Skip to main content
Biophysical Reviews logoLink to Biophysical Reviews
. 2016 Nov 4;8(4):269–277. doi: 10.1007/s12551-016-0227-5

Foreword to ‘Quantitative and analytical relations in biochemistry’—a special issue in honour of Donald J. Winzor’s 80th birthday

Damien Hall 1,2,, Stephen E Harding 3,
PMCID: PMC5425807  PMID: 28510020

Abstract

The purpose of this special issue is to honour Professor Donald J. Winzor’s long career as a researcher and scientific mentor, and to celebrate the milestone of his 80th birthday. Throughout his career, Don has been renowned for his development of clever approximations to difficult quantitative relations governing a range of biophysical measurements. The theme of this special issue, ‘Quantitative and analytical relations in biochemistry’, was chosen to reflect this aspect of Don’s scientific approach.

Keywords: Donald J Winzor, Physical chemistry, Analytical ultracentrifugation, Quantitative and analytical relations in biochemistry


In this special issue, Don’s colleagues, scientific collaborators and former students have contributed an array of articles concerned with biosensor analysis (Karlsson 2016), analytical ultracentrifugation (Harding et al. 2016), dynamic light scattering (Stetefeld et al. 2016), turbidimetric analysis (Zhao et al. 2016), protein–ribonucleic acid interaction (Jones 2016), computer modelling of amyloid formation (Schor et al. 2016), protein denaturation (Chebotareva et al. 2016), statistical mechanics (Wills 2016), structural phage biology (Arisaka et al. 2016), neutron scattering (Scott 2016), nuclear magnetic resonance analysis of mass transfer (Shishmarev and Kuchel 2016), electrophoresis (Laue et al. 2016), thermodynamic nonideality (Rao et al. 2016) and protein–DNA interactions (Munro et al. 2016). Prefatory comments have also been added by two long-time research collaborators and fellow passengers through time, Professor Emeritus William H. Sawyer (Sawyer 2016) and Dr. Allen P. Minton (Minton 2016).

As Editors of the special issue, we would like to thank all of the contributors for their efforts in meeting the special issue deadline and to Don for his help in filling in the gaps with information about his life and career. To one of us (DH), Don was a PhD supervisor (1995–2000)1, whilst to the other (SH), Don has been a long-time scientific collaborator2; however, to both, Don has been a stalwart companion, good friend and genuinely interesting character (Fig. 1). In this introduction, we describe a roughly chronological account of Don’s scientific history and finish with a few personal recollections about Don’s life and career (Winzor 2016).

Fig. 1.

Fig. 1

Example of an ideal interaction. Don Winzor (left) enjoying a beer with Allen Minton (middle) and Stephen Harding (right) in Nottingham

Early life and university

Don was born in 1935 in Gawler, South Australia, and, as the only son, seemed destined to follow the family tradition of becoming a farmer. However, his completion of primary education as dux of the school led Don’s parents to consider higher education as an alternative career pathway. Excellent results at high school propelled him into science at the University of Adelaide, from which Don graduated in 1956 with First Class Honours in Physical Chemistry. The award of a CSIRO postgraduate scholarship then paved the way for PhD studies (1957–1959) with J. M. (Mike) Creeth on the physicochemical characterisation of ovalbumin by diffusion, sedimentation velocity and moving boundary electrophoresis. The award of a PhD in 1960 was followed by a DSc in 1976.

Time at the CSIRO Wheat Research Unit (1960–1967)

On completing his PhD studies in 1959, Don moved from Adelaide to Sydney to take up an appointment as Research Scientist in the newly formed CSIRO Wheat Research Unit. His first 2 years there introduced him to the fractionation of gluten proteins by cellulose ion-exchange chromatography, and also to the librarian, Felicity, who became his wife. A honeymoon in Europe preceded their arrival at Cornell University for Don to take up a 12-month research associateship with Harold Scheraga in the Department of Chemistry. The year at Cornell resulted in the development of frontal size-exclusion chromatography for the quantitative characterisation of protein self-association (Winzor and Scheraga 1963, 1964). That work, a logical extension of his classical training in ultracentrifugation and electrophoresis, gained a degree of international acclaim for Don as a biophysical chemist at the age of 28 years old. Upon his return to Sydney, Michael Tracey, Leader of the Wheat Research Unit, encouraged Don to continue basic research into the characterisation of interacting systems. Those academic pursuits, performed in conjunction with Laurie Nichol, continued to the end of 1967 (Nichol and Winzor 1964; Winzor and Nichol 1965; Nichol et al. 1967a, b, c), at which stage Don moved to Brisbane to take up an appointment as Senior Lecturer in the Department of Biochemistry at the University of Queensland.

Transfer to academia

For someone faced with the prospect of presenting the first lecture in biochemistry that he had ever attended, the transition to academia posed considerable challenges. On the research side, it was facilitated by an affiliation with the Enzyme Biology Group headed by Colin Masters, who introduced Don to interactions involving glycolytic enzymes and proteins associated with the actin filaments of muscle myofibrils. At around this time publications with his first PhD student (Steven Lovell) began to appear (Masters et al. 1969; Clarke et al. 1970, 1976; Masters and Winzor 1971; Lovell and Winzor 1974, 1976, 1977). These efforts were rewarded with promotion to Reader in Biochemistry in 1972, a position held until Don was appointed to a Personal Chair in 1992. Upon retirement in 2000, he was granted Professor Emeritus status.

Characterisation of protein interactions

Collaborative studies with Laurie Nichol continued, with demonstrations of ways in which to unravel and quantify the interactions responsible for various types of self-association (Nichol et al. 1984) and binding behaviour (Baghurst et al. 1971; Nichol et al. 1972, 1973; Winzor et al. 1977; Tellam et al. 1978; Ward and Winzor 1982; Nichol and Winzor 1985; Cann and Winzor 1987; Gow et al. 1990; Wilson et al. 1997; Winzor et al. 1998). The effect of ligand-induced self-association on the size of acceptor species (Nichol and Winzor 1976) was used to explore the crosslinking theory of lymphocyte activation (Sawyer and Winzor 1976), as well as to characterise an antigen–antibody interaction from the precipitin curve (Winzor et al. 1989). Sigmoidal binding behaviour reflecting self-association of the ligand (rather than the acceptor) was also demonstrated (Sculley et al. 1981; Cann et al. 1981; Hogg and Winzor 1984a), as was the need to consider kinetic as well as thermodynamic consequences of the parking problem in the characterisation of non-specific protein interaction with linear polymer chains (Munro et al. 1998, 2000; Winzor 2013; Hall and Winzor 1999). Although most physico-chemical characterisations of interactions refer to in vitro systems, results from a study of the redistribution of desipramine in rats following the administration of a desipramine-specific monoclonal antibody allowed for estimation of the operative binding constant for the antigen–antibody interaction in vivo (Pond et al. 1991).

A 12-month sabbatical in Oxford

Sabbatical leave with Alexander (Sandy) Ogston for the 1974 calendar year in the Department of Biochemistry at Oxford University proved to be an extremely productive research experience for Don, providing as it did the first opportunity for direct collaboration (rather than by correspondence). The first problem tackled was the development of a set of quantitative expressions for the characterisation of all possible combinations of solute–matrix, solute–ligand and matrix–ligand interactions by quantitative affinity chromatography (Nichol et al. 1974), a technique introduced by Don in conjunction with Pat Andrews (University of Reading) the year before. A solution to the problem of allowance for the effects of osmotic gel shrinkage on the characterisation of protein self-association by gel chromatography then followed (Baghurst et al. 1975). This period also involved Don’s introduction to the statistical-mechanical treatment of thermodynamic nonideality on the basis of excluded volume (Ogston and Winzor 1975) a point of view which was later used to invalidate the Adams–Fujita approach to nonideality in the characterisation of protein self-association by equilibrium techniques (Scott and Winzor 2009).

Quantitative affinity chromatography

A noted odd feature of early quantitative affinity chromatography studies was the seeming conformity of experimental data with the behaviour predicted for a univalent solute (acceptor), despite the presence of multiple acceptor sites for interaction with matrix affinity sites (Brinkworth et al. 1975). Initially, this behaviour was attributed (somewhat intuitively) to spatial considerations, which suggested that further acceptor attachment after the initial event was unlikely to occur (Nichol et al. 1974). This potential paradox was resolved by quantitative solution to the problem of acceptor multivalence in its interaction with matrix-bound affinity sites (Nichol et al. 1981; Winzor et al. 1982), a problem solved initially by means of Flory reacted-site probability theory but later handled using a different approach (Lollar and Winzor 2014). That development then led to a counterpart of the Scatchard equation for a multivalent ligand (Hogg and Winzor 1984b, 1985); a mathematical statement of the condition under which the affinity chromatographic behaviour became symptomatic of a univalent solute (Kalinin et al. 1995); and to the binding equation of which the multivalent counterpart of the Scatchard equation was its linear transform (Harris et al. 1995). Around this time, some potential breakthroughs for the study of high-affinity interactions were also developed by Don, including the refinement of a recycling affinity chromatography procedure (Hogg et al. 1991) and the derivation of analytical expressions in terms of total ligand concentration (Winzor et al. 1992). The feasibility of measuring rate constants by affinity chromatography was also demonstrated (Winzor et al. 1991; Munro et al. 1993, 1994). Solution of the solute multivalency problem in quantitative affinity chromatography paved the way for partition equilibrium studies of interactions of glycolytic enzymes with muscle myofibrils, which established their significance at physiological ionic strength (Kuter et al. 1983). However, the results of these studies cast doubt over the existence of glycolytic enzyme complexes on the myofibrillar matrix as a means of enhancing glycolysis by direct metabolite transfer between enzymes. This doubt arose from the demonstration of active-site involvement in the aldolase–myofibril interaction (Harris and Winzor 1987) resulting from competition between some glycolytic enzymes for the myofibrillar matrix (Harris and Winzor 1989a), and by the demonstration that calcium ions were a non-competitive inhibitor of aldolase adsorption, with a binding constant corresponding to that for the Ca2+–troponin interaction (Harris and Winzor 1989b).

ELISA and biosensor studies of interactions

Don used his experiences gained from characterising solute partition in affinity chromatography to improve the quantification of antigen–antibody interactions by ELISA (Hogg et al. 1987; Winzor 2011), as well as by the biosensor technologies that emerged in the early 1990s (Ward et al. 1995; O’Shannessy and Winzor 1996; Hall and Winzor 1997, 1999; Hall et al. 1997; Edwards et al. 1998), methods in which the measured binding constant was previously based on 1:1 reaction stoichiometry, and referred to the solute–matrix interaction rather than that between reactants in the solution phase.

Effects of thermodynamic nonideality

Don performed a series of studies in which the effects of thermodynamic nonideality were either taken into account on the statistical-mechanical basis of excluded volume (Nichol et al. 1978, 1979; Siezen et al. 1981) or used for the detection of conformational changes in enzyme kinetic and ligand binding reactions (Nichol et al. 1983, 1985; Ford et al. 1983; Winzor et al. 1984; Shearwin and Winzor 1988a, b). Additional methods for evaluating second virial coefficients were also illustrated (Van Damme et al. 1989; Shearwin and Winzor 1990a). Excluded volume considerations were used to justify interpretation of the acid expansion of ribonuclease and serum albumin in terms of displacement of an isomerisation equilibrium between native and expanded enzyme states (Shearwin and Winzor 1990b). Another finding of significance was the demonstration of pre-existence of the isomerisation equilibrium as the source of allosteric effects for rabbit muscle pyruvate kinase, the only system for which unequivocal distinction between the Monod and Koshland models of allostery had, up till that point, been achieved (Harris and Winzor 1988). Indeed, to the present day such non-ideal experiments are one of only a few experimental techniques available for uncovering the possible existence of protein isomerisation equilibria (Bergman and Winzor 1989a, b; Bergman et al. 1989; Lonhienne and Winzor 2001; Lonhienne et al. 2003a, b), which otherwise remain merely theoretical postulates.

The Wills–Winzor collaboration

In the midst of the above investigations of thermodynamic nonideality, the departure of Laurie Nichol from the research scene to take on the Vice-Chancellorship of The Australian National University (1985) heralded the commencement of an equally fruitful collaborative effort in which Don was joined by Peter Wills from the University of Auckland, with a goal of gaining a deeper understanding of thermodynamic nonideality and, thereby, of improving the rigour of procedures for incorporating it into the quantitative characterisation of protein interactions. In that regard, the early demonstration that the displacement of protein equilibria, by small as well as macromolecular cosolutes, found rational explanation in terms of excluded volume (Winzor and Wills 1986) suggested that molecular crowding and protein solvation could be considered as the same phenomena. However trying to establish an equivalence of the quantitative expressions for the two phenomena required a return to earlier theoretical treatments (Wills 2016). The outcome of those revisionary deliberations was the realisation that, for highly concentrated systems, differing analytical results would be gained based on the choice of concentration scale (molar or molal) to define the system thermodynamics. Under constraints of constant temperature and solvent chemical potential, the chemical potential of a solute is a molar quantity, whereas it is a molal parameter under constraints of constant temperature and pressure (Wills and Winzor 1992; Wills et al. 1993; Wills and Winzor 1993; Winzor and Wills 1995a).

Another problem addressed was the nature of thermodynamic activity being monitored in sedimentation equilibrium distributions. Those studies led not only to correction of the basic sedimentation equilibrium equation for a single solute (the density parameter is that of the solvent instead of the solution density), but also to identification of the thermodynamic activity as a molar quantity, a finding of considerable significance in that the statistical-mechanical treatment of thermodynamic nonideality in terms of excluded volume refers specifically to molar thermodynamic activity (Wills and Winzor 1992; Wills et al. 1993; Wills et al. 2000a).

Solution to the question of the identity of the nature of the thermodynamic activity being monitored in a given experiment prompted a series of investigations with thermodynamic nonideality as the central theme (Wills et al. 1995; Hall et al. 1995). Notable outcomes included: (i) a reinterpretation of the concept of osmotic stress (Winzor and Wills 1995b; Poon et al. 1997), (ii) development of a direct analysis of sedimentation equilibrium distributions for self-associating proteins that takes account of thermodynamic nonideality on the statistical-mechanical basis of excluded volume (Wills et al. 1996), (iii) a demonstration that sedimentation equilibrium distributions for a non-interacting protein in the presence of a small cosolute provides a simpler way of obtaining information analogous to that derived from the Eisenberg treatment of partial specific volume measurements (Jacobsen et al. 1996), (iv) an analysis of sedimentation distributions for non-interacting proteins that quantifies second virial coefficients for solute self-interaction (Wills et al. 2000a; Winzor et al. 2001) and (v) development of a global analysis of sedimentation equilibrium distributions reflecting nonideal association of dissimilar reactants (Wills et al. 2000b).

Research in retirement

The downsizing of research resources that comes with retirement has slowed down Don’s research efforts somewhat. Since retiring, Don has received funding for a number of short-term visits to the University of Nottingham in a Biotechnology and Biological Sciences Research Council (UK) grant to the National Centre for Macromolecular Hydrodynamics. In retirement, the Wills–Winzor team has expanded to include Steve Harding, Dave Scott and Trushar Patel. During this period, Don’s major research theme has, once again, swung to thermodynamic nonideality (Chebotareva et al. 2001, 2005; Wills and Winzor 2001, 2002, 2009; Winzor and Wills 2003, 2006, 2007; Winzor et al. 2007a; Wills et al. 2012). Methods for the prediction of the thermodynamic osmotic 2nd virial coefficient from structure and charge information were developed with the routine COVOL (Harding et al. 1998, 1999) and the scope expanded to cover the determination of second virial coefficients from static light scattering measurements. Results from those methods (Deszczynski et al. 2006; Winzor et al. 2007b) as well as spectral studies (Wills and Winzor 2011) have allowed the testing of theory associated with the correct definition of virial coefficients (Wills et al. 2015). Attention has also turned to the analogous situation in small-angle X-ray scattering studies (Scott et al. 2010, 2011, 2013), which have been used to deduce the structure of components of the basement membrane protein system (Patel et al. 2011, 2012, 2014). Of very late, in a return to his research starting point as a PhD student, Don has been investigating the concentration dependence of diffusion coefficients for globular proteins, extending theories for Brownian motion in dynamic light scattering (Harding and Johnson 1985a, b) and boundary spreading in the analytical ultracentrifuge (Scott et al. 2014), along with the possible use of an approximate steady-state condition in sedimentation velocity experiments (Creeth 1964) to measure the diffusion coefficients of proteins exhibiting a sufficiently large negative concentration dependence of the sedimentation coefficient (Scott et al. 2015).

Concluding remarks

Don’s research style has been marked by a number of notable features. One has been his prodigious output, involving around ten papers a year for the last 60 years. Another is his tireless work ethic, which may/may not have its origins in his farming upbringing. Don’s always eager and energetic contributions, whether it be by simple discussion, conference organisation, taking on reviewing jobs or the critiquing of a colleague’s or student’s written work, has impressed all those who have experienced it. Balancing this zealous work ethic has been a healthy dose of humanity, which has shown itself in the form of kindness to workmates and colleagues in times of need. Indeed, with regards to this last point, we suspect that there are many working scientists today who, in part, owe their research success/tenured position/visa status/society membership to a well written letter of support, carefully crafted piece of advice or incisive critique written by Professor Donald J. Winzor. Third and fourth on the list of Don’s admirable qualities have been his very real enthusiasm and passion for his topic of physical biochemistry and his courage to, when necessary, dig in and argue heatedly over a point of science or conduct that he believed to be incorrect. Indeed, in a scientific world increasingly defined by slick operators who focus more on the psychology of collaboration and academic progression than the actual topic of scientific investigation, this last quality will be sorely missed when Don eventually decides to retire (from his semi-retirement).

In closing, we note that Don and his wife Felicity have three daughters (Catherine, Christine and Carol), along with six grandchildren. We (DH and SH), along with all the contributors to this special issue, wish Don and his family the very best for the future and congratulate him on his just passed 80th birthday. We hope that this special issue acts as a suitable commemoration to a scientific career defined by a desire to understand biochemical phenomenon in a quantitative and analytical fashion.

All the best Don in the years to come.

Acknowledgements

DH would like to thank Dr. Nami Hirota for both helpful discussions and for comments on an early version of this paper. The work of DH was jointly funded by an Australian National University Senior Research Fellowship and an Osaka University Cross-Appointment as an Associate Professor. SH acknowledges the University of Nottingham for assistance in the form of a Professorship and the United Kingdom’s Biotechnology and Biological Sciences Research Council (BBSRC) for various grants in aid of research.

Compliance with ethical standards

Conflict of interest

Damien Hall declares that he has no conflicts of interest. Stephen E. Harding declares that he has no conflicts of interest.

Ethical approval

This article does not contain any studies with human participants or animals performed by any of the authors.

Footnotes

1

Don’s undergraduate lecturing and postgraduate supervision were notable for a number of things. The lecture courses were quite hard; indeed, upon receiving a B-grade for one of Don’s undergraduate courses, DH was congratulated on achieving the highest mark in four years (we note that such a situation would not be allowed in today’s universities). As a postgraduate supervisor, Don was unusual in his tendency to only take on one student at a time, his encouragement/requirement of students to sit postgraduate coursework and the enjoyable Friday lab meetings at the staff club, nearly always accompanied by a jug of beer or two (when Don was not travelling overseas - indeed see Fig. 1 for an example of both.

2

The origins of the relationship between SH and Don stems from the common link of Mike Creeth, Don’s PhD supervisor at the University of Adelaide and SH’s postdoctoral mentor at the Lister Institute of Preventive Medicine in London. Mike was the nexus of their mutual interest in analytical ultracentrifugation. A nice exploration of this linkage can be found in their joint essay on his life (Harding and Winzor 2010).

This article is part of a special issue on ‘Analytical Quantitative Relations in Biochemistry’ edited by Damien Hall and Stephen Harding.

Contributor Information

Damien Hall, Email: damien.hall@anu.edu.au, Email: damien.hall@protein.osaka-u.ac.jp.

Stephen E. Harding, Email: steve.harding@nottingham.ac.uk

References

  1. Arisaka F, Yap ML, Kanamaru S, Rossmann MG (2016) Molecular assembly and structure of the bacteriophage T4 tail. Biophys Rev. doi:10.1007/s12551-016-0230-x [DOI] [PMC free article] [PubMed]
  2. Baghurst PA, Nichol LW, Richards RJ, Winzor DJ. Differential chromatographic study of macromolecular changes governed by environmental factors. Nature (London) 1971;234:299. doi: 10.1038/234299a0. [DOI] [PubMed] [Google Scholar]
  3. Baghurst PA, Nichol LW, Ogston AG, Winzor DJ. Quantitative interpretation of concentration-dependent migration in gel chromatography of reversibly polymerizing solutes. Biochem J. 1975;147:575. doi: 10.1042/bj1470575. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Bergman DA, Winzor DJ. Space-filling effects of inert solutes as probes for the detection and study of substrate-mediated conformational changes by enzyme kinetics: theoretical considerations. J Theor Biol. 1989;137:171. doi: 10.1016/S0022-5193(89)80204-8. [DOI] [PubMed] [Google Scholar]
  5. Bergman DA, Winzor DJ. Thermodynamic nonideality in enzyme catalysis: effect of albumin on the reduction of pyruvate by lactate dehydrogenase. Eur J Biochem. 1989;186:91. doi: 10.1111/j.1432-1033.1989.tb15086.x. [DOI] [PubMed] [Google Scholar]
  6. Bergman DA, Shearwin KE, Winzor DJ. Effects of thermodynamic nonideality on the kinetics of ester hydrolysis by α-chymotrypsin: a model system with preexistence of the isomerization equilibrium. Arch Biochem Biophys. 1989;274:55. doi: 10.1016/0003-9861(89)90414-1. [DOI] [PubMed] [Google Scholar]
  7. Brinkworth RI, Masters CJ, Winzor DJ. Evaluation of equilibrium constants for the interaction of lactate dehydrogenase isoenzymes with reduced nicotinamide-adenine dinucleotide by affinity chromatography. Biochem J. 1975;151:631. doi: 10.1042/bj1510631. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Cann JR, Winzor DJ. Frontal gel chromatography of interacting systems: theoretical and experimental evaluation of the shapes of elution profiles for systems of the type A + B ⇆ C. Arch Biochem Biophys. 1987;256:78. doi: 10.1016/0003-9861(87)90427-9. [DOI] [PubMed] [Google Scholar]
  9. Cann JR, Nichol LW, Winzor DJ. Micellarization of chlorpromazine: implications in the binding of the drug to brain tubulin. Mol Pharmacol. 1981;20:244. [PubMed] [Google Scholar]
  10. Chebotareva NA, Harding SE, Winzor DJ. Ultracentrifugal studies of the effect of molecular crowding by trimethylamine N-oxide on the self-association of muscle glycogen phosphorylase b. Eur J Biochem. 2001;268:506. doi: 10.1046/j.1432-1327.2001.01838.x. [DOI] [PubMed] [Google Scholar]
  11. Chebotareva NA, Kurganov BI, Harding SE, Winzor DJ. Effect of osmolytes on the interaction of flavin adenine dinucleotide with muscle glycogen phosphorylase b. Biophys Chem. 2005;113:61. doi: 10.1016/j.bpc.2004.07.040. [DOI] [PubMed] [Google Scholar]
  12. Chebotareva NA, Roman SG, Kurganov BI (2016) Dissociative mechanism for irreversible thermal denaturation of oligomeric proteins. Biophys Rev. doi:10.1007/s12551-016-0220-z [DOI] [PMC free article] [PubMed]
  13. Clarke FM, Masters CJ, Winzor DJ. The different adsorption of aldolase isoenzymes in rat brain. Biochem J. 1970;118:325. doi: 10.1042/bj1180325. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Clarke FM, Lovell SJ, Masters CJ, Winzor DJ. Beef muscle troponin: evidence for multiple forms of troponin-T. Biochim Biophys Acta. 1976;427:617. doi: 10.1016/0005-2795(76)90205-1. [DOI] [PubMed] [Google Scholar]
  15. Creeth JM. Approximate ‘steady state’ condition in the ultracentrifuge. Proc R Soc Lond A. 1964;282:403. doi: 10.1098/rspa.1964.0242. [DOI] [Google Scholar]
  16. Deszczynski M, Harding SE, Winzor DJ. Negative second virial coefficients as predictors of protein crystal growth: evidence from sedimentation equilibrium studies that refutes the designation of those light scattering parameters as osmotic virial coefficients. Biophys Chem. 2006;120:106. doi: 10.1016/j.bpc.2005.10.003. [DOI] [PubMed] [Google Scholar]
  17. Edwards PR, Maule CH, Leatherbarrow RJ, Winzor DJ. Second-order kinetic analysis of IAsys biosensor data: its use and applicability. Anal Biochem. 1998;263:1. doi: 10.1006/abio.1998.2814. [DOI] [PubMed] [Google Scholar]
  18. Ford CL, Winzor DJ, Nichol LW, Sculley MJ. Effects of thermodynamic nonideality in ligand binding studies. Biophys Chem. 1983;18:1. doi: 10.1016/0301-4622(83)80021-0. [DOI] [PubMed] [Google Scholar]
  19. Gow A, Winzor DJ, Smith R. Preferential ligand binding to multi-state acceptor systems: the unexplored paradox of acceptor self-association that is ligand-mediated but detrimental to ligand binding. J Theor Biol. 1990;145:407. doi: 10.1016/S0022-5193(05)80119-5. [DOI] [PubMed] [Google Scholar]
  20. Hall DR, Winzor DJ. Use of a resonant mirror biosensor to characterize the interaction of carboxypeptidase A with an elicited monoclonal antibody. Anal Biochem. 1997;244:152. doi: 10.1006/abio.1996.9867. [DOI] [PubMed] [Google Scholar]
  21. Hall DR, Winzor DJ. Parking problems as potential sources of deviation from predicted kinetic behaviour in biosensor studies with small immobilized antigens. Int J Biochromatogr. 1999;4:175. [Google Scholar]
  22. Hall DR, Jacobsen MP, Winzor DJ. Stabilizing effect of sucrose against irreversible denaturation of rabbit muscle lactate dehydrogenase. Biophys Chem. 1995;57:47. doi: 10.1016/0301-4622(95)00044-X. [DOI] [PubMed] [Google Scholar]
  23. Hall DR, Gorgani NN, Altin JG, Winzor DJ. Theoretical and experimental considerations of the pseudo-first-order approximation in conventional kinetic analysis of IAsys biosensor data. Anal Biochem. 1997;253:145. doi: 10.1006/abio.1997.2358. [DOI] [PubMed] [Google Scholar]
  24. Harding SE, Johnson P. The concentration-dependence of macromolecular parameters. Biochem J. 1985;231:543. doi: 10.1042/bj2310543. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Harding SE, Johnson P. Physicochemical studies on turnip-yellow-mosaic virus. Homogeneity, relative molecular masses, hydrodynamic radii and concentration-dependence of parameters. Biochem J. 1985;231:549. doi: 10.1042/bj2310549. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Harding SE, Winzor DJ. James Michael Creeth, 1924–2010. Macromol Biosci. 2010;10:696–699. doi: 10.1002/mabi.201000073. [DOI] [PubMed] [Google Scholar]
  27. Harding SE, Horton JC, Winzor DJ. COVOL: an answer to your thermodynamic non-ideality problems? Biochem Soc Trans. 1998;26:737. doi: 10.1042/bst0260737. [DOI] [PubMed] [Google Scholar]
  28. Harding SE, Horton JC, Jones S, Thornton JM, Winzor DJ. COVOL: an interactive program for evaluating second virial coefficients from the triaxial shape or dimensions of rigid macromolecules. Biophys J. 1999;76:2432. doi: 10.1016/S0006-3495(99)77398-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Harding SE, Gillis RB, Adams GG (2016) Assessing sedimentation equilibrium profiles in analytical ultracentrifugation experiments on macromolecules: from simple average molecular weight analysis to molecular weight distribution and interaction analysis. Biophys Rev. doi:10.1007/s12551-016-0232-8 [DOI] [PMC free article] [PubMed]
  30. Harris SJ, Winzor DJ. Enzyme kinetic evidence of active-site involvement in the interaction between aldolase and muscle myofibrils. Biochim Biophys Acta. 1987;911:121. doi: 10.1016/0167-4838(87)90279-2. [DOI] [PubMed] [Google Scholar]
  31. Harris SJ, Winzor DJ. Thermodynamic nonideality as a probe of allosteric mechanisms: preexistence of the isomerization equilibrium for rabbit muscle pyruvate kinase. Arch Biochem Biophys. 1988;265:458. doi: 10.1016/0003-9861(88)90150-6. [DOI] [PubMed] [Google Scholar]
  32. Harris SJ, Winzor DJ. Equilibrium partition studies of the myofibrillar interactions of glycolytic enzymes. Arch Biochem Biophys. 1989;275:185. doi: 10.1016/0003-9861(89)90363-9. [DOI] [PubMed] [Google Scholar]
  33. Harris SJ, Winzor DJ. Effect of calcium ion on the interaction of aldolase with rabbit muscle myofibrils. Biochim Biophys Acta. 1989;999:95. doi: 10.1016/0167-4838(89)90035-6. [DOI] [PubMed] [Google Scholar]
  34. Harris SJ, Jackson CM, Winzor DJ. The rectangular hyperbolic binding equation for multivalent ligands. Arch Biochem Biophys. 1995;316:20. doi: 10.1006/abbi.1995.1004. [DOI] [PubMed] [Google Scholar]
  35. Hogg PJ, Winzor DJ. Evidence for the preferential interaction of micellar chlorpromazine with human serum albumin. Biochem Pharmacol. 1984;33:1998. doi: 10.1016/0006-2952(84)90563-X. [DOI] [PubMed] [Google Scholar]
  36. Hogg PJ, Winzor DJ. Quantitative affinity chromatography: further developments in the analysis of experimental results from column chromatography and partition equilibrium studies. Arch Biochem Biophys. 1984;234:55. doi: 10.1016/0003-9861(84)90323-0. [DOI] [PubMed] [Google Scholar]
  37. Hogg PJ, Winzor DJ. Effects of ligand multivalency in binding studies: a general counterpart of the Scatchard analysis. Biochim Biophys Acta. 1985;843:159. doi: 10.1016/0304-4165(85)90134-5. [DOI] [PubMed] [Google Scholar]
  38. Hogg PJ, Johnstone SC, Bowles MR, Pond SE, Winzor DJ. Evaluation of equilibrium constants by solid-phase immunoassay: the binding of paraquat to its elicited mouse monoclonal antibody. Mol Immunol. 1987;24:797. doi: 10.1016/0161-5890(87)90064-2. [DOI] [PubMed] [Google Scholar]
  39. Hogg PJ, Jackson CM, Winzor DJ. Use of quantitative affinity chromatography for characterizing high-affinity interactions: binding of heparin to antithrombin III. Anal Biochem. 1991;192:303. doi: 10.1016/0003-2697(91)90540-A. [DOI] [PubMed] [Google Scholar]
  40. Jacobsen MP, Wills PR, Winzor DJ. Thermodynamic analysis of the effects of small inert cosolutes in the ultracentrifugation of noninteracting proteins. Biochemistry. 1996;35:13173. doi: 10.1021/bi960939q. [DOI] [PubMed] [Google Scholar]
  41. Jones S (2016) Protein-RNA Interactions: structural biology and computational modeling techniques. Biophys Rev. doi:10.1007/s12551-016-0223-9 [DOI] [PMC free article] [PubMed]
  42. Kalinin NL, Ward LD, Winzor DJ. Effects of solute multivalence on the evaluation of binding constants by biosensor technology: studies with concanavalin A and interleukin-6 as partitioning proteins. Anal Biochem. 1995;228:238. doi: 10.1006/abio.1995.1345. [DOI] [PubMed] [Google Scholar]
  43. Karlsson R (2016) Biosensor binding data and its applicability to determination of active concentration. Biophys Rev. doi:10.1007/s12551-016-0219-5 [DOI] [PMC free article] [PubMed]
  44. Kuter MR, Masters CJ, Winzor DJ. Equilibrium partition studies of the interaction between aldolase and myofibrils. Arch Biochem Biophys. 1983;225:384. doi: 10.1016/0003-9861(83)90043-7. [DOI] [PubMed] [Google Scholar]
  45. Laue T (2016) Charge matters. Biophys Rev. doi:10.1007/s12551-016-0229-3 [DOI] [PMC free article] [PubMed]
  46. Lollar P, Winzor DJ. Reconciliation of classical and reacted-site probability approaches to allowance for ligand multivalence in binding studies. J Mol Recognit. 2014;27:73. doi: 10.1002/jmr.2335. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Lonhienne TGA, Winzor DJ. Interpretation of the reversible inhibition of adenosine deaminase by small cosolutes in terms of molecular crowding. Biochemistry. 2001;240:9618. doi: 10.1021/bi010857o. [DOI] [PubMed] [Google Scholar]
  48. Lonhienne TGA, Jackson CM, Winzor DJ. Thermodynamic non-ideality as an alternative source of the effect of sucrose on the thrombin-catalyzed hydrolysis of peptide p-nitroanilide substrates. Biophys Chem. 2003;103:259. doi: 10.1016/S0301-4622(02)00322-8. [DOI] [PubMed] [Google Scholar]
  49. Lonhienne TGA, Reilly PEB, Winzor DJ. Further evidence for the reliance of catalysis by rabbit muscle pyruvate kinase upon isomerization of the ternary complex between enzyme and products. Biophys Chem. 2003;104:189. doi: 10.1016/S0301-4622(02)00366-6. [DOI] [PubMed] [Google Scholar]
  50. Lovell SJ, Winzor DJ. Effects of phosphate on the dissociation and enzymic stability of rabbit muscle lactate dehydrogenase. Biochemistry. 1974;13:3527. doi: 10.1021/bi00714a018. [DOI] [PubMed] [Google Scholar]
  51. Lovell SJ, Winzor DJ. Rabbit muscle myogen. Interactions with phosphate as the source of non-enantiography in moving-boundary electrophoresis. Biochem J. 1976;157:699. doi: 10.1042/bj1570699. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Lovell SJ, Winzor DJ. Self-association of troponin. Biochem J. 1977;167:131. doi: 10.1042/bj1670131. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Masters CJ, Winzor DJ. The molecular size of enzymically active aldolase A. Biochem J. 1971;121:735. doi: 10.1042/bj1210735. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Masters CJ, Sheedy RJ, Winzor DJ, Nichol LW. Reversible adsorption of enzymes as a possible allosteric control mechanism. Biochem J. 1969;112:806. doi: 10.1042/bj1120806. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Minton AP (2016) Big ideas from “small science”. Biophys Rev. doi:10.1007/s12551-016-0225-7 [DOI] [PMC free article] [PubMed]
  56. Munro PD, Winzor DJ, Cann JR. Experimental and theoretical studies of rate constant evaluation by affinity chromatography: determination of rate constants for the interaction of saccharides with concanavalin A. J Chromatogr A. 1993;646:3. doi: 10.1016/S0021-9673(99)87002-2. [DOI] [Google Scholar]
  57. Munro PD, Winzor DJ, Cann JR. Allowance for kinetics of solute partitioning in the determination of rate constants by affinity chromatography. J Chromatogr A. 1994;659:267. doi: 10.1016/0021-9673(94)85068-2. [DOI] [Google Scholar]
  58. Munro PD, Jackson CM, Winzor DJ. On the need to consider kinetic as well as thermodynamic consequences of the parking problem in quantitative studies of nonspecific binding between proteins and linear polymer chains. Biophys Chem. 1998;71:185. doi: 10.1016/S0301-4622(98)00104-5. [DOI] [PubMed] [Google Scholar]
  59. Munro PD, Jackson CM, Winzor DJ. Consequences of the non-specific binding of a protein to a linear polymer: reconciliation of stoichiometric and equilibrium titration data for the thrombin–heparin interaction. J Theor Biol. 2000;203:407. doi: 10.1006/jtbi.2000.1099. [DOI] [PubMed] [Google Scholar]
  60. Munro PD, Ackers GK, Shearwin KE (2016) Aspects of protein-DNA interactions: A review of quantitative thermodynamic theory for modelling synthetic circuits utilising LacI and CI repressors, IPTG and the reporter gene lacZ. Biophys Rev. doi:10.1007/s12551-016-0231-9 [DOI] [PMC free article] [PubMed]
  61. Nichol LW, Winzor DJ. The determination of equilibrium constants from transport data on rapidly reacting systems of the type A + B ⇆ C. J Phys Chem. 1964;68:2455. doi: 10.1021/j100791a012. [DOI] [Google Scholar]
  62. Nichol LW, Winzor DJ. Ligand-induced polymerization. Biochemistry. 1976;15:3015. doi: 10.1021/bi00659a012. [DOI] [PubMed] [Google Scholar]
  63. Nichol LW, Winzor DJ. The binding of a ligand to an acceptor undergoing indefinite self-association. J Theor Biol. 1985;117:597. doi: 10.1016/S0022-5193(85)80241-1. [DOI] [PubMed] [Google Scholar]
  64. Nichol LW, Ogston AG, Winzor DJ. Reaction boundaries and elution profiles in column chromatography. J Phys Chem. 1967;71:726. doi: 10.1021/j100862a037. [DOI] [Google Scholar]
  65. Nichol LW, Ogston AG, Winzor DJ. Evaluation of gel filtration data on systems interacting chemically and physically. Arch Biochem Biophys. 1967;121:517. doi: 10.1016/0003-9861(67)90107-5. [DOI] [PubMed] [Google Scholar]
  66. Nichol LW, Jackson WJH, Winzor DJ. A theoretical study of the binding of small molecules to a polymerizing protein system: a model for allosteric effects*. Biochemistry. 1967;6:2449. doi: 10.1021/bi00860a022. [DOI] [PubMed] [Google Scholar]
  67. Nichol LW, Jackson WJH, Winzor DJ. Preferential binding of competitive inhibitors to the monomeric form of α-chymotrypsin. Biochemistry. 1972;11:585. doi: 10.1021/bi00754a017. [DOI] [PubMed] [Google Scholar]
  68. Nichol LW, Smith GD, Winzor DJ. Interpretation of migration patterns for interacting mixtures of reactants that travel in opposite directions. J Phys Chem. 1973;77:2912. doi: 10.1021/j100642a601. [DOI] [Google Scholar]
  69. Nichol LW, Ogston AG, Winzor DJ, Sawyer WH. Evaluation of equilibrium constants by affinity chromatography. Biochem J. 1974;143:435. doi: 10.1042/bj1430435. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Nichol LW, Siezen RJ, Winzor DJ. The study of multiple polymerization equilibria by glass bead exclusion chromatography with allowance for thermodynamic non-ideality effects. Biophys Chem. 1978;9:47. doi: 10.1016/0301-4622(78)87014-8. [DOI] [PubMed] [Google Scholar]
  71. Nichol LW, Siezen RJ, Winzor DJ. Chromatographic evidence of the self-association of oxyhemoglobin in concentrated solutions: its biological implications. Biophys Chem. 1979;10:17. doi: 10.1016/0301-4622(79)80002-2. [DOI] [PubMed] [Google Scholar]
  72. Nichol LW, Ward LD, Winzor DJ. Multivalency of the partitioning species in quantitative affinity chromatography: evaluation of the site-binding constant for the aldolase–phosphate interaction from studies with cellulose phosphate as the affinity matrix. Biochemistry. 1981;20:4856. doi: 10.1021/bi00520a008. [DOI] [PubMed] [Google Scholar]
  73. Nichol LW, Sculley MJ, Ward LD, Winzor DJ. Effects of thermodynamic nonideality in kinetic studies. Arch Biochem Biophys. 1983;222:574. doi: 10.1016/0003-9861(83)90555-6. [DOI] [PubMed] [Google Scholar]
  74. Nichol LW, Sculley MJ, Jeffrey PD, Winzor DJ. Indefinite self-association of a solute in linear and branched arrays. J Theor Biol. 1984;109:285. doi: 10.1016/S0022-5193(84)80007-7. [DOI] [PubMed] [Google Scholar]
  75. Nichol LW, Owen EA, Winzor DJ. Effect of thermodynamic nonideality in kinetic studies: evidence for reversible unfolding of urease during urea hydrolysis. Arch Biochem Biophys. 1985;239:147. doi: 10.1016/0003-9861(85)90821-5. [DOI] [PubMed] [Google Scholar]
  76. Ogston AG, Winzor DJ. Treatment of thermodynamic nonideality in equilibrium studies on associating solutes. J Phys Chem. 1975;79:2496. doi: 10.1021/j100590a011. [DOI] [Google Scholar]
  77. O’Shannessy DJ, Winzor DJ. Interpretation of deviations from pseudo-first-order kinetic behavior in the characterization of ligand binding by biosensor technology. Anal Biochem. 1996;236:275. doi: 10.1006/abio.1996.0167. [DOI] [PubMed] [Google Scholar]
  78. Patel TR, Besong TMD, Patel N, Meier M, Harding SE, Winzor DJ, Stetefeld J. Evidence for self-association of a miniaturized version of agrin from hydrodynamic and small-angle X-ray scattering measurements. J Phys Chem B. 2011;115:11286. doi: 10.1021/jp206377b. [DOI] [PubMed] [Google Scholar]
  79. Patel TR, Reuten R, Xiong S, Meier M, Winzor DJ, Koch M, Stetefeld J. Determination of a molecular shape for netrin-4 from hydrodynamic and small angle X-ray scattering measurements. Matrix Biol. 2012;31:135. doi: 10.1016/j.matbio.2011.11.004. [DOI] [PubMed] [Google Scholar]
  80. Patel TR, Bernards C, Meier M, McEleney K, Winzor DJ, Koch M, Stetefeld J. Structural elucidation of full-length nidogen and the laminin–nidogen complex in solution. Matrix Biol. 2014;33:60. doi: 10.1016/j.matbio.2013.07.009. [DOI] [PubMed] [Google Scholar]
  81. Pond SM, Pentel PR, Keyler DE, Winzor DJ. Determination of the in vivo antigen–antibody affinity constant from the redistribution of desipramine in rats following administration of a desipramine-specific monoclonal antibody. Biochem Pharmacol. 1991;41:473. doi: 10.1016/0006-2952(91)90552-G. [DOI] [PubMed] [Google Scholar]
  82. Poon J, Bailey M, Winzor DJ, Davidson BE, Sawyer WH. Effects of molecular crowding on the interaction between DNA and the Escherichia coli regulatory protein TyrR. Biophys J. 1997;73:3257. doi: 10.1016/S0006-3495(97)78350-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Rao, A, Cölfen, H (2016) Mineralization and non-ideality: On nature's foundry. Biophys Rev. doi:10.1007/s12551-016-0228-4 [DOI] [PMC free article] [PubMed]
  84. Sawyer WH, Winzor DJ. Thermodynamic requirements for the cross-linking theory of lymphocyte activation: the interpretation of dose response curves. Immunochemistry. 1976;13:141. doi: 10.1016/0019-2791(76)90282-2. [DOI] [PubMed] [Google Scholar]
  85. Sawyer WH (2016) A note on the career of Donald J Winzor. Biophys Rev. doi:10.1007/s12551-016-0224-8 [DOI] [PMC free article] [PubMed]
  86. Schor M, Mey ASJS, MacPhee CJ (2016) Analytical methods for structural ensembles and dynamics of intrinsically disordered proteins. Biophys Rev. doi:10.1007/s12551-016-0234-6 [DOI] [PMC free article] [PubMed]
  87. Scott DJ (2016) Accounting for thermodynamic non-ideality in the Guinier region of small angle scattering data of proteins. Biophys Rev. doi:10.1007/s12551-016-0235-5 [DOI] [PMC free article] [PubMed]
  88. Scott DJ, Winzor DJ. Comparison of methods for characterizing nonideal solute self-association by sedimentation equilibrium. Biophys J. 2009;97:886. doi: 10.1016/j.bpj.2009.05.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Scott DJ, Wills PR, Winzor DJ. Allowance for the effect of protein charge in the characterization of nonideal solute self-association by sedimentation equilibrium. Biophys Chem. 2010;149:83. doi: 10.1016/j.bpc.2010.04.003. [DOI] [PubMed] [Google Scholar]
  90. Scott DJ, Patel TR, Besong DMT, Stetefeld J, Winzor DJ. Examination of the discrepancy between size estimates for ovalbumin from small-angle X-ray scattering and other physicochemical measurements. J Phys Chem B. 2011;115:10725. doi: 10.1021/jp2006149. [DOI] [PubMed] [Google Scholar]
  91. Scott DJ, Patel TR, Winzor DJ. A potential for overestimating the absolute magnitudes of second virial coefficients by small-angle X-ray scattering. Anal Biochem. 2013;435:159. doi: 10.1016/j.ab.2012.12.014. [DOI] [PubMed] [Google Scholar]
  92. Scott DJ, Harding SE, Winzor DJ. Concentration dependence of translational diffusion coefficients for globular proteins. Analyst. 2014;139:6242. doi: 10.1039/C4AN01060D. [DOI] [PubMed] [Google Scholar]
  93. Scott DJ, Harding SE, Winzor DJ. Evaluation of diffusion coefficients by means of an approximate steady-state condition in sedimentation velocity distributions. Anal Biochem. 2015;490:20. doi: 10.1016/j.ab.2015.08.017. [DOI] [PubMed] [Google Scholar]
  94. Sculley MJ, Nichol LW, Winzor DJ. Interactions between micellar ligand systems and acceptors: forms of binding curves. J Theor Biol. 1981;90:365. doi: 10.1016/0022-5193(81)90317-9. [DOI] [PubMed] [Google Scholar]
  95. Shearwin KE, Winzor DJ. Substrate as a source of thermodynamic nonideality in enzyme kinetic studies: invertase-catalyzed hydrolysis of sucrose. Arch Biochem Biophys. 1988;260:532. doi: 10.1016/0003-9861(88)90478-X. [DOI] [PubMed] [Google Scholar]
  96. Shearwin KE, Winzor DJ. Effect of sucrose on the dimerization of α-chymotrypsin: allowance for thermodynamic nonideality arising from the presence of a small inert solute. Biophys Chem. 1988;31:287. doi: 10.1016/0301-4622(88)80034-6. [DOI] [PubMed] [Google Scholar]
  97. Shearwin KE, Winzor DJ. Thermodynamic nonideality in macromolecular solutions: evaluation of parameters for the prediction of covolume effects. Eur J Biochem. 1990;190:523. doi: 10.1111/j.1432-1033.1990.tb15605.x. [DOI] [PubMed] [Google Scholar]
  98. Shearwin KE, Winzor DJ. Thermodynamic nonideality as a probe of reversible protein unfolding effected by variations in pH and temperature: studies of ribonuclease. Arch Biochem Biophys. 1990;282:297. doi: 10.1016/0003-9861(90)90120-N. [DOI] [PubMed] [Google Scholar]
  99. Shishmarev D, Kuchel PW (2016) NMR magnetization-transfer analysis of rapid membrane transport in human erythrocytes. Biophys Rev. doi:10.1007/s12551-016-0221-y [DOI] [PMC free article] [PubMed]
  100. Siezen RJ, Nichol LW, Winzor DJ. Exclusion chromatography of concentrated hemoglobin solutions: comparison of the self-association behavior of the oxy and deoxy forms of the α2β2 species. Biophys Chem. 1981;14:221. doi: 10.1016/0301-4622(81)85023-5. [DOI] [PubMed] [Google Scholar]
  101. Stetefeld J, McKennaa SA, Patel TR (2016) Dynamic light scattering: A practical guide and applications in Biomedical Sciences. Biophys Rev. doi:10.1007/s12551-016-0218-6 [DOI] [PMC free article] [PubMed]
  102. Tellam RO, Winzor DJ, Nichol LW. The role of zinc in the stabilization of the dimeric form of bacterial α-amylase. Biochem J. 1978;173:185. doi: 10.1042/bj1730185. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Van Damme M-PI, Murphy WH, Comper WD, Preston BN, Winzor DJ. Evaluation of nonideality from gel chromatographic partition coefficients: a technique with greater versatility than equilibrium dialysis. Biophys Chem. 1989;33:115. doi: 10.1016/0301-4622(89)80014-6. [DOI] [PubMed] [Google Scholar]
  104. Ward LD, Winzor DJ. Activation of rabbit muscle lactate dehydrogenase by phosphate: active enzyme gel chromatography and enzyme kinetic studies. Arch Biochem Biophys. 1982;216:329. doi: 10.1016/0003-9861(82)90218-1. [DOI] [PubMed] [Google Scholar]
  105. Ward LD, Howlett GJ, Hammacher A, Weinstock J, Yasukawa K, Simpson RJ, Winzor DJ. Use of a biosensor with surface plasmon resonance detection for the determination of binding constants: measurement of interleukin-6 binding to the soluble interleukin-6 receptor. Biochemistry. 1995;34:2901. doi: 10.1021/bi00009a021. [DOI] [PubMed] [Google Scholar]
  106. Wills PR (2016) A Hilly path through the thermodynamics and statistical mechanics of protein solutions. Biophys Rev. doi:10.1007/s12551-016-0226-6 [DOI] [PMC free article] [PubMed]
  107. Wills PR, Winzor DJ. Thermodynamic non-ideality and sedimentation equilibrium. In: Harding SE, Rowe AJ, Horton JC, editors. Ultracentrifugation in biochemistry and polymer science. Cambridge UK: Royal Society of Chemistry; 1992. pp. 311–330. [Google Scholar]
  108. Wills PR, Winzor DJ. Thermodynamic analysis of “preferential solvation” in protein solutions. Biopolymers. 1993;33:1627. doi: 10.1002/bip.360331012. [DOI] [Google Scholar]
  109. Wills PR, Winzor DJ. Studies of solute self-association by sedimentation equilibrium: allowance for effects of thermodynamic non-ideality beyond the consequences of nearest-neighbor interactions. Biophys Chem. 2001;91:253. doi: 10.1016/S0301-4622(01)00174-0. [DOI] [PubMed] [Google Scholar]
  110. Wills PR, Winzor DJ. Exact theory of sedimentation equilibrium made useful. Prog Colloid Polym Sci. 2002;119:113. doi: 10.1007/3-540-44672-9_16. [DOI] [Google Scholar]
  111. Wills PR, Winzor DJ. Direct allowance for the effects of thermodynamic nonideality in the quantitative characterization of protein self-association by osmometry. Biophys Chem. 2009;145:64. doi: 10.1016/j.bpc.2009.09.001. [DOI] [PubMed] [Google Scholar]
  112. Wills PR, Winzor DJ. Allowance for thermodynamic nonideality in the characterization of protein interactions by spectral techniques. Biophys Chem. 2011;158:21. doi: 10.1016/j.bpc.2011.04.011. [DOI] [PubMed] [Google Scholar]
  113. Wills PR, Comper WD, Winzor DJ. Thermodynamic nonideality in macromolecular solutions: interpretation of virial coefficients. Arch Biochem Biophys. 1993;300:206. doi: 10.1006/abbi.1993.1029. [DOI] [PubMed] [Google Scholar]
  114. Wills PR, Georgalis Y, Dijk J, Winzor DJ. Measurement of thermodynamic nonideality arising from volume-exclusion interactions between proteins and polymers. Biophys Chem. 1995;57:37. doi: 10.1016/0301-4622(95)00043-W. [DOI] [PubMed] [Google Scholar]
  115. Wills PR, Jacobsen MP, Winzor DJ. Direct analysis of solute self-association by sedimentation equilibrium. Biopolymers. 1996;38:119. doi: 10.1002/(SICI)1097-0282(199601)38:1<119::AID-BIP10>3.0.CO;2-C. [DOI] [Google Scholar]
  116. Wills PR, Hall DR, Winzor DJ. Interpretation of thermodynamic non-ideality in sedimentation equilibrium experiments on proteins. Biophys Chem. 2000;84:217. doi: 10.1016/S0301-4622(00)00124-1. [DOI] [PubMed] [Google Scholar]
  117. Wills PR, Jacobsen MP, Winzor DJ. Analysis of sedimentation equilibrium distributions reflecting nonideal macromolecular associations. Biophys J. 2000;79:2178. doi: 10.1016/S0006-3495(00)76466-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. Wills PR, Scott DJ, Winzor DJ. Allowance for effects of thermodynamic nonideality in sedimentation equilibrium distributions reflecting protein dimerization. Anal Biochem. 2012;422:28. doi: 10.1016/j.ab.2011.12.010. [DOI] [PubMed] [Google Scholar]
  119. Wills PR, Scott DJ, Winzor DJ. The osmotic second virial coefficient for protein self-interaction: use and misuse to describe thermodynamic nonideality. Anal Biochem. 2015;490:55. doi: 10.1016/j.ab.2015.08.020. [DOI] [PubMed] [Google Scholar]
  120. Wilson EK, Scrutton NS, Cölfen H, Harding SE, Jacobsen MP, Winzor DJ. An ultracentrifugal approach to quantitative characterization of the molecular assembly of a physiological electron-transfer complex. Eur J Biochem. 1997;243:393. doi: 10.1111/j.1432-1033.1997.0393a.x. [DOI] [PubMed] [Google Scholar]
  121. Winzor DJ. Allowance for antibody bivalence in the characterization of interactions by ELISA. J Mol Recognit. 2011;24:139. doi: 10.1002/jmr.1054. [DOI] [PubMed] [Google Scholar]
  122. Winzor DJ. A historical perspective of the biophysics of the thrombin–heparin system: an example of nonspecific binding and the consequent parking problem in action. Biophys Rev. 2013;5:173. doi: 10.1007/s12551-013-0103-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Winzor DJ (2016) Six decades of research in physical biochemistry. Biophys Rev. doi:10.1007/s12551-016-0222-x [DOI] [PMC free article] [PubMed]
  124. Winzor DJ, Nichol LW. Effects of concentration-dependence in gel filtration. Biochim Biophys Acta. 1965;104:1. doi: 10.1016/0304-4165(65)90213-8. [DOI] [PubMed] [Google Scholar]
  125. Winzor DJ, Scheraga HA. Studies of chemically reacting systems on Sephadex. I. Chromatographic demonstration of the Gilbert theory. Biochemistry. 1963;2:1263. doi: 10.1021/bi00906a016. [DOI] [PubMed] [Google Scholar]
  126. Winzor DJ, Scheraga HA. Studies of chemically reacting systems on Sephadex. II. Molecular weights of monomers in rapid association equilibrium. J Phys Chem. 1964;68:338. doi: 10.1021/j100784a022. [DOI] [Google Scholar]
  127. Winzor DJ, Wills PR. Effects of thermodynamic nonideality on protein interactions: equivalence of interpretations based on excluded volume and preferential solvation. Biophys Chem. 1986;25:343. doi: 10.1016/0301-4622(86)80016-3. [DOI] [PubMed] [Google Scholar]
  128. Winzor DJ, Wills PR. Thermodynamic nonideality and protein solvation. In: Gregory RB, editor. Protein–solvent interactions. New York: Marcel Dekker; 1995. pp. 483–520. [Google Scholar]
  129. Winzor DJ, Wills PR. Thermodynamic nonideality of enzyme solutions supplemented with inert solutes: yeast hexokinase revisited. Biophys Chem. 1995;57:103. doi: 10.1016/0301-4622(95)00051-X. [DOI] [PubMed] [Google Scholar]
  130. Winzor DJ, Wills PR. Allowance for thermodynamic non-ideality in the characterization of protein self-association by frontal exclusion chromatography: hemoglobin revisited. Biophys Chem. 2003;104:345. doi: 10.1016/S0301-4622(03)00003-6. [DOI] [PubMed] [Google Scholar]
  131. Winzor DJ, Wills PR. Molecular crowding effects of linear polymers in protein solutions. Biophys Chem. 2006;119:186. doi: 10.1016/j.bpc.2005.08.001. [DOI] [PubMed] [Google Scholar]
  132. Winzor DJ, Wills PR. Characterization of weak protein dimerization by direct analysis of sedimentation equilibrium distributions: the INVEQ approach. Anal Biochem. 2007;368:168. doi: 10.1016/j.ab.2007.04.030. [DOI] [PubMed] [Google Scholar]
  133. Winzor DJ, Tellam R, Nichol LW. Determination of the asymptotic shapes of sedimentation velocity patterns for reversibly polymerizing solutes. Arch Biochem Biophys. 1977;178:327. doi: 10.1016/0003-9861(77)90200-4. [DOI] [PubMed] [Google Scholar]
  134. Winzor DJ, Ward LD, Nichol LW. Quantitative considerations of the consequences of an interplay between ligand binding and reversible adsorption of a macromolecular solute. J Theor Biol. 1982;98:171. doi: 10.1016/0022-5193(82)90257-0. [DOI] [PubMed] [Google Scholar]
  135. Winzor DJ, Ford CL, Nichol LW. Thermodynamic nonideality as a probe of macromolecular isomerizations: application to the acid expansion of bovine serum albumin. Arch Biochem Biophys. 1984;234:15. doi: 10.1016/0003-9861(84)90319-9. [DOI] [PubMed] [Google Scholar]
  136. Winzor DJ, Stevens A, Augusteyn RC. Evaluation of equilibrium constants from precipitin curves: interaction of α-crystallin with an elicited monoclonal antibody. Arch Biochem Biophys. 1989;268:221. doi: 10.1016/0003-9861(89)90583-3. [DOI] [PubMed] [Google Scholar]
  137. Winzor DJ, Munro PD, Cann JR. Experimental and theoretical studies of rate constant evaluation for the solute–matrix interaction in affinity chromatography. Anal Biochem. 1991;194:54. doi: 10.1016/0003-2697(91)90150-R. [DOI] [PubMed] [Google Scholar]
  138. Winzor DJ, Munro PD, Jackson CM. Study of high-affinity interactions by quantitative affinity chromatography: analytical expressions in terms of total ligand concentration. J Chromatogr. 1992;597:57. doi: 10.1016/0021-9673(92)80096-D. [DOI] [PubMed] [Google Scholar]
  139. Winzor DJ, Jacobsen MP, Wills PR. Direct analysis of sedimentation equilibrium distributions reflecting complex formation between dissimilar reactants. Biochemistry. 1998;37:2226. doi: 10.1021/bi972211v. [DOI] [PubMed] [Google Scholar]
  140. Winzor DJ, Carrington LE, Harding SE. Analysis of thermodynamic non-ideality in terms of protein solvation. Biophys Chem. 2001;93:231. doi: 10.1016/S0301-4622(01)00223-X. [DOI] [PubMed] [Google Scholar]
  141. Winzor DJ, Patel CN, Pielak GJ. Reconsideration of sedimentation equilibrium distributions reflecting the effects of small inert cosolutes on the dimerization of α-chymotrypsin. Biophys Chem. 2007;130:89. doi: 10.1016/j.bpc.2007.07.006. [DOI] [PubMed] [Google Scholar]
  142. Winzor DJ, Deszczynski M, Harding SE, Wills PR. Nonequivalence of second virial coefficients from sedimentation equilibrium and static light scattering studies of protein solutions. Biophys Chem. 2007;128:46. doi: 10.1016/j.bpc.2007.03.001. [DOI] [PubMed] [Google Scholar]
  143. Zhao R, So M, Maat H, Ray N, Arisaka F, Goto Y, Carver J and Hall D (2016) Measurement of amyloid formation by turbidity assay – seeing through the cloud. Biophys Rev. doi:10.1007/s12551-016-0233-7 [DOI] [PMC free article] [PubMed]

Articles from Biophysical Reviews are provided here courtesy of Springer

RESOURCES