university of zurich - uzh · medicine and biochemistry to physical chemistry and material...

40
University of Zurich Zurich Open Repository and Archive Winterthurerstr. 190 CH-8057 Zurich http://www.zora.uzh.ch Year: 2008 A survey of the year 2006 literature on applications of isothermal titration calorimetry Okhrimenko, O; Jelesarov, I Okhrimenko, O; Jelesarov, I (2008). A survey of the year 2006 literature on applications of isothermal titration calorimetry. Journal of Molecular Recognition, 21(1):1-19. Postprint available at: http://www.zora.uzh.ch Posted at the Zurich Open Repository and Archive, University of Zurich. http://www.zora.uzh.ch Originally published at: Journal of Molecular Recognition 2008, 21(1):1-19.

Upload: others

Post on 01-Jun-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

University of ZurichZurich Open Repository and Archive

Winterthurerstr. 190

CH-8057 Zurich

http://www.zora.uzh.ch

Year: 2008

A survey of the year 2006 literature on applications of isothermaltitration calorimetry

Okhrimenko, O; Jelesarov, I

Okhrimenko, O; Jelesarov, I (2008). A survey of the year 2006 literature on applications of isothermal titrationcalorimetry. Journal of Molecular Recognition, 21(1):1-19.Postprint available at:http://www.zora.uzh.ch

Posted at the Zurich Open Repository and Archive, University of Zurich.http://www.zora.uzh.ch

Originally published at:Journal of Molecular Recognition 2008, 21(1):1-19.

Okhrimenko, O; Jelesarov, I (2008). A survey of the year 2006 literature on applications of isothermal titrationcalorimetry. Journal of Molecular Recognition, 21(1):1-19.Postprint available at:http://www.zora.uzh.ch

Posted at the Zurich Open Repository and Archive, University of Zurich.http://www.zora.uzh.ch

Originally published at:Journal of Molecular Recognition 2008, 21(1):1-19.

Page 2: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

A survey of the year 2006 literature on applications of isothermaltitration calorimetry

Abstract

Isothermal titration calorimetry (ITC) is a fast and robust method to determine the energetics ofassociation reactions in solution. The changes in enthalpy, entropy and heat capacity that accompanybinding provide unique insights into the balance of forces driving association of molecular entities. ITCis used nowadays on a day-to-day basis in hundreds of laboratories. The method aids projects both inbasic and practice-oriented research ranging from medicine and biochemistry to physical chemistry andmaterial sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding. In thisreview, we discuss selected results and ideas that have accumulated in the course of the year 2006, thefocus being on biologically relevant systems. Theoretical developments, novel applications and studiesthat provide a deeper level of understanding of the energetic principles of biological function areprimarily considered. Following the appearance of a new generation of titration calorimeters, recentpapers provide instructive examples of the synergy between energetic and structural approaches inbiomedical and biotechnological research.

Page 3: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

A survey of the year 2006 literature on applications of isothermal titration

calorimetry

Oksana Okhrimenko and Ilian Jelesarov*

Biochemisches Institut der Universität Zürich, Winterthurerstrasse 190, CH-8057 Zürich,

Switzerland

* Correspondence to: Ilian Jelesarov, Biochemisches Institut der Universität Zürich,

Winterthurerstrasse 190, CH-8057 Zürich, Switzerland. E-mail: [email protected]

Page 4: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

SUMMARY

Isothermal titration calorimetry (ITC) is a fast and robust method to determine the

energetics of association reactions in solution. The changes in enthalpy, entropy and heat

capacity that accompany binding provide unique insights into the balance of forces driving

association of molecular entities. ITC is used nowadays on a day-to-day basis in hundreds of

laboratories. The method aids projects both in basic and practice-oriented research ranging from

medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the

range of studies utilizing ITC data is steadily expanding. In this review we discuss selected

results and ideas that have accumulated in the course of the year 2006, the focus being on

biologically relevant systems. Theoretical developments, novel applications and studies that

provide a deeper level of understanding of the energetic principles of biological function are

primarily considered. Following the appearance of a new generation of titration calorimeters,

recent papers provide instructive examples of the synergy between energetic and structural

approaches in biomedical and biotechnological research.

Keywords: thermodynamics, calorimetry, molecular recognition, ligand binding, enthalpy,

entropy, heat capacity

Page 5: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

INTRODUCTION

Recent achievements in structural biology and bioinformatics have deepened our

understanding of the molecular principles of biological function. Biological systems evolve and

are maintained in accord with the laws of chemistry and physics. Indeed, at the molecular level,

highly complex biological processes can be regarded as variations of non-covalent and covalent

interactions. Therefore, descriptive and qualitative information has to be supplemented by

quantitative approaches, both experimental and theoretical, in order to understand why biological

macromolecules function the way they do. Thermodynamics reveals the essential energetic

principles of biological phenomena. It provides a general formalism for describing a system in

terms of its energy. Thanks to the development of commercially available precision instruments,

we now witness a real revival of calorimetry in biological sciences.

Isothermal titration calorimetry (ITC) directly quantifies the heat effects accompanying

association between molecular entities. The principle is compellingly simple. Molecule M1 is

added in small aliquots to the calorimetric cell, which contains the binding partner, M2, and the

heat change associated with each addition is measured. The concentrations are selected so that

the theoretical saturation of the available binding sites varies from < 0.1 to > 2-3 (or more)

during the titration experiment. Equilibrium distribution between free M1, free M2 and their

complex M1:M2 is attained after each addition. The total heat change accumulated after each

addition is proportional to the amount of M1:M2, and, is therefore proportional to the degree of

saturation Y = [M1:M2]/[M 2]tot, very much like the change of a spectroscopic signal at partial

saturation is proportional to the fractional population of molecules in the bound state relative to

the fully (i. e. 100 %) associated state of the system. In contrast to spectroscopic signals,

however, at constant pressure and temperature the accumulated heat is also proportional to the

molar enthalpy of association, ∆HA. (Index “A” stands for “association”.) It follows that

deconvolution of the binding isotherm obtained by ITC allows the simultaneous determination of

the binding constant, ( )[ ]1DA MY1

YK1

K−

== , and of ∆HA. Hence, in a single experiment all

relevant thermodynamic parameters describing the association reaction can be determined, since

the free energy change, ∆GA, is related to KA by ∆GA = −RTlnKA and ∆SA = (∆HA − ∆GA)/T. If

the experiment is repeated at several temperatures, the heat capacity change, ∆Cp, can also be

calculated from ∆Cp = dHA/dT. However, the measured heat changes do not only describe the

binding event per se in terms of the energy accumulated in bonds between M1 and M2. While we

are interested to characterize the energetics of a given reaction (formation of a complex) the

heats measured in the ITC experiment also contain contributions from accompanying processes:

Page 6: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

changes in hydration, changes in the number of ions bound to the free components and the

complex, proton uptake/release, conformational transitions, etc, all of which are manifested by

heat changes. The raw data have to be corrected for “unspecific heats” by performing blank

titrations of ligand into buffer and buffer into the macromolecule. Binding-induced changes of

protonation states are relatively frequent, yet cannot be inferred from structural data. pKa

changes are easily detected by ITC if ∆HA is measured as a function of the ionization enthalpy of

the buffer (∆Hbuf) and the data are plotted according to ∆HA = ∆Hbind + nH+∆Hbuf. The slope

yields the number of proton(s), nH+, being taken up or released, and the intercept, ∆Hbind, is the

protonation-independent binding enthalpy [1]. Still, it is important to keep in mind that ∆GA,

∆HA and ∆SA are apparent parameters, and any discussion of their magnitude and any

interpretations invoking details on the underlying physical process should be made with caution.

There are many in-depth reviews covering the theoretical foundation of ITC, the practical

aspects of the method, and the information content of the determined binding parameters [1-7].

Heat release or uptake is a universal property of chemical processes, so that the

experiment does not require introduction of artificial reporter groups or immobilization, and can

be performed under natural conditions. These simple features of the technique explain the

growing interest of the biologically oriented scientific community in using ITC for analyzing the

behaviour and function of biological macromolecules under conditions as close as possible to

their natural milieu. Less than 20 years after the appearance of a new generation of

microcalorimeters, ITC is now used on a day-to-day basis in hundreds of laboratories. The early

excitement of following a binding reaction by monitoring heat changes has been replaced by the

wish of many investigators to explore the method to its limits. ITC is a powerful tool for solving

problems in basic and practice-orientedapplied research. Not surprisingly, the number of

publications reporting ITC data is steadily growing, more complex systems are being explored,

and new applications are constantly being devised.

We have identified 400 papers published in the year 2006 that describe data from ITC

experiments. This number is higher than in the surveys of the years 2002-2005 [8-11] but this

does not necessarily mean that more research projects currently utilize ITC since it might only

reflect the use of new search strategies and increased maintenance of bibliographic databases.

We have searched the literature for “titration calorimetry” and “ITC” appearing as keywords, or

in the text. Records including the abbreviation ITC in a different context have been removed.

Also, papers reviewing or referring to previously published ITC studies have not been

considered, with a few exceptions which provide a deeper level of analysis or improved

understanding of data from a more general perspective [12-14]. However robust and meticulous

Page 7: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

the search strategy, and however powerful the search engines, it is impossible to use a single

satisfactory criterion for the inclusion of papers. Our hope is that we have not missed too many

papers and we would like to apologize to colleagues who may be disappointed not to see their

work included in the references.

For the sake of clarity, and in the tradition of previous surveys, the reference list is

divided into subsections:

(i) General subjects and articles cited in the text [1-56].

(ii) Protein-protein. Association between folded proteins is considered [15, 19, 21, 23,

24, 27-33, 35-38, 57-129].

(iii) Protein-peptide. Papers reporting data on binding of unstructured peptides to proteins

are grouped in this section [39, 40, 130-159].

(iv) Protein/peptide-small ligand. The term “small ligands” lumps together various small-

molecular weight ligands, which form a heterogeneous group: nucleotides, sugars

(mono-or oligosaccharides), co-factors, fatty acids, detergents, drugs, etc [12, 16, 34,

41, 42, 51. 160-269].

(v) Protein/peptide-metal[270-300].

(vi) Protein/peptide-nucleic acid [13, 26, 43, 55, 56, 301-316].

(vii) Protein/peptide-polymer [317-322].

(viii) Protein/peptide-lipid [323-334]

(ix) Nucleic acid-small ligands (drugs)[14, 25, 44, 335-353]

(x) Enzyme kinetics [48, 49, 354, 355].

(xi) Miscellaneous. The subjects of the papers in this section are very heterogeneous:

guest-host studies, chelation, complexation between inorganic molecules,

coordination chemistry, micellization, phase transitions, applications in material

sciences, pharmaceutics, etc [45, 356-420].

In the limited space of an annual review, it is impossible to discuss all the studies that have

used ITC data to gain new insights into a specific problem. In the following we will discuss

studies, that opened new paths for applying ITC to complex problems and which illustrate how

ITC can be used to relate energetic changes to the mechanism of underlying biological processes.

Chemists and physicists who deal with inorganic molecules will find relevant papers in the

“Miscellaneous” section. Although this review is focused on biologically oriented topics, many

of these “miscellaneous” studies may provide new ideas on how to apply ITC to the biological

realm.

Page 8: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

NEW THEORETICAL DEVELOPMENTS AND NON-STANDARD APPLICATIONS

Analysis of cooperative ligand binding by ITC

Allosteric communication between binding sites is a widespread phenomenon in proteins

and underlies important biological phenomena such as signal transduction, regulation of catalytic

activities, formation of multifunctional assemblages, etc. Binding of one ligand can increase

(positive cooperativity) or decrease (negative cooperativity) the affinity for another ligand.

Velazques-Campoy et al. have presented an exact analysis of heterotropic binding for arbitrary

values of the cooperativity parameter α [15]. For binding of ligands A and B to protein P, the

change in free energy related to the cooperative interaction can be defined as ∆g = −RTlnα =

∆GAB − ∆GA − ∆GB, where ∆GAB, ∆GA and ∆GB are the Gibbs free energy change for formation

of the ternary complex PAB, and the bimolecular complexes PA and PB, respectively. The

cooperative parameter α must be understood as a genuine equilibrium constant which can have

values α = 0 (full negative cooperativity), 0 < α < 1 (negative cooperativity), α = 1 (no

cooperativity) or α > 1 (positive cooperativity). In addition to ∆G, the enthalpy and entropy can

be obtained from an ITC experiment. Therefore, the enthalpic (∆h = ∆HAB − ∆HA − ∆HB) and

entropic (∆s = ∆SAB − ∆SA − ∆SB) contributions to ∆g can be measured. Since ∆h and ∆s are

manifestations of the molecular origins of allostery, the potential of ITC in characterization of

heterotropic interactions has been recognized early and simplified formal approaches to data

analysis have been developed. Typically, binding of ligand A to the macromolecule in the

presence of ligand B in the cell is measured and the data are analyzed by the standard procedure

for single ligand binding to obtain apparent values for the equilibrium constant, Kapp, and for the

enthalpy, ∆Happ, characterizing binding of A in the presence of B. From a series of experiments

with varying concentration of B, the interaction parameters α and ∆h can be estimated by

analyzing the dependence of Kapp and ∆Happ on the concentration of B. Alternatively, α and ∆h

can be obtained by comparing Kapp and ∆Happ measured in the absence of B and in the presence

of saturating concentrations of B. The disadvantages of these approaches are that the free

concentration of B throughout the titration, is not precisely known, many experiments must be

performed and it is difficult to vary the concentration in B over a wide range, especially for

reaching saturation levels when binding of B is weak. Furthermore, depending on the ratio of the

binding constants, the molar enthalpies, and the concentrations used in the experiment, the

Wiseman-type isotherms can be asymmetrical around the inflection mid-point, and well defined

steps may appear, which make it impossible to use a standard analysis. Complicated isotherms

arising from cooperative interactions can be characterized by Bayesian analysis according to a

step-wise binding model to determine the macroscopic constants describing the step-wise

Page 9: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

equilibria between stoichiometric states of a receptor [16]. Earlier, exact analytical treatment of

heterotropic binding equilibria have been developed only for the trivial case of α = 1 (no

cooperativity) and for α = 0 (displacement; full negative cooperativity; [17, 18]). Velazques-

Campoy et al. [15] discuss simulations and real experiments with ternary complexes formed

between ferredoxin-NADP+-reductase and the substrates NADP+, ferredoxin and flavodoxin.

They demonstrate that, although mathematically more complex, the analysis provides a

thermodynamically rigorous description of the coupled equilibria. Only three experiments have

to be performed: (i) Binding of A to determine KA and ∆HA, (ii) Binding of B to determine KB

and ∆HB, and (iii) Binding of A in the presence of an arbitrarily chosen concentration of B to

determine α and ∆h. The paper hints at rules for designing experiments and discusses the

conditions which must hold in order to obtain interaction parameters with a high level of

confidence. This work increases the arsenal of tools available to calorimetrists for analyzing the

energetics of ternary interactions. By avoiding the uncertainties that characterize traditional

approaches, it provides a sound foundation for investigating the structural and energetic

mechanisms of allostery.

Velasquez-Campoy [19] has also published a formal description of another complicated

equilibrium phenomenon; namely multiple ligand binding to one-dimensional lattice-like

macromolecule bearing N/l potential binding sites, N and l being the number of homogeneously

distributed lattice points (structural units), and the number of lattice points occupied by the

ligand, respectively. In biology, this situation is not uncommon, since DNA, polysaccharides and

some proteins can be regarded as a one-dimensional lattice. However, the biophysical

characterization of binding equilibria of this type has been attempted only occasionally. The

McGhee-von Hippel general closed form of the Scatchard equation describes the dependence of

the degree of saturation on the free ligand concentration, N, l, the microscopic binding constant,

k, and in the case of cooperative binding, the cooperativity parameter ω [20]. The McGhee-von

Hippel formalism has been used previously to solve specific problems, including analysis of ITC

data. The paper of Velazques-Campoy [19], however, represents a systematic effort to explore

the potential of the derived analytical expression. Next to k, ω and the ligand binding enthalpy,

ITC experiments reveal also the interaction enthalpy associated with the interactions between

neighboring ligands on the lattice, thus providing a deeper level of understanding of the

cooperative process. Numerous simulations are presented in order to illustrate the influence of

different combinations of binding parameters on the shape of the binding isotherms. As some of

the examples show and as the author pointed out, there are situations where analysis is difficult

or uncertain when the data are not very precise, or if additional information on the properties of

Page 10: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

the system is not available. It remains to be seen whether this method will find applications in

near future. The systems devised within the framework of the McGhee-von Hippel theory and

the presented extension are very challenging, bearing in mind their size, their largely unspecific

character (implying low affinity and enthalpy), and possibly slow kinetics which are all limiting

factors in ITC experiments.

Conformational transitions in macromolecules

Although ITC is typically used to characterize ligand binding, the method has also been

applied in studies of conformational transitions of macromolecules. Examples are protein

fragment complementation [21], protein folding or peptide coil-to-helix transitions triggered by

pH [22], salts, metals, and ligands, formation of oligomeric proteins by association of subunits

[23, 24], etc. Nucleic acid duplex formation from the isolated strands also represents large scale

conformational transitions amenable to ITC studies. [25]. Some studies published in 2006 are

worth mentioning. Protein fragment complementation studies provide a unique possibility to

investigate the energetic principles governing the hierarchical/modular organization of proteins

in terms of inter-domain versus intra-domain interactions. A good example of the approach can

be found in the paper by Shuman et al [21]. Calmodulin (CaM) is built up from four EF-hand

motifs, which are individually stabilized by Ca2+ and are arranged sequentially in two globular

domains. The reconstitution of the domains from the isolated EF-hand motifs was studied by

ITC. The pair of C-terminal EF-hands (EF3 and EF4) assembles with higher affinity than the N-

terminal pair (EF1 and EF2) but the difference is relatively small (a factor of 15). More

intriguing is the fact that the similar free energy changes are the result of very different enthalpy-

entropy balance. Both complexes are driven by enthalpy changes (at the selected conditions) but

the enthalpy of formation of EF3:EF4 is four times higher than that of EF1:EF2. In contrast,

formation of EF1:EF2 is entropically favored as well, while the total entropy changes oppose

association of FF3 and EF4. The pronounced energetic differences cannot be readily explained

by structural data, since the packing interaction at the interface of the EF-hand motifs within

each CaM domain is essentially identical. The likely explanation is structural differences in the

isolated EF-hand motifs, which are only partly folded, as judged by CD spectroscopy. Even so,

the picture is not totally clear since EF1 and EF2 experience more pronounced refolding within

the EF1:EF2 complex. Intuitively, association-coupled folding is expected to make the reaction

more exothermic and decrease the entropy, but the experiment shows the opposite.

Page 11: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

Unfortunately, the paper does not state whether the tabulated parameters have been corrected for

possible contribution from proton uptake or release.

The strength and energetic signature of interactions between the non-covalently

associated subunits of a protein may give clues about the physical-chemical mechanisms that

have led to the appearance and diversification of oligomeric proteins in the face of the

evolutionary pressure. ITC provides a useful alternative to study the energetics of

homooligomers by performing a dilution experiment [2]. The assembled oligomer at high

concentration is injected into the cell containing buffer. Initially, the protein concentration in the

cell is low and the monomeric state is populated predominantly. After additional injections, the

monomer concentration increases and the equilibrium shifts towards the oligomeric state. The

resulting titration curve can be deconvoluted to extract the equilibrium constant and the enthalpy

of oligomerization. This approach has been used by Luke et al. to characterize the assembly of

the heptameric GroES protein from the constituting monomers [23]. As for other proteins in the

family, both the entropy and enthalpy of association are positive, indicating that dehydration of

the inter-subunit interface is the major thermodynamic factor driving the assembly of the

heptamer.

Scolnik et al. have compared the energetics of coil-to-helix transition of poly (L-Glu)24

and poly (D-Glu)24 peptides [22]. Hydrochloric acid was titrated into the calorimetric cell to

lower pH from neutral to acidic. The shift of equilibrium of the peptides from the coil state (at

neutral pH) to the α-helical state (at low pH) was monitored by the associated heat changes. The

onset of the coil-to-helix transition was ~0.3 pH units higher, and proceeded with considerably

higher cooperativity and molar enthalpy in the case of poly (D-Glu)24. This means that poly (D-

Glu)24 helices are more stable than their chiral isomers and are more populated at lower degree of

deprotonation of Glu side chains. The most intriguing observation, however, was that the

differences disappear when the experiment is performed in D2O-H2O (4:1 v/v) mixtures. These

results represent one of the first experimental confirmations that L-amino acids are slightly more

hydrated (i.e. more hydrophilic) than the corresponding D mirror images. Further investigations

are needed to clarify whether the hypothesized minor structural and energetic differences

between peptides (and other compounds) composed of L and D enantiomers could have

represented a thermodynamic advantage in the chiral selection at the early stages of evolution of

organic matter.

Almost all ITC studies of conformational transitions in macromolecules have

concentrated on extracting thermodynamic parameters. Recently, Fanghänel et al. have

investigated more systematically the potential of ITC to yield data on protein folding kinetics

Page 12: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

[27]. As a model system they have selected RNase A, one of the best kinetically characterized

proteins to date. The idea of the experiment is to monitor the time-dependent changes of the

differential power signal after folding of the protein has been initiated at the desired buffer

conditions and temperature. For that purpose, after thermal equilibration, the injection syringe is

removed, the protein is injected manually into the cell, the system is reassembled and rotation is

started again. These manipulations disturb the thermal equilibrium, and meaningful data can be

collected after re-equilibration, which in these author’s hands takes 200 to 400 s. In a blank

experiment, the same manipulations are repeated but plain buffer is injected into the cell. The

difference of the two differential power-versus-time traces reflects the time-dependent heat

changes associated with the folding process. At conditions where the native state of RNAse is

marginally stabilized and folding is a simple mono-exponential process, a very reasonable

agreement was obtained between the time constants extracted from ITC data and spectroscopic

data. However, at strongly stabilizing conditions the time constants estimated by ITC for the bi-

phasic folding trace considerably overestimated the spectroscopic data. Further results with

RNAse and short proline-containing peptides demonstrate the potential feasibility of the method

in studies of processes involving proline cis-trans isomerization. There is an additional benefit

from studying folding by ITC. Knowing the time constant(s) from kinetic analysis, the

differential power-versus-time trace can be extrapolated to time zero and the molar enthalpy of

folding can be obtained by integration. Hence, the energetics of protein folding can be

characterized in the temperature range where the native state dominates, without the need of long

extrapolation of thermal melting data. In principle, folding intermediates populated at low

temperatures could be detected in this way. Altogether the paper is intriguing in that it introduces

new tools to study protein folding. However, some notes of caution are in order. First, as in other

kinetic experiments, meaningful time constants can be calculated only when the half time of the

studied process is substantially larger than the dead time of the measurement, which is on the

order of several minutes in this case. Second, a thorough thermal baseline equilibration and

negligible baseline drift should be warranted within the “productive” data collection interval,

both in the genuine and in the control experiments.

Another publication presenting kinetic data acquired by ITC deals with reconstitution of

the heterodimeric protein monellin from its constituent monomeric polypeptide chains [24]. The

experiment for such a system is straightforward: One component is titrated into the cell

containing the counterpart and the heat flow is followed after the system has re-equilibrated. The

reported ITC traces yield data which agree well with spectroscopic data. Unfortunately, little

Page 13: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

experimental details are given in the paper, making it difficult to assess the general applicability

of the experimental setup.

DIVERSE APPLICATIONS UTILIZING STANDARD PROTOCOLS

Protein-ligand binding

By far most of the published papers are devoted to characterization of protein-ligand

interactions. Two general types of studies can be distinguished. First, in-depth thermodynamic

description of a particular complex, or of closely related complexes is attempted [for example

see refs. 28-34]. Second, ITC is used to map binding sites and clarify binding modes by

determining the dissociation constants of wild type complexes and designed variants thereof [for

example see refs. 35-42]. An illustration of the latter type of study is the paper by Honnappa et

al. [37]. The interaction between the carboxy-terminal domain of EB1 (EB1c) and the N-

terminal CAP-Gly domain of the dynactin large subunit p150glued (p150n) was studied. The X-ray

structure of the EB1c-p150n complex was solved from a crystal form that proved to be identical

to the one analyzed previously by others. In the asymmetric unit, two p150n molecules contact

the EB1c dimer, thus defining two major contacts between EB1c and p150n, referred to as A and

B. Contact A has been described extensively by Hayashi et al. and put forward as the

biologically relevant contact [37]. On the basis of sequence alignments and earlier biochemical

data, Honnappa et al. hypothesized that contact B is more likely to be the relevant binding spot

in solution. ITC experiments revealed that mutations removing intermolecular interactions within

contact B greatly diminish or abolish binding. The result was confirmed by “positive design”

experiments. An alanine side chain on p150n points to a conserved hydrophobic cavity on EB1c

within contact B and the designed Ala-to-Met replacement caused a 20 fold increase of binding

affinity. In the X-ray structure of the mutant complex the methionine side chain fits into the

EB1c hydrophobic cavity, thereby expelling the four water molecules that are present in the wild

type structure.

Although the described approach is no doubt valid, two potential problems must be

considered. First, straightforward interpretation of the results of mutational studies is only

possible if extensive controls are undertaken to establish that the mutant proteins are correctly

folded. Second, KD changes induced by mutations are often listed and discussed in terms of

corresponding enthalpy and entropy changes, although controls for possible changes in

protonation and other nonspecific effects have not been carried out or are not mentioned. The

absence of such controls may jeopardize conclusions that are drawn regarding the contribution of

individual binding spots to the stability of a complex.

Page 14: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

Nucleic acid recognition

Protein-DNA binding is an enigmatic case of macromolecular recognition: The DNA

duplex is structurally monotonous, yet DNA-binding proteins utilize a variety of structural and

thermodynamic strategies for binding. Association is structurally promoted by α-helices, the face

or edges of β-sheets, and loops. Both the major and minor DNA grooves can accommodate

proteins. A protein-DNA complex can be enthalpically or entropically driven at physiological

temperatures, the magnitude of ∆H and ∆S varying in a broad range. Certain classes of proteins

use N-terminal or C-terminal, unstructured extensions to contact DNA. In 2006 a “Tale of tails”

has been narrated by Crane-Robinson et al. [13] who reviewed results from Peter Privalov’s

laboratory [see also ref. 305]. The DNA binding properties of homeodomains, HMG boxes, and

members of the HMGA family, all containing DNA-contacting tails, have been characterized.

The strategy is the following. First, ITC is used in combination with differential scanning

calorimetry to obtain the binding parameters “free” of contributions from temperature-induced

and binding-induced partial refolding. Second, the binding parameters in 1 M NaCl are

determined to separate the electrostatic and non-electrostatic contributions to the energy terms.

Third, the contribution of the DNA-contacting tails is filtered out by comparing the energetic

signatures of complexes formed by proteins containing the extension or truncated domains. The

analysis led to the conclusion that there was no unifying principle. Homeodomain arms

contribute mostly non-electrostatic binding free energy. The purely electrostatic interactions

involving the HMG-box tails confer not only stabilization but neutralize asymmetrically the

phosphate charges in the major groove and promote bending of the DNA duplex. The so-called

AT hooks bind in the minor groove driven in part by electrostatics, yet the main driving force is

the favorable entropy change, i.e. the hydrophobic effect.

Two interesting papers on protein-DNA association are discussed below in the

“Structural parameterization” section.

In view of the ever-growing interest in the development of DNA-binding drugs in

anticancer research, we would like to draw attention to the paper by Jonathan Chairs [14]. This

author compiled a large amount of calorimetric data on binding of small molecules to DNA. The

analysis reveals a very different energetic signature for intercalating and groove-binding

compounds. The binding of intercalators is enthalpically driven, while groove-binding is

entropically driven. Possible molecular origins of the distinctive energetic patterns are discussed.

Furthermore, the author contrasts the enthalpy-entropy fingerprints of small molecules and

DNA-binding proteins.

Page 15: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

While dozens of studies address the energetics of protein-DNA recognition, much less is

known about protein binding to RNA. This may change since several reports in 2006 describe

the thermodynamics of RNA binding by proteins and small molecular weight compounds [for

example see references 43 and 44]. Following the increase of structural data accumulated in

recent years, it seems likely that the number of biophysical and calorimetric studies of RNA

binding reactions will also increase.

Lipid/membrane systems

As in other fields of biomolecular research, the applications of ITC in studies of lipid

phases, membranes and mixed protein-lipid/membrane systems also increase steadily. Given the

ongoing multidisciplinary efforts toward a better structural and functional characterization of

membrane proteins, there is no doubt that more ITC studies will be done in the future. A good

example is the work published by Keller et al. [45]. These authors derived a function that can

be used to assess membrane translocation of charged substances. The function is general in the

sense that it unifies ITC release and uptake titrations and makes possible the implementation of

different membrane partitioning models. Electrostatic effects are modeled within the framework

of the Guoy-Chapman theory. The merits of the method have been tested by characterizing the

flip-flop of SDS across the membrane of unilamellar POPC vesicles. Membrane translocation of

charged substances is a ubiquitous phenomenon in biology and the proposed methodology

widens the number of systems that can be analyzed without the need of introducing reporter

groups.

ENZYME KINETICS

The potential of calorimetry to study enzymatic reactions is well established [46]. The

rate of enzymatic reactions is proportional to the heat flow associated with formation or breakage

of chemical bonds. Therefore, the measurement is independent of the existence and properties of

spectroscopically active groups, or of the availability of substrate analogues bearing artificial

spectroscopic tags. The principle, data acquisition and data analysis of kinetic experiments using

ITC to extract Km and kcat of enzymatic reactions have been reviewed [47]. D’Amico et al. have

presented an extension of the standard methodology [48]. While studying the activity of α-

amylases, the authors injected enzyme aliquots into the cell containing the substrate at varying

concentrations instead of making serial additions of substrate to the enzyme solution. Using

enzyme-to-substrate molar ratios sufficiently low to avoid substrate depletion, the steady state

regime is reached. The differential power signal deflects from the thermal baseline (Pi) and

Page 16: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

rapidly reaches a plateau (Pr). The reaction rate is proportional to the heat exchange of the

reaction, i.e. to Pr − Pi. Consecutive enzyme additions in the steady state regime cause equally

sized “steps” in the differential power signal, allowing a straightforward check of the linear

dependence of the reaction rate on the enzyme concentration (which should be maintained in the

absence of product inhibition). Decreasing concentrations of substrate lead to a decrease of the Pr

− Pi differences. From the hyperbolic saturation plots of (Pr − Pi) versus [S], Km and Vmax can be

calculated in the usual way by regression analysis using the Michaelis-Menten equation.

Whether this experimental setup is superior to the “standard” protocol remains to be seen, but it

appears that it facilitates easy monitoring of product inhibition and requires less a priori

knowledge or educated guesses to design meaningful experiments, a clear advantage when

studying enzymatic reactions for the first time.

Søren Olsen has published the results of an interesting investigation dealing with the

capability of ITC to measure enzyme kinetics and activity in solutions which cannot be classified

as “ideal”, in the usual way this adjective is used in the biochemical literature [49]. Since

enzyme specificity tends to be very high, the heat flow observed in an ITC experiment conducted

in a complex solution should directly reflect the enzyme activity at fixed pH, temperature,

pressure and salt concentration. Yet the influence of additives that interfere either with the

enzymatic action per se (inhibitors, allosteric regulators, etc), or with the enzyme structure has

never been rigorously tested. The problem is important since enzymes work in a “crowded”

environment in the cell, and many biotechnological processes require addition of highly

concentrated supplements. Using a model system, Olsen studied the activity of hexokinase in

solutions containing very high concentrations of BSA as the crowding agent (up to 300 mg ml–1).

It is noteworthy that the sensitivity of the method was not influenced by the crowded

environment. However, Km, kcat and the enthalpy (∆Hreact) of glucose phosphorylation decreased

as the concentration of BSA increased. This could be interpreted by assuming that the kcat

decrease is probably due to a slower diffusion rate of the product away from the enzyme-product

complex. Considering the decrease of Km, the author discussed the decrease of the activity

coefficient of the enzyme-substrate complex relative to the activity coefficient of the free

enzyme caused by the volume/surface reduction of the former relative to the latter. In view of the

reduction of ∆Hreact, the high BSA concentration might have changed the relative hydration

properties of the substrate and product. Although the paper reaches no definitive conclusion

regarding the molecular origins of the observed effects, the results point to the importance of

hydration effects in modulating the kinetic and thermodynamic parameters of enzymatic

reactions in crowded environments. It is worth noting that the specificity constant of hexokinase

Page 17: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

(kcat/Km) is not influenced by the presence of high BSA concentrations. It is not clear whether

this study encountered another case of “compensation behavior”.

DRUG DESIGN

A prime example for the potential of ITC for understanding the principles of molecular

recognition is the field of drug design and optimization [50]. The retroviral protease is a major

target for development of therapeutic strategies against retrovirus-caused syndromes. One of the

hurdles in AIDS therapy is the fast rate of appearance of strains that are drug-resistant to

protease inhibitors (PI). The viral protease is subjected to mutations which compromise severely

its affinity for artificial substrates (i.e. PI) with viable catalytic activity. Most of these mutations

affect non-polar side chains and re-shape the binding pocket without altering the polarity and

charge distribution. As the consequence, inhibitors designed for structural complementarity to

the wild-type binding site are incapable to bind with high affinity, either because enthalpically-

rich interactions become sub-optimal, or because the entropy gain stemming from dehydration of

the binding interface is low within the geometrically altered binding pocket. Ernesto Freire has

published an insightful review on current attempts at overcoming the HIV-1 resistance to

protease inhibitors [12]. His description of the thermodynamic evolution of PI illustrates the

conceptual complementarity between structural information and thermodynamic data. There are

two general energetic mechanisms of preservation of drug potency (in terms of binding affinity)

in the face of the most frequently occurring active site mutations in the HIV-1 protease. First,

small losses in binding enthalpy and binding entropy may still preserve a reasonable affinity.

Second, large enthalpic losses must be compensated by entropy gain, or alternatively,

pronounced enthalpic optimization must over-compensate the entropic expenditure. Cihlar et al.

have reported an interesting illustration of this phenomenon [51]. They have modified two of the

closely related high-affinity HIV-1 protease drugs, amprenavir and TMC-126, by attaching a

phosphonic acid moiety to the drug scaffold in a position that is not buried in the protease-drug

complex. The modified drugs are less affected by common resistance-inducing protease

mutations in antiviral activity assays. ITC experiments revealed that the modification does not

alleviate the enthalpic losses experienced by amrenavir and TMC-126 in the mutated binding

pocket, yet leads to entropic gains. On the basis of structural analysis the authors suggest a

“solvent anchoring” thermodynamic mechanism: The solvent-exposed and favorably hydrated

phosphonate group stabilizes the protease-inhibitor complex, since the inhibitor, being “solvent

anchored” (i.e. being “fixed” in an approximately favorable orientation), can explore a wider

conformational space within the binding pocket. The result is a favorable increase of the binding

Page 18: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

entropy. This paper illustrates how an investigation that was conceived to answer certain

questions, may lead to totally new ideas when applied in different context.

The identification of binding sites and the mapping of hot spots within a binding pocket

are key steps in the development of drugs. A study by Ciulli et al. presents a very clear example

of a useful experimental approach [41]. They have used the E. coli ketopantoate reductase as a

model system to probe for the energetically important structural elements that confer affinity and

specificity in binding of the cofactor NADPH. The binding of NADPH-derived fragments to the

wild type protein and binding site mutants thereof was characterized by ITC. Two hot spots were

identified at the opposite ends of the binding cavity. The paper demonstrates the potential of

binding studies using small fragments for exploring the hierarchical structural organization that

determines ligand efficiency and to chemical scaffolds as targets of optimization.

STRUCTURE-ENERGY CORRELATIONS

The advantage of the calorimetric method over other techniques is that it provides a direct

decomposition of the binding free energy into temperature-dependent enthalpic and entropic

components. But even if measured with very high precision, the “generic” ∆HA, ∆SA and ∆Cp,A

(after corrections for (de)protonation and other non-specific effects) are still numbers that

contain no "molecular" information. Real progress in understanding the mechanisms of

macromolecular association can be achieved only if an intimate link is to be established between

energy and structure. Since non-polar and polar contacts exhibit different thermodynamic

signatures, the intuitive and relatively simple approach which has been pursued is to

parameterize the observed quantities in terms of the amount and type of surface buried in a

complex [52, 53]. The idea has been adopted from studies of protein energetics; namely that ∆H,

∆S, and ∆Cp of protein unfolding could be predicted with relatively good precision from the

increase in surface exposure upon disruption of the compact native state [ref. 54 and references

therein]. It is debatable whether the surface-energy correlations found for proteins really take

into account the physics of the underlying process, or whether they simply represent a useful

empirical observation. If such correlations could be established for binding reactions, a deeper

understanding of the forces holding the binding partners together would result and predictions of

the binding profile may become possible, which would be of considerable value in the area of

drug design. Although, there are many examples where structure-energy correlations hold [52,

53], there are also numerous reports about large discrepancies between measured and predicted

values. Several papers published in 2006 re-fuel the debate about how good structure-based

predictions of ∆Cp are, and discuss possible reasons for the observed “anomalous” ∆Cp values.

Page 19: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

Kozlov and Lohman report a dramatic example [55]. They have measured binding of the

E. coli SBB tetramer to single-stranded DNA. Previous work of these authors demonstrated large

contributions to the apparent ∆H and ∆Cp from temperature-dependent changes in the

protonation state of protein side chains and base unstacking of DNA. In the much wider

temperature interval studied now, ∆Cp exhibits a strong temperature dependence (which is

usually small and can be neglected). The negative slope of ∆H-versus-T plots decreases upon

temperature increase. Above 35 °C the slope becomes even positive, indicating that ∆Cp changes

not only in magnitude but also its sign. Furthermore, both ∆H and ∆Cp are significantly

influenced by the concentration and type of monovalent anions. The authors propose a

thermodynamic linkage model according to which the measured temperature dependence of ∆H

and ∆Cp can be rationalized by the existence of coupled equilibria: (i) The protein undergoes

temperature- and salt-dependent conformational transition; (ii) Anions preferentially bind to the

protein and are released from the protein-DNA complex; (iii) Net protonation of the protein

occurs. All these processes add enthalpy and heat capacity to the “genuine” ∆H and ∆Cp (an

estimation of the magnitude is attempted in the paper), yet cannot be captured by

parameterization schemes using “fixed” molecular structures. The discussed “anomalies” might

be widely present in protein-DNA recognition. Unfortunately, characterization of DNA binding

over a large temperature interval and at various solvent conditions has been attempted only

occasionally.

Experimental heat capacity changes larger than the structure-based predictions are typical

for protein-DNA complexes. The Klenow polymerase is no exception: ∆Cp determined for the

association to primed-template DNA is twice higher than the expected one on the basis of

surface burial/dehydration [56]. Large ∆Cp of a protein-DNA complex is thought to reflect

sequence-specific binding, since almost all of the nonsequence-specific complexes exhibit small

∆Cp. Interestingly, Klenow polymerase shows no sequential preferences, as all type I

polymerases. Rather, it recognizes the primed-template structure while looking for molecular

reasons which may explain an “extra” ∆Cp, the authors note that Rg of Klenow decreases by ~1

Ǻ upon DNA binding and that the protein conformation is more ordered in certain regions of the

co-crystal structure compared to the non-bound enzyme, indicating compaction within the

protein-DNA complex. However, the decrease in Rg is too small to account for the missing part

of ∆Cp stemming from additional surface burial.

Interesting thoughts on the putative links between structure and energetic changes can be

found in ref. [31]. Cognate and non-cognate complexes of immunity proteins (Im2, Im7, Im8 and

Im9) with the DNAse domains of colicins (E2, E7, E8 and E9) have been characterized by ITC.

Page 20: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

Structurally, the Im2/E7, Im7/E7 and Im8/E8 complexes exhibit a higher hydrogen bond density

at the binding interface than the Im9/E9 complex. Furthermore, although the interface areas of

Im7/E7 and Im9/E9 are comparable, the proportion of buried polar surface is higher for Im7/E7.

The specificity determining polar and charged side chains in Im7/E7 are replaced by non-polar

specificity-determining side chains in Im9/E9. Not surprisingly, enthalpic stabilization of the

more polar complexes is stronger, in line with the low enthalpic content of hydrophobic contacts

(at 25 °C). For all cognate pairs, the buried surface-based prediction underestimates the

experimental ∆Cp by a factor of 10. Intriguingly, the most non-polar pair (Im9/E9) has the

smallest ∆Cp, contra to the assumption that the dehydration of non-polar groups dominates the

negative ∆Cp of association. Five and seven water molecules are present at the Im9/E9 and

Im7/E7 interfaces, respectively. Bound water may possibly cause a ∆Cp decrease, but the effect

is too small to explain the experimental observations. The authors suggest as an additional source

of negative ∆Cp, global “rigidification” of the complex structure. The more polar cognate

complexes have a larger favorable ∆H and larger negative ∆Cp than the non-cognate pairs. The

situation is different for E9, where the cognate complex is entropically favored in comparison to

non-cognate complexes, due to the high degree of surface complementarity between non-polar

surfaces. The conclusion is that a high level of discrimination/specificity can be achieved by very

different thermodynamic mechanisms.

CONCLUDING REMARKS

ITC is a fast and robust method to determine the energetics of association reactions in

solution. New theoretical developments help to overcome the intrinsic limitations of the method

and to devise new applications. Both in basic biomedical and biotechnology-oriented research it

is now widely recognized that the magnitude of the changes in enthalpy, entropy and heat

capacity that accompany binding provides unique insights into the balance of forces underlying

biologically-relevant processes. There are conceptual difficulties in interpreting the observed

energetic changes in terms of discrete events at the molecular level and the correlations between

intermolecular contacts and their energetic content are still incompletely understood. The reasons

for this unfortunate situation can be found in the very nature of biological macromolecules, as

well as in the role played by ubiquitous water in tuning the energetics of chemical reactions in

the condensed aqueous phase. Macromolecules are large, cooperative systems which react to

small perturbations in a still largely unpredictable way. Large energetic changes can be caused

by structural changes that are on the limit of sensitivity of structural observations. Nevertheless,

Page 21: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

the year 2006 has brought instructive examples of the synergy between energetic and structural

approaches and it seems likely that the coming years will follow the same trend!

Acknowledgments

This work was funded in part by the Swiss National Center of Competence in Research

“Structural Biology”.

Page 22: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

REFERENCES General subjects

1. Ababou A, Ladbury JE. 2006. Survey of the year 2004:

Literature on applications of isothermal titration

calorimetry. J. Mol. Recognit. 19: 79-89.

2. Ababou A, Ladbury JE. 2007. Survey of the year 2005:

Literature on applications of isothermal titration

calorimetry. J. Mol. Recognit. 20: 4-14.

3. Baker BM, Murphy KP. 1998. Prediction of binding

energetics from structure using empirical

parameterization. Meth. Enzymol. 295: 294-315.

4. Beezer AE, Steenson TI, Tyrrell HJV. 1974.

Application of flow-microcalorimetry to analytical

problems .2. Urea-urease system. Talanta 21: 467-474.

5. Cliff MJ, Gutierrez A, Ladbury JE. 2004. A survey of

the year 2003 literature on applications of isothermal

titration calorimetry. J. Mol. Recognit. 17:513-523.

6. Cliff MJ, Ladbury JE. 2003. A survey of the year 2002

literature on applications of isothermal titration

calorimetry. J. Mol. Recognit. 16:383-391.

7. Cooper A. 1998. Microcalorimetry of protein-protein

interactions. In: Biocalorimetry: Applications of

calorimetry in the biological sciences, Ladbury JE,

Chowdhry BZ (eds). John Wiley & Sons Ltd.:

Chichester, U.K. pp. 103-111.

8. Freire E, Mayorga OL, Straume M. 1990. Isothermal

titration calorimetry. Anal. Chem. 62: A950-A959.

9. Holdgate GA, Ward WHJ. 2005. Measurements of

binding thermodynamics in drug discovery. Drug

Discovery Today 10: 1543-1550.

10. Indyk L, Fisher HF. 1998. Theoretical aspects of

isothermal titration calorimetry. Meth. Enzymol 295:

350-364.

11. Jelesarov I, Bosshard HR. 1999. Isothermal titration

calorimetry and differential scanning calorimetry as

complementary tools to investigate the energetics of

biomolecular recognition. J. Mol. Recognit. 12: 3-18.

12. Luque I, Freire E. 1998. Structure-based prediction of

binding affinities and molecular design of peptide

ligands. Meth. Enzymol. 295: 100-127.

13. McGhee JD, Hippel PHV. 1974. Theoretical aspects of

DNA-protein interactions - cooperative and non-

cooperative binding of large ligands to a one-

dimensional homogeneous lattice. J. Mol. Biol. 86: 469-

489.

14. O'Brien R, Haq I. 2004. Applications of biocalorimetry:

Binding, stability and enzyme kinetics. In:

Biocalorimetry 2: Applications of calorimetry in the

biological sciences, Ladbury JE, Doyle ML (eds).

John Wiley & Sons Ltd.: Chichester, U.K. pp. 3-34.

15. Robertson AD, Murphy KP. 1997. Protein structure and

the energetics of protein stability. Chem. Rev. 97: 1251-

1268.

16. Sigurskjold BW. 2000. Exact analysis of competition

ligand binding by displacement isothermal titration

calorimetry. Anal. Biochem. 277: 260-266.

17. Thomson JA, Ladbury JE. 2004. Isohermal titration

calorimetry: A tutorial. In: Biocalorimetry 2:

Applications of calorimetry in the biological sciences,

Ladbury JE, Doyle ML (eds). John Wiley & Sons Ltd.:

Chichester, U.K. pp. 37-58.

18. Todd MJ, Gomez J. 2001. Enzyme kinetics determined

using calorimetry: A general assay for enzyme activity?

Anal. Biochem. 296: 179-187.

19. Velazquez Campoy A, Freire E. 2005. ITC in the post-

genomic era...? Priceless. Biophys. Chem. 115: 115-

124.

20. Wiseman T, Williston S, Brandts JF, Lin LN. 1989.

Rapid measurement of binding constants and heats of

binding using a new titration calorimeter. Anal.

Biochem. 179: 131-137.

Protein-protein

21. Ames JB, Levay K, Wingard JN, Lusin JD, Slepak VZ.

2006. Structural basis for calcium-induced inhibition of

rhodopsin kinase by recoverin. J. Biol. Chem. 281:

37237-37245.

22. Bergqvist S, Croy CH, Kjaergaard M, Huxford T,

Ghosh G, Komives EA. 2006. Thermodynamics reveal

that helix four in the NLS of NF-kappaB p65 anchors

ikappabalpha, forming a very stable complex. J. Mol.

Biol. 360: 421-434.

23. Beuron F, Dreveny I, Xuemei Yuan X, Pye VE,

Mckeown C, Briggs LC, Cliff MJ, Kaneko Y, R. W,

Isaacson RL, Ladbury JE, Matthews SJ, Kondo H,

Zhang X, Freemont PS. 2006. Conformational changes

in the AAA ATPase p97–p47 adaptor complex. EMBO

J. 25: 1967-1976.

24. Botuyan MV, Lee J, Ward IM, Kim J-E, Thompson JR,

Chen J, Mer G. 2006. Structural basis for the

methylation state-specific recognition of histone H4-

K20 by 53BP1 and Crb2 in DNA repair. Cell.

127:1361-1373.

25. Bowman P, Galea CA, Lacy E, Kriwacki RW. 2006.

Thermodynamic characterization of interactions

between p27Kip1 and activated and non-activated

Cdk2: Intrinsically unstructured proteins as

Page 23: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

thermodynamic tethers. Biochim. Biophys. Acta. -

Proteins & Proteomics. 1764: 182-189.

26. Boyault C, Gilquin B, Zhang Y, Rybin V, Garman E,

Meyer-Klaucke W, Matthias P, Müller CW, Khochbin

S. 2006. HDAC6–p97/VCP controlled polyubiquitin

chain turnover. EMBO J. 25: 3357-3366.

27. Brautigam CA, Wynn RM, Chuang JL, Machius M,

Tomchick DR, Chuang DT. 2006. Structural insight

into interactions between dihydrolipoamide

dehydrogenase (E3) and E3 binding protein of human

pyruvate dehydrogenase complex. Structure. 14: 611-

621.

28. Brondijk THC, van Boxel GI, Mather OC, Quirk PG,

White SA, Jackson JB. 2006. The role of invariant

amino acid residues at the hydride transfer site of

proton-translocating transhydrogenase. J. Biol. Chem.

281: 13345-13354.

29. Bullock AN, Debreczeni JE, Edwards AM, Sundstrom

M, Knapp S. 2006. Crystal structure of the SOCS2-

elongin C-elongin B complex defines a prototypical

SOCS box ubiquitin ligase. Proc. Natl. Acad. Sci. U. S.

A. 103: 7637-7642.

30. Canton DA, Olsten MEK, Niederstrasser H, Cooper JA,

Litchfield DW. 2006. The role of ckip-1 in cell

morphology depends on its interaction with actin-

capping protein. J. Biol. Chem. 281: 36347-36359.

31. Chaudhury C, Brooks CL, Carter DC, Robinson JM,

Anderson CL. 2006. Albumin binding to fcrn: Distinct

from the FcRn-IgG interaction. Biochemistry. 45:4983-

4990.

32. Chen G, Wang, C., Fuqua C, Zhang LH, Chen L. 2006.

Crystal structure and mechanism of TraM2, a second

quorum-sensing antiactivator of agrobacterium

tumefaciens strain A6. J Bacteriol. 188: 8244-8251.

33. Chen YM, Ferrar TS, Lohmeier-Vogel EM, Morrice N,

Mizuno Y, Berenger B, Ng KKS, Muench DG,

Moorhead GBG. 2006. The PII signal transduction

protein of Arabidopsis thaliana forms an arginine-

regulated complex with plastid N-acetyl glutamate

kinase. J. Biol. Chem. 281: 5726-5733.

34. Chereau D, Dominguez R. 2006. Understanding the

role of the G-actin-binding domain of Ena/VASP in

actin assembly. J. Struct. Biol. 155: 195-201.

35. Choi H-J, Huber AH, Weis WI. 2006. Thermodynamics

of beta-catenin-ligand interactions: The roles of the N-

and C-terminal tails in modulating binding affinity. J.

Biol. Chem. 281: 1027-1038.

36. Chrencik JE, Brooun A, Kraus ML, Recht MI, Kolatkar

AR, Han GW, Seifert JM, Widmer H, Auer M, Kuhn P.

2006. Structural and biophysical characterization of the

EphB4/Ephrin B2 protein-protein interaction and

receptor specificity. J. Biol. Chem. 281: 28185-28192.

37. Couture J-F, Collazo E, Trievel RC. 2006. Molecular

recognition of histone H3 by the WD40 protein WDR5.

Nature Struct. Mol. Biol. 13: 698-703.

38. Crawley SW, de la Roche MA, Lee S-F, Li Z, Chitayat

S, Smith SP, Cote GP. 2006. Identification and

characterization of an 8-kDa light chain associated with

Dictyostelium discoideum Myob, a class I myosin. J.

Biol. Chem. 281: 6307-6315.

39. Demarest SJ, Hopp J, Chung J, Hathaway K,

Mertsching E, Cao X, George J, Miatkowski K,

LaBarre MJ, Shields M, Kehry MR. 2006. An

intermediate pH unfolding transition abrogates the

ability of IgE to interact with its high affinity receptor

FcεRIα. J. Biol. Chem. 281: 30755-30767.

40. Di Lello P, Jenkins LMM, Jones TN, Nguyen BD, Hara

T, Yamaguchi H, Dikeakos JD, Appella E, Legault P,

Omichinski JG. 2006. Structure of the Tfb1/p53

complex: Insights into the interaction between the

p62/Tfb1 subunit of TFIIH and the activation domain of

p53. Mol. Cel. 22: 731-740.

41. Dogan J, Lendel C, Hard T. 2006. Thermodynamics of

folding and binding in an affibody:affibody complex. J.

Mol. Biol. 359: 1305-1315.

42. Eldredge J, Berkowitz S, Corin AF, Day ES, Hayes D,

Meier W, Strauch K, Zafari M, Tadi M, Farrington GK.

2006. Stoichiometry of LTbetaR binding to LIGHT.

Biochemistry. 45: 10117-10128.

43. Ely LK, Beddoe T, Clements CS, Matthews JM, Purcell

AW, Kjer-Nielsen L, McCluskey J, Rossjohn J. 2006.

Disparate thermodynamics governing T cell receptor-

MHC-I interactions implicate extrinsic factors in

guiding MHC restriction. Proc. Natl. Acad. Sci. U. S. A.

103: 6641-6646.

44. Evans EJ, Castro MAA, O'Brien R, Kearney A, Walsh

H, Sparks LM, Tucknott MG, Davies EA, Carmo AM,

van der Merwe PA, Stuart DI, Jones EY, Ladbury JE,

Ikemizu S, Davis SJ. 2006. Crystal structure and

binding properties of the CD2 and CD244 (2B4)-

binding protein, CD48. J. Biol. Chem. 281: 29309-

29320.

45. Evans LDB, Stafford GP, Ahmed S, Fraser GM,

Hughes C. 2006. An escort mechanism for cycling of

export chaperones during flagellum assembly. Proc.

Natl. Acad. Sci. U. S. A. 103:17474-17479.

46. Fairlie WD, Perugini MA, Kvansakul M, Chen L,

Huang DCS, Colman PM. 2006. CED-4 forms a 2:2

heterotetrameric complex with CED-9 until specifically

displaced by EGL-1 or CED-13. Cell Death Differ. 13:

426-434.

47. Fanghanel J, Wawra S, Lucke C, Wildemann D, Fischer

G. 2006. Isothermal calorimetry as a tool to investigate

slow conformational changes in proteins and peptides.

Page 24: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

Anal. Chem. 78: 4517-4523.

48. Fellouse FA, Barthelemy PA, Kelley RF, Sidhu SS.

2006. Tyrosine plays a dominant functional role in the

paratope of a synthetic antibody derived from a four

amino acid code. J. Mol. Biol. 357: 100-114.

49. Ford B, Hornak V, Kleinman H, Nassar N. 2006.

Structure of a transient intermediate for GTP hydrolysis

by Ras. Structure. 14: 427-436.

50. Harjes E, Harjes S, Wohlgemuth S, Muller K-H,

Krieger E, Herrmann C, Bayer P. 2006. GTP-Ras

disrupts the intramolecular complex of C1 and RA

domains of Nore1. Structure. 14: 881-888.

51. Harrington AE, Morris-Triggs SA, Ruotolo BT,

Robinson CV, Ohnuma S, Hyvonen M. 2006. Structural

basis for the inhibition of activin signalling by

follistatin. EMBO J. 25: 1035-1045.

52. Honnappa S, Jahnke W, Seelig J, Steinmetz MO. 2006.

Control of intrinsically disordered stathmin by multisite

phosphorylation. J. Bio.l Chem. 281: 16078-16083.

53. Honnappa S, Okhrimenko O, Jaussi R, Jawhari H,

Jelesarov I, Winkler FK, Steinmetz MO. 2006. Key

interaction modes of dynamic +Tip networks. Mol.

Cell. 23: 663-671.

54. Houtman JCD, Yamaguchi H, Barda-Saad M, Braiman

A, Bowden B, Appella E, Schuck P, Samelson LE.

2006. Oligomerization of signaling complexes by the

multipoint binding of GRB2 to both LAT and SOS1.

Nature Struct. Mol. Biol. 13: 798-805.

55. Izadi-Pruneyre N, Huche F, Lukat-Rodgers GS,

Lecroisey A, Gilli R, Rodgers KR, Wandersman C,

Delepelaire P. 2006. The heme transfer from the soluble

HasA hemophore to its membrane-bound receptor

HasR is driven by protein-protein interaction from a

high to a lower affinity binding site. J. Biol. Chem. 281:

25541-25550.

56. Jenkins HT, Mark L, Ball G, Persson J, Lindahl G,

Uhrin D, Blom AM, Barlow PN. 2006. Human C4B-

binding protein, structural basis for interaction with

streptococcal M protein, a major bacterial virulence

factor. J. Biol. Chem. 281: 3690-3697.

57. Keeble AH, Kirkpatrick N, Shimizu S, Kleanthous C.

2006. Calorimetric dissection of colicin DNase-

immunity protein complex specificity. Biochemistry.

45: 3243-3254.

58. Kim MS, Kim J-A, Song HK, Jeon H. 2006. STAM-

AMSH interaction facilitates the deubiquitination

activity in the C-terminal AMSH. Biochem. Biophys.

Re.s Commun. 351: 612-618.

59. Kozlov G, Maattanen P, Schrag JD, Pollock S, Cygler

M, Nagar B, Thomas DY, Gehring K. 2006. Crystal

structure of the bb' domains of the protein disulfide

isomerase ERp57. Structure 14: 1331-1339.

60. Lamers MH, Georgescu RE, Lee S-G, O'Donnell M,

Kuriyan J. 2006. Crystal structure of the catalytic α-

subunit of E. coli replicative DNA polymerase III. Cell

126: 881-892.

61. Lazic A, Dolmer K, Strickland DK, Gettins PGW.

2006. Dissection of RAP-LRP interactions: Binding of

RAP and RAP fragments to complement-like repeats 7

and 8 from ligand binding cluster II of LRP. Arch.

Biochem. Biophys. 450: 167-175.

62. Lee S, Tsai YC, Mattera R, Smith WJ, Kostelansky

MS, Weissman AM, Bonifacino JS, Hurley JH. 2006.

Structural basis for ubiquitin recognition and

autoubiquitination by Rabex-5. Nature Struc.t Mol.

Biol. 13: 264-271.

63. Li JZ, Wu YK, Qian XG, Sha BD. 2006. Crystal

structure of yeast Sis1 peptide-binding fragment and

Hsp70 Ssa1 C-terminal complex. Biochem J. 398: 353-

360.

64. Loftus SR, Walker D, Mate MJ, Bonsor DA, James R,

Moore GR, Kleanthous C. 2006. Competitive

recruitment of the periplasmic translocation portal TolB

by a natively disordered domain of colicin E9. Proc.

Natl. Acad. Sci. U. S. A. 103: 12353-12358.

65. Lu J, Machius M, Dulubova I, Dai H, Sudhof TC,

Tomchick DR, Rizo J. 2006. Structural basis for a

Munc13-1 homodimer to Munc13-1/RIM heterodimer

switch. PloS Biol. 4: 1159-1172.

66. Lu L, Ma YY, Zhang JR. 2006. Streptococcus

pneumoniae recruits complement factor H through the

amino terminus yof CbpA. J. Biol. Chem. 281: 15464-

15474.

67. Luke K, Wittung-Stafshede P. 2006. Folding and

assembly pathways of co-chaperonin proteins 10:

Origin of bacterial thermostability. Arch. Biochem.

Biophys. 456: 8-18.

68. Lykken GL, Chen G, Brutinel ED, Chen L, Yahr TL.

2006. Characterization of ExsC and ExsD self-

association and heterocomplex formation. J. Bacteriol.

188: 6832-6840.

69. Machius M, Wynn RM, Chuang JL, Li J, Kluger R, Yu

D, Tomchick DR, Brautigam CA, Chuang DT. 2006. A

versatile conformational switch regulates reactivity in

human branched-chain α-ketoacid dehydrogenase.

Structure 14: 287-298.

70. Meher AK, Bal NC, Chary KVR, Arora A. 2006.

Mycobacterium tuberculosis H37Rv ESAT-6-CFP-10

complex formation confers thermodynamic and

biochemical stability. FEBS J. 273: 1445-1462.

71. Munshi UM, Kim J, Nagashima K, Hurley JH, Freed

EO. 2007. An Aalix fragment potently inhibits HIV-1

budding - characterization of binding to retroviral

YPXL late domains. J. Biol. Chem. 282: 3847-3855.

Page 25: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

72. Muskotal A, Kiraly R, Sebestyen A, Gugolya Z, Vegh

BM, Vonderviszt F. 2006. Interaction of FliS flagellar

chaperone with flagellin. FEBS Lett. 580: 3916-3920.

73. Nakamura F, Pudas R, Heikkinen O, Permi P,

Kilpelainen I, Munday AD, Hartwig JH, Stossel TP,

Ylanne J. 2006. The structure of the GPIb-filamin a

complex. Blood 107: 1925-1932.

74. Nezami AG, Poy F, Eck MJ. 2006. Structure of the

autoinhibitory switch in formin mDia1. Structure 14:

257-263.

75. Olia AS, Al-Bassam J, Winn-Stapley DA, Joss L,

Casjens SR, Cingolani G. 2006. Binding-induced

stabilization and assembly of the phage p22 tail

accessory factor gp4. J. Mol. Biol. 363: 558-576.

76. Padovani D, Labunska T, Banerjee R. 2006. Energetics

of interaction between the G-protein chaperone, MeaB,

and B-12-dependent methylmalonyl-CoA mutase. J.

Biol. Chem. 281: 17838-17844.

77. Park S-Y, Lowder B, Bilwes AM, Blair DF, Crane BR.

2006. Structure of FliM provides insight into assembly

of the switch complex in the bacterial flagella motor.

Proc. Natl. Acad. Sci. U. S. A. 103: 11886-11891.

78. Patel CN, Smith VF, Randall LL. 2006.

Characterization of three areas of interactions

stabilizing complexes between SecA and SecB, two

proteins involved in protein export. Protein Sci. 15:

1379-1386.

79. Penengo L, Mapelli M, Murachelli AG, Confalonieri S,

Magri L, Musacchio A, Di Fiore PP, Polo S, Schneider

TR. 2006. Crystal structure of the ubiquitin binding

domains of Rabex-5 reveals two modes of interaction

with ubiquitin. Cell 124: 1183-1195.

80. Reyes-Turcu FE, Horton JR, Mullally JE, Heroux A,

Cheng X, Wilkinson KD. 2006. The ubiquitin binding

domain ZnF UBP recognizes the C-terminal diglycine

motif of unanchored ubiquitin. Cell 124: 1197-1208.

81. Rozak DA, Alexander PA, He YN, Chen YH, Orban J,

Bryan PN. 2006. Using offset recombinant polymerase

chain reaction to identify functional determinants in a

common family of bacterial albumin binding domains.

Biochemistry 45: 3263-3271.

82. Ryan DP, Sunde M, Kwan AHY, Marianayagam NJ,

Nancarrow AL, vanden Hoven RN, Thompson LS,

Baca M, Mackay JP, Visvader JE, Matthews JM. 2006.

Identification of the key LMO2-binding determinants

on Ldb1. J. Mol. Biol. 359: 66-75.

83. Sampietro J, Dahlberg CL, Cho US, Hinds TR,

Kimelman D, Xu W. 2006. Crystal structure of a β-

catenin/BCL9/Tcf4 complex. Mol Cell 24: 293-300.

84. Saunders RM, Holt MR, Jennings L, Sutton DH,

Barsukov IL, Bobkov A, Liddington RC, Adamson EA,

Dunn GA, Critchley DR. 2006. Role of vinculin in

regulating focal adhesion turnover. Eur. J. Cell Biol.

85:4 87-500.

85. Schonichen A, Alexander M, Gasteier JE, Cuesta FE,

Fackler OT, Geyer M. 2006. Biochemical

characterization of the diaphanous autoregulatory

interaction in the formin homology protein FHOD1. J.

Biol. Chem. 281: 5084-5093.

86. Schuetz A, Allali-Hassani A, Martin F, Loppnau P,

Vedadi M, Bochkarev A, Plotnikov AN, Arrowsmith

CH, Min JR. 2006. Structural basis for molecular

recognition and presentation of histone H3 by WDR5.

EMBO J. 25: 4245-4252.

87. Scolnik Y, Portnaya I, Cogan U, Tal S, Haimovitz R,

Fridkin M, Elitzur AC, Deamer DW, Shinitzky M.

2006. Subtle differences in structural transitions

between poly-L- and poly-D-amino acids of equal

length in water. Phys. Chem. Chem. Phys. 8 :333-339.

88. Sharma VA, Kan E, Sun Y, Lian Y, Cisto J, Frasca V,

Hilt S, Stamatatos L, Donnelly JJ, Ulmer JB, Barnett

SW, Srivastava IK. 2006. Structural characteristics

correlate with immune responses induced by HIV

envelope glycoprotein vaccines. Virology 352: 131-144.

89. Shiroishi M, Kuroki K, Tsumoto K, Yokota A, Sasaki

T, Amano K, Shimojima T, Shirakihara Y, Rasubala L,

van der Merwe PA, Kumagai I, Kohda D, Maenaka K.

2006. Entropically driven MHC class I recognition by

human inhibitory receptor leukocyte Ig-like receptor B1

(LILRB1/ILT2/CD85j). J. Mol. Biol. 355: 237-248.

90. Shuman CF, Jiji R, Akerfeldt KS, Linse S. 2006.

Reconstitution of calmodulin from domains and

subdomains: Influence of target peptide. J. Mol. Biol.

358: 870-881.

91. Sitar T, Popowicz GM, Siwanowicz I, Huber R, Holak

TA. 2006. Structural basis for the inhibition of insulin-

like growth factors by insulin-like growth factor-

binding proteins. Proc. Natl. Acad. Sci. U. S. A. 103:

13028-13033.

92. Spadaccini R, Reidt U, Dybkov O, Will C, Frank R,

Stier G, Corsini L, Wahl MC, Luhrmann R, Sattler M.

2006. Biochemical and NMR analyses of an SF3b155-

p14-U2AF-RNA interaction network involved in

branch point definition during pre-mRNA splicing.

RNA. 12: 410-425.

93. Sudo Y, Yamabi M, Kato S, Hasegawa C, Iwamoto M,

Shimono K, Kamo N. 2006. Importance of specific

hydrogen bonds of archaeal rhodopsins for the binding

to the transducer protein. J. Mo.l Biol. 357: 1274-1282.

94. Tidow H, Veprintsev DB, Freund SMV, Fersht AR.

2006. Effects of oncogenic mutations and DNA

response elements on the binding of p53 to p53-binding

protein 2 (53bp2). .J Biol. Chem. 281: 32526-32533.

95. Tong KI, Katoh Y, Kusunoki H, Itoh K, Tanaka T,

Page 26: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

Yamamoto M. 2006. Keap1 recruits Neh2 through

binding to ETGE and DLG motifs: Characterization of

the two-site molecular recognition model. Mol. Cell

Biol. 26: 2887-2900.

96. Tso S-C, Kato M, Chuang JL, Chuang DT. 2006.

Structural determinants for cross-talk between pyruvate

dehydrogenase kinase 3 and lipoyl domain 2 of the

human pyruvate dehydrogenase complex. J. Biol.

Chem. 281: 27197-27204.

97. Velazquez-Campoy A. 2006. Ligand binding to one-

dimensional lattice-like macromolecules: Analysis of

the McGhee-von Hippel theory implemented in

isothermal titration calorimetry. Anal Biochem 348: 94-

104.

98. Velazquez-Campoy A, Goni G, Peregrina JR, Medina

M. 2006. Exact analysis of heterotropic interactions in

proteins: Characterization of cooperative ligand binding

by isothermal titration calorimetry. Biophys J. 91:

1887-1904.

99. Wang X, Baloh RH, Milbrandt J, Garcia KC. 2006.

Structure of artemin complexed with its receptor

GFRα3: Convergent recognition of glial cell line-

derived neurotrophic factors. Structure 14:1083-1092.

100. Wei J, Fain S, Harrison C, Feig LA, Baleja JD. 2006.

Molecular dissection of Rab11 binding from coiled-coil

formation in the Rab11-FIP2 C-terminal domain.

Biochemistry 45: 6826-6834.

101. Wilder PT, Lin J, Bair CL, Charpentier TH, Yang D,

Liriano M, Varney KM, Lee A, Oppenheim AB, Adhya

S, Carrier F, Weber DJ. 2006. Recognition of the tumor

suppressor protein p53 and other protein targets by the

calcium-binding protein S100b. Biochim. Biophys. Acta

- Molecular Cell Research. 1763: 1284-1297.

102. Woo JS, Imm JH, Min CK, Kim KJ, Cha SS, Oh BH. 2006.

Structural and functional insights into the B30.2/SPRY

domain. EMBO J. 25: 1353-1363.

103. Xu Y, Xing Y, Chen Y, Chao Y, Lin Z, Fan E, Yu JW,

Strack S, Jeffrey PD, Shi Y. 2006. Structure of the

protein phosphatase 2A holoenzyme. Cell 127: 1239-

1251.

104. Xue W-F, Szczepankiewicz O, C. Bauer M, Thulin E,

Linse S. 2006. Intra- versus intermolecular interactions

in monellin: Contribution of surface charges to protein

assembly. J. Mol. Biol. 358: 1244-1255.

105. Yakubovskaya E, Chen Z, Carrodeguas JA, Kisker C,

Bogenhagen DF. 2006. Functional human

mitochondrial DNA polymerase γ. forms a

heterotrimer. J. Biol. Chem. 281: 374-382.

106. Yamniuk AP, Ishida H, Vogel HJ. 2006. The

interaction between calcium- and integrin-binding

Protein 1 and the αIIb integrin cytoplasmic domain

involves a novel C-terminal displacement mechanism.

J. Biol. Chem. 281: 26455-26464.

107. Yao Z-P, Zhou M, Kelly SE, Seeliger MA, Robinson

CV, Itzhaki LS. 2006. Activation of ubiquitin ligase

SCFSkp2 by Cks1: Insights from hydrogen exchange

mass spectrometry. J. Mol. Biol. 363: 673-686.

108. Zaccheo OJ, Prince SN, Miller DM, Williams C, Fred

Kemp C, Brown J, Yvonne Jones E, Catto LE, Crump

MP, Bassim Hassan A. 2006. Kinetics of insulin-like

growth factor II (IGF-II) interaction with domain 11 of

the human IGF-II/mannose 6-phosphate receptor:

Function of CD and AB loop solvent-exposed residues.

J. Mol. Biol. 359: 403-421.

109. Zhao G, Zhou X, Wang L, Li G, Kisker C, Lennarz WJ,

Schindelin H. 2006. Structure of the mouse peptide N-

Glycanase-HR23 complex suggests co-evolution of the

endoplasmic reticulum-associated degradation and

DNA repair pathways. J. Biol. Chem. 281: 13751-

13761.

Protein-peptide

110. Andre I, Kesvatera T, Jonsson B, Linse S. 2006. Salt

enhances calmodulin-target interaction. Biophys. J. 90:

2903-2910.

111. Chan CS, Howell JM, Workentine ML, Turner RJ.

2006. Twin-arginine translocase may have a role in the

chaperone function of NarJ from Escherichia coli.

Biochem. Biophys. Res. Commun. 343: 244-251.

112. Chrencik JE, Brooun A, Recht MI, Kraus ML, Koolpe

M, Kolatkar AR, Bruce RH, Martiny-Baron G, Widmer

H, Pasquale EB, Kuhn P. 2006. Structure and

thermodynamic characterization of the EphB4/Ephrin-

B2 antagonist peptide complex reveals the determinants

for receptor specificity. Structure 14: 321-330.

113. Cosgrove MS, Bever K, Avalos JL, Muhammad S,

Zhang X, Wolberger C. 2006. The structural basis of

sirtuin substrate affinity. Biochemistry 45: 7511-7521.

114. Daum S, Fanghanel J, Wildemann D, Schiene-Fischer

C. 2006. Thermodynamics of phosphopeptide binding

to the human peptidyl prolyl cis/trans isomerase Pin1.

Biochemistry 45: 12125-12135.

115. Francois JA, Kumaran S, Jez JM. 2006. Structural basis

for interaction of o-acetylserine sulfhydrylase and

serine acetyltransferase in the arabidopsis cysteine

synthase complex. Plant Cell 18: 3647-3655.

116. Guilloreau L, Damian L, Coppel Y, Mazarguil H,

Winterhalter M, Faller P. 2006. Structural and

thermodynamical properties of CuII amyloid-beta16/28

complexes associated with Alzheimer's disease. J. Biol.

Inorg. Chem. 11: 1024-1038.

117. Harkiolaki M, Gilbert RJC, Jones EY, Feller SM. 2006.

The C-terminal SH3 domain of CRKL as a dynamic

Page 27: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

dimerization module transiently exposing a nuclear

export signal. Structure 14: 1741-1753.

118. He X-l, Dukkipati A, Garcia KC. 2006. Structural

determinants of natriuretic peptide receptor specificity

and degeneracy. J. Mol. Biol. 361: 698-714.

119. Hoelz A, Janz JM, Lawrie SD, Corwin B, Lee A,

Sakmar TP. 2006. Crystal structure of the sh3 domain

of β-PIX in complex with a high affinity peptide from

PAK2. J. Mol. Biol. 358: 509-522.

120. Hu J, Hubbard SR. 2006. Structural basis for

phosphotyrosine recognition by the src homology-2

domains of the adapter proteins SH2-B and APS. J.

Mol. Biol. 361: 69-79.

121. Hu M, Gu LC, Li MY, Jeffrey PD, Gu W, Shi YG.

2006. Structural basis of competitive recognition of p53

and MDM2 by HAUSP/USP7: Implications for the

regulation of the p53-MDM2 pathway. PloS Biol. 4:

228-239.

122. Jäger M, Zhang Y, Bieschke J, Nguyen H, Dendle M,

Bowman ME, Noel JP, Gruebele M, Kelly JW. 2006.

Structure-function-folding relationship in a WW

domain. Proc. Natl. Acad. Sci. U. S. A. 103:10648-

10653.

123. Katragadda M, Magotti P, Sfyroera G, Lambris JD.

2006. Hydrophobic effect and hydrogen bonds account

for the improved activity of a complement inhibitor,

compstatin. J. Med. Chem. 49:4616-4622.

124. Lee BJ, Cansizoglu AE, Suel KE, Louis TH, Zhang Z,

Chook YM. 2006. Rules for nuclear localization

sequence recognition by Karyopherin β2. Cell 126:

543-558.

125. Li MS, Hazelbauer GL. 2006. The carboxyl-terminal

linker is important for chemoreceptor function. Mol.

Microbiol. 60: 469-479.

126. Lim NS, Kozlov G, Chang TC, Groover O, Siddiqui N,

Volpon L, De Crescenzo G, Shyu AB, Gehring K.

2006. Comparative peptide binding studies of the

PABC domains from the ubiquitin-protein isopeptide

ligase HYD and poly(A)-binding protein - implications

for HYD function. J. Biol. Chem. 281: 14376-14382.

127. Liu J, Li M, Ran X, Fan J-s, Song J. 2006. Structural

insight into the binding diversity between the human

Nck2 SH3 domains and proline-rich proteins.

Biochemistry 45: 7171-7184.

128. Liu J, Xing Y, Hinds TR, Zheng J, Xu W. 2006. The

third 20 amino acid repeat is the tightest binding site of

APC for beta-catenin. J. Mol. Biol. 360:.133-144.

129. Lo SC, Li XC, Henzl MT, Beamer LJ, Hannink M.

2006. Structure of the Keap1 : Nrf2 interface provides

mechanistic insight into Nrf2 signaling. EMBO J. 25:

3605-3617.

130. Marlow MS, Wand AJ. 2006. Conformational

dynamics of calmodulin in complex with the

calmodulin-dependent kinase kinase alpha calmodulin-

binding domain. Biochemistry 45: 8732-8741.

131. Martinell M, Salvatella X, Fernandez-Carneado J,

Gordo S, Feliz M, Menendez M, Giralt E. 2006.

Synthetic ligands able to interact with the p53

tetramerization domain. Towards understanding a

protein surface recognition event. ChemBiochem. 7:

1105-1113.

132. Martinez-Sanz J, Yang A, Blouquit Y, Duchambon P,

Assairi L, Craescu CT. 2006. Binding of human centrin

2 to the centrosomal protein hSfi1. FEBS J. 273: 4504-

4515.

133. Ojida A, Inoue M-a, Mito-oka Y, Tsutsumi H, Sada K,

Hamachi I. 2006. Effective disruption of

phosphoprotein-protein surface interaction using Zn(II)

dipicolylamine-based artificial receptors via two-point

interaction. J. Am. Chem. Soc. 128: 2052-2058.

134. Patel B, Gingras AR, Bobkov AA, Fujimoto LM,

Zhang M, Liddington RC, Mazzeo D, Emsley J,

Roberts GCK, Barsukov IL, Critchley DR. 2006. The

activity of the vinculin binding sites in talin is

influenced by the stability of the helical bundles that

make up the talin rod. J. Biol. Chem. 281: 7458-7467.

135. Roufik S, Gauthier SF, Leng X, Turgeon SL. 2006.

Thermodynamics of binding interactions between

bovine beta-lactoglobulin A and the antihypertensive

peptide beta-Lg f142-148. Biomacromolecules 7: 419-

426.

136. Thickman KR, Swenson MC, Kabogo JM, Gryczynski

Z, Kielkopf CL. 2006. Multiple u2af65 binding sites

within sf3b155: Thermodynamic and spectroscopic

characterization of protein-protein interactions among

pre-mRNA splicing factors. J. Mol. Biol. 356: 664-683.

137. Wear MA, Walkinshaw MD. 2006. Thermodynamics of

the cyclophilin-A/cyclosporin-A interaction: A direct

comparison of parameters determined by surface

plasmon resonance using BIAcore T100 and isothermal

titration calorimetry. Anal. Biochem. 359: 285-287.

138. Winstone TL, Workentine ML, Sarfo KJ, Binding AJ,

Haslam BD, Turner RJ. 2006. Physical nature of signal

peptide binding to DmsD. Arch. Biochem. Biophys.

455: 89-97.

139. Woo J-S, Suh H-Y, Park S-Y, Oh B-H. 2006. Structural

basis for protein recognition by B30.2/SPRY domains.

Mol Cell. 24: 967-976.

140. Yang A, Miron S, Duchambon P, Assairi L, Blouquit

Y, Craescu CT. 2006. The N-terminal domain of human

centrin 2 has a closed structure, binds calcium with a

very low affinity, and plays a role in the protein self-

assembly. Biochemistry 45: 880-889.

141. Yu GW, Rudiger S, Veprintsev D, Freund S,

Page 28: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

Fernandez-Fernandez MR, Fersht AR. 2006. The

central region of HDM2 provides a second binding site

for p53. Proc. Natl. Acad. Sci. U. S. A. 103: 1227-1232.

Protein/peptide-small ligand

142. Aguilar RC, Longhi SA, Shaw JD, Yeh LY, Kim S,

Schon A, Freire E, Hsu A, McCormick WK, Watson

HA, Wendland B. 2006. Epsin N-terminal homology

domains perform an essential function regulating Cdc42

through binding Cdc42 GTPase-activating proteins.

Proc. Natl. Acad. Sc.i U. S. A. 103: 4116-4121.

143. Ahmad MF, Ramakrishna T, Raman B, Rao CM. 2006.

Fibrillogenic and non-fibrillogenic ensembles of SDS-

bound human alpha-synuclein. J. Mol. Biol. 364: 1061-

1072.

144. Andujar-Sanchez M, Martinez-Rodriguez S, Heras-

Vazquez FJL, Clemente-Jimenez JM, Rodriguez-Vico

F, Jara-Perez V. 2006. Binding studies of hydantoin

racemase from Sinorhizobium meliloti by calorimetric

and fluorescence analysis. Biochim. Biophys. Acta -

Proteins & Proteomics. 1764: 292-298.

145. Banerjee T, Kishore N. 2006. Binding of 8-

anilinonaphthalene sulfonate to dimeric and tetrameric

concanavalin A: Energetics and its implications on

saccharide binding studied by isothermal titration

calorimetry and spectroscopy. J. Phys. Chem. B 110:

7022-7028.

146. Banerjee T, Singh SK, Kishore N. 2006. Binding of

naproxen and amitriptyline to bovine serum albumin:

Biophysical aspects. J. Phys. Chem. B 110: 24147-

24156.

147. Baraquet C, Theraulaz L, Guiral M, Lafitte D, Mejean

V, Jourlin-Castelli C. 2006. TorT, a member of a new

periplasmic binding protein family, triggers induction

of the Tor respiratory system upon trimethylamine N-

oxide electron-acceptor binding in Escherichia coli. J.

Biol. Chem. 281: 38189-38199.

148. Barratt E, Bronowska A, Vondrasek J, Cerny J,

Bingham R, Phillips S, Homans SW. 2006.

Thermodynamic penalty arising from burial of a ligand

polar group within a hydrophobic pocket of a protein

receptor. J. Mol. Biol. 362: 994-1003.

149. Barrientos LG, Matei E, Lasala F, Delgado R,

Gronenborn AM. 2006. Dissecting carbohydrate-

cyanovirin-N binding by structure-guided mutagenesis:

Functional implications for viral entry inhibition. PEDS

19: 525-535.

150. Begg GE, Holman SR, Stokes PH, Matthews JM,

Graham RM, Iismaa SE. 2006. Mutation of a critical

arginine in the GTP-binding site of transglutaminase 2

disinhibits intracellular cross-linking activity. J. Biol.

Chem. 281: 12603-12609.

151. Bhor VM, Dev S, Vasanthakumar GR, Kumar P, Sinha

S, Surolia A. 2006. Broad substrate stereospecificity of

the Mycobacterium tuberculosis 7-keto-8-

aminopelargonic acid synthase: Spectroscopic and

kinetic studies. J. Biol. Chem. 281: 25076-25088.

152. Boraston AB, Healey M, Klassen J, Ficko-Blean E,

Lammerts van Bueren A, Law V. 2006. A structural

and functional analysis of alpha-glucan recognition by

family 25 and 26 carbohydrate-binding modules reveals

a conserved mode of starch recognition. J. Biol. Chem.

281: 587-598.

153. Borges JC, Ramos CHI. 2006. Spectroscopic and

thermodynamic measurements of nucleotide-induced

changes in the human 70-kDa heat shock cognate

protein. Arch. Biochem. Biophys. 452: 46-54.

154. Brandner B, Kurkela R, Vihko P, Kungl AJ. 2006.

Investigating the effect of VEGF glycosylation on

glycosaminoglycan binding and protein unfolding.

Biochem. Biophys. Res. Commun. 340: 836-839.

155. Bravman T, Bronner V, Lavie K, Notcovich A, Papalia

GA, Myszka DG. 2006. Exploring "One-shot" kinetics

and small molecule analysis using the ProteOn XPR36

array biosensor. Anal. Biochem. 358: 281-288.

156. Bunney TD, Harris R, Gandarillas NL, Josephs MB,

Roe SM, Sorli SC, Paterson HF, Rodrigues-Lima F,

Esposito D, Ponting CP, Gierschik P, Pearl LH,

Driscoll PC, Katan M. 2006. Structural and mechanistic

insights into Ras association domains of phospholipase

C epsilon. Mol. Cell 21: 495-507.

157. Buschiazzo A, Goytia M, Schaeffer F, Degrave W,

Shepard W, Grégoire C, Chamond N, Cosson A,

Berneman A, Coatnoan N, Alzari PM, Minoprio P.

2006. Crystal structure, catalytic mechanism, and

mitogenic properties of Trypanosoma cruzi proline

racemase. Proc. Natl. Acad. Sci. U. S. A. 103: 1705-

1710.

158. Buts L, Garcia-Pino A, Imberty A, Amiot N, Boons GJ,

Beeckmans S, Versees W, Wyns L, Loris R. 2006.

Structural basis for the recognition of complex-type

biantennary oligosaccharides by pterocarpus angolensis

lectin. FEBS J. 273: 2407-2420.

159. Calvo E, Mans BJ, Andersen JF, Ribeiro JMC. 2006.

Function and evolution of a mosquito salivary protein

family. J. Biol. Chem. 281: 1935-1942.

160. Celej MS, Dassie SA, Gonzalez M, Bianconi ML,

Fidelio GD. 2006. Differential scanning calorimetry as

a tool to estimate binding parameters in multiligand

binding proteins. Anal. Biochem. 350: 277-284.

161. Chavelas EA, Zubillaga RA, Pulido NO, Garcia-

Hernandez E. 2006. Multithermal titration calorimetry:

Page 29: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

A rapid method to determine binding heat capacities.

Biophys. Chem. 120: 10-14.

162. Chen M, Bridges A, Liu J. 2006. Determination of the

substrate specificities of N-acetyl-D-

glucosaminyltransferase. Biochemistry 45: 12358-

12365.

163. Cihlar T, He G-X, Liu X, Chen JM, Hatada M,

Swaminathan S, McDermott MJ, Yang Z-Y, Mulato

AS, Chen X, Leavitt SA, Stray KM, Lee WA. 2006.

Suppression of HIV-1 protease inhibitor resistance by

phosphonate-mediated solvent anchoring. J. Mol. Biol.

363: 635-647.

164. Cioci G, Mitchell EP, Chazalet V, Debray H, Oscarson

S, Lahmann M, Gautier C, Breton C, Perez S, Imberty

A. 2006. Beta-propeller crystal structure of Psathyrella

velutina lectin: An integrin-like fungal protein

interacting with monosaccharides and calcium. J. Mol.

Biol. 357: 1575-1591.

165. Ciulli A, Williams G, Smith AG, Blundell TL, Abell C.

2006. Probing hot spots at protein-ligand binding sites:

A fragment-based approach using biophysical methods.

J. Med. Chem. 49: 4992-5000.

166. Davis R, Bienvenue D, Swierczek SI, Gilner DM,

Rajagopal L, Bennett B, Holz RC. 2006. Kinetic and

spectroscopic characterization of the E134A- and

E134D-altered dapE-encoded N-succinyl-L,L-

diaminopimelic acid desuccinylase from Haemophilus

influenzae. J. Biol. Inorg. Chem. 11: 206-216.

167. Deka RK, Brautigam CA, Yang XF, Blevins JS,

Machius M, Tomchick DR, Norgard MV. 2006. The

PnrA (Tp0319; TmpC) lipoprotein represents a new

family of bacterial purine nucleoside receptor encoded

within an ATP-binding cassette (ABC)-like operon in

Treponema pallidum. J. Biol. Chem. 281: 8072-8081.

168. Denadai AML, Santoro MM, Lopes MTP, Chenna A,

de Sousa FB, Avelar GM, Gomes MRT, Guzman F,

Salas CE, Sinisterra RD. 2006. A supramolecular

complex between proteinases and beta-cyclodextrin that

preserves enzymatic activity: Physicochemical

characterization. Biodrugs 20: 283-291.

169. Devenport M, Alvarenga PH, Shao L, Fujioka H,

Bianconi ML, Oliveira PL, Jacobs-Lorena M. 2006.

Identification of the Aedes aegypti peritrophic matrix

protein AeIMUCI as a heme-binding protein.

Biochemistry 45: 9540-9549.

170. Diaz-Benjumea R, Laxman S, Hinds T, Beavo JA,

Rascon A. 2006. Characterization of a novel c-AMP-

specific cyclic nucleotide phosphodiestersa

(TrcPDEB1) from Trypanosoma cruzi. Biochem. J.

399: 305-314.

171. Egloff MP, Malet H, Putics A, Heionen M, Dutarte H,

Frangeul A, Gruez A, Campanacci V, Cambilau C,

Ziebhur J, Ahola T, Canardl B. 2006. Stuctural and

functional basis for ADP-ribose and poly(ADP-ribose)

binding by viral macro domains. J. Virol. 80:8493-

8502.

172. Engel M, Hindie V, Lopez-Garcia LA, Stroba A,

Schaeffer F, Adrian I, Imig J, Idrissova L, Nastainczyk

W, Zeuzem S, Alzari PM, Hartmann RW, Piiper A,

Biondi RM. 2006. Allosteric activation of the protein

kinase Pdk1 with low molecular weight compounds.

EMBO J. 25: 5469-5480.

173. Ficko-Blean E, Boraston AB. 2006. The interaction of a

carbohydrate-binding module from a Clostridium

perfringens N-acetyl-beta-hexosaminidase with its

carbohydrate receptor. J. Biol. Chem. 281: 37748-

37757.

174. Frazier RA, Papadopoulou A, Green RJ. 2006.

Isothermal titration calorimetry study of epicatechin

binding to serum albumin. J. Pharmaceut. Biomed.

Anal. 41: 1602-1605.

175. Freire E. 2006. Overcoming HIV-1 resistance to

protease inhibitors. Drug Discovery Today. 3: 281-286.

176. Gasper R, Scrima A, Wittinghofer A. 2006. Structural

insights into HypB, a GTP-binding protein that

regulates metal binding. J. Biol. Chem. 281: 27492-

27502.

177. Golovko MY, Rosenberger TA, Fargeman NJ,

Feddersen S, Cole NB, Pribill I, Berger J, Nussbaum

RL, Murphy EJ. 2006. Acyl-CoA synthetase activity

links wild-type but not mutant α-synuclein to brain

arachidonate metabolism. Biochemistry 45: 6956-6966.

178. Gopalakrishnapai J, Gupta G, Karthikeyan T, Sinha S,

Kandiah E, Gemma E, Oscarson S, Surolia A. 2006.

Isothermal titration calorimetric study defines the

substrate binding residues of calreticulin. Biochem.

Biophys. Res. Commun. 351: 14-20.

179. Grunau A, Paine MJ, Ladbury JE, Gutierrez A. 2006.

Global effects of the energetics of coenzyme binding:

NADPH controls the protein interaction properties of

human cytochrome p450 reductase. Biochemistry 45:

1421-1434.

180. Gupta S, Das L, Datta AB, Poddar A, Janik ME,

Bhattacharyya B. 2006. Oxalone and lactone moieties

of podophyllotoxin exhibit properties of both the B and

C rings of colchicine in its binding with tubulin.

Biochemistry 45: 6467-6475.

181. Hajjar E, Perahia D, Debat H, Nespoulous C, Robert

CH. 2006. Odorant binding and conformational

dynamics in the odorant-binding protein. J. Biol. Chem.

281: 29929-29937.

182. Hara T, Durell SR, Myers MC, Appella DH. 2006.

Probing the structural requirements of peptoids that

inhibit HDM2-p53 interactions. J. Am. Chem. Soc. 128:

Page 30: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

1995-2004.

183. Harnsilawat T, Pongsawatmanit R, McClements DJ.

2006. Characterization of β-lactoglobulin-sodium

alginate interactions in aqueous solutions: A

calorimetry, light scattering, electrophoretic mobility

and solubility study. Food Hydrocoll. 20: 577-585.

184. Henshaw J, Horne-Bitschy A, van Bueren AL, Money

VA, Bolam DN, Czjzek M, Ekborg NA, Weiner RM,

Hutcheson SW, Davies GJ, Boraston AB, Gilbert HJ.

2006. Family 6 carbohydrate binding modules in beta-

agarases display exquisite selectivity for the non-

reducing termini of agarose chains. J. Biol. Chem. 281:

17099-17107.

185. Hori T, Ishijima J, Yokomizo T, Ago H, Shimizu T,

Miyano M. 2006. Crystal structure of anti-configuration

of indomethacin and leukotriene B4 12-

hydroxydehydrogenase/15-oxo-prostaglandin 13-

reductase complex reveals the structural basis of broad

spectrum indomethacin efficacy. J. Biochem (Tokyo).

140: 457-466.

186. Hyre DE, Le Trong I, Merritt EA, Eccleston JF, Green

NM, Stenkamp RE, Stayton PS. 2006. Cooperative

hydrogen bond interactions in the streptavidin-biotin

system. Protein Sci. 15:459-467.

187. Isin EM, Guengerich FP. 2006. Kinetics and

thermodynamics of ligand binding by cytochrome P450

3A4. J. Biol. Chem. 281: 9127-9136.

188. Itoh Y, Watanabe J, Fukada H, Mizuno R, Kezuka Y,

Nonaka T, Watanabe T. 2006. Importance of Trp59 and

Trp60 in chitin-binding, hydrolytic, and antifungal

activities of Streptomyces griseus chitinase c. Appl.

Microbiol. Biotechnol. 72: 1176-1184.

189. Jao S-C, Chen J, Yang K, Li W-S. 2006. Design of

potent inhibitors for Schistosoma japonica glutathione

s-transferase. Bioorg. Med. Chem. 14: 304-318.

190. Jimenez M, Andre S, Siebert H-C, Gabius H-J, Solis D.

2006. AB-type lectin (toxin/agglutinin) from mistletoe:

Differences in affinity of the two galactoside-binding

Trp/Tyr-sites and regulation of their functionality by

monomer/dimer equilibrium. Glycobiol. 16: 926-937.

191. Joly N, Schumacher J, Buck M. 2006. Heterogeneous

nucleotide occupancy stimulates functionality of phage

shock protein F, an AAA+ transcriptional activator. J.

Biol. Chem. 281: 34997-35007.

192. Karsten WE, Cook PF. 2006. An isothermal titration

calorimetry study of the binding of substrates and

ligands to the tartrate dehydrogenase from

Pseudomonas putida reveals half-of-the-sites reactivity.

Biochemistry 45: 9000-9006.

193. Kasimova MR, Grigiene J, Krab K, Hagedorn PH,

Flyvbjerg H, Andersen PE, Moller IM. 2006. The free

NADH concentration is kept constant in plant

mitochondria under different metabolic conditions.

Plant Cell. 18: 688-698.

194. Kasper C, Pickering DS, Mirza O, Olsen L, Kristensen

AS, Greenwood JR, Liljefors T, Schousboe A, Watjen

F, Gajhede M, Sigurskjold BW, Kastrup JS. 2006. The

structure of a mixed GluR2 ligand-binding core dimer

in complex with (S)-glutamate and the antagonist (S)-

NS1209. J. Mol. Biol. 357: 1184-1201.

195. Kathir KM, Kumar TKS, Yu C. 2006. Understanding

the mechanism of the antimitogenic activity of suramin.

Biochemistry 45: 899-906.

196. Kavanagh KL, Guo K, Dunford JE, Wu X, Knapp S,

Ebetino FH, Rogers MJ, Russell RGG, Oppermann U.

2006. The molecular mechanism of nitrogen-containing

bisphosphonates as antiosteoporosis drugs. Proc. Natl.

Acad. Sci. U. S. A. 103: 7829-7834.

197. Keramisanou D, Biris N, Gelis I, Sianidis G,

Karamanou S, Economou A, Kalodimos CG. 2006.

Disorder-order folding transitions underlie catalysis in

the helicase motor of SecA. Nature Struct. Mol. Biol.

13: 594-602.

198. Kim M-K, Lee JH, Kim H, Park SJ, Kim SH, Kang GB,

Lee YS, Kim JB, Kim KK, Suh SW, Eom SH. 2006.

Crystal structure of Visfatin/pre-B cell colony-

enhancing factor 1/nicotinamide

phosphoribosyltransferase, free and in complex with the

anti-cancer agent FK-866. J. Mol. Biol. 362: 66-77.

199. Kraut DA, Sigala PA, Pybus B, Liu CW, Ringe D,

Petsko GA, Herschlag D. 2006. Testing electrostatic

complementarity in enzyme catalysis: Hydrogen

bonding in the ketosteroid isomerase oxyanion hole.

PLoS Biol. 4: e99.

200. Krishna R, Manjunath GP, Kumar P, Surolia A,

Chandra NR, Muniyappa K, Vijayan M. 2006.

Crystallographic identification of an ordered C-terminal

domain and a second nucleotide-binding site in RecA:

New insights into allostery. Nucleic Acids Res. 34:

2186-2195.

201. Lacal J, Busch A, Guazzaroni M-E, Krell T, Ramos JL.

2006. The TodS-TodT two-component regulatory

system recognizes a wide range of effectors and works

with DNA-bending proteins. Proc. Natl. Acad. Sci. U.

S. A. 103: 8191-8196.

202. Lawrence SH, Luther KB, Schindelin H, Ferry JG.

2006. Structural and functional studies suggest a

catalytic mechanism for the phosphotransacetylase

from Methanosarcina thermophila. J. Bacteriol. 188:

1143-1154.

203. Li X, Chow D-C, Tu S-C. 2006. Thermodynamic

analysis of the binding of oxidized and reduced FMN

cofactor to Vibrio harveyi NADPH-FMN

oxidoreductase FRP apoenzyme. Biochemistry 45:

Page 31: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

14781-14787.

204. Li X, Thompson KSJ, Godber B, Cooper MA. 2006.

Quantification of small molecule-receptor affinities and

kinetics by acoustic profiling. Assay & Drug Develop

Technol. 4:565-573.

205. Lim J-H, Kim M-S, Kim H-E, Yano T, Oshima Y,

Aggarwal K, Goldman WE, Silverman N, Kurata S, Oh

B-H. 2006. Structural basis for preferential recognition

of diaminopimelic acid-type peptidoglycan by a subset

of peptidoglycan recognition proteins. J. Biol. Chem.

281: 8286-8295.

206. Lin C-H, Huang C-F, Chen W-Y, Chang Y-Y, Ding W-

H, Lin M-S, Wu S-H, Huang R-N. 2006.

Characterization of the interaction of galectin-1 with

sodium arsenite. Chem. Res. Toxicol. 19: 469-474.

207. Lund H, Christensen BP, Nielsen AD, Westh P. 2006.

Proton exchange coupled to the specific binding of

alkylsulfonates to serum albumins. Biochim. Biophys.

Acta - Proteins & Proteomics. 1764: 1243-1251.

208. McCartney L, Blake AW, Flint J, Bolam DN, Boraston

AB, Gilbert HJ, Knox JP. 2006. Differential recognition

of plant cell walls by microbial xylan-specific

carbohydrate-binding modules. Proc. Natl. Acad. Sci.

U. S. A. 103: 4765-4770.

209. McClard RW, Holets EA, MacKinnon AL, Witte JF.

2006. Half-of-sites binding of orotidine 5'-phosphate

and alpha-d-5-phosphorylribose 1-diphosphate to

orotate phosphoribosyltransferase from Saccharomyces

cerevisiae supports a novel variant of the Theorell-

Chance mechanism with alternating site catalysis.

Biochemistry 45: 5330-5342.

210. McElroy CA, Manfredo A, Gollnick P, Foster MP.

2006. Thermodynamics of tryptophan-mediated

activation of the Trp RNA-binding attenuation protein.

Biochemistry 45: 7844-7853.

211. McLaughlin SH, Sobott F, Yao Z-p, Zhang W, Nielsen

PR, Grossmann JG, Laue ED, Robinson CV, Jackson

SE. 2006. The co-chaperone p23 arrests the Hsp90

ATPase cycle to Trap client proteins. J. Mol. Biol. 356:

746-758.

212. Melkko S, Dumelin CE, Scheuermann J, Neri D. 2006.

On the magnitude of the chelate effect for the

recognition of proteins by pharmacophores scaffolded

by self-assembling oligonucleotides. Chemistry &

Biology. 13: 225-231.

213. Milano SK, Kim Y-M, Stefano FP, Benovic JL,

Brenner C. 2006. Nonvisual arrestin oligomerization

and cellular localization are regulated by inositol

hexakisphosphate binding. J. Biol. Chem. 281: 9812-

9823.

214. Mintzer E, Sargsyan H, Bittman R. 2006.

Lysophosphatidic acid and lipopolysaccharide bind to

the PIP2-binding domain of gelsolin. Biochim. Biophys.

Acta - Biomembranes. 1758: 85-89.

215. Mitkevich VA, Kononenko AV, Petrushanko IY,

Yanvarev DV, Makarov AA, Kisselev LL. 2006.

Termination of translation in eukaryotes is mediated by

the quaternary eRF1*eRF3*GTP*Mg2+ complex. The

biological roles of eRF3 and prokaryotic RF3 are

profoundly distinct. Nucleic Acids Res. 34: 3947-3954.

216. Miyanaga A, Koseki T, Miwa Y, Mese Y, Nakamura S,

Kuno A, Hirabayashi J, Matsuzawa H, Wakagi T,

Shoun H, Fushinobu S. 2006. The family 42

carbohydrate-binding module of family 54 alpha-L-

arabinofuranosidase specifically binds the

arabinofuranose side chain of hemicellulose. Biochem.

J. 399: 503-511.

217. Moinuddin SGA, Youn B, Bedgar DL, Costa MA,

Helms GL, Kang C, Davin LB, Lewis NG. 2006.

Secoisolariciresinol dehydrogenase: Mode of catalysis

and stereospecificity of hydride transfer in

podophyllum peltatum. Org. Biomol. Chem. 4: 808-

816.

218. Moll D, Prinz A, Gesellchen F, Drewianka S,

Zimmermann B, Herberg FW. 2006. Biomolecular

interaction analysis in functional proteomics. J. Neural.

Trans. 113: 1015-1032.

219. Moosavi-Movahedi Z, Safarian S, Zahedi M, Sadeghi

M, Saboury AA, Chamani J, Bahrami H, Ashraf-

Modarres A, Moosavi-Movahedi AA. 2006.

Calorimetric and binding dissections of HSA upon

interaction with bilirubin. Protein J. 25: 193-201.

220. Morgunova E, Illarionov B, Sambaiah T, Haase I,

Bacher A, Cushman M, Fischer M, Ladenstein R. 2006.

Structural and thermodynamic insights into the binding

mode of five novel inhibitors of lumazine synthase

from Mycobacterium tuberculosis. FEBS J. 273: 4790-

4804.

221. Mueller-Dieckmann C, Kernstock S, Lisurek M, von

Kries JP, Haag F, Weiss MS, Koch-Nolte F. 2006. The

structure of human ADP-ribosylhydrolase 3 (ARH3)

provides insights into the reversibility of protein ADP-

ribosylation. Proc. Natl. Acad. Sci. U. S. A. 103: 15026-

15031.

222. Muralidhara BK, Negi S, Chin CC, Braun W, Halpert

JR. 2006. Conformational flexibility of mammalian

cytochrome P450 2B4 in binding imidazole inhibitors

with different ring chemistry and side chains. Solution

thermodynamics and molecular modeling. J. Biol.

Chem. 281: 8051-8061.

223. Nie Y, Smirnova I, Kasho V, Kaback HR. 2006.

Energetics of ligand-induced conformational flexibility

in the lactose permease of Escherichia coli. J. Biol.

Chem. 281: 35779-35784.

Page 32: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

224. Ozen C, Malek JM, Serpersu EH. 2006. Dissection of

aminoglycoside-enzyme interactions: A calorimetric

and NMR study of neomycin B binding to the

aminoglycoside phosphotransferase (3')-IIIa. J. Am.

Chem. Soc. 128: 15248-15254.

225. Padovani D, Banerjee R. 2006. Assembly and

protection of the radical enzyme, methylmalonyl-CoA

mutase, by its chaperone. Biochemistry 45: 9300-9306.

226. Pasternak O, Bujacz GD, Fujimoto Y, Hashimoto Y,

Jelen F, Otlewski J, Sikorski MM, Jaskolski M. 2006.

Crystal structure of Vigna radiata cytokinin-specific

binding protein in complex with zeatin. Plant Cell. 18:

2622-2634.

227. Perry JL, Goldsmith MR, Williams TR, Radack KP,

Christensen T, Gorham J, Pasquinelli MA, Toone EJ,

Beratan DN, Simon JD. 2006. Binding of warfarin

influences the acid-base equilibrium of H242 in Sudlow

site I of human serum albumin. Photochem. Photobiol.

82: 1365-1369.

228. Pey AL, Martinez A, Charubala R, Maitland DJ, Teigen

K, Calvo A, Pfleiderer W, Wood JM, Schallreuter KU.

2006. Specific interaction of the diastereomers 7(R)-

and 7(S)-tetrahydrobiopterin with phenylalanine

hydroxylase: Implications for understanding

primapterinuria and vitiligo. FASEB J. 20: 2130-2132.

229. Pokorna M, Cioci G, Perret S, Rebuffet E, Kostlanova

N, Adam J, Gilboa-Garber N, Mitchell EP, Imberty A,

Wimmerova M. 2006. Unusual entropy-driven affinity

of Chromobacterium violaceum lectin CV-IIL toward

fucose and mannose. Biochemistry 45: 7501-7510.

230. Pop SM, Gupta N, Raza AS, Ragsdale SW. 2006.

Transcriptional activation of dehalorespiration:

Identification of redox-active cysteines regulating

dimerization and DNA binding. J. Biol. Chem. 281:

26382-26390.

231. Pretz MG, Albers S-V, Schuurman-Wolters G, Tampe

R, Driessen AJM, van der Does C. 2006.

Thermodynamics of the ATPase cycle of GlcV, the

nucleotide-binding domain of the glucose ABC

transporter of Sulfolobus solfataricus. Biochemistry 45:

15056-15067.

232. Qin J, Chai G, Brewer JM, Lovelace LL, Lebioda L.

2006. Fluoride inhibition of enolase: Crystal structure

and thermodynamics. Biochemistry 45: 793-800.

233. Rappl C, Barbier P, Bourgarel-Rey V, Gregoire C, Gilli

R, Carre M, Combes S, Finet JP, Peyrot V. 2006.

Interaction of 4-arylcoumarin analogues of

combretastatins with microtubule network of HBL100

cells and binding to tubulin. Biochemistry 45: 9210-

9218.

234. Roy S, Katayama D, Dong A, Kerwin BA, Randolph

TW, Carpenter JF. 2006. Temperature dependence of

benzyl alcohol- and 8-anilinonaphthalene-1-sulfonate-

induced aggregation of recombinant human interleukin-

1 receptor antagonist. Biochemistry 45: 3898-3911.

235. Sabin C, Mitchell EP, Pokorna M, Gautier C, Utille J-P,

Wimmerova M, Imberty A. 2006. Binding of different

monosaccharides by lectin PA-IIL from Pseudomonas

aeruginosa: Thermodynamics data correlated with x-

ray structures. FEBS Lett. 580: 982-987.

236. Sanchez J-F, Lescar J, Chazalet V, Audfray A, Gagnon

J, Alvarez R, Breton C, Imberty A, Mitchell EP. 2006.

Biochemical and structural analysis of Helix pomatia

agglutinin. A hexameric lectin with a novel fold. J.

Biol. Chem. 281: 20171-20180.

237. Schon A, Madani N, Klein JC, Hubicki A, Ng D, Yang

X, Smith AB, Sodroski J, Freire E. 2006.

Thermodynamics of binding of a low-molecular-weight

CD4 mimetic to HIV-1 gp120. Biochemistry 45: 10973-

10980.

238. Schrift GL, Waldron TT, Timmons MA, Ramaswamy

S, Kearney WR, Murphy KP. 2006. Molecular basis for

nucleotide-binding specificity: Role of the exocyclic

amino group "N2" In recognition by a guanylyl-

ribonuclease. J. Mol. Biol. 355: 72-84.

239. Schujman GE, Guerin M, Buschiazzo A, Schaeffer F,

Llarull LI, Reh G, Vila AJ, Alzari PM, Mendoza D.

2006. Structural basis of lipid biosynthesis regulation in

gram-positive bacteria. EMBO J. 25: 4074-4083.

240. Sengupta K, Kamdar RP, D'Souza JS, Mustafi SM, Rao

BJ. 2006. GTP-induced conformational changes in

translin: A comparison between human and Drosophila

proteins. Biochemistry 45: 861-870.

241. Singh SK, Kishore N. 2006. Elucidating the binding

thermodynamics of 8-anilino-1-naphthalene sulfonic

acid with the A-state of alpha-lactalbumin: An

isothermal titration calorimetric investigation.

Biopolymers 83: 205-212.

242. Singh SK, Kishore N. 2006. Thermodynamic insights

into the binding of Triton X-100 to globular proteins: A

calorimetric and spectroscopic investigation. J. Phys.

Chem. B 110: 9728-9737.

243. Sobhany M, Negishi M. 2006. Characterization of

specific donor binding to alpha-1,4-N-

acetylhexosaminyltransferase EXTL2 using isothermal

titration calorimetry. Meth Enzymol. 416: 3-12.

244. Steuber H, Zentgraf M, Podjarny A, Heine A, Klebe G.

2006. High-resolution crystal structure of aldose

reductase complexed with the novel sulfonyl-

pyridazinone inhibitor exhibiting an alternative active

site anchoring group. J. Mol. Biol. 356: 45-56.

245. Swaminathan CP, Brown PH, Roychowdhury A, Wang

Q, Guan R, Silverman N, Goldman WE, Boons G-J,

Mariuzza RA. 2006. Dual strategies for peptidoglycan

Page 33: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

discrimination by peptidoglycan recognition proteins

(PGRPs). Proc. Natl. Acad. Sci. U. S. A. 103: 684-689.

246. Taurog RE, Jakubowski H, Matthews RG. 2006.

Synergistic, random sequential binding of substrates in

cobalamin-independent methionine synthase.

Biochemistry 45: 5083-5091.

247. Tellez-Sanz R, Cesareo E, Nuccetelli M, Aguilera AM,

Baron C, Parker LJ, Adams JJ, Morton CJ, Lo Bello M,

Parker MW, Garcia-Fuentes L. 2006. Calorimetric and

structural studies of the nitric oxide carrier S-

nitrosoglutathione bound to human glutathione

transferase P1-1. Protein Sci. 15: 1093-1105.

248. Teran W, Krell T, Ramos JL, Gallegos M-T. 2006.

Effector-repressor interactions, binding of a single

effector molecule to the operator-bound ttgr homodimer

mediates derepression. J. Biol. Chem. 281: 7102-7109.

249. Toke O, Monsey JD, DeKoster GT, Tochtrop GP, Tang

C, Cistola DP. 2006. Determinants of cooperativity and

site selectivity in human ileal bile acid binding protein.

Biochemistry 45:727-737.

250. Trosset J-Y, Dalvit C, Knapp S, Fasolini M, Veronesi

M, Mantegani S, Gianellini LM, Catana C, Sundstrom

M, Stouten PFW, Moll JK. 2006. Inhibition of protein-

protein interactions: The discovery of druglike beta-

catenin inhibitors by combining virtual and biophysical

screening. Proteins 64: 60-67.

251. Wright E, Serpersu EH. 2006. Molecular determinants

of affinity for aminoglycoside binding to the

aminoglycoside nucleotidyltransferase(2'')-Ia.

Biochemistry 45: 10243-10250.

252. Xi J, Liu R, Rossi MJ, Yang J, Loll PJ, Dailey WP,

Eckenhoff RG. 2006. Photoactive analogues of the

haloether anesthetics provide high-resolution features

from low-affinity interactions. ACS Chem Biol. 1: 377-

384.

253. Ye M, English AM. 2006. Binding of

polyaminocarboxylate chelators to the active-site

copper inhibits the GSNO-reductase activity but not the

superoxide dismutase activity of Cu,Zn-superoxide

dismutase. Biochemistry 45: 12723-12732.

254. Yu Y, Sweeney MD, Saad OM, Leary JA. 2006.

Potential inhibitors of chemokine function: Analysis of

noncovalent complexes of CC chemokine and small

polyanionic molecules by ESI FT-ICR mass

spectrometry. J. Am. Soc. Mass. Spectrom. 17: 524-535.

255. Yuan C, Rieke CJ, Rimon G, Wingerd BA, Smith WL.

2006. Partnering between monomers of

cyclooxygenase-2 homodimers. Proc. Natl. Acad. Sci.

U. S. A. 103: 6142-6147.

256. Zhao Y, White MA, Muralidhara BK, Sun L, Halpert

JR, Stout CD. 2006. Structure of microsomal

cytochrome P450 2B4 complexed with the antifungal

drug bifonazole: Insight into P450 conformational

plasticity and membrane interaction. J. Biol. Chem.

281: 5973-5981.

257. Zimmermann K, Heck M, Frank J, Kern J, Vass I,

Zouni A. 2006. Herbicide binding and thermal stability

of photosystem II isolated from Thermosynechococcus

elongatus. Biochim. Biophys. Acta. 1757: 106-114.

Protein/peptide-metal

258. Babini E, Bertini I, Capozzi F, Chirivino E, Luchinat C.

2006. A structural and dynamic characterization of the

EF-hand protein CLSP. Structure 14: 1029-1038.

259. Bevers LE, Hagedoorn P-L, Krijger GC, Hagen WR.

2006. Tungsten transport protein a (WtpA) in

Pyrococcus furiosus: The first member of a new class

of tungstate and molybdate transporters. J. Bacteriol.

188: 6498-6505.

260. Charbonnier J-B, Christova P, Shosheva A, Stura E, Le

Du MH, Blouquit Y, Duchambon P, Miron S, Craescu

CT. 2006. Crystallization and preliminary x-ray

diffraction data of the complex between human centrin

2 and a peptide from the protein XPC. Acta Crystallogr.

F. 62: 649-651.

261. Choi DW, Do YS, Zea CJ, McEllistrem MT, Lee S-W,

Semrau JD, Pohl NL, Kisting CJ, Scardino LL, Hartsel

SC, Boyd ES, Geesey GG, Riedel TP, Shafe PH,

Kranski KA, Tritsch JR, Antholine WE, DiSpirito AA.

2006. Spectral and thermodynamic properties of Ag(I),

Au(III), Cd(II), Co(II), Fe(III), Hg(II), Mn(II), Ni(II),

Pb(II), U(IV), and Zn(II) binding by methanobactin

from Methylosinus trichosporium OB3b. J. Inorg.

Biochem. 100: 2150-2161.

262. Conner SH, Kular G, Peggie M, Shepherd S,

Schuttelkopf AW, Cohen P, Van Aalten DMF. 2006.

Tak1-binding protein 1 is a pseudophosphatase.

Biochem. J. 399: 427-434.

263. Contarino MR, Sergi M, Harrington AE, Lazareck A,

Xu J, Chaiken I. 2006. Modular, self-assembling

peptide linkers for stable and regenerable carbon

nanotube biosensor interfaces. J. Mol. Recognit. 19:

363-371.

264. Couture J-F, Hauk G, Thompson MJ, Blackburn GM,

Trievel RC. 2006. Catalytic roles for carbon-oxygen

hydrogen bonding in set domain lysine

methyltransferases. J. Biol. Chem. 281: 19280-19287.

265. Divsalar A, Saboury AA, Moosavi-Movahedi AA.

2006. Conformational and structural analysis of bovine

beta lactoglobulin-A upon interaction with Cr+3.

Protein J. 25: 157-165.

266. Erickson JR, Moerland TS. 2006. Functional

characterization of parvalbumin from the arctic cod

Page 34: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

(Boreogadus saida): Similarity in calcium affinity

among parvalbumins from polar teleosts. Comp.

Biochem. Physiol. A 143: 228-233.

267. Gerhardt WW, Weck M. 2006. Investigations of metal-

coordinated peptides as supramolecular synthons. J.

Org. Chem. 71: 6333-6341.

268. Grossoehme NE, Akilesh S, Guerinot ML, Wilcox DE.

2006. Metal-binding thermodynamics of the histidine-

rich sequence from the metal-transport protein IRT1

Arabidopsis haliana. Inorg. Chem. 45: 8500-8508.

269. Henzl MT, Agah S. 2006. Divalent ion-binding

properties of the two avian beta-parvalbumins. Proteins

62: 270-278.

270. Hilge M, Aelen J, Vuister GW. 2006. Ca2+ regulation in

the Na+/Ca2+ exchanger involves two markedly different

Ca2+ sensors. Mol. Cel. 22: 15-25.

271. Kittleson JT, Loftin IR, Hausrath AC, Engelhardt KP,

Rensing C, McEvoy MM. 2006. Periplasmic metal-

resistance protein CusF exhibits high affinity and

specificity for both CuI and AgI. Biochemistry 45:

11096-11102.

272. Kliegman JI, Griner SL, Helmann JD, Brennan RG,

Glasfeld A. 2006. Structural basis for the metal-

selective activation of the manganese transport

regulator of Bacillus subtilis. Biochemistry 45: 3493-

3505.

273. Lamb HK, Mee C, Xu W, Liu L, Blond S, Cooper A,

Charles IG, Hawkins AR. 2006. The affinity of a major

Ca2+ binding site on Grp78 is differentially enhanced by

ADP and ATP. J. Biol. Chem. 281: 8796-8805.

274. Le TQ, Gochin M, Featherstone JDB, Li W, DenBesten

PK. 2006. Comparative calcium binding of leucine-rich

amelogenin peptide and full-length amelogenin. Eur. J.

Oral Sci. 114 Suppl 1: 320-326.

275. Martinez-Rodriguez S, Andujar-Sanchez M, Clemente

Jimenez JM, Jara-Perez V, Rodriguez-Vico F, Heras-

Vazquez FJL. 2006. Thermodynamic and mutational

studies of l-N-carbamoylase from Sinorhizobium

meliloti CECT 4114 catalytic centre. Biochimie. 88:

837-847.

276. Mikhaylova M, Sharma Y, Reissner C, Nagel F,

Aravind P, Rajini B, Smalla K-H, Gundelfinger ED,

Kreutz MR. 2006. Neuronal Ca2+ signaling via

caldendrin and calneurons. Biochim. Biophys. Acta -

Molecular Cell Research. 1763: 1229-1237.

277. Park JH, Pulvermuller A, Scheerer P, Rausch S, Giel A,

Hohne W, Wolfrum U, Hofmann KP, Ernst OP, Choe

H-W, Krau N. 2006. Insights into functional aspects of

centrins from the structure of N-terminally extended

mouse centrin 1. Vision Res. 46: 4568-4574.

278. Park S-Y, Borbat PP, Gonzalez-Bonet G, Bhatnagar J,

Pollard AM, Freed JH, Bilwes AM, Crane BR. 2006.

Reconstruction of the chemotaxis receptor-kinase

assembly. Nature Struct. Mol. Biol. 13: 400-407.

279. Rangarajan ES, Proteau A, Wagner J, Hung M-N,

Matte A, Cygler M. 2006. Structural snapshots of

Escherichia coli histidinol phosphate phosphatase along

the reaction pathway. J. Biol. Chem. 281: 37930-37941.

280. Roos G, Buts L, Van Belle K, Brosens E, Geerlings P,

Loris R, Wyns L, Messens J. 2006. Interplay between

ion binding and catalysis in the thioredoxin-coupled

arsenate reductase family. J. Mol. Biol. 360: 826-838.

281. Saboury AA, Atri MS, Sanati MH, Moosavi-Movahedi

AA, Hakimelahi GH, Sadeghi M. 2006. A

thermodynamic study on the interaction between

magnesium ion and human growth hormone.

Biopolymers. 81: 120-126.

282. Sanchez-Bautista S, Marin-Vicente C, Gomez-

Fernandez JC, Corbalan-Garcia S. 2006. The C2

domain of PKCα is a Ca2+ -dependent PtdIns(4,5)P2

sensing domain: A new insight into an old pathway. J.

Mol. Biol. 362: 901-914.

283. Sheckler LR, Henry L, Sugita S, Sudhof TC, Rudenko

G. 2006. Crystal structure of the second LNS/LG

domain from neurexin 1α: Ca2+ binding and the effects

of alternative splicing. J. Biol. Chem. 281: 22896-

22905.

284. Sivaraja V, Kumar TKS, Rajalingam D, Graziani I,

Prudovsky I, Yu C. 2006. Copper binding affinity of

S100A13, a key component of the FGF-1 nonclassical

copper-dependent release complex. Biophys. J. 91:

1832-1843.

285. Streiff JH, Allen TW, Atanasova E, Juranic N, Macura

S, Penheiter AR, Jones KA. 2006. Prediction of volatile

anesthetic binding sites in proteins. Biophys. J. 91:

3405-3414.

286. Wang Z, Feng LS, Matskevich V, Venkataraman K,

Parasuram P, Laity JH. 2006. Solution structure of a

Zap1 zinc-responsive domain provides insights into

metalloregulatory transcriptional repression in

Saccharomyces cerevisiae. J. Mol. Biol. 357: 1167-

1183.

287. Wei Y, Fu D. 2006. Binding and transport of metal ions

at the dimer interface of the Escherichia coli metal

transporter YiiP. J. Biol. Chem. 281: 23492-23502.

288. Zerovnik E, Skerget K, Tusek-Znidaric M, Loeschner

C, Brazier MW, Brown DR. 2006. High affinity copper

binding by stefin B (cystatin B) and its role in the

inhibition of amyloid fibrillation. FEBS J. 273: 4250-

4263.

Protein/peptide-nucleic acid

Page 35: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

289. Allen MD, Grummitt CG, Hilcenko C, Min SY, Tonkin

LM, Johnson CM, Freund SM, Bycroft M, Warren AJ.

2006. Solution structure of the nonmethyl-CpG-binding

CXXC domain of the leukaemia-associated MLL

histone methyltransferase. EMBO J. 25: 4503-4512.

290. Buczek P, Horvath MP. 2006. Structural reorganization

and the cooperative binding of single-stranded telomere

DNA in Sterkiella nova. J. Biol. Chem. 281: 40124-

40134.

291. Buczek P, Horvath MP. 2006. Thermodynamic

characterization of binding Oxytricha nova single

strand telomere DNA with the alpha protein N-terminal

domain. J. Mol. Biol. 359: 1217-1234.

292. Cicero DO, Nadra AD, Eliseo T, Dellarole M, Paci M,

de Prat-Gay G. 2006. Structural and thermodynamic

basis for the enhanced transcriptional control by the

human papillomavirus strain-16 E2 protein.

Biochemistry 45: 6551-6560.

293. Crane-Robinson C, Dragan AI, Privalov PL. 2006. The

extended arms of DNA-binding domains: A tale of

tails. Trends Biochem. Sci. 31: 547-552.

294. Datta K, Wowor AJ, Richard AJ, LiCata VJ. 2006.

Temperature dependence and thermodynamics of

Klenow polymerase binding to primed-template DNA.

Biophys. J. 90: 1739-1751.

295. Dragan AI, Li Z, Makeyeva EN, Milgotina EI, Liu Y,

Crane-Robinson C, Privalov PL. 2006. Forces driving

the binding of homeodomains to DNA. Biochemistry

45: 141-151.

296. Ferreira JM, Sheardy RD. 2006. Enthalpy of the b-to-z

conformational transition of a DNA oligonucleotide

determined by isothermal titration calorimetry. Biophys.

J. 91:3383-3389.

297. Fisher RJ, Fivash MJ, Stephen AG, Hagan NA, Shenoy

SR, Medaglia MV, Smith LR, Worthy KM, Simpson

JT, Shoemaker R, McNitt KL, Johnson DG, Hixson

CV, Gorelick RJ, Fabris D, Henderson LE, Rein A.

2006. Complex interactions of HIV-1 nucleocapsid

protein with oligonucleotides. Nucleic Acids Res. 34:

472-484.

298. Fodor E, Ginsburg A. 2006. Specific DNA binding by

the homeodomain Nkx2.5(C56S): Detection of

impaired DNA or unfolded protein by isothermal

titration calorimetry. Proteins 64: 13-18.

299. Kim I, Liu CW, Puglisi JD. 2006. Specific recognition

of HIV TAR RNA by the dsRNA binding domains

(dsRBD1-dsRBD2) of PKR. J. Mol. Biol. 358: 430-442.

300. Kozlov AG, Lohman TM. 2006. Effects of monovalent

anions on a temperature-dependent heat capacity

change for Escherichia coli SSB tetramer binding to

single-stranded DNA. Biochemistry 45: 5190-5205.

301. Kumaran S, Kozlov AG, Lohman TM. 2006.

Saccharomyces cerevisiae replication protein a binds to

single-stranded DNA in multiple salt-dependent modes.

Biochemistry 45: 11958-11973.

302. Kuznetsov SV, Kozlov AG, Lohman TM, Ansari A.

2006. Microsecond dynamics of protein-DNA

interactions: Direct observation of the

wrapping/unwrapping kinetics of single-stranded DNA

around the E. coli SSB tetramer. J. Mol. Biol. 359: 55-

65.

303. Lebbink JHG, Georgijevic D, Natrajan G, Fish A,

Winterwerp HHK, Sixma TK, de Wind N. 2006. Dual

role of MutS glutamate 38 in DNA mismatch

discrimination and in the authorization of repair. EMBO

J. 25: 409-419.

304. Liew CW, Rand KD, Simpson RJY, Yung WW,

Mansfield RE, Crossley M, Proetorius-Ibba M, Nerlov

C, Poulsen FM, Mackay JP. 2006. Molecular analysis

of the interaction between the hematopoietic master

transcription factors GATA-1 and PU.1. J. Biol. Chem.

281: 28296-28306.

305. Madl T, Van Melderen L, Mine N, Respondek M,

Oberer M, Keller W, Khatai L, Zangger K. 2006.

Structural basis for nucleic acid and toxin recognition

of the bacterial antitoxin CcdA. J. Mol. Biol. 364: 170-

185.

306. McKenna SA, Kim I, Liu CW, Puglisi JD. 2006.

Uncoupling of RNA binding and PKR kinase activation

by viral inhibitor RNAs. J. Mol. Biol. 358: 1270-1285.

307. Mishra SH, Shelley CM, Barrow DJ, Jr., Darby MK,

Germann MW. 2006. Solution structures and

characterization of human immunodeficiency virus

REV responsive element IIB RNA targeting zinc finger

proteins. Biopolymers 83: 352-364.

308. Oddo C, Freire E, Frappier L, de Prat-Gay G. 2006.

Mechanism of DNA recognition at a viral replication

origin. J. Biol. Chem. 281: 26893-26903.

309. Singh M, D'Silva L, Holak TA. 2006. DNA-binding

properties of the recombinant high-mobility-group-like

at-hook-containing region from human BRG1 protein.

Biol. Chem. 387:1469-1478.

Protein/peptide-polymer

310. Goncalves E, Kitas E, Seelig J. 2006. Structural and

thermodynamic aspects of the interaction between

heparan sulfate and analogues of melittin. Biochemistry

45: 3086-3094.

311. Guzey D, McClements DJ. 2006. Characterization of β-

lactoglobulin-chitosan interactions in aqueous

solutions: A calorimetry, light scattering,

electrophoretic mobility and solubility study. Food

Page 36: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

Hydrocoll. 20: 124-131.

312. Hsu C-Y, Lin H-Y, Thomas JL, Wu B-T, Chou T-C.

2006. Incorporation of styrene enhances recognition of

ribonuclease a by molecularly imprinted polymers.

Biosens. Bioelectron. 22: 355-363.

313. Mangold SL, Cloninger MJ. 2006. Binding of

monomeric and dimeric concanavalin a to mannose-

functionalized dendrimers. Org. Biomol. Chem. 4:

2458-2465.

314. Najmudin S, Guerreiro CIPD, Carvalho AL, Prates

JAM, Correia MAS, Alves VD, Ferreira LMA, Romao

MJ, Gilbert HJ, Bolam DN, Fontes CMGA. 2006.

Xyloglucan is recognized by carbohydrate-binding

modules that interact with beta-glucan chains. J. Biol.

Chem. 281: 8815-8828.

315. Wang J, Rabenstein DL. 2006. Interaction of heparin

with two synthetic peptides that neutralize the

anticoagulant activity of heparin. Biochemistry 45:

15740-15747.

Protein/peptide-lipid

316. Carneiro FA, Lapido-Loureiro PA, Cordo SM, Stauffer

F, Weissmuller G, Bianconi ML, Juliano MA, Juliano

L, Bisch PM, Poian ATD. 2006. Probing the interaction

between vesicular stomatitis virus and

phosphatidylserine. Eur. Biophys. J. 35: 145-154.

317. Elegbede AI, Srivastava DK, Hinderliter A. 2006.

Purification of recombinant annexins without the use of

phospholipids. Protein Express. Purif. 50: 157-162.

318. Fang Y, Tong GC, Means GE. 2006. Structural changes

accompanying human serum albumin's binding of fatty

acids are concerted. Biochim. Biophys. Acta - Proteins

& Proteomics. 1764: 285-291.

319. Kong X, Wang X, Misra S, Qin J. 2006. Structural

basis for the phosphorylation-regulated focal adhesion

targeting of type Iγ phosphatidylinositol phosphate

kinase (PIPKIγ) by talin. J. Mol. Biol. 359: 47-54.

320. Lai AL, Park H, White JM, Tamm LK. 2006. Fusion

peptide of influenza hemagglutinin requires a fixed

angle boomerang structure for activity. J. Biol. Chem.

281: 5760-5770.

321. Mosbahi K, Walker D, James R, Moore GR,

Kleanthous C. 2006. Global structural rearrangement of

the cell penetrating ribonuclease colicin E3 on

interaction with phospholipid membranes. Protein Sci.

15: 620-627.

322. Nomura K, Corzo G. 2006. The effect of binding of

spider-derived antimicrobial peptides, oxyopinins, on

lipid membranes. Biochim. Biophys. Acta. 1758: 1475-

1482.

323. Sauer I, Nikolenko H, Keller S, Abu Ajaj K, Bienert M,

Dathe M. 2006. Dipalmitoylation of a cellular uptake-

mediating apolipoprotein E-derived peptide as a

promising modification for stable anchorage in

liposomal drug carriers. Biochim. Biophys. Acta -

Biomembranes. 1758:552-561.

324. Shalev DE, Rotem S, Fish A, Mor A. 2006.

Consequences of N-acylation on structure and

membrane binding properties of dermaseptin derivative

K4-S4-(1-13). J. Biol. Chem. 281: 9432-9438.

325. Tanaka M, Dhanasekaran P, Nguyen D, Ohta S, Lund-

Katz S, Phillips MC, Saito H. 2006. Contributions of

the N- and C-terminal helical segments to the lipid-free

structure and lipid interaction of apolipoprotein A-I.

Biochemistry 45: 10351-10358.

326. Wei SY, Wu JM, Kuo YY, Chen HL, Yip BS, Tzeng

SR, Cheng JW. 2006. Solution structure of a novel

tryptophan-rich peptide with bidirectional antimicrobial

activity. J. Bacteriol. 188:328-334.

327. Xiang J, Fan JB, Chen N, Chen J, Liang Y. 2006.

Interaction of cellulase with sodium dodecyl sulfate at

critical micelle concentration level. Coll. Surf. B 49:

175-180.

328. Zhang W, Krishnan N, Becker DF. 2006. Kinetic and

thermodynamic analysis of Bradyrhizobium japonicum

puta-membrane associations. Arch. Biochem. Biophys.

445: 174-183.

Nucleic acid-small ligand

329. Barbieri CM, Pilch DS. 2006. Complete

thermodynamic characterization of the multiple

protonation equilibria of the aminoglycoside antibiotic

paromomycin: A calorimetric and natural abundance

15N NMR study. Biophys. J. 90: 1338-1349.

330. Buchmueller KL, Bailey SL, Matthews DA, Taherbhai

ZT, Register JK, Davis ZS, Bruce CD, O'Hare C,

Hartley JA, Lee M. 2006. Physical and structural basis

for the strong interactions of the ImPy central pairing

motif in the polyamide f-ImPyIm. Biochemistry 45:

13551-13565.

331. Cai P, Huang Q, Jiang D, Rong X, Liang W. 2006.

Microcalorimetric studies on the adsorption of DNA by

soil colloidal particles. Coll. Surf. B. 49: 49-54.

332. Chaires JB. 2006. A thermodynamic signature for drug-

DNA binding mode. Arch. Biochem. Biophys. 453: 26-

31.

333. Duff MR, Tan WB, Bhambhani A, Perrin BS, Thota J,

Rodger A, Kumar CV. 2006. Contributions of

hydroxyethyl groups to the DNA binding affinities of

anthracene probes. J. Phys. Chem. B. 110: 20693-

Page 37: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

20701.

334. Freyer MW, Buscaglia R, Nguyen B, Wilson WD,

Lewis EA. 2006. Binding of netropsin and 4,6-

diamidino-2-phenylindole to an A2T2 DNA hairpin: A

comparison of biophysical techniques. Anal. Biochem.

355: 259-266.

335. Gilbert SD, Stoddard CD, Wise SJ, Batey RT. 2006.

Thermodynamic and kinetic characterization of ligand

binding to the purine riboswitch aptamer domain. J.

Mol. Biol. 359: 754-768.

336. Giri P, Hossain M, Kumar GS. 2006. Molecular aspects

on the specific interaction of cytotoxic plant alkaloid

palmatine to poly(A). Int. J. Biol. Macromol. 39: 210-

221.

337. Giri P, Hossain M, Kumar GS. 2006. RNA specific

molecules: Cytotoxic plant alkaloid palmatine binds

strongly to poly(A). Bioorg. Med. Chem. Lett. 16:

2364-2368.

338. Kim W, Yamasaki Y, Kataoka K. 2006. Development

of a fitting model suitable for the isothermal titration

calorimetric curve of DNA with cationic ligands. J.

Phys. Chem. B 110: 10919-10925.

339. Lah J, Carl N, Drobnak I, Sumiga B, Vesnaver G. 2006.

Competition of some minor groove binders for a single

DNA binding site. Acta Chim. Slov. 53: 284-291.

340. Mikulecky PJ, Feig AL. 2006. Heat capacity changes

associated with DNA duplex formation: Salt- and

sequence-dependent effects. Biochemistry 45: 604-616.

341. Modukuru NK, Snow KJ, Perrin JBS, Bhambhani A,

Duff M, Kumar CV. 2006. Tuning the DNA binding

modes of an anthracene derivative with salt. J.

Photochem. Photobiol. A 177: 43-54.

342. Muller M, Weigand JE, Weichenrieder O, Suess B.

2006. Thermodynamic characterization of an

engineered tetracycline-binding riboswitch. Nucleic

Acids Res. 34: 2607-2617.

343. Nguyen B, Stanek J, Wilson WD. 2006. Binding-linked

protonation of a DNA minor-groove agent. Biophys J.

90: 1319-1328.

344. Patel MM, Anchordoquy TJ. 2006. Ability of spermine

to differentiate between DNA sequences - preferential

stabilization of A-tracts. Biophys. Chem. 122: 5-15.

345. Rajput C, Rutkaite R, Swanson L, Haq I, Thomas JA.

2006. Dinuclear monointercalating Ru-I complexes that

display high affinity binding to duplex and quadruplex

DNA. Chemistry. 12: 4611-4619.

346. Sato K, Onoguchi M, Hosokawa K, Maeda M. 2006.

Affinity capillary electrophoresis of DNA for detection

of single-nucleotide polymorphisms and point

mutations: Comprehensive study for optimization of the

weak affinity. J. Chromatogr A 1111: 120-126.

347. Soto AM, Rentzeperis D, Shikiya R, Alonso M, Marky

LA. 2006. DNA intramolecular triplexes containing dT

-> dU substitutions: Unfolding energetics and ligand

binding. Biochemistry 45: 3051-3059.

348. Tan JF, Too HP, Hatton TA, Tam KC. 2006.

Aggregation behavior and thermodynamics of binding

between poly(ethylene oxide)-block-poly(2-

(diethylamino)ethyl methacrylate) and plasmid DNA.

Langmuir 22: 3744-3750.

349. Thomas JR, Liu X, Hergenrother PJ. 2006. Biochemical

and thermodynamic characterization of compounds that

bind to RNA hairpin loops: Toward an understanding of

selectivity. Biochemistry 45: 10928-10938.

350. Zhu DM, Evans RK. 2006. Molecular mechanism and

thermodynamics study of plasmid DNA and cationic

surfactants interactions. Langmuir 22: 3735-3743.

Enzyme kinetics

351. Andujar-Sanchez M, Las Heras-Vazquez FJ, Clemente-

Jimenez JM, Martinez-Rodriguez S, Camara-Artigas A,

Rodriguez-Vico F, Jara-Perez V. 2006. Enzymatic

activity assay of D-hydantoinase by isothermal titration

calorimetry. Determination of the thermodynamic

activation parameters for the hydrolysis of several

substrates. J. Biochem. Biophys. Meth. 67: 57-66.

352. D'Amico S, Sohier JS, Feller G. 2006. Kinetics and

energetics of ligand binding determined by

microcalorimetry: Insights into active site mobility in a

psychrophilic alpha-amylase. J. Mol. Biol. 358: 1296-

1304.

353. Olsen SN. 2006. Applications of isothermal titration

calorimetry to measure enzyme kinetics and activity in

complex solutions. Thermochim. Acta. 448: 12-18.

354. Poduch E, Bello AM, Tang S, Fujihashi M, Pai EF,

Kotra LP. 2006. Design of inhibitors of orotidine

monophosphate decarboxylase using bioisosteric

replacement and determination of inhibition kinetics. J.

Med. Chem. 49: 4937-4945.

Miscellaneous

355. Alvarez-Lorenzo C, Yanez F, Barreiro-Iglesias R,

Concheiro A. 2006. Imprinted soft contact lenses as

norfloxacin delivery systems. J. Contr. Rel. 113: 236-

244.

356. Arseneault M, Lafleur M. 2006. Isothermal titration

calorimetric study of calcium association to lipid

bilayers: Influence of the vesicle preparation and

composition. Chem. Phys. Lip. 142: 84-93.

357. Backfolk K, Rosenholm JB, Husband J, Eklund D.

2006. The influence of surface chemical properties of

Page 38: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

kaolin surfaces on the adsorption of poly(vinyl alcohol).

Coll. Surf. A 275: 133-141.

358. Ballester P, Oliva AI, Costa A, Deya PM, Frontera A,

Gomila RM, Hunter CA. 2006. Dabco-induced self-

assembly of a trisporphyrin double-decker cage:

Thermodynamic characterization and guest recognition.

J. Am. Chem. Soc. 128: 5560-5569.

359. Barbosa S, Taboada P, Castro E, Mosquera V. 2006.

Influence of SDS and two anionic hydrotropes on the

micellized state of the triblock copolymer E71G7E71.

J. Coll. Interf. Sci. 296: 677-684.

360. Berger BW, Gendron CM, Lenhoff AM, Kaler EW.

2006. Effects of additives on surfactant phase behavior

relevant to bacteriorhodopsin crystallization. Protein

Sci. 15: 2682-2696.

361. Beyer K, Leine D, Blume A. 2006. The demicellization

of alkyltrimethylammonium bromides in 0.1 M sodium

chloride solution studied by isothermal titration

calorimetry. Coll. Surf. B 49: 31-39.

362. Bharadwaj S, Montazeri R, Haynie DT. 2006. Direct

determination of the thermodynamics of self-assembly

of polyelectrolytes and its implications to the

electrostatic layer-by-layer assembly of thin films.

Langmuir 22: 6093-6101.

363. Brule C, Holmer S, Krechanin S, Laali KK. 2006.

Sterically crowded azulene-based dication salts as novel

guests: Synthesis and complexation studies with crown

ethers and calixarenes in solution and in the gas phase.

Org. Biomol. Chem. 4: 3077-3084.

364. Bu H, Kjoniksen A-L, Elgsaeter A, Nystrom B. 2006.

Interaction of unmodified and hydrophobically

modified alginate with sodium dodecyl sulfate in dilute

aqueous solution: Calorimetric, rheological, and

turbidity studies. Coll. Surf. A 278: 166-174.

365. Camara-Campos A, Hunter CA, Tomas S. 2006.

Cooperativity in the self-assembly of porphyrin ladders.

Proc. Natl. Acad. Sci. U. S. A. 103: 3034-3038.

366. Castro E, Taboada P, Mosquera V. 2006.

Characterization of triblock copolymer E67S15E67-

surfactant interactions. Chem. Phys. 325: 492-498.

367. Chakraborty T, Chakraborty I, Ghosh S. 2006. Sodium

carboxymethylcellulose-CTAB interaction: A detailed

thermodynamic study of polymer-surfactant interaction

with opposite charges. Langmuir 22: 9905-9913.

368. Charlot A, Heyraud A, Guenot P, Rinaudo M, Auzely-

Velty R. 2006. Controlled synthesis and inclusion

ability of a hyaluronic acid derivative bearing beta-

cyclodextrin molecules. Biomacromolecules 7: 907-

913.

369. Charman SA, Perry CS, Chiu FCK, McIntosh KA,

Prankerd RJ, Charman WN. 2006. Alteration of the

intravenous pharmacokinetics of a synthetic ozonide

antimalarial in the presence of a modified cyclodextrin.

J. Pharmaceut. Sci. 95: 256-267.

370. Dai S, Tam KC. 2006. Effect of cosolvents on the

binding interaction between poly(ethylene oxide) and

sodium dodecyl sulfate. J. Phys. Chem. B 110: 20794-

20800.

371. Dai S, Tam KC. 2006. Isothermal titration calorimetric

studies on the interaction between sodium dodecyl

sulfate and polyethylene glycols of different molecular

weights and chain architectures. Coll. Surf. A 289: 200-

206.

372. Danil de Namor AF, Abbas I, Hammud HH. 2006.

Anion complexation by calix[3]thieno[1]pyrrole: The

medium effect. J. Phys. Chem. B 110: 2142-2149.

373. Feitosa E, Saverio Brazolin MR, Zumstein Georgetto

Naal RM, Perpetua Freire de Morais Del Lama M,

Lopes JR, Loh W, Vasilescu M. 2006. Structural

organization of cetyltrimethylammonium sulfate in

aqueous solution: The effect of Na2SO4. J. Coll. Interf.

Sci. 299: 883-889.

374. Fielden J, Gunning PT, Long D-L, Nutley M, Ellern A,

Kogerler P, Cronin L. 2006. Anion control of

isomerism, crystal packing and binding properties in a

mononuclear zinc complex. Polyhedron 25: 3474-3480.

375. Gianni P, Barghini A, Bernazzani L, Mollica V,

Pizzolla P. 2006. Aggregation of cesium

perfluorooctanoate on poly(ethylene glycol) oligomers

in water. J. Phys. Chem. B 110: 9112-9121.

376. Gianni P, Bernazzani L, Guido CA, Mollica V. 2006.

Calorimetric investigation of the aggregation of lithium

perfluorooctanoate on poly(ethyleneglycol) oligomers

in water. Thermochim. Acta. 451: 73-79.

377. Gomez CG, Chambat G, Heyraud A, Villar M, Auzely-

Velty R. 2006. Synthesis and characterization of a β-

CD-alginate conjugate. Polymer 47: 8509-8516.

378. Gomy C, Schmitzer AR. 2006. Rational design of new

polymerizable oxyanion receptors. J. Org. Chem. 71:

3121-3125.

379. Goobes R, Goobes G, Campbell CT, Stayton PS. 2006.

Thermodynamics of statherin adsorption onto

hydroxyapatite. Biochemistry 45: 5576-5586.

380. Gorman-Lewis D, Fein JB, Jensen MP. 2006.

Enthalpies and entropies of proton and cadmium

adsorption onto Bacillus subtilis bacterial cells from

calorimetric measurements. Geochim. Cosmochim. Acta

70: 4862-4873.

381. Haldar B, Chakrabarty A, Mallick A, Mandal MC, Das

P, Chattopadhyay N. 2006. Fluorometric and isothermal

titration calorimetric studies on binding interaction of a

telechelic polymer with sodium alkyl sulfates of

varying chain length. Langmuir 22: 3514-3520.

382. Hamedi MH, Laurent B, Grolier J-PE. 2006.

Page 39: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

Construction of solid-liquid phase diagrams in ternary

systems by titration calorimetry. Thermochim. Acta

445: 70-74.

383. Heitmann LM, Taylor AB, Hart PJ, Urbach AR. 2006.

Sequence-specific recognition and cooperative

dimerization of N-terminal aromatic peptides in

aqueous solution by a synthetic host. J. Am. Chem. Soc.

128: 12574-12581.

384. Hoare T, Pelton R. 2006. Titrametric characterization of

pH-induced phase transitions in functionalized

microgels. Langmuir 22: 7342-7350.

385. Hofs B, Voets IK, de Keizer A, Cohen Stuart MA.

2006. Comparison of complex coacervate core micelles

from two diblock copolymers or a single diblock

copolymer with a polyelectrolyte. Phys. Chem. Chem.

Phys. 8: 4242-4251.

386. Ishikawa E, Yamase T. 2006. 31P NMR and isothermal

titration calorimetry studies on polyoxomolybdates-

catalyzed hydrolysis of ATP. J. Inorg. Biochem. 100:

344-350.

387. Jadhav VD, Schmidtchen FP. 2006. Judging on host-

guest binding mode uniqueness: Association entropy as

an indicator in enantioselection. Org. Lett. 8: 2329-

2332.

388. Kano K, Kitagishi H, Dagallier C, Kodera M, Matsuo

T, Hayashi T, Hisaeda Y, Hirota S. 2006. Iron

porphyrin-cyclodextrin supramolecular complex as a

functional model of myoglobin in aqueous solution.

Inorg. Chem. 45: 4448-4460.

389. Keller S, Heerklotz H, Blume A. 2006. Monitoring lipid

membrane translocation of sodium dodecyl sulfate by

isothermal titration calorimetry. J. Am. Chem. Soc. 128:

1279-1286.

390. Keller S, Heerklotz H, Jahnke N, Blume A. 2006.

Thermodynamics of lipid membrane solubilization by

sodium dodecyl sulfate. Biophys. J. 90: 4509-4521.

391. Kresheck GC. 2006. The temperature dependence of the

heat capacity change for micellization of nonionic

surfactants. J. Coll. Interf. Sci. 298: 432-440.

392. Lah J, Bester-Roga Ccaron M, Perger T-M, Vesnaver

G. 2006. Energetics in correlation with structural

features: The case of micellization. J. Phys. Chem. B

110: 23279-23291.

393. Li Y, Taulier N, Rauth AM, Wu XY. 2006. Screening

of lipid carriers and characterization of drug-polymer-

lipid interactions for the rational design of polymer-

lipid hybrid nanoparticles (PLN). Pharmaceut. Res. 23:

1877-1887.

394. Li YJ, Cao MW, Wang YL. 2006. Alzheimer amyloid

beta(1-40) peptide: Interactions with cationic gemini

and single-chain surfactants. J. Phys. Chem. B 110:

18040-18045.

395. Li YJ, Li PX, Dong CH, Wang XY, Wang YL, Yan

HK, Thomas RK. 2006. Aggregation properties of

cationic gemini surfactants with partially fluorinated

spacers in aqueous solution. Langmuir 22: 42-45.

396. Liu Y, Ma YH, Chen Y, Guo DS, Li Q. 2006.

Molecular recognition thermodynamics of pyridine

derivatives by sulfonatocalixarenes at different pH

values. J. Org. Chem. 71: 6468-6473.

397. Meier M, Blatter XL, Seelig A, Seelig J. 2006.

Interaction of verapamil with lipid membranes and p-

glycoprotein: Connecting thermodynamics and

membrane structure with functional activity. Biophys. J.

91: 2943-2955.

398. Nielsen KA, Cho W-S, Lyskawa J, Levillain E, Lynch

VM, Sessler JL, Jeppesen JO. 2006. Tetrathiafulvalene-

calix[4]pyrroles: Synthesis, anion binding, and

electrochemical properties. J. Am. Chem. Soc. 128:

2444-2451.

399. Ojida A, Honda K, Shinmi D, Kiyonaka S, Mori Y,

Hamachi I. 2006. Oligo-Asp Tag/Zn(II) complex probe

as a new pair for labeling and fluorescence imaging of

proteins. J. Am. Chem. Soc. 128: 10452-10459.

400. Perry CS, Charman SA, Prankerd RJ, Chiu FCK,

Scanlon MJ, Chalmers D, Charman WN. 2006. The

binding interaction of synthetic ozonide antimalarials

with natural and modified beta-cyclodextrins. J.

Pharmaceut. Sci. 95: 146-158.

401. Portnaya I, Cogan U, Livney YD, Ramon O, Shimoni

K, Rosenberg M, Danino D. 2006. Micellization of

bovine beta-casein studied by isothermal titration

microcalorimetry and cryogenic transmission electron

microscopy. J. Agr. Food Chem. 54: 5555-5561.

402. Prasad M, Chakraborty I, Rakshit AK, Moulik SP.

2006. Critical evaluation of micellization behavior of

nonionic surfactant MEGA 10 in comparison with ionic

surfactant tetradecyltriphenylphosphonium bromide

studied by microcalorimetric method in aqueous

medium. J. Phys. Chem. B 110: 9815-9821.

403. Rekharsky MV, Yamamura H, Kawai M, Osaka I,

Arakawa R, Sato A, Ko YH, Selvapalam N, Kim K,

Inoue Y. 2006. Sequential formation of a ternary

complex among dihexylammonium, cucurbit[6]uril, and

cyclodextrin with positive cooperativity. Org. Lett. 8:

815-818.

404. Rodriguez-Perez AI, Rodriguez-Tenreiro C, Alvarez-

Lorenzo C, Taboada P, Concheiro A, Torres-

Labandeira JJ. 2006. Sertaconazole/hydroxypropyl-

beta-cyclodextrin complexation: Isothermal titration

calorimetry and solubility approaches. J. Pharmaceut.

Sci. 95: 1751-1762.

405. Sasaki S-i, Kotegawa Y, Tamiaki H. 2006.

Trifluoroacetyl-chlorin as a new chemosensor for

Page 40: University of Zurich - UZH · medicine and biochemistry to physical chemistry and material sciences. Not surprisingly, the range of studies utilizing ITC data is steadily expanding

alcohol/amine detection. Tetrahedron Lett. 47: 4849-

4852.

406. Sessler JL, Gross DE, Cho WS, Lynch VM,

Schmidtchen FP, Bates GW, Light ME, Gale PA. 2006.

Calix[4]pyrrole as a chloride anion receptor: Solvent

and countercation effects. J. Am. Chem. Soc. 128:

12281-12288.

407. Sinn CG, Antonietti M, Dimova R. 2006. Binding of

calcium to phosphatidylcholine-phosphatidylserine

membranes. Coll. Surf. A 282-283:410-419.

408. Soderman O, Jonsson B, Olofsson G. 2006. Titration of

fatty acids solubilized in cationic and anionic micelles.

Calorimetry and thermodynamic modeling. J. Phys.

Chem. B 110: 3288-3293.

409. South CR, Higley MN, Leung KCF, Lanari D, Nelson

A, Grubbs RH, Stoddart JF, Weck M. 2006. Self-

assembly with block copolymers through metal

coordination of SCS-Pd-II pincer complexes and

pseudorotaxane formation. Chemistry 12: 3789-3797.

410. Suksai C, Gomez SF, Chhabra A, Liu JY, Skepper JN,

Tuntulani T, Otto S. 2006. Controlling the morphology

of aggregates of an amphiphilic synthetic receptor

through host-guest interactions. Langmuir 22: 5994-

5997.

411. Sun D-Z, Li L, Qiu X-M, Liu F, Yin B-L. 2006.

Isothermal titration calorimetry and 1H NMR studies

on host-guest interaction of paeonol and two of its

isomers with beta-cyclodextrin. Int. J. Pharmaceut.

316: 7-13.

412. Torres JD, Faria EA, Prado AGS. 2006.

Thermodynamic studies of the interaction at the

solid/liquid interface between metal ions and cellulose

modified with ethylenediamine. J. Haz. Mat. 129: 239-

243.

413. Tsamaloukas A, Szadkowska H, Heerklotz H. 2006.

Thermodynamic comparison of the interactions of

cholesterol with unsaturated phospholipid and

sphingomyelins. Biophys. J. 90: 4479-4487.

414. Venter JP, Kotze AF, Auzely-Velty R, Rinaudo M.

2006. Synthesis and evaluation of the mucoadhesivity

of a CD-chitosan derivative. Int. J. Pharmaceut. 313:

36-42.

415. Wang C, Ravi P, Tam KC. 2006. Morphological

transformation of [60]fullerene-containing poly(acrylic

acid) induced by the binding of surfactant. Langmuir

22: 2927-2930.

416. Wang H, Wang YL, Yan H, Zhang J, Thomas RK.

2006. Binding of sodium dodecyl sulfate with linear

and branched polyethyleneimines nn aqueous solution

at different pH values. Langmuir 22: 1526-1533.

417. Wang L-H, Guo D-S, Chen Y, Liu Y. 2006.

Thermodynamics of interactions between organic

ammonium ions and sulfonatocalixarenes. Thermochim.

Acta. 443: 132-135.

418. Wang Y, Bai G, Marques EF, Yan H. 2006. Phase

behavior and thermodynamics of a mixture of cationic

gemini and anionic surfactant. J. Phys. Chem. B 110:

5294-5300.

419. Wangsakan A, Chinachoti P, McClements DJ. 2006.

Isothermal titration calorimetry study of the influence

of temperature, pH and salt on maltodextrin-anionic

surfactant interactions. Food Hydrocoll. 20: 461-467.

420. Zhang R, Tang M, Bowyer A, Eisenthal R, Hubble J.

2006. Synthesis and characterization of a D-glucose

sensitive hydrogel based on CM-dextran and

concanavalin A. React. Funct. Polym. 66: 757-767.