imbrogno and Georges Belfort p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p29
Multiscale Materials Modeling in an Industrial Environment
Horst Weiß, Peter Deglmann, Pieter J. in ‘t Veld, Murat Cetinkaya,
and Eduard Schreiner p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p65
The Modern Temperature-Accelerated Dynamics Approach
Richard J. Zamora, Blas P. Uberuaga, Danny Perez, and Arthur F. Voter p p p p p p p p p p p p p p87
Charged Polymer Membranes for Environmental/Energy Applications
Jovan Kamcev and Benny D. Freeman p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 111
The Evolution of Process Safety: Current Status and Future Direction
M. Sam Mannan, Olga Reyes-Valdes, Prerna Jain, Nafiz Tamim,
and Monir Ahammad p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 135
Design of Responsive and Active (Soft) Materials Using Liquid Crystals
Emre Bukusoglu, Marco Bedolla Pantoja, Peter C. Mushenheim,
Xiaoguang Wang, and Nicholas L. Abbott p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 163
Thiol-Disulfide Exchange Reactions in the Mammalian Extracellular
Environment
Michael C. Yi and Chaitan Khosla p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 197
A Selection of Recent Advances in C1 Chemistry
Carl Mesters p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 223
The Energy-Water-Food Nexus
D.L. Keairns, R.C. Darton, and A. Irabien p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 239

vii
CH07-FrontMatter ARI 14 May 2016 8:27

Status of Solid State Lighting Product Development and Future Trends


for General Illumination
Thomas M. Katona, P. Morgan Pattison, and Steve Paolini p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 263
Switchable Materials for Smart Windows
Yang Wang, Evan L. Runnerstrom, and Delia J. Milliron p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 283
Nanoparticle-Based Modulation of the Immune System
Ronnie H. Fang and Liangfang Zhang p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 305
Shape-Controlled Metal Nanocrystals for Heterogeneous Catalysis
Aleksey Ruditskiy, Hsin-Chieh Peng, and Younan Xia p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 327
Annu. Rev. Chem. Biomol. Eng. 2016.7:395-414. Downloaded from www.annualreviews.org

Computer Simulations of Ion Transport in Polymer Electrolyte


Access provided by University of Nebraska - Lincoln on 06/10/16. For personal use only.

Membranes
Santosh Mogurampelly, Oleg Borodin, and Venkat Ganesan p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 349
Polymer Thin Films and Surface Modification by Chemical Vapor
Deposition: Recent Progress
Nan Chen, Do Han Kim, Peter Kovacik, Hossein Sojoudi, Minghui Wang,
and Karen K. Gleason p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 373
Thermodynamics of Bioreactions
Christoph Held and Gabriele Sadowski p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 395
Complex Fluids and Hydraulic Fracturing
Alexander C. Barbati, Jean Desroches, Agathe Robisson, and Gareth H. McKinley p p p p 415
Biomanufacturing of Therapeutic Cells: State of the Art, Current
Challenges, and Future Perspectives
Kyung-Ho Roh, Robert M. Nerem, and Krishnendu Roy p p p p p p p p p 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
Polymer Fluid Dynamics: Continuum and Molecular Approaches
R.B. Bird and A.J. Giacomin p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 479
Progress in the Development of Oxygen Reduction Reaction Catalysts for
Low-Temperature Fuel Cells
Dongguo Li, Haifeng Lv, Yijin Kang, Nenad M. Markovic,
and Vojislav R. Stamenkovic p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 509
The Carbon-Water Interface: Modeling Challenges and Opportunities
for the Water-Energy Nexus
Alberto Striolo, Angelos Michaelides, and Laurent Joly p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 533
New Vistas in Chemical Product and Process Design
Lei Zhang, Deenesh K. Babi, and Rafiqul Gani p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 557
Advances in Nanoimprint Lithography
Matthew C. Traub, Whitney Longsine, and Van N. Truskett p p p p p p p p p p p p p p p p p p p p p p p p p p p 583

viii Contents
CH07-FrontMatter ARI 14 May 2016 8:27

Theoretical Heterogeneous Catalysis: Scaling Relationships


and Computational Catalyst Design
Jeffrey Greeley p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 605
Engineering Delivery Vehicles for Genome Editing
Christopher E. Nelson and Charles A. Gersbach p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 637
Lewis Acid Zeolites for Biomass Conversion: Perspectives and Challenges
on Reactivity, Synthesis, and Stability
Helen Y. Luo, Jennifer D. Lewis, and Yuriy Román-Leshkov p p p p p p p p p p p p p p p p p p p p p p p p p p p p 663

Indexes
Annu. Rev. Chem. Biomol. Eng. 2016.7:395-414. Downloaded from www.annualreviews.org
Access provided by University of Nebraska - Lincoln on 06/10/16. For personal use only.

Cumulative Index of Contributing Authors, Volumes 3–7 p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 693


Cumulative Index of Article Titles, Volumes 3–7 p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 696

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

Contents ix

You might also like