12
Application of Isothermal Titration Calorimetry (ITC) to Biomolecular Interactions

Graeme L. Conn

Department of Biochemistry, Emory University School of Medicine, Atlanta, GA, USA

12.1 Introduction

12.1.1 Why Measure Binding of Biological Molecules?

Proteins, nucleic acids (DNA and RNA) and other biomolecules do not work in isolation. Rather, the many biological functions of these molecules are defined by their often complex networks of intermolecular exchanges with a diverse array of other biomolecules and bioactive ligands. Enzymes and macromolecular machines like the ribosome, for example, must specifically assemble all required subunits, essential co‐enzymes and/or co‐substrates in order to efficiently catalyse turnover of their bound substrates. Further, enzyme activity may be controlled or localised through interaction with other regulatory binding partners. Similarly, biomolecules that play structural or transport roles in the cell must precisely associate to form their complex macromolecular assemblies. More broadly, regulation of cellular processes from genome replication and stability to each step of gene expression involves a complex and precisely coordinated collection of biomolecular interactions. As a result, understanding what drives biomolecular interactions is at the basis of understanding biology itself.

12.1.2 Approaches for Analysis of Biomolecular Interactions

Many approaches are available to study biomolecular interactions, each with its associated strengths and limitations. These approaches range from qualitative identification of potential binding partners to detailed, quantitative analyses of binding affinities and kinetics. The choice of an optimal approach is often dictated by the specific system under study, including availability or otherwise of purified biomolecules in a suitable form, as well as the ultimate goal of the experiment. A yeast two‐hybrid screen, phage display or co‐immunoprecipitation (e.g. from complex mixtures or cell lysate) may be a good starting point to identify potential interactions that can be further validated by other direct binding methods. Co‐elution of purified binding partners from a size exclusion column might also serve as a simple qualitative assay for binding and the impact of changes (such as amino acid substitution or nucleic acid sequence changes) in one or both binding partners.

Classical biophysical approaches for quantitatively analysing binding often require a source of purified components, though a number of methods are equally applicable to complex mixtures, such as cell lysates, or even to interactions within whole cells. All methods also require a means of detecting one of the binding partners or the binding process itself. For protein–nucleic acid interactions, approaches such as filter binding and electromobility shift assay (EMSA) have a long‐standing record of successful use. Such experiments typically use radioactive phosphorus (32P)‐labelling for its high sensitivity though fluorescent labelling of the nucleic acid is also possible. A limitation of these methods is that detection of binding requires separation of bound and free ligand: this process can perturb the binding equilibrium, potentially leading to inaccuracies in the measured affinities and also making the approach less well suited for low affinity measurements.

Where fluorescently labelled proteins/polypeptides, nucleic acids or small molecule ligands are available, a variety of other methods becomes possible, including fluorescence anisotropy/polarisation (FP), Förster resonance energy transfer (FRET), fluorescence correlation spectroscopy (FCS), analytical ultracentrifugation (AU) and, more recently, microscale thermophoresis (MST). While each of these approaches has advantages and limitations, a key benefit of fluorescence‐based approaches, and many other classical biophysical methods (including isothermal titration calorimetry, ITC), is that they are performed in free solution at thermodynamic equilibrium. It is also notable that many fluorescence‐based approaches are equally applicable to studies of purified components as well as interactions within complex mixtures or whole cells.

For in vitro studies of purified binding partners, a key advantage of approaches like FP is that they do not require separation of bound and free ligand allowing binding to be quantified without perturbing the equilibrium and making them suitable for measurement of low affinity interactions. A further specific advantage of FP is the ability to set up the experiment in a straightforward manner as a competition assay in which a single fluorescently labelled ligand is pre‐bound to the target molecule and this interaction is disrupted by titration of unlabelled competitor ligands. This approach can be used, for example, to relatively rapidly determine apparent affinities for libraries of protein/nucleic acid variants or small molecules for a given target. More recently, MST has also gained significantly in popularity as a versatile approach to study interactions of purified components and within complex mixtures. In MST, a solution or mixture containing the binding partners is placed with specialised thin capillaries and an infrared laser used to induce a microscopic temperature gradient that perturbs two parameters that both typically change upon binding: thermophoresis, or the movement of the molecule in the temperature gradient, and a temperature‐related fluorescence intensity change that depends on the fluorophore's microenvironment. Key advantages of MST are that it is broadly applicable, performed in free solution, without the need for immobilisation, and requires only that one molecule is fluorescently labelled.

Finally, amongst the most widely used approaches for detailed, quantitative measurements of binding are surface plasmon resonance (SPR) and the conceptually similar ‘dip and read’ method, biolayer interferometry (BLI), though, as noted above, MST is also rapidly gaining in popularity. The power of these approaches lies in their capacity to directly measure kinetics of binding, i.e. rates of association (k on ) and dissociation (k off ), from which the equilibrium binding affinity (K d ) can be calculated. The main disadvantage of these approaches is the need to immobilise one of the binding partners on a sensor or tip surface (typically the smaller binding partner as the signal depends on changes in the immobilised mass on the surface). Nonetheless, the ability to define binding kinetics can provide important insights into the nature of binding interactions that are not usually possible with other methods, which only measure or approximate equilibrium binding affinities. However, as a widely applicable and quantitative method for analysis of biomolecular interactions, ITC has firmly established its place at the forefront of these approaches since its initial development several decades ago [1,2]. Indeed, ITC is often referred to as the ‘gold standard’ for analysing intermolecular interactions.

12.1.3 Why Use ITC?

ITC has two key advantages: it is performed entirely in solution, meaning that there is no need to immobilise one of the interacting partners (as is the case for SPR or BLI), and it detects a universal signal, meaning no label is required (as is the case for EMSA, FP or MST). However, probably the foremost benefit of using ITC is that it allows the user to determine a complete thermodynamic profile of binding from a single experiment. Specifically, ITC can be used to directly determine binding affinity (K a = 1/K d ), stoichiometry (n) and enthalpy (ΔH). From these, Gibbs free energy (ΔG) and entropy (ΔS) of binding can also be calculated (see Section 12.2). Measuring the enthalpy of binding provides insight into how favourable interactions contribute to binding, as it reflects the total contributions of hydrogen bonding as well as hydrophobic and electrostatic interactions. While generally more complex to interpret, entropy of binding, on the other hand, reports on the collective contributions of binding‐induced changes in molecular freedom of motion and reorganisation of bound or free water molecules, ions, etc. Such information may not be readily discernible, even from high resolution structures, making ITC a powerful and often essential component of detailed structure–function studies of biomolecules. The insights ITC can provide on thermodynamics and kinetics of binding can also be particularly valuable in the process of drug development in identifying the best candidate(s) for further optimisation of their physicochemical properties from a group of leads that might otherwise appear equivalent on their binding affinities alone [35].

Following the introduction of the first commercial microcalorimeters over three decades ago, these instruments are now routinely found both in individual research labs as well as in shared equipment facilities in academic and industrial settings. More recent improvements in instrumentation have increased sensitivity while reducing the amount of sample required, opening the method to the study of harder to produce or purify biomolecules. In practical terms, while not every biomolecular interaction can be analysed by ITC, the method has without question found exceptionally wide application in the study of discrete biological systems and is applicable to measurement of many different types of binding interaction. The following sections will describe the basic principles of ITC theory, current instrumentation typically available at most research institutions and the application of ITC to the study of biological macromolecular interactions. Next, detailed protocols and a trouble‐shooting guide will be provided, which are exemplified by two case studies with one example each of macromolecule–macromolecule and macromolecule–small molecule interaction.

Finally, before delving into the theory and practice of ITC, a brief note on nomenclature. For simplicity of the descriptions that follow, the molecule (titrand) that is placed in the ITC sample cell will be referred to as the ‘macromolecule’ and its binding partner (titrant) that is placed in the automated pipette used for injection (titration) during the experiment will be referred to as the ‘ligand’. However, while often the most practical arrangement, in principle there is no requirement for a larger protein or nucleic acid macromolecule to be the binding partner placed in the sample cell, and the ‘ligand’ may equally be a second protein, DNA/RNA or small molecule. While this chapter will focus on the use of ITC for analysis of biomolecular binding interactions, the power of the approach and instrumentation extends to advanced methods for global ITC data analysis and measuring binding and measuring enzyme kinetics [68].

12.2 Principles and Theory of ITC

12.2.1 What Is ITC and What Does It Measure?

ITC is a non‐destructive, label‐free analytical technique that is performed entirely in solution. In principle, ITC can be applied to the study of any intermolecular interaction provided heat is either evolved (exothermic) or taken up (endothermic) upon binding. This ‘universal signal’ detected in ITC means that there is no requirement for labelling with radioisotope, chromophore or fluorophore, or for immobilisation of one binding partner to a surface. While the nature of the signal detected is one of ITC's principal advantages, it also results in an important limitation: all other processes occurring in the ITC cell upon titration of ligand will contribute to the signal. Although small background signals (such as heats arising from mechanical mixing of the ligand into the cell solution) can be readily accounted for, other more significant changes due to mismatches in solution conditions between macromolecule and ligand samples can cause significant problems. Complex binding interactions, e.g. where protein–ligand interaction leads to conformational change on a different timescale, or binding‐induced aggregation, will also complicate ITC analysis and interpretation and may require other approaches for a complete resolution.

While some limitations exist, another key advantage of ITC is that experiments can be performed under a wide variety of different experimental conditions. Such flexibility may be useful to tailor solution conditions such that they are optimal for the specific biomolecules under study, which may require specific ionic strength or other additives like glycerol for increased solubility and stability. Flexibility in terms of solution and temperature parameters also allows the potential to adjust experimental conditions to maximise the signal detected. Finally, variations in factors like pH, ionic strength or temperature can be used to directly learn more about the binding interaction. For example, experiments performed under different conditions of ionic strength can provide insight into the role of electrostatic interactions in binding, while variations in solution pH can provide information on the role of ionisation state and the mechanism of binding. In short, time invested in optimising the conditions used for a particular binding interaction is likely to be worthwhile and potentially highly informative on the nature of the interaction. With the speed and reduced sample requirement of the most recent versions of modern ITC instruments (see Section 12.2.2 below), as well as the ability in some cases to automate many aspects of data collection and initial analysis, such optimisations can be performed relatively quickly. Additionally, as ITC is performed in solution and is non‐destructive, precious samples can in principle be recovered and used for additional ITC experiments (if the two binding partners are readily separable) or for other purposes.

Despite its flexibility and broad applicability, there are unfortunately still some instances where ITC may not be a useful approach. As noted above, the signal measured in ITC is proportional to binding enthalpy, meaning that interactions with very low ΔH may not be detectable. In such cases, the signal might be increased by altering the temperature at which titrations are performed or increasing the concentrations of both macromolecule and ligand to proportionally increase the signal generated. However, in practical terms (e.g. protein solubility) it may not always be possible to adjust the necessary parameters sufficiently. Additionally, due to the nature of the relationship between K d and reactant concentrations in ITC, simply increasing concentrations may lead to other complications in data analysis. Nonetheless, with highly purified and carefully prepared samples, ITC is generally applicable to most biological interactions with binding affinities in the low nM to high μM range (K d  ∼ 10−8 to 10−4 M). Weaker binding affinities can also be measured reliably, though a full thermodynamic profile of binding may not be obtainable [9,10]. Other approaches involving displacement of a weaker binding ligand can be used to extend this useful range to much tighter interactions (K d  ∼ 10−9 to 10−12 M) [11].

12.2.2 Overview of Current ITC Microcalorimeters

Commercial ITC microcalorimeters have been available for several decades and are produced today by a number of different manufacturers. Those made by MicroCal (currently part of Malvern Panalytical) and TA Instruments are among those most commonly found in most academic or industrial laboratories. The MicroCal/Malvern Panalytical VP‐ITC and Auto‐iTC200 will be specifically referred to in descriptions later in this chapter, as these are the instruments with which the author is most familiar. However, the sample preparation and requirements, general principles of ITC experiment design and many practical aspects of performing a binding analysis will be more broadly applicable to newer instruments from MicroCal/Malvern Panalytical (such as the PEAQ‐ITC range) or those from other manufacturers.

The basic principles of ITC instruments have not changed significantly in the last decade or so but two notable advancements in the technology have arisen during this time. First, the volumes of sample required to run an ITC titration have been substantially reduced compared to early instruments with the development of small cell (∼200 μl) instruments such as the PEAQ‐ITC and iTC200 (MicroCal/Malvern Panalytical) and Affinity ITC and Nano ITC (TA Instruments). While these instruments have the same limited throughput and requirement for substantial user intervention in an experimental setup as older instruments, the reduced volumes of sample needed (both macromolecule and ligand) have opened the method to use greater effort to produce biomolecules. Importantly, despite the reduced sample requirements, these instruments can detect comparable or smaller heats of binding and operate with similar low levels of noise as instruments with larger cells. The second major recent advance has been the development of automation built around these same small‐cell ITC instruments. Automation has the advantages of eliminating some of the steps that typically cause problems for novice users, such as proper sample loading, as well as allowing for unattended operation for multiple ITC experiments.

Manual instruments such as the iTC200 or the older workhorse VP‐ITC are most likely to be found in individual labs and may be shared by a relatively small number of users. Significant expertise in sample handling and instrument use, particularly sample loading, are critical for obtaining high quality data with these instruments. Automation brings additional cost for purchase and maintenance, meaning that instruments like the Auto‐iTC200 are more often found in managed shared facilities (e.g. [12]). Such ITC facilities can accommodate a much larger group of users with a wide range of expertise, supported by an instrument that is very easy to use. Whatever the specific type of instrument available, the principle of its operation and ultimate experimental output will be largely the same.

A typical ITC instrument (Figure 12.1a) consists of two identical coin‐shaped or cylindrical cells made of a chemically inert and highly thermally conducting metal (such as gold or Hasteloy alloy). Each cell has a long, narrow neck through which it is filled with either the macromolecule sample or a reference solution (typically water). In the case of the sample cell, this long neck is also where the automated titration pipette is placed for injection of the ligand, the second component of the binding reaction. Within the instrument, the two cells are held under conditions of constant pressure (isobaric) and temperature (isothermal) within an adiabatic jacket, meaning that no heat is transferred into or out of this enclosed system. Under application of constant cooling, power is applied to heaters on each cell to maintain them at constant and equal temperature.

Image described by caption and surrounding text.

Figure 12.1 Overview of ITC instrument design and measurement. (a) Simplified schematic of a power compensation ITC instrument. Two coin‐shaped cells within an adiabatic jacket hold the reference and macromolecule solutions while the ligand (titrant) is loaded in a specialised automated stirring pipette at the beginning of the experiment. Constant power is applied to the reference cell heater while the sample cell heater is controlled by a feedback loop that adjusts power to compensate for heat evolved or taken up during the binding reaction within the cell. The variation power supplied to the sample cell is calibrated to maintain the two cells at equal temperature (i.e. ΔT = 0) and is the signal detected in ITC. (b) Ligand is injected into the cell in a series of injections of equal volume. In the example shown for a typical experiment on a MicroCal/Malvern Panalytical instrument, a negative deflection from baseline is due to an exothermic binding reaction: heat is evolved, meaning that power to the sample cell heater is reduced to compensate. (c) Values from integration of peaks in the raw ITC thermogram are plotted against the molar ratio of binding reaction components and fit to a specific binding model to yield molar enthalpy (ΔH), binding affinity (K a ) and molecular stoichiometry (n).

When a sample of the ligand is injected from the automated pipette, a binding reaction occurs in the sample cell and heat is either evolved (for an exothermic reaction) or taken up (for an endothermic reaction). In the more common case of an exothermic binding reaction, the heat increase in the cell due to binding is sensed by the instrument and a proportional reduction made to the power applied to the sample cell. In other words, the total heating power from the binding reaction and sample cell heater are maintained at a constant total level, with any heat input from binding compensated by a drop in the heat input from the cell heater. The effect of this power compensation is to maintain the sample and reference cells at an equal temperature (ΔT = 0). Similarly, for an endothermic binding reaction, increased power will be applied to the cell heater to maintain the constant total heating power with the same final result.

The difference in power that needs to be applied to the sample cell heater is the direct output measurement in ITC. Further, the power difference is directly related to the molar enthalpy change associated with the binding process, allowing determination of binding thermodynamics. The specific measurement recorded by an ITC instrument is the time‐dependent sample cell power input (i.e. power per unit of time, such as μcal/s or μJ/s) needed to keep the temperature of the sample and reference cells equal. This output is usually plotted in real‐time during the course of the ITC experiment and appears as a series of sharp changes (peaks) in applied power, each corresponding to one injection of the ligand. In the case of an exothermic binding reaction, less power needs to be applied and the peaks therefore appear as negative deflections from the baseline. The example titration shown in Figure 12.1b corresponds to an exothermic binding reaction.

At the end of the ITC experiment the peaks corresponding to heater power are integrated with respect to time, from the beginning of the ligand injection to the point where the applied power returns to the pre‐injection baseline value. This process of peak integration gives the total heat exchanged per injection and these heat effects can then be analysed as a function of the molar ratio of the binding partners to give the thermodynamic parameters of the interaction under study (Figure 12.1c). The entire ITC experiment and recording of sample cell power take place under computer control, and software associated with the instrument will usually generate an initial automated analysis of the raw data. As described further below, it is usually necessary to manually check and adjust the peak integrations and to perform additional analysis steps before the most accurate fit of the interaction thermodynamic parameters is obtained.

12.2.3 ITC and Binding Theory

The power of ITC comes from its ability to provide a complete thermodynamic analysis of a binding reaction from a single experiment. Specifically, the user can directly determine the affinity (K a ), stoichiometry (n) and enthalpy (ΔH) of binding. From these directly determined parameters, the Gibbs free energy (ΔG) and entropy (ΔS) of binding can also be calculated. The following descriptions present a simplified view of the theory behind ITC analysis.

In an ITC experiment, the heat evolved (q) in the cell of volume V o upon binding of the macromolecule (M) and titrated ligand (L) is proportional to the concentration of the macromolecule–ligand complex (ML) and related to ΔH of binding by Eq. 12.1. This can be expressed as the ratio of concentrations of complex and total macromolecule as in Eq. 12.2:

12.1 equation
12.2 equation

In the simplest case of a macromolecule with a single binding site for one ligand, the extent of reversible binding, i.e. concentration of the ML complex, is given by the equilibrium binding constant (Eq. 12.3). Note that some ITC fitting software (e.g. Origin® supplied by MicroCal/Malvern Panalytical) generates the association equilibrium binding constant, K a (units per molar, M−1), whereas biologists typically discuss binding in terms of the dissociation binding constant, K d (units of molar, M). Free energy of binding can be calculated from the K a (or K d ) measured by ITC using Eq. 12.4, where R is the gas constant and T is temperature (in K):

12.3 equation
12.4 equation

In terms of binding affinities and binding energies, it follows from these relationships that a larger value of K a (or smaller value of K d ) refers to tighter binding between the macromolecule and ligand, as well as a more negative free energy of binding, ΔG. Finally, ΔG can be partitioned out into its enthalpic (directly determined by ITC) and entropic (ΔS) components in Eq. 12.5, allowing determination of each of these thermodynamic parameters:

12.5 equation

12.3 Protocols for Design, Implementation and Analysis of ITC Experiments

12.3.1 Where to Begin? Planning a First ITC Experiment

Having made the decision to study a binding interaction using ITC, the novice user (and even experienced ITC practitioner) is faced with a plethora of potential experimental variables when considering an analysis of a new system. For example, in sample preparation, which specific protein, nucleic acid or other biomolecule(s)/ligand(s) are the optimal ones to begin with? How much is needed and how should each component be purified? What experimental conditions, instrument settings and approaches for data analysis should be used?

Before beginning any new analysis, the most important first consideration is how much of each sample is needed and whether can these be feasibly produced at the level of purity needed for ITC. The answer to this question depends on both the instrument available and the system itself. For the MicroCal/Malvern Panalytical VP‐ITC (∼1.4 ml cell volume) or standard cell configurations of TA Instruments' ITC microcalorimeters (1 ml cell volume), up to 2 ml of macromolecule sample and 0.5 ml of ligand sample is needed for each titration experiment. For automated small‐cell (∼200 μl cell volume) instruments such as the Auto‐iTC200, these volumes are considerably less at 400 and 120 μl for macromolecule and ligand samples, respectively. For manually loaded small cell instruments, such as the MicroCal/Malvern iTC200 or TA Instruments' small‐cell configuration of the Nano ITC, there may up to a further 50% reduction in the volumes needed.

As discussed further below, macromolecule and ligand concentrations are important variables in optimising ITC experiments that also directly impact the quantity of material needed for each titration. Typically, the concentration of the macromolecule should be >10 times the expected K d with the ligand sample at least ∼10 times more concentrated, such that binding approaches saturation by the midpoint of the titration. However, for a new system with no prior knowledge of binding affinity, a first experiment with macromolecule and ligand concentrations of 10 and 100 μM, respectively, is a reasonable starting point (for a 1 : 1 binding stoichiometry). For a 50 kDa protein, this cell concentration corresponds to 0.5 mg/ml, or ∼0.2 to 1 mg total protein depending on the sample cell size in the instrument to be used. For the same protein placed in the pipette, approximately two to three times more is required. For this reason, the harder to produce, more costly and/or most challenging to concentrate component is more often placed in the instrument sample cell. Of course, no matter which instrument is available or the optimal concentrations for the sample, significantly more material will be needed to complete a rigorous experiment complete with appropriate controls and titration replicates. However, if generating sufficient material for a pilot experiment is readily feasible, one should be able to quickly determine whether ITC will ultimately produce useful data for the system under study, thus making it worthwhile to further the investment in time and effort to produce the samples.

Most ITC instruments are supplied with software capable of simulating ITC experiments from a set of parameters (n, ΔH and K d ) provided by the user, allowing some optimisation of starting concentrations to be performed before even beginning a first experiment. Simulating ITC experiments can be particularly useful if some information is already known about the system in question, such as the approximate K d from another binding measurement method or reasonably guessed from experiments with closely related molecules such as protein homologues. The simulation software is also a useful and simple way for the novice user to begin to get a feel for how sample concentration (both absolute and the ratio of macromolecule and ligand concentrations), K d and ΔH influence the shape of the final binding isotherm obtained.

The preferred starting concentrations for an ITC experiment are commonly determined by consideration of a unitless parameter c, which is the product of the predicted affinity of the system and the total macromolecule concentration:

12.6 equation

Optimally, experiments should be performed in the range 5 < c < 100. However, in most cases all three parameters measured by ITC (i.e. K a , n and ΔH) can be determined with confidence for 1 < c < 1000 in a single experiment. The macromolecule concentration should therefore be determined with these ranges of c values in mind. For example, for a binding reaction with a K d of 10−6 M (or 1 μM), ideally a concentration of 5–500 μM should be used. For a given binding affinity, the curvature binding isotherm is dependent on the value of c (Figure 12.2). This makes it possible to manipulate the shape of the binding isotherm by adjusting the concentrations of the cell sample to optimise the quality of the data obtained. Of course, obtaining the fully optimal concentration may not always be possible: very low reactant concentrations might result in too small an experimental signal (which depends on ΔH of binding) whereas very high concentrations might result in heats that are too high or may simply be impractical or impossible for biological macromolecules (e.g. if a protein has a tendency to aggregate at a high concentration).

Image described by caption and surrounding text.

Figure 12.2 Experimental design and impact of c on the ITC binding isotherm. Simulation of ITC binding curves using the VP‐ITC control software for a single set of user‐defined parameters input through the panel shown on the left. In these examples the values of stoichiometry, n, K d , and ΔH are fixed at 1, 1 μM and − 10 kcal/mol, respectively. Each plot was generated by editing the value of c and accepting the system generated values for macromolecule (cell) and ligand (pipette) sample. For the suboptimal extreme values of c, additional warnings are flagged (red box) for too low or too high heats for c = 1 and c = 1000, respectively. However, experimentally derived estimates of K d and ΔH from a pilot ITC experiment could be used to more usefully optimise experimental conditions for subsequent experiments.

What happens outside of this optimal range of c values? For very tight binding reactions (where conditions of c > 1000 are likely) only ΔH and n can be accurately determined directly from an ITC experiment. Additional approaches, such as using competition experiments with a pre‐bound competing ligand, may however still allow the user to determine a K d [11]. In contrast, for weak binding events it may still be possible to obtain useful data working in the ‘low c range’, by using a very high concentration of ligand to drive binding to near saturation at the end of the titration [ 9, 10]. In this scenario, only the binding affinity can be measured with confidence. ΔH will not be well defined by the experiment and ΔH and n will be correlated in the fit of the data to a binding model. However, in cases where n is known with confidence and can be fixed in the fitting procedure if necessary, such experiments can still be very informative.

In summary, the availability of some prior knowledge (or reasonable guesses) of the system to be studied, use of ITC experiment simulation software and consideration of the c parameter can facilitate the early identification of suitable sample concentrations for an optimal ITC experiment. However, as noted above, in the absence of knowledge about a new system, a first guess of 10 and 100 μM concentrations in cell and pipette, respectively, is a reasonable starting point for an anticipated 1 : 1 binding stoichiometry for a typical macromolecular interaction. From initial estimates of n, K d and ΔH from this first pilot experiment, the same simulation tools can then be used to rapidly identify optimal concentrations for subsequent titrations.

12.3.2 Sample Requirements and Preparing for an ITC Experiment

The importance of sample preparation and planning for an ITC experiment cannot be overstated. The quality of data produced, and thus reliability of insights into binding, are only ever as good as the reagents used to generate them. The following sections delve deeper into many of the necessary practical considerations before embarking on an ITC experiment, but it is worth noting three particularly critical points from the outset:

  • All samples – protein, nucleic acids, small molecules, etc. – should be as purified and homogeneous as is practical.
  • Solutions containing the macromolecule and ligand must be as closely matched as possible to avoid excessive heats arising from simple mixing of the two different solutions.
  • Ideally, the active concentrations of both components in the binding reaction should be very well defined. However, the accuracy of measurement of the ligand concentration is most critical as this directly impacts the accuracy of determining all three binding parameters (n, K a and ΔH).

While no one set of protocols can cover all possible scenarios for a given ITC experiment, the following sections aim to serve as general guides for best practice in accomplishing each of the above goals. An important early consideration in experiment design is which molecule to place in the cell and which in the automated injection pipette. While in principle the same binding parameters should be determined regardless, practical reasons often dictate how the experiment is set up. For example, since the macromolecule sample requires considerably lower concentration, less total amount of material and uncertainties in concentration have less impact on the binding parameters determined, protein or other biomolecules prone to aggregation/precipitation may be best placed in the sample cell. Those molecules for which concentration can be very accurately measured may be best considered as the ligand and placed in the pipette.

12.3.2.1 Sample Preparation

The quality of data produced by an ITC experiment is critically dependent on the quality of the samples used to generate them. Accordingly, all macromolecules, ligands, etc. should be as purified as possible, free of degradation products and other contaminants, and ideally conformationally homogeneous. While each individual macromolecule or ligand will require its own specific protocol for purification, the following general considerations should be kept in mind when optimising sample preparation for ITC.

  • Protein samples should be both chemically and conformationally as pure as possible. This can typically be achieved with two (or more) complementary chromatographic methods, such as affinity chromatography followed by gel filtration chromatography. While a vast range of options are available, use of 6×His‐tagged protein with initial purification by Ni2+‐affinity chromatography is most common. Gel filtration as the final step of purification is particularly useful as it can remove misfolded or aggregated target protein in addition to other differently sized contaminants and identify potential mixtures of oligomeric states of the protein that could complicate later analysis of binding data.
  • Proteins are also often expressed with an additional short sequence tag, such as hexahistidine, or as fusions with complete protein domains, like maltose binding protein (MBP), glutathione‐S‐transferase (GST) and many more. Such additions are routinely used to simplify and expedite purification, aid protein expression/solubility, or often both [13]. While appended tags or domains do not generally affect protein function, their possible impact on protein interactions or the potential to induce changes in protein oligomerisation should be kept in mind. Ideally, for example, for fusion proteins like GST, which can promote dimerisation, the GST tag should be removed during the purification protocol to eliminate potential confounding effects on the binding analyses.
  • Chemically synthesised nucleic acids (DNA and RNA) and longer in vitro transcribed RNAs should be purified by high performance liquid chromatography (HPLC), fast protein liquid chromatography (FPLC) (e.g. gel filtration chromatography) or polyacrylamide gel electrophoresis (PAGE). The latter approach is typically done under denaturing conditions (e.g. 50% urea) but requires refolding, which may be challenging for some larger structured RNAs. Native gel purification is also possible.
  • Small molecules and other ligands should also be as pure as possible. From commercial suppliers, purchase the highest purity grade available. For in‐house syntheses, standard isolation procedures such as chromatographic methods, recrystallisation, etc., should be used with confirmation of identity and purity by mass spectrometry and nuclear magnetic resonance (NMR).
  • Final samples for use in the ITC experiment should be free of aggregates and other particles. The final stage of sample preparation may involve concentration of protein or other macromolecules, e.g. in a centrifugal concentrator. Such devices can produce concentration gradients of the macromolecule, which may promote aggregation. The user should therefore be alert to signs of problems during this step. Before use, it may be desirable to filter or centrifuge the sample at high speed in a microfuge tube. Finally, the sample quality can be reassessed before use by re‐running a small quantity on a gel filtration column or using a light scattering method like dynamic light scattering (DLS).
  • Immediately before use, degassing of samples is often recommended to avoid formation of air bubbles during the ITC experiment. This may be particularly problematic for experiments performed at temperatures higher than used for sample preparation or storage (e.g. within the Auto‐iTC200 autosampler tray). With the VP‐ITC, degassing can be easily accomplished using the ThermoVac device typically supplied with this instrument. In our experience, however, sample degassing is not always necessary to obtain consistent high quality titrations but may be considered as an additional optimisation step if needed.

12.3.2.2 Buffer Considerations and Matching of Solutions

Many factors must be considered when deciding on the ‘optimal’ solution conditions for a given ITC experiment. Of course, a primary requirement is to use solution conditions – buffer/pH, ionic strength, reducing agent, organic solvent, or other additives, cofactors, etc. – that are best suited for the long‐term (> several hours to days) solubility and stability of the macromolecule(s) or ligand to be studied. In the absence of homogeneous, well‐behaved samples, no ITC experiment is going to succeed! Keeping this in mind, however, it is also important to note that some limitations on solution conditions do exist and, most critical for performing a successful ITC experiment, the ability is needed to precisely match the solution conditions for the macromolecule and ligand.

  • Other than the two components of the binding interaction being studied, the two solutions containing the macromolecule and ligand should be as close to identical as possible. Even very small differences in composition or pH between the two solutions can cause significant artefactual signals in the ITC measurements due to large heats of dilution or ionisation. How this matching of macromolecule and ligand solutions is accomplished depends, in part, on the molecules under study.
  • A commonly used approach for macromolecules is exhaustive dialysis against the experimental buffer solution to ensure exact matching of the sample(s). For example, two or three dialysis steps are required with >100‐ to 1000‐fold volume difference between the sample and dialysis solution using dialysis tubing or disposable units (e.g. Slide‐A‐Lyzer™ or Spectra/Por® Float‐A‐Lyzer®) with an appropriate molecular weight cut‐off (MWCO). A convenient approach is to perform two dialyses steps of three to four hours during the day prior to running the ITC experiment with a third overnight dialysis before the final steps of sample preparation the following morning. In cases where both binding partners are macromolecules, precise matching can be accomplished by placing both samples in separate dialysis tubing/units within the same large volume of dialysis buffer. Save the final dialysis buffer at the end of this process to preparing the ITC instrument and running control experiments, e.g. a ligand into buffer titration. As an alternative to dialysis, buffer exchange can also be accomplished using centrifugal concentration devices.
  • With the shorter run times of columns now available, gel filtration chromatography can also be an excellent way to ensure buffers are well matched without the need for further dialysis. The reduced time in sample preparation may be particularly helpful in cases where proteins or other macromolecules are prone to aggregation or degradation. The final step of macromolecule purification can be performed directly in the buffer to be used for the ITC experiment and some of the buffer retained for dilution of other binding partners or for control titrations.
  • Other ligands, such as small molecules, can also be dialysed along with macromolecules if membranes or dialysis units with a suitable MWCO are available. Alternatively, a commonly used approach for small molecules is to dissolve them in the buffer solution from the final step of dialysis used for the other binding partner (typically a macromolecule). Note, however, that care should be taken when using lyophilised samples (e.g. small molecules, peptides or nucleic acids) as salts present before lyophilisation will be retained in the final sample. These will be challenging to match in the other sample.
  • While ITC experiments can be performed in most commonly used biological buffers, inadequate matching of pH for the macromolecule and ligand samples can cause significant issues. After preparing samples, it can therefore be useful to check that the final pH values are well matched (±0.05 pH units). The impact of pH discrepancies may also be exacerbated when using buffers with a large enthalpy of ionisation (ΔH ion ), such as Tris. Choosing a buffering system with a low (ΔH ion ), such as phosphate, may therefore be preferable so long as it is compatible with the samples under study. Alternatively, comparison of binding enthalpies and affinities in multiple buffers with different ΔH ion values can be informative on the nature of binding. For example, for binding reactions with a coupled protonation event, Tris buffer would yield a large enthalpic signal and measurements in multiple buffers can be used to quantify protons taken up or released upon binding [14,15].
  • If reducing agents are required for protein stability, the best choices are tris(2‐carboxyethyl)phosphine (TCEP) or up to a few mM concentration of 2‐mercaptoethanol (ß‐mercaptoethanol). Use fresh stocks and, as noted above, make sure the concentration of reducing agent is closely matched between the macromolecule and ligand sample solutions. The commonly used reducing agent dithiothreitol (DTT) should usually be avoided if possible as it can cause serious problems with ITC baseline stability. However, specific compatibilities and tolerated reducing agent concentrations vary a little from instrument to instrument and should be checked before performing experiments.
  • Small molecules may require a small amount of organic solvent (e.g. DMSO) in the final sample solution to maintain their solubility. In such cases, the same percentage of organic solvent should be added to the other binding partner (typically the macromolecule in the sample cell).

Even carefully following these suggested guidelines, slight differences in buffer composition between the ligand and macromolecule are possible due to co‐solvents, salts or pH, particularly where samples are prepared independently (i.e. not co‐dialysed), dissolved from lyophilised material or have components like organic solvents that need to be added immediately prior to experiments. In such cases, a test titration(s) of ligand into the macromolecule sample buffer and/or the ligand sample buffer into the macromolecule before performing the binding experiment may be advisable to check for large heats of dilution that could mask the binding signal. Finally, it is important to keep in mind that many of these steps used for buffer exchange and matching can result in potentially large changes in sample concentration. As a result, concentration measurements should be made following these steps, ideally on the final samples to be used in the ITC experiment.

12.3.2.3 Importance of Determining Accurate Sample Concentrations

The accuracy of data obtained from an ITC experiment is dependent on precise determination of sample concentrations. In particular, error in determining the concentration of the ligand in the pipette is directly proportional to the error of measurements of n, K a and ΔH, and therefore the calculated free energy (ΔG) and entropic (ΔS) contribution to binding. In contrast, errors in cell component concentration only affect the accuracy of the determined stoichiometry. In cases where the concentration of one component is more reliably known, this may be an important consideration for experiment design: placing the component with the most accurate concentration in the pipette will result in the highest accuracy of the data produced in the ITC experiment.

Concentrations of each sample should be determined using an applicable method such as absorbance measurement, where a precise measured or calculated extinction coefficient is known, colorimetric assays such as BCA for proteins or HPLC with appropriate standards for small molecules. Ideally the concentrations of both binding partners should be accurately determined after buffer matching and final sample preparation. Two additional steps, described in the experimental protocols below, can also modestly impact the final sample concentrations. First, it is commonly recommended to degas the samples to avoid signal artefacts due to air bubbles or release of dissolved gases during the titration, particularly at higher temperatures. However, this also has the potential to reduce the sample volume via evaporation and thus increase concentration. Of course, since typically a little more sample is prepared than needed, the concentration can simply be rechecked to ensure there is no significant change. Second, rinsing of the sample cell will leave some residual buffer in the cell which will dilute the sample (by ∼1% to 2%). While this is harder to accurately account for, as noted above, this change will only impact the determined binding stoichiometry and can likely be ignored in most cases.

12.3.3 Running the ITC Experiment

In each ITC experiment, a high concentration solution of ligand is injected (titrated) from an automated pipette (or syringe) in a series of equal injections into a more dilute sample of the macromolecule in the instrument cell. The final task, with final purified samples in hand, is therefore placing these two sample solutions in their respective components of the ITC instrument. Correctly loading the samples, for example without introducing any air bubbles, is essential for a successful experiment. How this is accomplished depends on the type of instrument available and is described below for both a manual VP‐ITC and the fully automated Auto‐iTC200 MicroCal/Malvern Panalytical instruments. Newer manual instruments than the VP‐ITC are likely to have partial automation that simplifies the sample loading process. Additionally, while the same general goals and principles apply, the practical procedures for setting up instruments from other manufacturers, such as the Nano ITC from TA Instruments, may also vary considerably (links to additional web resources for Nano ITC cell and syringe loading are provided under Website Resources at the end of this chapter).

Prior to beginning any experiment with precious samples, the user should make sure the ITC instrument is in good working order. Check any instrument logs for the date of last use and any notes from previous users to make sure the instrument was cleaned following manufacturer recommended or local instructions for cell and pipette cleaning. For automated instruments, visually inspect the fine tubing that carries samples and other reagents during the filling process for signs of precipitated materials or blockages. Running a test titration, which could simply be a small number of water‐into‐water injections, is a good idea in either type of instrument to confirm the instrument is operating as expected with a stable baseline close to the set reference power at the beginning of the test titration. Watching the pipette and cell loading processes on an automated instrument during this test titration is a good way to identify any potential issues before placing precious samples on the instrument.

12.3.3.1 Sample Loading and Experiment Set Up on a VP‐ITC

Sample loading on the VP‐ITC involves the use of a custom air‐tight Hamilton syringe with a protective plastic sheath around the metal needle. Check that this sheath extends very slightly (∼1 mm) beyond the tip of the metal needle to avoid scratching the surface of the ITC cell. Approximately 2 ml of the macromolecule and 0.5 ml of the ligand solutions are required for the filling processes for the cell and pipette, respectively. Also keep aside some of the final macromolecule dialysis buffer (10–20 ml) for control titrations and rinsing the cell during sample loading.

  • The correct technique for sample cell filling is important to avoid introducing air bubbles that will impact the quality of the titration experiment. While the final addition of the macromolecule solution is most critical, using the following technique is advisable each time liquid is placed in the sample cell: (i) Fill the Hamilton syringe with liquid (∼2 ml), then invert and tap to move air bubbles to the top of the barrel and expel them along with a very small amount of the liquid. (ii) Place the loaded Hamilton syringe carefully into the sample cell opening and lower until the tip gently touches the bottom of the cell. Raise the Hamilton syringe very slightly so that it is no longer touching the cell surface and then depress the plunger to fill the cell using a steady flowrate.
  • Using the method described above, wash out the sample cell two to three times with water and then rinse two to three times with buffer (from the final dialysis). The sample cell is the more central of the two cell openings and is surrounded by a plastic overflow reservoir.
  • Carefully load the macromolecule solution into the sample cell using the same technique. The final ∼0.3 to 0.5 ml should be added in several short, rapid bursts that will dislodge any small air bubbles from the top of the cell. Lift the Hamilton syringe fully out of the sample cell and then slowly reinsert into the same opening at a sight angle until the tip touches a ledge located near the top of the opening (where the cell stem contacts the plastic overflow reservoir). Draw up any liquid above this ledge by pulling up the Hamilton syringe plunger.
  • Fill the reference cell with distilled water via the second opening adjacent to the sample cell. When filled with water, the reference cell does not need to be emptied and refilled between experiments; a newly refilled reference cell should be good for several days. However, if the experimental sample contains solvents such as DMSO, the reference cell should be filled with the same percentage of solvent in distilled water, or if the experimental buffer is of a particularly high ionic strength then the buffer may be used to fill the reference cell. This is necessary to maintain the heat capacities of the solutions in the reference and sample cells at a sufficiently close value. If reference solutions other than water are used, the reference cell should be emptied and thoroughly rinsed with distilled water immediately after the experiments are completed.
  • Next, prepare the automated injection pipette by attaching a plastic syringe to the pipette fill port using clear tubing. Place the tip of the pipette needle into a microfuge tube filled with distilled water and use the attached plastic syringe to draw through the water after opening the fill port (note that opening and closing the pipette fill port is controlled by the instrument software). Repeat with the experimental buffer and then draw air through the pipette to remove all liquid. The plastic syringe can be removed and emptied between these steps if necessary.
  • To fill the pipette with ligand solution, place the needle into a microfuge tube containing the ligand and draw the solution into the injection pipette until it is completely filled (i.e. when a small amount of ligand solution is visible in the tubing attached to the fill port). Close the pipette fill port, detach the tubing and plastic syringe and then purge/refill the pipette (using the instrument software) to remove any bubbles. Remove the ligand solution tube and gently dab the needle with a Kimwipe to remove any residual liquid from the outside of the pipette. Finally, place the injection pipette carefully into the sample cell.

Identifying the optimal experimental parameters for a given system is likely to take some trial and error. However, the following guidelines and suggestions should provide a suitable starting point for most experiments.

  • The temperature at which the experiment will be performed is an important consideration (and as discussed elsewhere can be used in optimising experiments for challenging systems to study). While the VP‐ITC can be run at temperatures between 2 and 80 °C, a first experiment at 25–37 °C is most common. Factors that might influence the final choice of experimental temperature include ensuring optimal protein activity, solubility and/or stability, best matching the biologically relevant temperature for the system under study or the need to correlate ITC analyses with data from other experiments performed at a specific temperature.
  • The number and volume of injections are also important (and coupled) variables for the experimenter to consider. A typical first experiment on a new system the VP‐ITC will likely use a relatively low number of injections of higher volume, e.g. 25–30 × 10 μl. The first injection is often inaccurate due to mixing of the macromolecule and ligand solutions at the pipette tip during the relatively long instrument baseline equilibration process before the titration begins. The first injection is therefore often set to a smaller volume (e.g. 2 μl) and this data point is subsequently ignored in the fitting process. Some experiments may also be more optimally performed with a larger number of small injections (e.g. 70 × 4 μl), including interactions with a strong heat signal or where binding is weak (low c range) as a larger number of injections will give more data points for accurate fitting. For systems with intrinsically low heat signals, fewer injections (minimally 10–15 for a reasonable binding isotherm) of larger volume may help, combined with increasing sample concentrations or a testing effect of altering temperature or solution conditions, e.g. pH or ionic strength.
  • Most other instrument settings can be used at their default values, at least for a first experiment on a new macromolecule–ligand system. These include the reference power, stirring speed, and spacing and duration of each injection. In optimising ITC experiments the user can choose to increase spacing of injections if needed: it is critical that the system has sufficient time to equilibrate so that the heat signal returns to its baseline value before the next injection begins. Alternatively, for experiments with many small volume injections, where equilibration is rapid, the total time for the experiment can be significantly reduced by decreasing the injection spacing. Stirring speed (default 310 rpm) can be reduced if there are any indications of denaturation (e.g. of protein samples) or increased for efficient mixing of more viscous samples (e.g. containing glycerol). The optimal reference power depends on the nature (exothermic versus endothermic) and scale of heats generated by binding while the duration of injection has only minimal impact on the peak profile and is usually left at its default value (0.5 μl/s).

At the end of the titration, the macromolecule–ligand mixture can be retrieved from the sample cell using the Hamilton syringe. Clean out the ITC cell with water and repeat the procedure for preparing the cell for sample loading to perform additional experiments. Ideally, at least one or two repeats of each macromolecule–ligand titration should be performed for reproducibility, in addition to any control experiments of ligand solution into buffer. At the end of a set of experiments (or between each experiment if the sample is heavily precipitated), clean the VP‐ITC by manually filling and emptying the cell several times with detergent solution (e.g. 5% Contrad‐70) and then rinsing using distilled water. Alternatively, the ThermoVac system, if available, can be used to flush larger volumes of these solutions through the cell. For more vigorous cleaning the cell can be filled with detergent of a higher concentration (up to 20% Contrad‐70), for an extended time (60 minutes) and/or at an elevated temperature (65 °C).

12.3.3.2 Sample Loading and Experiment Set Up on an Auto‐iTC200

The Auto‐iTC200 instrument (and other semi‐ or fully automated systems) greatly improves the consistency of ITC experiments by automating sample loading, which is one of the most challenging aspects of performing an experiment for novice users. As with the VP‐ITC, the software supplied with the instrument is used to define the experimental parameters and control the running of each ITC titration, but an additional benefit is the ability to queue many experiments for unattended operation. Samples are supplied in up to four 96‐well deep well blocks and the cell can also be filled using a macromolecule solution from five 30 ml tubes. While this means that the Auto‐iTC200 could, in principle, run hundreds of experiments without user intervention, more commonly experiments run in smaller groups with periodic vigorous cleaning (which can also be automated) and manual checking of data quality.

  • Each titration experiment on the Auto‐iTC200 requires 400 μl of macromolecule and 120 μl of ligand solutions. These volumes are needed for the automated liquid handling system to accurately fill the cell and automated pipette, respectively. They are significantly higher than the actual volumes used in the experiment (approximately 200 μl in the cell and 40 μl in the pipette) and also more than is required to manually load a standalone iTC200 instrument.
  • Configuring the system software with the planned experiment organisation (number of sample groups and number of experiments in each group) and selecting the desired run parameters before loading the sample block(s) is advisable. While sample loading on the Auto‐iTC200 is greatly simplified by using 96‐well deep well blocks it is, of course, imperative than each sample solution or buffer is placed in its correct well. The final tab of the experiment setup shows a ‘map’ of the block and allows the user to select the site of the first sample (all subsequent sample positions are updated based on that selection).
  • In the simplest setup, each titration uses only two samples, the macromolecule and ligand. However, up to four wells per titration will be needed if the experimental protocol includes a pre‐rinse of the cell with buffer and the user selects to save the final macromolecule–ligand mixture to a clean empty well. Figure 12.3 shows examples of 96‐well block maps for each of these scenarios with a small number of experiments.
  • Once the complete set of samples is mapped on to the block, each solution is simply dispensed into its correct position, taking care to not introduce air bubbles, and the block is sealed with an adhesive film. Specialised films should be used that are designed for the Auto‐iTC200's liquid handling cannulas to pierce to access the solution. The sealed block is then placed in the correct position in the temperature‐controlled storage tray (4–25 °C).
  • The experimental parameters that can be customised for each experiment are the same as those described for the VP‐ITP: total number, volume, duration and spacing of injections, experiment temperature (4–40 °C), reference power and stir speed. However, the defaults for these user optimisable settings are specific to the Auto‐iTC200 with its smaller capacity cell and pipette. A reasonable starting point for a new experiment on an Auto‐iTC200 is to use 16 injections of 2.5 μl (with default duration and spacing), with a reference power of 10 μCal/s and stir speed of 1000 rpm.
  • A system cleaning procedure with water is performed automatically between each experiment. Protocols offering more rigorous cleaning of the cell and/or pipette can also be selected for some or all experiments and are recommended to be performed periodically even with well‐behaved samples (e.g. every 5–10 titrations).
Image described by caption and surrounding text.

Figure 12.3 Example block layouts for an Auto‐iTC200 run. The final tab in the experiment setup of the Auto‐iTC200 software shows a map of the sample locations based on user‐provided details of sample number and experiment type. This map should be carefully reviewed to ensure correct order of sample loading. The well for the macromolecule sample of the first titration in any run is user‐selectable on this screen; all remaining sample locations are then mapped to the subsequent wells. Two examples are shown: (left) an experiment with 2 groups of 6 titrations (12 experiments in total) where the macromolecule and ligand in the cell at the end of the experiment will be discarded; (right) a more complex method involving a pre‐rinse of the cell with the experiment buffer and saving the final cell solution to a new well at the end of each titration. It is important to note that regardless of the method selected and the order in which the reagents are used, the macromolecule and ligand sample are always placed in the first two positions for each titration. Once the samples are loaded and the run started, the complete set of experiments will be performed by the instrument without further input from the user.

12.3.3.3 Dealing with Ligand ‘Heats of Dilution’

Towards the end of the ITC experiment, the final few additions of ligand should result in relatively small peaks of equal height corresponding to the ‘heat of dilution’. This heat arises from the mechanical mixing of the ligand into a sample containing the macromolecule already fully saturated by ligand binding. As discussed in more detail above, close matching of the macromolecule and ligand solutions is essential to avoid large heats of dilution that could mask the actual binding signal. These non‐binding related heats need to be subtracted from each of the points in the macromolecule–ligand titration before further data analysis. This can be accomplished in several ways:

  1. Perform a full titration control experiment (i.e. with the same number, volume and spacing of injections) of ligand solution into buffer and then subtract the integrated peak values from this control from those in the experiment during data analysis (see Section 12.3.4). This process can be automated within the Auto‐iTC200 software by designating control and experimental titrations so the reference run is subtracted from the macromolecule–ligand titration during the automated analysis.
  2. Perform a separate, more limited titration of ligand into buffer to calculate a single average value for the corresponding heat of dilution and then manually subtract this value from each point in the experimental titration. This can be accomplished using the Simple Math option in the Origin software supplied with the VP‐ITC or Auto‐iTC200.
  3. A reasonable estimate of the heat of dilution can also often be obtained directly from the experimental (i.e. ligand into macromolecule) titration by averaging the heats associated with the final few (e.g. two or three) injections. Subtraction of this value from the experimental heats can then be accomplished as for approach 2 above. This method of estimating the heat of dilution makes the critical assumption that the system has essentially reached the point of saturation of macromolecule–ligand binding, which might not always be the case, e.g. where binding is weak or the concentrations of macromolecule and ligand used were not optimal. However, this method can be very useful if a ligand into buffer titration is not performed for some reason or if the heats of dilution obtained by the other approaches above are inconsistent with final heats in the experimental titration. (It is important to keep in mind that dilution of ligand into buffer is not precisely the same as dilution into solution containing the macromolecule, especially when relatively high concentrations are being used.)
  4. Finally, more recent software supplied with some instruments (or developed independently by other researchers) allows heats of dilution to be fitted directly in the data analysis step. When available, this approach is likely to provide the simplest and most robust way to accurately account for these additional heats.

After heats of dilution have been accounted for by one of these approaches, the enthalpy of binding at saturation should approach zero and the binding isotherm can be fitted using an appropriate model to determine binding parameters (see Section 12.3.4 below).

12.3.3.4 Troubleshooting Some Common Issues

Several problems with ITC experiments are possible that can directly impact data quality. Some of these may be more common in one type of instrument than another, e.g. through the introduction of air bubbles into the sample cell due to a poor cell filling technique on a manual instrument that is less likely with an automated system (though these can have their own sample loading issues if not well maintained). Other problems with data quality may have their origin in problems with the samples themselves. Fortunately, many of these common problems give rise to identifiable signatures in the resulting titrations (Figure 12.4) that can point the user in the right direction to resolve them.

Image described by caption and surrounding text.

Figure 12.4 Signatures of some common problems in ITC experiments. (a) Schematic of an ideal titration in which initial large peaks of approximately equal size become smaller in the middle of the experiment (as binding sites become progressively more occupied) and finally ending with several peaks of equal value close to zero, corresponding to heats of dilution after saturation of binding. Poor‐quality titrations may be indicative of (b) mismatch of macromolecule and ligand solutions resulting in large final peaks and a small difference (Δ) between peaks at the start and end of the experiment, (c) too high ligand concentration in the syringe relative to the macromolecule in the cell leading to rapid saturation of the binding signal, (d) too low ligand concentration in the syringe relative to the macromolecule in the cell meaning that binding is not saturated before the end of the titration, (e) slow binding events or inadequate spacing (s) of injections, (f) air bubbles introduced by a poor cell filling technique or spontaneous formation during the experiment, (g) a change in heat capacity of the system (e.g. enzymatic reaction or potentially a mismatch of buffers), (h) a slow equilibrium (e.g. degradation or aggregation) or other processes in addition to binding, or (i) two or more simultaneous processes occurring upon mixing with opposite heat effects.

An ideal titration will begin a series of approximately equal, large peaks that become reduced as binding is saturated with further injections of ligand during the experiment. The final peaks, corresponding to the heats of dilution of ligand after binding is fully saturated, should be close to zero (Figure 12.4a). Underfilling of the cell or pipette will lead to erratic peaks and a shifting baseline that will likely make data unusable. A strongly downward stepping or shifting baseline may be indicative of an underfilled pipette, while upward changes may point to an underfilled or dirty cell. Such instrument issues are not likely to be common on a well‐maintained and clean instrument but can happen on an Auto‐iTC200 instrument if some of the components or fine tubing needed to move the samples around become dirty or clogged. After ensuring that the instrument has sufficient supplies of water, detergent cleaning solution, methanol and N2 gas, watching the processes of cell/pipette loading and cleaning should pin‐point any problem areas.

Wildly oscillating or moving baselines, long times for equilibration of baseline or baseline values far from the set reference power may be resolved by thorough cleaning of the sample cell but could also be signs that the instrument has more serious issues that require expert service.

Other issues with sample quality or experiment design may also manifest in the following ways:

  • Figure 12.4b shows large heats of dilution at the end of the titration. Large mismatches of the macromolecule and ligand buffer solutions will result in large heats of dilution that can potentially swamp the genuine binding heats. Check a buffer–buffer titration to confirm and improve buffer matching, ideally by exhaustive codialysis. A similar issue will arise on the Auto‐iTC200 if the methanol used to dry the tubing after system cleaning is not adequately removed before the next run. Check the tubing for blockages, ensure the system is being supplied with N2 gas used to blow out residual methanol and/or watch the sample loading and system cleaning processes for signs of problems with liquid handling.
  • Figure 12.4c shows binding that is too rapidly saturated – a very small number of large peaks are followed by many small peaks. This may arise from inappropriate choice(s) of ligand and/or macromolecule concentration such that the relative of concentration in the pipette is too high (»10 times more than the cell concentration for an expected 1 : 1 binding). Check concentration measurements and either increase macromolecule concentration or decrease ligand concentration. Also consider the possibility that the binding stoichiometry may not be as expected, meaning that the relative concentrations of ligand and macromolecule should be adjusted (e.g. ligand concentration 20 times or more than the macromolecule for 2 : 1 stoichiometry).
  • Figure 12.4d shows binding that fails to saturate – small equal peaks corresponding to heats of dilution are not observed before the end of the titration. This may arise for the same reasons as described above for too rapid saturation when relative concentration of the ligand is too low («10 times more than the macromolecule concentration for an expected 1 : 1 binding).
  • Figure 12.4e shows peaks that do not return to the baseline between experiments. This will occur if binding results in broad peaks or there is inadequate spacing of injections used in the titration method (note the difference in spacing, s, in Figure 12.4a and c).
  • Figure 12.4f shows random spikes that appear during the titration. Air bubbles can cause spurious peaks during the experiment and may impact data quality depending on their number and location: spikes in the baseline regions can simply be ignored during peak integration and discarding a small number of ligand injection peaks impacted by such artefacts should not greatly reduce data fitting quality. Air bubbles could arise due to the poor cell filling technique with a manual instrument (see Section 12.3.3.1) or from pipette filling issues on the Auto‐iTC200. In the latter case, inspect the pipette and watch the automated filling process for signs of problems during docking with the fill port adaptor, leaks from the pipette after filling, etc. Bubbles can also form spontaneously if samples are not adequately degassed, particularly when there is a significant difference in temperature between the sample storage and experiment.
  • Figure 12.4g shows an evenly stepping baseline. This can occur when there is a change in heat capacity of the system, e.g. if an enzymatic reaction is occurring in the cell or potentially through a mismatch of the solutions containing the macromolecule and ligand. Some modification to the molecules or solutions containing them in the ITC experiment may be needed.
  • Figure 12.4h shows initially sharp peaks but with a long ‘tail’ before returning to the baseline. Binding of the macromolecule and ligand may be accompanied by a second slow process (e.g. degradation or aggregation) in addition to binding. Alteration of the samples themselves (macromolecule or ligand) or the solution conditions may be needed; first, check for aggregation before and after the experiment (e.g. by visual inspection or using DLS or gel filtration chromatography).
  • Figure 12.4i shows peaks that are followed by an ‘overshoot’ in the opposite direction before returning to baseline. One or more processes with opposite heat effects may be occurring upon mixing in addition to macromolecule–ligand binding. This situation could also arise due to a dirty sample cell, which is easily resolved by following a procedure for thorough cell cleaning with detergent.

12.3.4 ITC Data Analysis and Fitting to a Binding Model

Having performed a complete titration of ligand into macromolecule, along with any associated control experiments, the next steps are to perform initial adjustments of the raw data (e.g. baseline) and then to fit the measured heats to an appropriate model for the binding reaction, e.g. a single set of identical sites, two sets of independent sites, sequential binding, etc. A more detailed description of this process and the binding models for fitting ITC data can be found elsewhere (e.g. [16]). This fitting procedure allows the direct determination of binding affinity (K d ), stoichiometry (n) and enthalpy (ΔH) of binding with the goal of modelling the experimental data with the least error and using the simplest model that is in accordance with any prior information on the system under study. Free energy and entropy of binding (ΔG and ΔS, respectively) can then be calculated according to the equations of Section 12.2.3. Such data analysis and fitting of binding to one of several possible models is typically done in specially modified software supplied with the ITC instrument. With the VP‐ITC and Auto‐iTC200 this analysis is done with a customised version of Origin software.

  • First, open the analysis software, e.g. using the MicroCal Data Analysis Launcher, and load in the desired raw .itc titration file(s). For the VP‐ITC, a single file is loaded with the Read Data button located in a menu on the top left of the screen; Auto‐iTC200 data can be loaded as single titrations using the iTC200 option or as multiple files using the Auto‐iTC200 option from the launcher menu. The descriptions below cover the basic steps for analysing a single titration on either instrument.
  • After opening the titration file, the Origin software will display the raw thermogram and perform an initial automated analysis of the data, including baseline fitting and peak integration, before presenting a view of the integrated ITC data in the deltaH window. Inspect the data points for signs of any potential problems such as significantly outlying values, which should be investigated in the raw titration data.
  • Go to the raw titration data by switching to the RawITC window. Inspect the automatic fit baseline and integrated peaks for signs of any problems arising from air bubbles, abrupt baseline changes that are not well accounted for or other possible artefacts (see Section 12.3.3.4). A menu is displayed in the top left of the screen that allows the user to manually adjust the baseline points and the integration window around the peak. To do this, select Adjust Integrations and click on the first peak where adjustments are needed. A zoomed view of the peak base and fit baseline will be displayed along with new options to move the baseline points and reintegrate the peaks. The integration window can also be narrowed (e.g. to remove noise or artefacts outside of the peaks) by moving the blue vertical lines displayed on each side of the peak. Adjusting the baseline and integration window can often significantly improve the quality of the peak integrations and therefore quality of the final fit to the data. However, care should be taken in using these options as such decisions can be quite subjective and effort should be made to apply any changes consistently to each of the peaks in the thermogram.
  • Return to the deltaH window and load the ligand–buffer control titration if this method is to be used to account for heats of dilution (see Section 12.3.3.3). Inspect the control titration and, if necessary, adjust as described above and then use the option in the deltaH window menu to Subtract Reference Set.
  • Two other options on the deltaH window menu can be used at this point, if needed. Modify Concentrations can be used to adjust the concentration of the macromolecule or ligand, e.g. if analysis performed after the experiment was started revealed that the final concentrations differed from those entered in the instrument control software. This option can also be used as a starting point for further investigation if initial fitting of the data suggests the active concentration of the macromolecule and/or ligand was not as expected, e.g. a significantly non‐integer value of n is obtained. Finally, use Remove Bad Data Point to remove the first data point in titrations where a smaller injection volume was used as well as any other point(s) where problems cannot be satisfactorily resolved using manual adjustment of the baseline and integration window. This step is performed last as further modifications of the baseline, integrations or concentrations will reset the plot to include the deleted points.
  • If using Simple Math to account for heats of dilution (see Section 12.3.3.3 ), perform an initial correction at this point. Take careful note of the sign and value entered as this change is permanently applied to all integrated data points.
  • Next, select the data fitting model: One Set of Sites, Two Sets of Sites, Sequential Binding Sites, Competitive Binding or Dissociation. In the absence of other information, begin with the simplest option for one binding site. Clicking on the desired binding model will open a new dialogue window in which initial values of the fit parameters will be displayed and can be adjusted if information from other methods is available. There is also an option to fix one or more of these values, if desired. For the one binding site model, the parameters displayed will be: stoichiometry (n), enthalpy (ΔH) and binding affinity (K a ). Click the button for single or multiple iterations of the fitting procedure and repeat until the fit has converged, i.e. when fit values and Chi2 no longer change.
  • Return to the deltaH window where the data, fit curve and binding parameters will be displayed. Inspect the quality of the fit to the integrated heats and make any further necessary adjustments, as described above. In particular, if using Simple Math to account for heats of dilution, check for systematic error in the fit to the final points, which would suggest too low or too high a value was entered. After any further adjustments, repeat the fit to the binding model. Finally, the MicroCal Origin software has a feature that allows the user to quickly generate a publication quality figure of the ITC experiment using the Final Figure option from the Analysis menu.

12.4 Example Applications of ITC to Analysis of Biomolecular Interactions

12.4.1 Case Study 1: Enthalpy versus Entropy Driven Transcription Factor (TF) Binding to Distinct DNA Sequences

Interaction of transcription factors (TFs) with their DNA binding sites is a key regulatory process in gene expression and an essential activity in living cells [17]. TFs directly control RNA transcription from the DNA genetic material by promoting (activator TF) or blocking (repressor TF) recruitment of RNA polymerase to the DNA [18]. Through these activities, TFs ensure both the correct level of expression in the correct cell type in multicellular organisms. TFs can act as either homodimeric complexes or as monomers, but always contain at least one DNA binding domain (DBD) that is responsible for directing specific interaction of the TF with its associated target DNA.

TF interaction with DNA can be driven by so‐called ‘direct read‐out’ through formation of direct hydrogen bonds and hydrophobic interactions between protein and DNA, as well as ‘indirect read‐out’ of DNA shape and electrostatics. These contacts can also be made directly between the TF and DNA or mediated by water molecules, and each can therefore make distinct contributions to the enthalpy and entropy of binding. The ability of TFs to bind to multiple related sequences with biologically relevant affinities has been rationalised based on ‘enthalpy–entropy compensation’ arising from relatively subtle changes in the contributions of these different types of interaction. However, it is less clearly understood how some monomeric TFs are apparently able to bind more than one distinct DNA sequence, which would require more significant adjustment in the number and type of interactions necessary to maintain binding. To unravel the basis of this phenomenon in molecular detail, Morgunova et al. [19] performed a detailed mechanistic study of DNA binding by the posterior homeodomain protein HOXB13 and several other monomeric TFs, using X‐ray crystallography, molecular dynamics (MD) simulations and ITC. These studies used two distinct high affinity DNA sequences CTCGTAAA (DNATCG) and CCAATAAA (DNACAA), which differ by the underlined nucleotides and to which HOXB13 binds with similar affinity despite their significant sequence difference.

X‐ray crystal structures of HOXB13 bound to each DNA sequence revealed little difference in the HOXB13 protein structure in each complex, while the two DNAs differed in the bending of their backbone at the location of the distinguishing nucleotides (Figure 12.5a). Additionally, there were few differences observed in HOXB13's interactions with each DNA (Figure 12.5b). Together with extensive mutagenesis experiments, these structures suggested that the differences in protein–DNA interactions or commonalities in DNA structure cannot fully explain the dual sequence preferences of HOXB13 and other TFs. Rather, the major difference between the structures appeared to be the role of bridging water molecules in the interaction of HOXB13 with each DNA. Specifically, water molecules in the DNACAA complex appeared more stably bound and making interactions that were likely to correspond to stronger enthalpic contributions. In contrast, water molecules in the HOXB13‐DNATCG complex appeared significantly more disordered, suggesting a high entropy state. These ideas were further supported by MD simulations and free energy perturbation calculations. However, to directly test their hypothesis of distinct enthalpic and entropic minima driving TF binding to distinct DNA sequences, the authors turned to ITC analysis with its power to provide a complete thermodynamic profile of HOXB13 binding.

Image described by caption and surrounding text.

Figure 12.5 Enthalpic and entropic minima control HOXB13 binding to two distinct DNA sequences. (a) Superposition of the HOXB13 DBD bound to DNATCG (tan protein/blue DNA) and DNACAA (red protein/green DNA). The two protein structures are essentially identical with a root mean square deviation of 0.81 Å for the 57 residues shown. DNA residues that differ between the DNAs are shown in orange and exhibit a distinct bending of the DNA backbone at this site. (b) Schematic of interactions formed between HOXB13 DBD and the two DNA sequences. Interactions with DNA backbone phosphate/deoxyribose and bases are indicated with dashed and solid lines, respectively. Shading denotes the regions with distinct DNA sequence. The TCG site lacks direct contacts to the DNA bases, whereas the CAA site is recognised by direct contacts by Gln‐265 and Ile‐262. Most other contacts are similar in both structures. (c) ITC analysis of HOXB13‐DNA interactions for the two distinct sequences and (d) resulting thermodynamic parameters of binding. While both K d and free energy are essentially identical for both, the DNA sequences differ strongly in the ΔH and TΔS contributions to binding. (e) Model for the presence of two distinct optimal (low ΔG) DNA binding sequences, one driven by low enthalpy (left) and the other by high entropy (right), distinguished by differences in ordering of water molecules (blue spheres) and thus the interactions they mediate.

Source: This figure is adapted from Morgunova E. et al., eLife, e32963, 2018 [19].

For use in ITC experiments, HOXB13 was expressed in Rosetta 2(DE3)‐pLysS Escherichia coli cells using an auto‐induction protocol and purified by Ni2+‐affinity chromatography [ 19,20]. DNAs were chemically synthesised by a commercial supplier as single strands and the complementary sequences mixed and annealed to form each double‐stranded DNA. Both HOXB13 and each DNA were prepared in a solution containing 20 mM HEPES buffer pH 7.5, 300 mM NaCl, 10% glycerol and 2 mM TCEP, and titration experiments were performed on an iTC200 microcalorimeter at the Protein Science Facility at Karolinska Institute, Sweden. Each experiment was performed at 25 °C and involved 23 injections of HOXB13 ‘ligand’ (150 μM) into DNA solution (12–16 μM) in the sample cell. All data were analysed using the customised Origin software supplied with the instrument and binding parameters (n, K d , ΔH) determined by a non‐linear least square fit of the data to the one binding site model as implemented in the package. The TΔS and ΔG of binding were obtained using Eq. 12.5. Each titration was performed in triplicate and average values for the determined binding parameters were reported and reproduced here.

ITC titrations of HOXB13 into each DNA revealed a very similar affinity and free energy of binding for both sequences, as anticipated (Figure 12.5c and d). The data also revealed, however, the strongly differing contributions of enthalpy and entropy in each case. For DNACAA, ΔH was significantly increased whereas ΔS was strongly reduced compared to DNATCG (Figure 12.5d). The authors were thus able to draw the conclusion from these direct experimental measurements that HOXB13 binding to one DNA sequence (DNACAA) represents an enthalpically driven minimum in binding free energy while the other (DNATCG) an entropically driven minimum (Figure 12.5e). This novel finding also demonstrates the power of ITC to provide a complete picture of binding thermodynamics to complement other biophysical approaches like X‐ray crystallography in understanding the molecular basis of biomolecular function.

12.4.2 Case Study 2: Identification of a New Motif Essential for Co‐substrate S‐adenosyl‐L‐methionine (SAM) Binding by an rRNA Methyltransferase

RNA post‐translational modifications such as nucleobase or ribose sugar methylations occur in many different types of RNA in both prokaryotes and eukaryotes. In ribosomal RNA (rRNA), for example, such modifications are critical for correct RNA folding and ribosome subunit assembly, as well as ribosome function in protein synthesis [21,22]. Additionally, rRNA methylation can directly block the activity of ribosome‐targeting antibiotics and is an important contributor to the escalating problem of bacterial resistance to antibiotics such as macrolides and aminoglycosides [2325]. In some other rarer cases, specific rRNA modifications can also be required for proper antibiotic binding and therefore antimicrobial activity. One example of this scenario is the ribosome‐targeting antibiotic capreomycin, which requires the ribose 2′‐O‐methylations incorporated by the 16S/23S rRNA methyltransferase TlyA for its antimicrobial activity [26]. Given the ubiquity of RNA modifications in biology and their important contributions to current biomedical problems, such as antibiotic resistance, identifying the enzymes responsible and defining their mechanisms of action, is an active area of investigation. An important aspect of such mechanistic understanding of rRNA methyltransferase enzyme activity is defining how they bind their obligatory co‐substrate for the methylation reaction, S‐adenosyl‐L‐methionine (SAM). In a recent study from our lab on the capreomycin resistance‐related rRNA methyltransferase TlyA from Mycobacterium tuberculosis, we used ITC to define the role of a new protein motif critical for SAM binding by this enzyme [27].

At the outset of our study on TlyA, a key goal was to determine a high resolution crystal structure of the intact enzyme. However, when crystals of the full‐length protein could not be obtained, we used a collection of proteases to identify stable TlyA fragments than might be more amenable to crystallisation. Limited proteolysis with the endopeptidase GluC, which selectively cleaves the polypeptide backbone to the C‐terminal side of accessible glutamic acid residues, identified one such stable fragment. Further analyses suggested that GluC cleavage occurred primarily at glutamic acid 59 (E59), immediately adjacent to a short tetrapeptide linker between the TlyA N‐terminal domains (NTD) and C‐terminal domains (CTD), …59E↓RAWVS64…, where the site of cleavage is denoted by the arrow and the underlined sequence is the interdomain linker. A new construct was designed for expression of the isolated TlyA CTD (residues 64‐268) beginning with the first residue of α‐helix 1 (α1) in the CTD. Using this new construct, a crystal structure of the TlyA CTD was determined at 1.7 Å resolution. The TlyA CTD adopts a canonical Class I methyltransferase fold [28], including all the expected conserved motifs required for SAM binding. Despite this, we were unable to obtain a TlyA‐CTD:SAM complex structure by co‐crystallisation or soaking of preformed TlyA CTD crystals. We therefore used ITC to gain complementary insights into SAM binding by TlyA.

For these ITC analyses, 6× His‐tagged wild‐type TlyA and TlyA CTD proteins were expressed in E. coli BL21 (DE3) and purified using His‐Trap HP and Superdex 75 16/60 columns on an ÄKTA Purifier FPLC for Ni2+‐affinity and gel filtration column chromatography, respectively. TlyA proteins were exhaustively dialysed against 50 mM Tris buffer (pH 7.5) containing 120 mM NaCl and 10% glycerol and concentrated using a centrifugal concentration device to 60–100 μM. The final dialysis buffer solution was used to resuspend SAM at 1.5 mM final concentration. As binding was expected to be relatively weak, ITC experiments were performed at the highest practically achievable protein concentration and much higher ligand (SAM) concentration to drive binding towards saturation at the end of the titration (i.e. a ‘low c’ ITC experimental design). All ITC experiments were performed at 25 °C and used 16 injections of 2.4 μl of SAM into the cell containing the protein. After subtraction of residual heats derived from the average values of the final two or three injections, ITC data were fit to the model for one‐binding site. The equilibrium dissociation constants (K d ) reproduced here (Table 12.1) are the averages of at least three experiments with their associated standard deviation.

Table 12.1 SAM binding by TlyA proteins. a

Binding affinity
Protein K d (μM) Fold‐change b
TlyA 23.4 ± 2.9
CTD No binding
RAWVCTD 20 ± 1.1
TlyA‐R60A 87.2 ± 8.5 3.6
TlyA‐R60E 62.8 ± 23 2.7
TlyA‐A61V 21.1 ± 4.3 1.0
TlyA‐W62A 234 ± 56 10
TlyA‐W62F 98.5 ± 27 4.3
TlyA‐V63A 470 ± 19 20

a Data are from Witek et al., J. Biol. Chem., 2017 [27].

b Relative to wild‐type TlyA, indicated for RAWV linker substitutions only.

Wild‐type TlyA bound SAM with low micromolar affinity as expected (Figure 12.6a and Table 12.1), comparable to many other rRNA methyltransferase enzymes. In contrast, the isolated TlyA CTD protein failed to show any binding to SAM (Figure 12.6b), despite its crystal structure showing all expected conserved SAM binding motifs being present. While this result was surprising, it was consistent with our inability to obtain a TlyA CTD‐SAM co‐crystal structure. TlyA cleaved by GluC was examined next, both as the cleaved NTD/CTD mixture and the isolated CTD purified by gel filtration chromatography. Both samples were shown by ITC to retain the wild‐type SAM binding affinity (data not shown). Together, these results suggested that some element outside the CTD is critical for SAM binding and a new protein expression construct was created to express the CTD protein including the four amino acids of the interdomain linker at its N‐terminus (residues 60‐268; RAWVCTD). Remarkably, addition of the RAWV tetrapeptide sequence fully restored SAM binding to the wild‐type TlyA affinity (Figure 12.6c and Table 12.1).

Image described by caption and surrounding text.

Figure 12.6 Identification of a novel motif critical for TlyA methyltransferase binding to SAM. ITC analysis of SAM binding to (a) full‐length N‐terminally 6× His‐tagged TlyA (His‐TlyA), (b) the isolated TlyA C‐terminal Class I methyltransferase domain (CTD) and (c) the same isolated CTD but with the four‐amino acid (RAWV) TlyA interdomain linker appended on its N‐terminus (RAWVCTD). (d) Crystal structure of the RAWVCTD protein with the RAWV interdomain linker in its ‘helix’ conformation. The zoomed region shows the RAWV linker in 2mFo‐DFc omit electron density contoured at 1σ. (e) Overlay of the RAWVCTD structures with RAWV interdomain linker in its ‘helix’ (cyan) and ‘loop’ (tan) conformations.

Source: This figure is adapted from Witek et al., J. Biol. Chem., 2017 [27].

These results suggested that the RAWV sequence of the TlyA interdomain linker might form a critical but previously unappreciated motif necessary for SAM binding by the TlyA family of rRNA methyltransferases. Site‐directed mutagenesis was used to make one or two substitutions of each of the RAWV amino acids and the impact of these changes were tested by measuring SAM binding by ITC (Table 12.1). Consistent with sequence analysis of the RAWV tetrapeptide, which showed W62 and V63 to be most strongly conserved, changes at these two residues had the greatest impact on SAM binding affinity. X‐ray crystallographic studies of the TlyA RAWVCTD protein revealed that the RAWV tetrapeptide is able to adopt two distinct conformations: a ‘helix’ form that extends α1 at the beginning of the CTD or a ‘loop’ structure that closely corresponded to the conformation of the equivalent sequence in several hemolysin proteins that had previously been used for homology modelling of TlyA [29].

Interactions made by V63 and other changes in the SAM binding pocket, which were only observed in the RAWVCTD ‘helix’ structure, also offered insight into a potential molecular basis for the role of the new RAWV motif in SAM binding (Figure 12.6d and e). The newly identified ability of the RAWV interdomain linker to adopt two distinct structural conformations, as well as its critical importance for SAM binding as revealed by ITC, suggested that this short sequence may play a critical functional role in substrate recognition by TlyA. Specifically, the question of how TlyA can specifically recognise and modify the 2′‐OH of two cytosine residues within distinct structural contexts, 23S rRNA C1920 in the 50S subunit and 16S rRNA C1409 in the 30S subunit, may have its answer in a common substrate binding‐induced interdomain linker conformational change that is linked to SAM binding affinity and thus TlyA activity. Fully delineating the molecular mechanisms of TlyA and other rRNA methyltransferase activities will require further investigations, of which ITC will likely be a critical component.

12.5 Concluding Remarks

ITC is firmly established as one of the best choices of approach to study many different types of biomolecular interaction. With its power to fully define the thermodynamics of binding in a single experiment, ITC has found broad application in studies of basic biological processes as well as in drug discovery. Further advances in instrument technology and the associated software have also simultaneously reduced the previously high sample quantity demands and simplified the processes of performing and analysing ITC data. As such, ITC is an approach that should be among the first considered for any study that aims to understand the nature and role of a biomolecular interaction.

Acknowledgements

I thank Dr Debayan Dey and Mr Zane Laughlin (Emory University) and the anonymous reviewers for their helpful comments and suggestions on this chapter. Our research is currently supported by NIH/NIAID grant R01‐AI088025.

References

  1. 1 Velazquez‐Campoy, A., Leavitt, S.A., and Freire, E. (2004). Characterization of protein‐protein interactions by isothermal titration calorimetry. Methods Mol. Biol. 261: 35–54.
  2. 2 Velazquez‐Campoy, A., Ohtaka, H., Nezami, A. et al. (2004). Isothermal titration calorimetry. Curr. Protoc. Cell Biol. 17–18.
  3. 3 Klebe, G. (2015). Applying thermodynamic profiling in lead finding and optimization. Nat. Rev. Drug. Discovery 14: 95–110.
  4. 4 Claveria‐Gimeno, R., Vega, S., Abian, O., and Velazquez‐Campoy, A. (2017). A look at ligand binding thermodynamics in drug discovery. Expert Opin. Drug Discovery 12: 363–377.
  5. 5 Di Trani, J.M., De Cesco, S., O'Leary, R. et al. (2018). Rapid measurement of inhibitor binding kinetics by isothermal titration calorimetry. Nat. Commun. 9: 893. PMCID: PMC5832847.
  6. 6 Zhao, H., Piszczek, G., and Schuck, P. (2015). SEDPHAT‐‐a platform for global ITC analysis and global multi‐method analysis of molecular interactions. Methods 76: 137–148. PMCID: PMC4380758.
  7. 7 Mazzei, L., Ciurli, S., and Zambelli, B. (2016). Isothermal titration Calorimetry to characterize enzymatic reactions. Methods Enzymol. 567: 215–236.
  8. 8 Falconer, R.J. (2016). Applications of isothermal titration calorimetry ‐ the research and technical developments from 2011 to 2015. J. Mol. Recognit. 29: 504–515.
  9. 9 Turnbull, W.B. and Daranas, A.H. (2003). On the value of c: can low affinity systems be studied by isothermal titration calorimetry? J. Am. Chem. Soc. 125: 14859–14866.
  10. 10 Tellinghuisen, J. (2008). Isothermal titration calorimetry at very low c. Anal. Biochem. 373: 395–397.
  11. 11 Velazquez‐Campoy, A. and Freire, E. (2006). Isothermal titration calorimetry to determine association constants for high‐affinity ligands. Nat. Protoc. 1: 186–191.
  12. 12 Yennawar, N.H., Fecko, J.A., Showalter, S.A., and Bevilacqua, P.C. (2016). A high‐throughput biological calorimetry core: steps to startup, run, and maintain a multiuser facility. Methods Enzymol. 567: 435–460.
  13. 13 Malhotra, A. (2009). Tagging for protein expression. Methods Enzymol. 463: 239–258.
  14. 14 Bradshaw, J.M. and Waksman, G. (1998). Calorimetric investigation of proton linkage by monitoring both the enthalpy and association constant of binding: application to the interaction of the Src SH2 domain with a high‐affinity tyrosyl phosphopeptide. Biochemistry 37: 15400–15407.
  15. 15 Baker, B.M. and Murphy, K.P. (1996). Evaluation of linked protonation effects in protein binding reactions using isothermal titration calorimetry. Biophys. J. 71: 2049–2055. PMCID: PMC1233671.
  16. 16 Freyer, M.W. and Lewis, E.A. (2008). Isothermal titration calorimetry: experimental design, data analysis, and probing macromolecule/ligand binding and kinetic interactions. Methods Cell Biol. 84: 79–113.
  17. 17 Levine, M. and Tjian, R. (2003). Transcription regulation and animal diversity. Nature 424: 147–151.
  18. 18 Smith, N.C. and Matthews, J.M. (2016). Mechanisms of DNA‐binding specificity and functional gene regulation by transcription factors. Curr. Opin. Struct. Biol. 38: 68–74.
  19. 19 Morgunova, E., Yin, Y., Das, P.K. et al. (2018). Two distinct DNA sequences recognized by transcription factors represent enthalpy and entropy optima. eLife 7: PMCID: PMC5896879.
  20. 20 Jolma, A., Yin, Y., Nitta, K.R. et al. (2015). DNA‐dependent formation of transcription factor pairs alters their binding specificity. Nature 527: 384–388.
  21. 21 Polikanov, Y.S., Melnikov, S.V., Soll, D., and Steitz, T.A. (2015). Structural insights into the role of rRNA modifications in protein synthesis and ribosome assembly. Nat. Struct. Mol. Biol. 22: 342–344. PMCID: PMC4401423.
  22. 22 Decatur, W.A. and Fournier, M.J. (2002). rRNA modifications and ribosome function. Trends Biochem. Sci. 27: 344–351.
  23. 23 Wachino, J. and Arakawa, Y. (2012). Exogenously acquired 16S rRNA methyltransferases found in aminoglycoside‐resistant pathogenic gram‐negative bacteria: an update. Drug Resist. Updat. 15: 133–148.
  24. 24 Conn, G.L., Savic, M., and Macmaster, R. (2009). DNA and RNA Modification Enzymes: Comparative Structure, Mechanism, Functions, Cellular Interactions and Evolution (ed. H. Grosjean), 524–536. Austin, TX: Landes Bioscience.
  25. 25 Long, K.S. and Vester, B. (2009). DNA and RNA Modification Enzymes: Comparative Structure, Mechanism, Functions, Cellular Interactions and Evolution (ed. H. Grosjean), 537–549. Austin, TX: Landes Bioscience.
  26. 26 Monshupanee, T., Johansen, S.K., Dahlberg, A.E., and Douthwaite, S. (2012). Capreomycin susceptibility is increased by TlyA‐directed 2'‐O‐methylation on both ribosomal subunits. Mol. Microbiol. 85: 1194–1203. PMCID: PMC3438285.
  27. 27 Witek, M.A., Kuiper, E.G., Minten, E. et al. (2017). A novel motif for S‐adenosyl‐l‐methionine binding by the ribosomal RNA methyltransferase TlyA from Mycobacterium tuberculosis. J. Biol. Chem. 292: 1977–1987. PMCID: PMC5290967.
  28. 28 Schubert, H.L., Blumenthal, R.M., and Cheng, X.D. (2003). Many paths to methyltransfer: a chronicle of convergence. Trends Biochem. Sci. 28: 329–335.
  29. 29 Arenas, N.E., Salazar, L.M., Soto, C.Y. et al. (2011). Molecular modeling and in silico characterization of Mycobacterium tuberculosis TlyA: possible misannotation of this tubercle bacilli‐hemolysin. BMC Struct. Biol. 11: 16. PMCID: PMC3072309.

Further Reading

  1. Ladbury, J.E. and Doyle, M.L. (2005). Biocalorimetry 2: Applications of Calorimetry in the Biological Sciences. Wiley.
  2. Privalov, P.L. (2012). Microcalorimetry of Macromolecules: The Physical Basis of Biological Structures. Wiley.

Website Resources

  1. ITC experiment set up and analysis using a Microcal/Malvern Panalytical VP ITC instrument:
  2. https://www.jove.com/video/2796/isothermal‐titration‐calorimetry‐for‐measuring‐macromolecule‐ligand.
  3. Sample cell and syringe loading on a TA Instruments Nano ITC:
  4. https://www.youtube.com/watch?v=VNf9ujlHrkc.
  5. https://www.youtube.com/watch?v=EDgjtABjLc4.
..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset