INTERACTION STUDIES BETWEEN IONS AND PROTEINS AT PHYSIOLOGICALLY RELEVANT CONCENTRATIONS

96 181 0
INTERACTION STUDIES BETWEEN IONS AND PROTEINS AT PHYSIOLOGICALLY RELEVANT CONCENTRATIONS

Đang tải... (xem toàn văn)

Tài liệu hạn chế xem trước, để xem đầy đủ mời bạn chọn Tải xuống

Thông tin tài liệu

INTERACTION STUDIES BETWEEN IONS AND PROTEINS AT PHYSIOLOGICALLY RELEVANT CONCENTRATIONS MIAO LINLIN (B. Sc) A THESIS SUBMITTED FOR THE DEGREE OF MASTER OF SCIENCE DEPARTMENT OF BIOLOGICAL SCIENCES NATIONAL UNIVERSITY OF SINGAPORE 2012 Acknowledgements I would like to take this opportunity to express my deepest and sincerest gratitude to my supervisor, A/P Song Jianxing, for his invaluable guidance, inspiration and patience in my project. His incredible enthusiasm for science made a deep impression on me and inspired me to complete this thesis. I am also very thankful to all the members in the structural biology labs, especially my labmates, Dr. Qin Haina, Ms Shaveta, Huan Xuelu, Wang Wei, for their valuable advice and kind help in my research. In particular, I would like to express my thanks to Dr. Fan Jingsong for NMR operation training and NMR data collection on the 800 MHz and 500 MHz spectrometer. I also want to thank Professor Elena Pasquale for generously giving us the gene of ephrinB2. In addition I want to extend my sincere thanks to my family, especially to my husband for his love, tolerance and encouragement. At last, I would like to give my sincere thanks to the Department of Biological Sciences for giving me the opportunity to study in this top university and thank all administrative and research staff for their help in both my life and research. I TABLE OF CONTENTS Chapter 1 Introduction ................................................................................................... 1 1.1 Salt-protein interaction ......................................................................................... 2 1.1.1 Hofmeister series ........................................................................................... 2 1.1.2 Phenomenology of the Hofmeister Series ..................................................... 3 1.1.3 The mechanism of Hofmeister Series ............................................................ 6 1.1.4 Specific protein-ion binding .......................................................................... 9 1.2 Nuclear Magnetic Resonance (NMR) spectroscopy .......................................... 11 1.2.1 NMR phenomenon....................................................................................... 11 1.2.2 Chemical shift .............................................................................................. 12 1.2.3 Molecular interaction studies by NMR ....................................................... 14 1.2.4 Study of protein-salt interactions using NMR ............................................. 14 1.3 Biological background ....................................................................................... 15 1.3.1 Cytoplasmic domain of ephrinB2 ................................................................ 15 1.3.2 WW4 domain ............................................................................................... 19 1.4 Objectives ........................................................................................................... 23 Chapter 2 Materials and methods ................................................................................ 24 2.1 DNA manipulation ............................................................................................. 25 2.1.1 Polymerase chain reaction (PCR) ................................................................ 25 2.1.2 Agarose gel electrophoresis and DNA fragment purification ..................... 25 2.1.3 DNA digestion and ligation ......................................................................... 26 2.1.4 Preparation of E. coli competent cell ........................................................... 26 2.1.5 Transformation of E. coli cells .................................................................... 27 2.1.6 Purification of plasmid ................................................................................ 27 II 2.1.7 DNA sequencing.......................................................................................... 27 2.3 Protein manipulation .......................................................................................... 28 2.3.1 Soluble protein (WW4) expression and purification ................................... 28 2.3.2 Insoluble protein expression and purification.............................................. 28 2.3.3 Preparation of isotope labeled proteins........................................................ 29 2.3.4 Determination of protein concentration ....................................................... 29 2.4 Circular Dichroism (CD) spectroscopy and sample preparation........................ 30 2.5 NMR experiments .............................................................................................. 30 2.5.1 1H-15N HSQC titration ................................................................................. 30 2.5.2 NMR structure determination ...................................................................... 32 2.5.3 Hydrogen/deuterium exchange .................................................................... 32 2.6 Calculation of electrostatic potentials ................................................................ 32 2.7 Data fitting of dissociation constants ................................................................. 33 Chapter 3 Studies of interactions between anions and intrinsically unstructured protein— cytoplasmic domain of ephrinB2 .................................................. 35 3.1 Solution conformation of the entire ephrinB2 cytoplasmic domain .................. 36 3.2 The effects of salts on the conformation studied by CD and NMR ................... 40 3.3 Selective binding of salts visualized by NMR ................................................... 41 3.4 Binding affinity .................................................................................................. 47 3.5 Residue specific anion binding .......................................................................... 49 3.6 Discussion .......................................................................................................... 51 Chapter 4 Studies of interactions between anions and well-folded protein WW4 ...... 54 4.1 Structural characterization of WW4 ................................................................... 55 4.2 Anion- and residue-specific binding .................................................................. 57 4.3 Quantitative assessment of binding affinity ....................................................... 65 III 4.4 Properties of binding surfaces for different anions ............................................ 67 References…………………………………………………………………………….74 Appendix……………………………………………………………………………...82 Publications…………………………………………………………………………...85 IV Summary The cytoplasmic domain of ephrinB proteins has been implicated to have important roles in bidirectional signalling pathways controlling pattern formation and morphogenesis. However, its structure remains unknown as it has been reported to be insoluble in buffer. Recently our group found that these “insoluble” proteins could be solubilized in unsalted water, so in my project we aim to study the structural characteristics of ephrinB2 cytoplasmic domain in water and salt’s effects to its structure. For the first time, we demonstrated that this cytoplasmic domain could be solubilized in unsalted water, with the N-terminal fragment highly unstructured while the C-terminal 33 residues adopt a similar conformation as the isolated ephrinB-33 whose NMR structure in buffer has been studied previously. This result raises another fundamental question as to whether being unstructured is because of the absence of salt ions. To answer this question, we systematically studied the effects of 14 different salts on the protein using CD and NMR HSQC titrations. The result shows that the addition of salts even up to 100 mM could not induce significant conformational change to the protein, indicating that ephrinB2 cytoplasmic domain is intrinsically unstructured. Surprisingly, during our research we found that the eight different anions of these salts could bind to the protein with high specificity at biologically relevant concentrations with high binding affinities. Ions are commonly believed to impose their effects on proteins by unspecific electrostatic screening. However, according to V our results, the binding seems to be both salt- and residue-specific, Besides, Na2SO4 turned out to be the strongest binder, with the apparent dissociation constant at about 2 mM. To further study the interaction characteristics between ions and proteins, we choose a small and well-folded protein WW4 domain, whose structure has been reported previously. Through NMR HSQC titrations, we reveal that the three anions, SO42-, Cl- and SCN-, could also bind to the well-folded protein at distinctive residues and affinities and SO42- is the tightest binder. Besides, we also interestingly found that with the existence of 20 mM sodium phosphate, the binding patterns of these three salts are totally changed and the binding affinities are largely reduced. However, the perturbation of binding patterns is not observed during titrations with the pre-existence of 150 mM sodium chloride. Our study reveals that the anion- and residue-specific binding not only happens for the unstructured protein but also for the well folded protein WW4. As all cellular processes occur in buffers, we suspect that ions have important roles in modulating protein functions, and many specific ion effects on proteins at low, physiologically relevant concentrations remains to be discovered yet. VI List of Tables Table 1 Residue-type specificity of salt binding to EphrinB2 51 Table 2 Apparent dissociation constants (Kd) for binding of salts to WW4 domain 66 Table 3 Exchange rates of backbone amide protons of WW4 70 VII List of Figures Figure 1.1 A typical ordering of cations and anions in Hofmeister series 3 Figure 1.2 Proposed mechanisms for specific anion effects 9 Figure 1.3 Chemical shift deviations of Hα, Cα and Cβ from random coil values 13 Figure 1.4 Function of Eph/ephrin bidirectional signaling pathway 17 Figure 1.5 Sequence alignment of the cytoplasmic domain of ephrinB proteins 17 Figure 1.6 Expanded region of the cytoplasmic functional subdomain ephrinB2 18 Figure 1.7 The diagram of the WWP1 protein domains 20 Figure 1.8 Sequence alignments of WW domains 21 Figure 1.9 Structures of the free and complexed WW4 domains 21 Figure 1.10 Binding of the 15N-labeled WW4 domain with Nogo-A 22 Figure 3.1 Preliminary structural characterization of the ephrinB2 cytoplasmic domain 37 Figure 3.2 Assignment of ephrinB2 cytoplasmic domain 38 Figure 3.3 NOEs plotted against amino acid sequence 39 Figure 3.4 Far-UV CD spectra of ephrinB2 cytoplasmic domain 40 Figure 3.5 Superimposition of two-dimensional 1H-15N NMR HSQC spectra 42 Figure 3.6 Chemical shift difference of 1H and 15N for residues of ephrinB2 cytoplasmic domain 43 Figure 3.7 Electrostatic surface of ephrinB2 cytoplasmic domain 45 Figure 3.8 Superimposition of two-dimensional 1H-15N NMR HSQC spectra with 46 NaCl, MgCl2, KCl, CaCl2 Figure 3.9 Superimposition of two-dimensional 1H-15N NMR HSQC spectra with 47 NaF, LiF and KCl Figure 3.10 Residue-specific chemical shift difference and Kd 48 Figure 3.11 Accessibility of ephrinB2 cytoplasmic domain 50 Figure 4.1 CD characterization of WW4 56 Figure 4.2 NMR HSQC titrations with Na2SO4, NaCl and NaSCN 57 Figure 4.3 NMR HSQC titrations of WW4 58 VIII Figure 4.4 NMR HSQC titrations with Na2HPO4. 59 Figure 4.5 NMR HSQC titrations by Na2SO4 in the pre-existence of 150 mM NaCl 61 Figure 4.6 NMR HSQC titrations with MgSO4 62 Figure 4.7 NMR HSQC titrations with KCl 63 Figure 4.8 NMR HSQC titrations with KSCN 64 Figure 4.9 Apparent dissociation constants for representative amide protons. 65 Figure 4.10 Binding sites on WW4 68 Figure 4.11 H/D exchange experiments of the WW4 domain 69 Figure 4.12 The electrostatic potential of WW4 at pH 6.4 and 4.0 70 IX Notions and Abbreviations cDNA Complementary DNA CD Circular Dichroism DNA Deoxyribonucleic Acid E. coli Escherichia coli Eph Erythropoietin-producing hepatocellular carcinoma HPLC High-performance liquid chromatography HSQC Heteronuclear Single Quantum Coherence IPTG Isopropyl β-D-1-thiogalactopyranoside LB Luria Bertani min Minute M(mM) Mole/L (Millimole/L) NMR Nuclear Magnetic Resonance NOE Nuclear Overhauser Effect PBS Phosphate-buffered Saline PCR Polymerase Chain Reaction ppm Parts Per Million TFA Trifluoroacetic acid WWP1 WW domain-containing protein 1 X Chapter 1 Introduction 1 1.1 Salt-protein interaction Life emerged from an inanimate, inorganic world where inorganic salts were key components and played important roles in directing the origin and evolution of life (Rode et al. 2007). Besides, in vivo, proteins are all exposed to an environment containing moderate to high concentrations of different ions and cosolutes. So ion-specific effects abound in both chemical and biochemical systems and the presence of ions could affect the stability, solubility, binding, enzyme activity, as well as crystallization of proteins (Saluja et al. 2009; Hochachka & Somero, 1984;Gokarn et al. 2009; Ries-Kautt et al. 1989). These effects exhibit a classical ranking called the Hofmeister series. 1.1.1 Hofmeister series Over 100 years ago Hofmeister and co-workers studied the ability of different salts to alter the solubility of proteins from egg whites and blood serum (Hofmeister, 1888; Kunz et al. 2004). They discovered that the egg white protein would precipitate out of solution at different concentrations of salt specifically related to the ion identities and this order of salt-protein interactions is then known as the “Hofmeister series” (Figure 1.1). Ions on the left side are known as kosmotropes, which tend to precipitate proteins from solution and stabilize proteins. On the other hand, the ions on the right side, chaotropes, are known to promote denaturation of proteins and increase protein solubility. Chloride is usually believed to be neutral between the two types of behavior. 2 At first most people focus their research on anions, as the effects of cations in promoting the precipitation of proteins are not as pronounced as anions. However, recent years more and more studies showed that the effect of cations on crystallization of proteins is comparable to anions, especially at low concentrations (Carbonnaux et al. 1995). A reasonable interpretation given (Kunz, 2009) was that both cations and anions can interact with proteins with specificities. People even found that specific co-ion of both anions and cations could bind to proteins and affect the solubility (Bénas et al. 2002). HOFMEISTER SERIES N(CH3)4 CO32- SO42- + NH4 + Cs + S2O32- H2PO4- Cations Rb+ K+ Na+ F- ClAnions kosmotropic surface tension harder to make cavity solubility hydrocarbons salt out (aggregate) protein denaturation protein stability Li+ Br- NO3- Mg2+ Ca2+ I- ClO4- SCN- chaotropic surface tension easier to make cavity solubility hydrocarbons salt in (solubilize) protein denaturation protein stability Figure 1.1 A typical ordering of cations and anions in Hofmeister series (Kunz, 2010) 1.1.2 Phenomenology of the Hofmeister series After the discovery of Hofmeister series, substantial attention has been paid to these phenomena because of their ubiquitous effects in both biology and chemistry ranging from protein folding/stability to enzymatic activities, as well as colloidal 3 assembly (Kunz, 2010). The effects on proteins were most extensively investigated as the series was got by studying salts’ effect on the solubility of egg white proteins. 1.1.2.1 Effect of salts on the stability of proteins In a thermodynamic study on the B1 domain of protein L (ProtL) by fluorescence spectroscopy, circular dichroism and differential scanning calorimetry, Xavier Tadeo et al. (2007) demonstrated the stabilization by kosmotropes to ProtL’s thermal denaturation and destabilization by chaotropes. The solubility of a basic protein, Peptibody A (PbA), was also found to be affected by ions. Besides, the solubility of PbA turned to be more affected by anions than by cations (Saluja et al. 2009). Enzymatic activity might be also mediated by the conformational changes induced by ions. Kosmotropes, which are commonly believed to stabilize the native conformation, usually can enhance enzyme activity (Hochachka & Somero, 1984). 1.1.2.2 Effect of salts on protein interactions Ions have strong effects on the solution interactions of proteins, not only protein-protein interactions, but also DNA-protein interactions. Gokarn et al. (2009) reported that the reversible oligomerization of a fusion protein peptibody A (PbA) is modulated by specific anion-protein interactions. Pedersen et al. (2006) also investigated the effects of different salts on fibrillation of glucagon. They found that ions could interact directly with glucagon fibrils like structural ligands and thus assist the formation of fibrils. 4 O’Brien et al. (1998) studied salts’ effects on the interactions of the TATA binding protein (TBP) with DNA. They found that only at high salt concentrations the protein-DNA interaction could be detected and ions appeared to influence the binding. A compelling explanation was that cations were incorporated in the interface between the electrostatically negative regions on the protein and the negatively charged DNA (O’Brien et al. 1998). 1.1.2.3 Effect of salts on enzyme activity It has been known for more than 40 years that the catalytic activity of enzymes is affected by specific ion types (Warren and Cheatum, 1966). A remarkable example was given by Hall & Darke (1995). They found that the catalytic efficiency of the herpes simplex virus type I protease undergoes 420-fold increase in the presence of 0.8 M sodium phosphate and 860-fold increase in the presence of 0.8 M sodium citrate. A recent study (Salis et al. 2007) also showed that Candida rugosa lipase is fully inactivated at 2 M concentration of NaSCN, while Na2SO4 can activate the lipase and NaCl acts quasi-neutrally. 1.1.2.4 Effect of salts on protein crystallization Salts are commonly used in protein crystallization by a salting-out process, which depends strongly on salt types. Gilliand (1988) compiled over 1000 crystal forms of more than 600 biological macromolecules and found that almost one third of the single crystals were obtained with the presence of ammonium sulfate, while acetate 5 and chloride were scarcely used. Chakrabarti (1993) also systematically analyzed the binding of 52 phosphate and sulfate ions in 34 different protein crystal structures and listed detailedly which amino acids could bind to the anions in these crystal structures. Ries-Kautt et al. (1989) investigated the effect of different ions on hen egg white lysozyme and got a reverse order of the Hofmeister series. On the contrary, in the study of the effects of anions on acidic Hypoderma lineatum collagenase, Carbonnaux (1995) got the result in agreement with the Hofmeister series. Thus they proposed that the effects of anions on protein crystallization may be dependent on the protein’s net charge. 1.1.3 The mechanism of Hofmeister series In view of the ubiquity of Hofmeister series, understanding the molecular-level mechanism is quite important for both biological and chemical systems. However, although numerous efforts have been paid to it, the underlying mechanism is still elusive and no unified theory has been achieved yet. 1.1.3.1 Water structure maker/breaker The classical view is that these interactions are induced by ions through the changed in bulk water structure (Collins and Washabaugh, 1985). It was proposed that the kosmotropes could enhance water structure surrounding ions, leading to the strengthening of the hydrophobic-bonding network of bulk water and thus stabilize the protein. On the contrary, the chaotropes could break water structure and thereby 6 denature the proteins. However, recent experimental findings exposed the water structure maker/breaker model under criticism. Through dynamic measurement of water molecules, Omta et al. (2003) found that ions could not influence the water structure outside the hydration shell of the ion. 1.1.3.2 Dispersion force taken into account In more recent studies, Ninham and colleagues (Ninham, 1997; Boström, 2001, 2002, 2003) took the nonelectrostatic and electrodynamic fluctuation (called “dispersion potential”) into account and treated it at the same level as the classical electrostatic forces. DLVO (Derjaguin-Landau-Verwey-Overbeek) theory is commonly used to describe the electrostatic forces using Poisson-Boltamann equation. However, in this method, the ions are only treated as point charges and thus the ion specificity is lost. So Ninham et al. (1997) introduced the dispersion potential into the theory and use this model to interpret specific ion effects on many chemical and biological phenomena. 1.1.3.3 Direct ion-protein interaction Recently, a more acceptable theory is direct ion-protein interactions (Zhang and Cremer, 2006). Through a series of literatures (Collins, 1995, 1997, 2004, 2006), Collins proposed the landmark theory “Law of Matching Water Affinities”: oppositely charged ions in free solution form inner sphere ion pairs spontaneously only when they have equal water affinities (Collins, 2004). They hold the view that the effects of 7 ions on water structures are limited to the first hydration shell and it provide a good basis for the idea of direct ion-protein interaction. According to the model, the weakly hydrated anions, like SCN-, match well with the weakly hydrated side chains of Arg, Lys, His, which are derivatives of ammonium and all positively charged. The strongly hydrated cations, by contrast, could interact more efficiently with the negatively charged and strongly hydrated side-chain carboxylates and backbone carbonyls. Collins also proposed that besides the charged side chains, the weakly hydrated anions could also interact with other polar and non-polar groups, which are considered weakly hydrated (Collins, 2004). Paul Cremer and his colleagues (Zhang and Cremer, 2010) also support the view of direct ion-protein interaction and they proposed a model of three interactions between anions and a elastin-like polypeptides (ELPs) (Figure 1.2). As shown, they thought the kosmotropes X- could polarize first-hydration-shell water involved in hydrogen bonding to the carbonyl of amide backbone (Figure 1.2a) and this could be manifest by hydration entropy values of the anion. The chaotropes, in contrast, would weaken the hydrophobic hydration of the protein by increasing the surface tension of the anions (Figure 1.2b). On the other hand, the chaotropic anions could also bind directly to the amide moieties (Figure 1.2c) and cause salting-in effects. This direct binding is a saturation effect. 8 Figure 1.2 Proposed mechanisms for specific anion effects on the LCST of ELP V5-120 (Cho, 2008). (a) Direct interactions of anions with water involved in hydrogen bonding to the amide. Kosmotropic anions polarize these water molecules and thereby weaken the hydrogen bonding of water to the macromolecule, a salting-out effect. (b) The blue lines represent the hydrophobically hydrated regions of the biomacromolecule. The cost of such hydration increases as salt is added to solution. (c) Direct ion binding of chaotropic anions to the amide moieties along the backbone of the polypeptide should cause a salting-in effect. 1.1.4 Specific protein-ion binding The study of interactions between ions and proteins has last for decades and in most studies, the interaction turned to be specific. However, these specific interactions seem to be highly dependent on salt concentrations. A typical example is the study of cloud-point temperature by Zhang and Cremer (2009). At low salt concentrations (< 9 0.5 M), the cloud-point of lysozyme follows an inverse Hofmeister series: ClO4- > SCN- > I- > NO3- > Br- > Cl-; while at high salt concentrations (0.8-1.5 M) it follows a direct Hofmeister series: SCN- < I- < ClO4- < Br- < NO3- < Cl- (Zhang and Cremer, 2009). In most previous studies, people focus their attention on moderate to high (> 0.1 M) salt concentrations as it is commonly believed that non-specific, electrostatic interactions are dominant at low salt concentrations (< 0.1 M), while at higher salt concentrations, the electrostatic interactions would be screened and thus the specific ion effect could be detected (Kunz, 2010). However, numerous studies related to ion-specific effects on protein solubility (Saluja et al. 2009), stability (Zhang and Cremer, 2009), self-association (Munishkina et al. 2004) at the low concentration range have been reported. A recent study (Gokarn et al. 2011) using effective charge measurements of hen-egg white lysozyme found that even at low salt concentrations (< 0.1 M), anions could selectively and preferentially bind to protein surface. Chakrabarti (1993) also systematically analyzed the binding of 52 phosphate and sulfate ions in 34 different protein crystal structures. In his review, he listed detailedly which amino acids could bind to the anions and even gave the possible geometric parameters for protein-anion interactions as well as hydrogen bonds between ions and amino acids. In view of all these evidences, further high-resolution view of ion-protein interactions is quite necessary to understand the ion-specific effects at low salt concentrations. 10 1.2 Nuclear Magnetic Resonance (NMR) spectroscopy Nuclear magnetic resonance (NMR) spectroscopy is a versatile and powerful technique to study the three dimensional structure, dynamics and interactions of macromolecules such as protein and nucleic acid at atomic resolution. It has been extensively developed since the first solution protein structure was determined from NOE derived distance restraints in 1985 by Kurt Wüthrich (Williamson et al. 1985), who was awarded a Nobel Prize in 2002. At the present time, the two major techniques to determine structures of macromolecules at atomic resolution are NMR spectroscopy and X-ray crystallography. Compared to X-ray crystallography which requires single crystals, NMR experiments are carried out in solution, where the buffer conditions such as pH, temperature and salt concentrations could be easily adjusted for different aims. What’s more, NMR could not only provide structural data but also conformational change, folding and intermolecular interaction information. 1.2.1 NMR phenomenon The phenomenon of nuclear magnetic resonance occurs when certain nuclei in a static magnetic field are exposed to a second oscillating magnetic field. The fundamental property nuclear spin (I), which is the angular momentum quantum number, decides whether a nucleus could have the NMR phenomenon. Basing on the number, the nuclei could be divided into three groups. One group has no spin with I=0 (e.g. O16), the second with integral spins (e.g. I=1, 2, 3) and the third with fractional 11 spins (e.g. I=1/2, 3/2, 5/2). For the first group, the nuclei are NMR inactive, while for nuclei with I>1/2, they are NMR active but difficult to detect. The most widely used in NMR are the nuclei with I=1/2 (e.g. 1H, 15 N, 13C), which could give interpretable signals. In the presence of an external magnetic field, spin of 1/2 will orient aligned with or opposed to the external field. There is slight energy difference between these two spin states, which is dependent on the strength of the external magnetic field. The spins in parallel to the external field with lower energy could be excited to antiparallel state by absorbing the energy difference in the form of electromagnetic radiation while the antiparallel state could also jump to parallel state by emitting energy. The difference between the populations of these two spin states contributes to the NMR signal. The sensitivity of NMR signal is also related to magnetogyrio ratio (γ) as the magnetic moment (μ) of the nucleus is μ=γIh, in which h is the Plan’s constant and γ (magnetogyric ratio) is a proportional constant for each particular type of nucleus. For the three most widely used nuclei (1H, 15 N and 13 C), 1H has the highest NMR sensitivity because 1H has the largest γ (γ1H/γ13C is about 4 andγ1H/γ15N is about 10) (Wüthrich, 1986). 1.2.2 Chemical shift When an atom is placed in an external magnetic field, the magnetic field at the nucleus is not equal to the externally applied magnetic field as the electrons circulate around the nucleus and cause a small magnetic field which is opposite the external field. In a molecule, the electron environment around the nuclei is diverse according 12 to the type of the nucleus and the bonds it forms. Thus the exact resonance frequency of each spin is different from the standard depending on its chemical environment and this difference is called chemical shift. Chemical shift is one important parameter to identify individual nucleus and assign the resonances in the spectrum to its site in chemical structure (Wüthrich, 1986). In a well-defined protein structure, even the same elements often have different chemical shifts because of different chemical environments. What’s more, chemical shift is also quite helpful to determine protein secondary structure (Wishart et al. 1991; Wishart and Sykes, 1994). The conformational shift of Hα, Cα and Cβ, which is the difference between the experimental chemical shift and random coil chemical shift, could be used to indicate the existence of α-helix or β-sheet (Figure 1.3). Figure 1.3 Chemical shift deviations of Hα, Cα and Cβ from random coil values (Wishart and Sykes, 1994). 13 1.2.3 Molecular interaction studies by NMR NMR spectroscopy is a quite useful and efficient tool to study molecular interactions. The binding of a ligand to the protein will change the chemical environment of the spins and thus cause chemical shift changed in the NMR spectrum. The amplitude of the changes is largely dependent on the distance between the spin and the binding site and therefore we could easily map out the binding surface of the protein and ligand. Besides, from the NMR titration experiments we could obtain the chemical shift difference (CSD) of each spin upon the binding of ligands and subsequently calculate the dissociation constants. As NMR spectroscopy is quite sensitive, even very weak interactions that could not be detected by other methods such as Isothermal Titration Calorimetry (ITC) could be studied using NMR. The technique “SAR by NMR” for rational design of drug molecules was developed basing on the chemical shift mapping using NMR (Shuker et al. 1996). 1.2.4 Study of protein-salt interactions using NMR Many investigations have been reported to prove that NMR is quite helpful to study weak protein-salt and protein-cosolute interactions (Jolivart et al. 1998; Foord et al. 1998) as well as water-protein interactions (Huang and Melacini, 2006). As the chemical shifts in NMR spectrum are quite sensitive to perturbation of proton environment, Jolivart et al. (1998) studied the interactions between thiocyanate and bovine pancreatic trypsin inhibitor (BPTI) using NMR spectroscopy following several experimental approaches. They first monitored the chemical shift variations of BPTI 14 protons upon addition of salts and found that up to 30 protons of 20 residues were significantly perturbed. Then the influence of thiocyanate on the electrostatic potential surrounding BPTI was studied using NOESY spectra selective at the water frequency (Jolivart et al. 1998). Similarly, Tadeo et al. (2007) also studied the chemical shift perturbation upon addition of salts using NMR 1H-15N HSQC spectra and found that the anions could strongly interact with protein surface, causing significant perturbation of amide protein chemical shift. Moreover, hydrogen exchange is also widely used to study the interactions between salts, osmolytes and proteins (Jolivart et al., 1998; Foord et al., 1998; Jaravine et al., 2000), as the NH resonances will exchange to ND in D2O and the NH signal decays. So in this thesis, I employed NMR spectroscopy to study the interactions between ions and proteins at low salt concentrations. 1.3 Biological background 1.3.1 Cytoplasmic domain of ephrinB2 Eph receptors, the largest family of receptor tyrosine kinases, together with their ephrin ligands form a complicated cell communication system which plays important roles in normal physiology as well as disease pathogenesis. The signaling system mediated by Eph-ephrin has been reported to have diverse functions ranging from neural development (Flanagan and Vanderhaeghen, 1998; O’Leary and Wilkinson, 1999; Xu et al. 1999) to pattern formation and morphogenesis (Holder and Klein, 1999; Schmucker and Zipursky, 2001). 15 In humans, fourteen Eph receptors have been identified, which are divided into two subclasses: nine EphA ((EphA1-8 and EphA10) receptors that bind to five glycosylphosphatidylinositol (GPI)-linked ephrin A ligands (ephrin A1-A5) and five EphB (EphB1-4 and EphB6) receptors that bind to three transmembrane ephrin B ligands (ephrin B1-B3) (Pasquale EB, 2005). The bindings within each class (EphA-ephrinA, EphB-ephrinB) are promiscuous, while there is also limited cross-binding between the two classes (e.g. EphA4-Ephrin B2). A distinction of the Eph-ephrin signaling pathway is the ability to generate bidirectional signals, meaning that both Eph receptors and ephrin ligands could either send or receive signals (Pasquale EB, 2005). The forward signaling that affect the Eph receptor-expressing cells is dependent on the tyrosine kinase domain, which could mediate the phosphorylation as well as the associations of the receptors and other proteins. The reverse signaling that affect the ephrin-expressing cells depends on the tyrosine phosphorylation of the cytoplasmic region of ephrins. The Eph-ephrin bidirectional signaling has been implicated to function in the growth and migration/invasion of cancer cells in culture, tumor growth, invasiveness, angiogenesis in vivo (Pasquale, 2010), intestinal homeostasis (Clevers and Batlle, 2006) as well as bone homeostasis (Zhao et al., 2006). In particular, Eph-ephrin bidirectional signaling in the nervous system have been extensively studied (Figure1.4). To date, most studies about the reverse signaling are focused on ephrin-B while the mechanisms of reverse signaling for ephrin-A are less understood. The proteins of the ephrin-B family are anchored to the membrane with an 16 extracellular ligand binding domain, a transmembrane domain and a cytoplasmic dimain (Figure 1.5). The cytoplasmic domain has about 83 residues and is quite conserved among different members in the family, especially the last 33 residues, which has about 90-100% identity (Figure 1.5). This striking conservation strongly indicates an important role of this cytoplasmic domain of ephrin-B proteins. Figure 1.4 Function of Eph/ephrin bidirectional signaling pathway in the nervous system (Scicolone et al. 2009). Figure 1.5 Sequence alignment of the cytoplasmic domain of ephrinB proteins. 17 Many functional studies have also identified that this domain is involved in bidirectional signaling pathways (Adams and Klein, 2000; Brückner et al. 1999; Lu et al. 2001; Cowan and Henkemeyer, 2001). It was reported that the phosphorylated N-terminal region of ephrin-B2301-333 could bind to the SH2 domain of the Grb4 protein and subsequently mediate cytoskeletal dynamics (Adams and Klein, 2000) (Figure 1.6). The C-terminal part of ephrin-B2301-333 was also identified to interact with PDZ domain of a novel PDZ-RGS protein and further control the signaling for cellular guidance (Xu et al., 1999) (Figure 1.6). Figure 1.6 Expanded region of the cytoplasmic functional subdomain ephrinB2301-333 with all five functionally important tyrosine residues (Tyr304, Tyr311, Tyr316, Tyr330, and Tyr331). (Song et al. 2002) In view of the functions of the cytoplasmic domain, the study of the structural properties of the cytoplasmic domain will help to understand the mechanism of the bidirectional signaling mediated by interactions between ephrinB proteins and the Eph receptors. The solution structure of the last 33-residue cytoplasmic subdomain has been studied (Song et al. 2002), with the first 22 residues of the sub-domain adopting 18 a well-packed β-hairpin followed by largely unstructured 11 residues. However, as previously published (Song et al. 2002), the 83-residue cytoplasmic domain of ephrin-B2 lack structure formation and tend to aggregate in buffer. Recently in our lab we discovered that these buffer-insoluble proteins could actually be solubilized in unsalted water with solution pH several units away from the pI of the protein (Song, 2009). Thus in my project, I studied the cytoplasmic domain of ephrinB2 in unsalted water as well as the effects of different salts to the protein. 1.3.2 WW4 domain WW domain-containing E3 ubiquitin protein ligase 1 (WWP1), also known as TGIF-interacting ubiquitin ligase (Seo et al. 2004), is a multifunction protein widely expressed in different organisms. In human WWP1 appears to have important roles in context-dependent manner in cancers (Nguyen Huu et al. 2008), infectious diseases (Galinier et al. 2002), neurological diseases (Qin et al. 2008) and even aging (Cao et al. 2011). It has been reported to have functions as the E3 ligase targeting several PY motif-containing proteins, such as KLF5 (Chen et al. 2005; Zhao et al. 2010), Smad2 (Seo et al. 2004), and Nogo-A (Qin et al. 2008), and many non-PY motif containing proteins, such as EPS15 (Chen and Matesic, 2007), KLF2 (Zhang et al. 2004) and Smad4 (Moren et al. 2005) (Figure 1.7B). WWP1 is involved in a variety of cellular processes including membrane protein trafficking, virus budding, protein degradation and signaling (Chen and Matesic, 2007; Harvy and Kumar, 1999). The gene WWP1 is quite conserved ranging from C. elegans, chicken to human. 19 The human WWP1 protein contains an N-terminal C2 domain, four WW domains and a catalytic HECT domain for ubiquitin ligase enzyme activity on the C-terminal (Chen and Matesic, 2007) (Figure 1.7A). The N-terminal domain mediates membrane and protein interactions (Plant et al. 1997; Wang et al. 2010). The four tandem WW domains are responsible for protein recognition and usually bind to Pro-rich polypeptide sequences (Sudol et al. 1995). Figure 1.7 (A) The diagram of the WWP1 protein domains. The C2 domain at N-terminus is responsible for membrane and protein binding. The four WW domains in the central region are responsible for the interaction with substrate proteins. The WW1 and WW3 are type I WW domains that recognize PY motifs. The HECT domain at the C-terminus is responsible for the ubiquitin–protein ligase activity. The Cystein-890 is the catalytic activation site. (B) WWP1 regulates different substrates in different cellular processes. (Zhi and Chen, 2012) The WW domains are quite conserved with three beta strands stabilized by many conserved aromatic and proline residues (Sudol, 1996), especially the WW4 domain (Figure 1.8). The conservation of WW4 ranging from C. elegans to humans suggests 20 the important role of this domain. Figure 1.8 Sequence alignments of the regions spanning the tandem WW domains 3 and 4. Arrows indicate the domain boundaries bordered by prolines, and beta-strand elements are boxed. Rows are labeled as follows: Dm, Su(dx) of Drosophila melanogaster; Ce, WWP1 of Caenorhabditis elegans; Mm1, Itch of Mus musculus; Mm2, WWP1 of M. musculus; Hm1, WWP1 of Homo sapiens; Hm2, WWP2 of H. sapiens; Hm3, AIP4 of H. sapiens. (Fedoroff et al. 2004) The three dimensional solution structure of the WW4 has been determined by our group using experimental NMR distance and dihedral angle constraints (Qin et al. 2008). As shown in Figure 3a, the WW4 adopts a WW domain typical three-stranded β-sheet, with a short 310-helix over residues Pro34-Asn36 on the C-terminal. The first β-strand ranges from Gly8 to Tyr13, the second from Arg19 to His24 and the third from Thr28-Ph31. Figure 1.9 Structures of the free and complexed WW4 domains. (a) Superimposition of the 10 selected NMR structures of the WW4 domain having the lowest target functions in ribbon mode. (b) Best docking solution of the WW4 domain (gray) in complex with the Nogo-A peptide (yellow). The side chains of the binding-important residues are displayed as sticks and labeled for both the WW4 domain and Nogo-A peptide. (Qin et al. 2008) 21 Binding studies between WW4 domain and Nogo-A peptide using 1H-15N HSQC titration (Qin et al. 2008) showed that there are mainly two binding regions, one over Trp9-Ile11 on the firstβ-strand and the other over Thr28-Thr30 on the third β-strand (Figure 1.10). The most significantly perturbed residues are Thr28, Thr29 and Thr30 on the third β-strand. The complex model of WW4 and Nogo-A peptide generated from docking also support the results of HSQC titration (Figure 1.9b). As the solution structure of WW4 domain has been well studied and the protein is relatively small with well-folded structures, we choose this protein as a model to study the interactions between anions and well-structured proteins. Figure 1.10 Binding of the 15N-labeled WW4 domain with Nogo-A (650-666). (a) Superimposition of the HSQC spectra of the 15N-labeled WW4 domain in the absence (blue) and presence (red) of Nogo-A(650-666) at a 1:4 WW4:peptide molar ratio. The residues with significant peak shifts upon binding are labeled in the spectra. (b) Residue-specific chemical shift difference (CSD) of WW4 induced by binding to Nogo-A(650-666). (Qin et al. 2008) 22 1.4 Objectives As mentioned above, the cytoplasmic domain of ephrinB2 has been reported to play important roles in the bidirectional signaling pathways. But the solution structure of the cytoplasmic domain remains unknown due to its insolubility in buffer. At first, my project aims to investigate the structural characteristics of the cytoplasmic domain and subsequently generate some mutants that could be buffer-soluble. However, during my research, we interestingly found that different anions could interact with the protein specifically. Thus the aims of this research are: 1. To study the structural characteristics of ephrinB2 cytoplasmic domain in unsalted water; 2. To study the effects of different salts to the intrinsic unstructured protein; 3. To study the characteristics of the interactions between anions and unstructured proteins; 4. To study the characteristics of the interactions between anions and well-folded protein WW4. 23 Chapter 2 Materials and Methods 24 2.1 DNA manipulation 2.1.1 Polymerase chain reaction (PCR) The gene of the cytoplasmic domain of ephrinB2 was amplified from the cDNA of ephrinB2 (from Elena B. Pasquale’s group) using the thermostable DNA polymerase pfu and a pair of primers (Forward: 5’-CGC GGA TCC AAG TAC CGG AGG A-3’, Reverse: 5’-CCG CTC GAG TCA GAC CTT GTA GTA A-3’). The PCR reaction mixture (μl) contains: 5μl 10×pfu Buffer (Promega), 0.5μl forward primer (100μM), 0.5μl reverse primer (100μM), 0.5μl pfu DNA polymerase (2-3U/μl), 0.5μl 10mM dNTPs, 1μl DNA templates (20-100ng/μl), 42μl distilled water top up to 50μl. The PCR is running on thermal cycler (Bio-Rad, USA) with a program as followed: 95˚C, 5min; 30cycles of 94 ˚C, 30sec, 55 ˚C, 30sec, and 72 ˚C, 30sec; finally 72 ˚C, 10min and then 4 ˚C hold. 2.1.2 Agarose gel electrophoresis and DNA fragment purification The PCR products were separated using electrophoresis on 2% agarose gel with 0.5μg/mL ethidium bromide (EB) in 1×TAE buffer (0.04M Tris-acetate, 1mM EDTA, pH 8.0) under constant voltage 100 V for 15minustes. The DNA ladder and 5×loading dye were from Promega. The band with correct size was cut off under UV and then purified using Qiagen Gel Extraction Kit according to the manual. 25 2.1.3 DNA digestion and ligation The purified PCR products were digested with nuclease restriction enzymes BamHI and XhoI from New England Biolabs based on the manual. After digestion at 37 ˚C for proper time, the digested DNA products were then separated on 2% agarose gel and further purified using Qiaquick Gel Extraction Kit (Qiagen). T4 DNA ligase from New England Biolabs was used to ligate the DNA and the pET32a vector. The ligation was performed at room temperature for one hour. 2.1.4 Preparation of E. coli competent cell The frozen stock of E. coli cells were spread on LB agar plates and cultured overnight at 37 ˚C. Then a single colony was picked and inoculated into 5-10ml LB media without antibiotics and grown at 37 ˚C overnight (~14hours) with shaking. The overnight culture was diluted into 100 ml LB at a ratio of 1:100 and cultured at 37 ˚C (200 rpm) until OD600 reached 0.4~0.6. The cells were then chilled on ice for 10 min and spun down at 2700 g for 10 min at 4 ˚C. The cell pellets were resuspended gently in 30 ml sterile, ice-cold MgCl2-CaCl2 solution (80 mM MgCl2, 20 mM CaCl2) and incubated on ice for 15 min. Then cells were collected again at 2700 g for 10 min at 4˚C. After decanting the supernatant, the cell pellets were resuspended in 4 ml ice-cold 100 mM CaCl2 with 15% glycerol and quickly frozen by liquid nitrogen as 50~100 μl aliquots and stored at -80 ˚C. 26 2.1.5 Transformation of E. coli cells The competent cells were thawed on ice and then mixed gently with the transforming DNA, either plasmid or ligation products (10-100ng). After incubated on ice for 20-30 min, the cells were heat-shocked at 42 ˚C for 90 sec and immediately chilled on ice for another 2 min. Then 600 μl LB was added into the tube and incubated at 37 ˚C for one hour with shaking. Subsequently the recovered cells were spread onto LB agar plated with ampicillin or other proper antibiotic and incubated at 37 ˚C overnight. 2.1.6 Purification of plasmid The plasmids were amplified in DH5α E coli. strain. Single bacterial colony containing the interested plasmid was picked and grown in 5mL LB with proper antibiotic at 37 ˚C overnight. Then the cells were harvested and the plasmids were extracted following the manual of QIAprep Spin Miniprep Kit (Qiagen). 2.1.7 DNA sequencing The plasmid DNA was first amplified using PCR. The reaction system contains: 2 μl of 5×sequencing buffer, 1μl of BigDye, 0.4 μl forward for reverse primer (10μM), 200-500ng DNA and MilliQ water to a final volume of 10μl. The sequencing reaction was performed on thermal cycler (Bio-Rad, USA) with a program as followed: 95 ˚C, 5min; 25 cycles of 94 ˚C, 30 sec, 50 ˚C, 10 sec, and 60 ˚C, 4 min; then hold at 4 ˚C. 27 2.3 Protein manipulation 2.3.1 Soluble protein (WW4) expression and purification The expression vector for WW4 is pGEX-4T-1 and the construct was transformed into E.coli BL21 (DE3) competent cell for expression. A single colony was first inoculated into 10 ml LB with Ampicillin and cultured at 37 ˚C (200rpm) overnight. The overnight culture was inoculated into 1000 ml and grown until OD600 reach 0.6. Then the cells were induced by 0.3mM IPTG for expression overnight at 20˚C. The next day the culture was harvested at 6000 rpm for 10 minutes. The cell pellet was suspended in phosphate buffer (20mM NaH2PO4, 150mM NaCl, pH 6.8) and further sonicated on ice thoroughly. The supernatant was separated from pellet at 18,000 rpm for 30 minutes and then bound with Glutathione Sepharose 4B, which had been equilibrated with PBS, for one hour at room temperature. Then the beads were washed with 10× column volumes (CV) PBS and subjected to thrombin cleavage at room temperature. The samples at each step during purification were collected and analyzed on 15% SDS-PAGE. The protein WW4 released from thrombin cleavage was further purified by reverse-phase HPLC on a C18 column and lyophilized. 2.3.2 Insoluble protein expression and purification The gene of human ephrinB2 cytoplasmic domain was cloned into pET32a (Novagen) between BamHI and XhoI and transformed to E. coli BL21 (DE3) for 28 expression as the method for soluble proteins. After sonication, the recombinant protein was found to be expressed in the inclusion body. So the protein was purified under denaturing conditions. After dissolved in mother buffer (8M urea), the protein solution was centrifuged at 18,000 rpm for 30 minutes. Then the supernatant was allowed to bind with Ni-NTA beads, which had been equilibrated with mother buffer, for one hour. Then the beads were washed with 5× column volumes washing buffer. Finally, the His-tagged protein was eluted with elution buffer and further purified by reverse-phase HPLC on a C4 column. The elution fractions from HPLC were then lyophilized and the powder was stored at -80 ˚C. 2.3.3 Preparation of isotope labelled proteins The preparation of 15 N or 15 N, 13 C-labeled proteins were similar to unlabeled sample except that the cells were grown in M9 medium with addition of 15 15 NH4Cl for N-labeled sample or 13C-glucose and 15NH4Cl for 15N, 13C-labeled sample. 2.3.4 Determination of protein concentration The concentration of protein samples were determined by measuring UV absorbance at 280nm, A280nm. The protein concentration could be obtained using the Beer-Lambert Law: A=εlC [where ε is the molar absorption coefficient (M-1cm-1), l is the path length (cm), C is protein concentration (M)]. For proteins containing Trp, Tyr, Cys, ε can be calculated using the equation: ε280nm (M-1cm-1) = (#Trp)(5,500)+(#Tyr)(1,490)+(#cystine)(125) 29 where #Trp, #Tyr, #cystine are the number of corresponding residues in the protein (Pace et al. 1995). 2.4 Circular Dichroism (CD) spectroscopy and sample preparation Far-UV CD spectra were recorded on a Hasco-810 spectropolarimeter equipped with a temperature controller. All spectra were acquired at 25 ˚C using a 0.1 cm path length cuvette with a scan speed of 50 nm/min, 0.1 nm spectral resolution and 1 nm band width. Each spectrum was generated by the average of three scans to reduce noise and random error. All protein samples were dissolved in MilliQ water with a concentration of 25 μM at different pH. To study the effects of different salts on the secondary structure of the proteins, different salt stock solutions with the sample pH as the protein samples were gradually added into the protein samples. 2.5 NMR experiments All NMR experiments were acquired at 25 ˚C on an 800 MHz Bruker Avance spectrometer equipped with pulse field gradient units and a shielded cryoprobe. All NMR data were processed with NMRPipe (Delagio, 1995) and then analyzed with NMRview (Johnson, 1994) on Linux workstation. 2.5.1 1H-15N HSQC titration To prepare NMR samples, the protein powder was dissolved in 450 μl Milli-Q 30 water with 50μl D2O for NMR spin-lock. For ephrinB2 cytoplasmic domain, the pH was adjusted to 4.0 with microliter amounts of diluted sodium hydroxide at a protein concentration of 300 μM; while for WW4, the pH was adjusted to 4.0 and 6.4 at a protein concentration of 250 μM. To characterize the interactions between ions and the proteins, a series of two-dimensional 1H-15N HSQC spectra of the 15 N-labeled protein samples were acquired at a concentration of 300 μM for ephrinB2 cytoplasmic domain and 250 μM for WW4. For the titration of ephrin-B2 cytoplasmic domain, the stock solution of 14 different kinds of salts (Na2SO4, NaH2PO4, NaF, NaCl, NaBr, NaNO3, NaI, NaSCN, MgCl2, KCl, CaCl2, GdmCl, LiF, KF) were prepared at a concentration of 1M with pH 4.0 and gradually added into the protein samples at varying salt concentrations (0 mM, 1 mM, 3 mM, 5 mM, 10 mM, 20 mM, 50 mM, 100 mM). The pH of the protein samples before and after adding salts were measured and there were no significant change of pH values caused by addition of salts. 15 N-labeled WW4 were titrated under three different solution conditions: pH 4.0 in pure water, pH 6.4 in water adjusted by addition of microliter amounts of diluted sodium hydroxide and pH 6.4 in 20 mM sodium phosphate buffer. The pH values of the stock solutions of three salts (Na2SO4, NaCl and NaSCN) were adjusted to 4.0 and 6.4 accordingly and titrated at varying salt concentrations of 0 mM, 1 mM, 3 mM, 6 mM, 10 mM, 15 mM, 20 mM, 30 mM, 40 mM, 60 mM, 80 mM, 100 mM, 125 mM, 150 mM and 200 mM. 31 2.5.2 NMR structure determination For the sequential assignments of ephrinB2 cytoplasmic domain, triple-resonance experiments HNCACA and CBCA(CO)NH were recorded on a 15 N, 13 C-labeled sample at a protein concentration of 800 μM. For NOE analysis and Hα chemical shifts, 15 15 N-edited HSQC-TOCSY and HSQC-NOESY spectra were acquired on the N-labeled sample at a protein concentration of 800 μM. The chemical shift of 1H was directly referenced to 2, 2-dimthyl-2-silapentanesulfonic acid (DSS), while the chemical shift of 15N was indirectly referenced to DSS (Wishart et al., 1995). 2.5.3 Hydrogen/deuterium exchange The 15 N-labeled NMR sample of WW4 in water at pH6.4, together with an unlabeled sample at the same protein concentration and same buffer condition, were lyophilized. The unlabeled sample powder was first dissolved in D2O (pD 6.4) for tuning, matching and shimming. Then the powder of the 15 N-labeled sample was dissolved in D2O (pD 6.4) and transferred to an NMR tube. After two minutes, sequential two-dimensional NMR HSQC spectra were recorded until 200 minutes when all NH peaks were almost completely exchanged. The decreasing amplitude of every HSQC peak was fitted to a single exponential to obtain the exchange rate constants (Kex). 2.6 Calculation of electrostatic potentials The structure of the last 33 residues of ephrinB2 cytoplasmic domain and WW4 32 were previously published (Song et al. 2002, Qin et al. 2008). As the whole ephrin B2 cytoplasmic domain is unstructured, we generated the extended structure for the first 50 residues plus the 16-residue His-tag using CYANA program (Güntert, 2004). The structures were first converted from PDB format to PQR format using PDB2PQR to be prepared for continuum electrostatics calculations (Dolinsky et al. 2004). Then the electrostatic potential surfaces of the proteins at different pH were calculated by Adaptive Poisson-Boltzmann Solver (APBS) using the PQR files (Baker et al. 2001). 2.7 Data fitting of dissociation constants To obtain the residue-specific apparent dissociation constants, all the HSQC spectra with and without salts in salt titration were superimposed together to identify the shifted peaks and then assign to corresponding residues. The shifted peaks were traced to get the differences in 1H and 15N chemical shifts of the HSQC peaks without salts and in presence of different salts at varying concentrations. Then we separately fitted the 1H and 15N chemical shift changes using on binding site model (Macomber, 1992) to obtain residue specific dissociation constants (Kd) using the DataFit curve fitting and data plotting software (Oakdale Engineering). The ions were treated as ligand and the protein and the ions form equilibrium: P+L PL. Thus at a given set of conditions: [ ][ [ ] ] 33 The function we used to fit the dissociation constants (Kd) is: [ ] [ ] [ [ ] ( [ ] ) [ ] ] [ ] ⁄ where δ is the chemical shift difference, [L] is the concentration of ions, [P] is the concentration of proteins. 34 Chapter 3 Studies of Interactions Between Anions and Intrinsically Unstructured Protein—Cytoplasmic Domain of EphrinB2 35 Results 3.1 Solution conformation of the entire ephrinB2 cytoplasmic domain The structure of the entire ephrinB2 cytoplasmic domain remains unknown due to its insolubility. During purification, I did find that most of the His-tagged recombinant protein was expressed in the inclusion body and only a small quantity was expressed in the supernatant, which precipitated quickly during the procedures of purification. Our previous studies have revealed that many buffer-insoluble proteins could be solubilized in pure water (Li et al. 2006, Song 2009). Thus we purified the cytoplasmic domain with 16 residues N-terminal His-tag under denaturing condition in the presence of 8 M urea and succeeded to solubilize the lyophilized protein powder in unsalted water at pH 4.0. To study the protein under physiological pH, I tried to adjust the pH to 6.0 with highly diluted sodium hydroxide. However, the protein was very sensitive to pH and started to precipitate quickly at pH 5.0. Thus all the subsequent studies were carried out under pH 4.0. In the preliminary structural studies using far-UV CD (Figure 3.1A) and NMR 1 H-15N HSQC spectroscopy (Figure 3.1B), the cytoplasmic domain turned out to be largely unstructured. In the far-UV CD spectrum, it is negative at 195 nm; while in the HSQC spectrum, it has quite narrow resonance lineshape and limited 1H chemical shift dispersion. Both these results indicate that this protein is highly unstructured without tight tertiary packing. 36 Figure 3.1 Preliminary structural characterization of the ephrinB2 cytoplasmic domain. (A) Far-UV CD spectrum of the ephrinB2 cytoplasmic domain at a protein concentration of 20 μM (pH 4.0) at 25˚C. (B) Two-dimensional 1H-15N NMR HSQC spectrum at a protein concentration of 300 μM (pH 4.0) at 25˚C. To further characterize the structural properties of ephrin-B2 cytoplasmic domain, I acquired a pair of triple resonance experiments HNCACB and CBCA(CO)NH to achieve the sequential assignment of the whole domain (Figure 3.2A). Then we obtained the chemical shifts for the Hα and NOEs through a pair of 15 N-edited HSQC-TOCSY and HSQC-NOESY. The Hα chemical shifts of the residues, especially the N-terminal part, are quite small and disordered, supporting previous conclusion that the domain is largely disordered (Figure 3.2B). The absence of medium- and long- range NOEs (Figure 3.3) is another evidence that the N-terminal part of the protein is predominantly disordered. 37 N(ppm) 15 Figure 3.2 Assignment of ephrinB2 cytoplasmic domain. (A) Sequential assignment of the entire ephrinB2 cytoplasmic domain. (B) Hα conformational shifts of the full-length domain (83 residues) plus a 16-residue His-tag (blue bars) and of the isolated last 33 residues called ephrinB-33 (red bars) previously reported (Song et al. 2002). The first 7 residues of the His-tag could not be assigned due to their missing side-chain resonances. Lys17 is the starting residue of the cytoplasmic domain while Cys67 is the starting residue of ephrinB-33. 38 Figure 3.3 NOEs plotted against amino acid sequence However, strikingly, we can observe that the last 33 residues (Cys67-Val99) have relatively large conformational shifts, almost the same as those of the isolated 33 residues (Figure 3.2B), except Cys67, which becomes the N-terminal in the isolated ephrinB-33. This result suggests that no matter in the entire domain or as isolate ephrinB-33, these 33 residues adopt a similar conformation. 39 3.2 The effects of salts on the conformation studied by CD and NMR To examine whether salts could induce any conformational change to the cytoplasmic domain in unsalted water, we assessed the effects of 8 different salts with the same cations but different anions (Na2SO4, NaH2PO4, NaF, NaCl, NaBr, NaNO3, NaI and NaSCN) at five different salt concentrations (0 mM, 1 mM, 10 mM, 50 mM and 100 mM) by far-UV CD. Figure 3.4 Far-UV CD spectra of the 83-residue cytoplasmic domain of the human ephrin-B2 at 20µM (pH 4.0) at 25 ˚C, in the absence and in the presence of 8 different salts at varying salt concentrations. The CD spectra in the presence of Na2SO4, NaH2PO4 and NaF are of high quality, but show little or no change with the addition of salts at four different concentrations compared to the reference spectrum (Figure 3.4). These results indicate that the 40 addition of salts at low concentrations could not induce folding or any conformational change to the protein. Because of high noise level, the CD spectra with NaBr, NaNO3, NaI and NaSCN are of low quality and the data are less conclusive. To further verify this conclusion, we subsequently studied the effects of salts using 1H-15N NMR HSQC spectrum. By gradually adding salts into the protein samples of the entire domain solubilized in unsalted water at pH 4.0, we acquired the HSQC spectra at different salts concentrations and then superimposed these spectra together. From the results we found that addition of salts could not induce significant increase to the dispersion of the spectra but only shifts of peaks (Figure 3.5), suggesting no conformational changes with the addition of salts, which is consistent with the results of far-UV CD. Thus, both the results of far-UV CD and NMR spectroscopy indicate that the entire ephrinB2 cytoplasmic domain is intrinsically unstructured. Being unstructured is an intrinsic property but not because of the absence of salt ions in the solution. 3.3 Selective binding of salts visualized by NMR As previously described, with the addition of salts, there was no increase of the HSQC spectral dispersion but only shifts of peaks. However, surprisingly we found that even though all salts could cause shifts of HSQC peaks, the salts effects were not uniform (Figure 3.5). Thus we separately mapped out the changes of 1H and 15 N chemical shifts by tracking the shifts of every HSQC peak with the increase of salt concentrations (Figure 3.6). 41 Figure 3.5 Superimposition of two-dimensional 1H-15N NMR HSQC spectra at a protein concentration of 300µM (pH 4.0) at 25 ˚C, in the absence and in the presence of 8 different salts at varying salt concentrations. 42 Figure 3.6 Chemical shift difference (CSD) of the amide proton (1H) and nitrogen (15N) for residues of 43 the ephrinB2 cytoplasmic domain induced by the addition of 8 salts at three different concentrations. From the figures we could observe that the patterns of the chemical shift difference (CSD) are quite diverse among different salts. For the salts on the right side of the Hofmeister series, NaCl, NaBr, NaNO3, NaI and NaSCN, the patterns are almost the same. However, for the salts on the left side of the series, Na2SO4, NaH2PO4 and NaF, the patterns are quite diverse. We also noticed that most of the residues with relatively large HSQC peak shifts upon the addition of salts are on the N-terminal part of the domain. Thus we calculated the electrostatic potentials of the extended structure of the first 50 residues (with the 16 residues N-terminal His-tag) (Figure 3.7C), as this fragment of the protein is highly unstructured. We also separately calculated the electrostatic potential for the NMR structure of the ephrinB-33 (Figure 3.7A, B) because the C-terminal 33 residues adopt similar conformation both as isolated ephrinB-33 or in the whole domain as we previous demonstrated in 3.1. The electrostatic surface of the N-terminal fragment up to Arg43 turns out to be quite positive, while the C-terminal ephrinB-33 has a relatively neutral surface. Thus we infer that one possible reason for the residues with large shifts focused on the N-terminal half during salt titrations is electrostatic attraction, in which the anions make great contributions. However, the attractions should be salt specific as the effects of different salts are quite diverse. 44 Figure 3.7 Electrostatic surface of ephrin-B2 cytoplasmic domain. (A) NMR structure of ephrinB-33 previously determined (Song et al. 2002) and its electrostatic surface (B, C) Electrostatic surface of an extended structure consisting of the first 50 residues of the ephrinB2 cytoplasmic domain plus the N-terminal His-tag. To determine the contributions of cations to the shifts of HSQC peaks during titration with different salts, we subsequently titrated the protein with chloride salts and fluoride salts with different cations [MgCl2, KCl, CaCl2, Guanidinium chloride (GdmCl), LiF and KF]. With the addition of salts, the patterns of HSQC peak shifts are almost the same among different chloride salts at low concentrations (Figure 3.8). This indicates that the salts effects are mostly induced by the anions, not the cations. For the fluoride salts, we got similar results (Figure 3.9). At higher concentrations (>50 mM), the peaks shifts are slightly different among salts with different cations, especially the last residue Val99. This may be because of relevant cation binding at high concentrations. The effects of cations are indirect compared to anions binding, as they are known to affect the clustering of anions in solution (Bian et al. 2011). Besides, 45 as Val99 is the last residue on the C-terminal, its unusual behavior might be due to the interactions between its free carboxyl group and the cations of the salts we used for titration. To sum up, all these results indicate that the effects that salts induce HSQC peak shifts are mostly triggered by anion binding, and this binding is anion specific. The effects of cations are quite weak and indirect. Figure 3.8 Superimposition of two-dimensional 1H-15N NMR HSQC spectra at a protein concentration of 300µM (pH 4.0) at 25 ˚C, in the presence of NaCl, MgCl 2, KCl, CaCl2 and CdmCl at different concentrations. 46 Figure 3.9 Superimposition of two-dimensional 1H-15N NMR HSQC spectra at a protein concentration of 300 µM (pH 4.0) at 25 ˚C, in the presence of NaF, LiF and KCl at different concentrations. 3.4 Binding affinity To obtain the binding affinity between the protein and ions, we plotted the chemical shift difference with the addition of ions as a function of salt concentrations. We employed the one binding site model to separately fit the 1H and 15 N chemical shifts for all residues with significant peak shifts. The residue-specific dissociation constants (Kd) obtained for eight salts with different anions are summarized in Appendix Table. Interestingly, all eight anions show high binding affinity with the ephrinB2 cytoplasmic domain, with majority of Kd values less than 50 mM. 47 Figure 3.10 Residue-specific chemical shift difference and apparent dissociation constants (Kd). Left panel: chemical shift difference (CSD) of 1H and 15N for all residues of three representative salts (Na2SO4, NaF and NaSCN) at 50 mM (blue bars) and 100mM (red circles). Residues with significant changes are labeled. Right panel: Experimental (red dots) and fitted (blue lines) values for CSD of 1H and 15N. We choose two representative residues for each salt: His25/Ser26 for Na2SO4, Ser26/Asp76 for NaF, His25/Ser26 for NaSCN. 48 Here we choose two representative residues for each salt (Na2SO4, NaF and NaSCN) and show their apparent dissociation constants (Kd) as well as fitted curves in Figure 3.10. Surprisingly, Na2SO4 turned out to be the strongest binder with an average Kd of 1 mM, consistent with previous studies which demonstrated that Na2SO4 is the strongest stabilizer for ribonuclease (Ramos and Baldwin, 2002). The result also rationalizes a recent report in which it was demonstrated that Na2SO4 has the strongest ability to reduce the effective charge of hen-egg white lysozyme (Gokarn et al. 2011). 3.5 Residue specific anion binding As previously described, the binding of ions to the protein is not uniform along the sequence. Most of the residues with significant shifts are focused on the N-terminal fragment of the protein, which is predominantly unstructured. On the other hand, the N-terminal fragment seems to have similar conformation as in the isolated ephrinB-33. Thus we wonder whether the binding difference is because of the difference of accessibility of all residues. So we calculated the accessibility of all side chains and amide protons based on the extended model of ephrinB2 cytoplasmic domain as well as the ephrinB-33 NMR structure (Figure 3.11), with all atoms included for calculation using CNS (Brunger et al. 1998). The amide protons of the last 33 residues seem to be relatively less accessible, which may be one reason for the less perturbation of these residues by ions. However, this could not be the main reason because the accessibility of the first 66 residues are almost the same, but the 49 ion-binding region mainly focus on residues Arg14-His29, which have continuously positive-charged surface (Figure 3.7). Figure 3.11 Accessibility of side chains (blue bars) and amide protons (red bars) of the extended model of residues 1-66 and ephrinB2-33 structure (Song et al. 2002), which were separately calculated but displayed together. All atoms were included for calculation with CNS using the method described by Lee & Richards (1971). The sequence of the whole ephrin-B2 cytoplasmic domain is shown in upper panel with residues differentially colored: blue for positively charged, red for negatively charged, purple for neutral polar and green for hydrophobic. Subsequently we categorized all the residues into four groups: positively charged, negatively charged, neutral polar and hydrophobic (Figure 3.11) and calculated the average Kd for each group to see whether the specific binding is related to the type of the residue. The results are summarized in table 1. As expected, the positively charged residues could bind stronger than other types of residues. For many salts like NaCl and NaSCN, the binding affinity of negatively charged residues and hydrophobic 50 residues are quite similar, suggesting that the binding between ions and residues mainly occur on the backbones, not side chains. Besides, we can also observe that the binding affinities of all eight anions studied to the hydrophobic residues are almost the same. Table 1 Residue-type specificity of salt binding to ephrinB2 Residue type1 Salt Na2SO4 NaH2PO4 NaF NaCl NaBr NaNO3 NaI NaSCN Charged, positive 1.2 (0.7)2 5.2 (6.9) 73.4 (13.4) 18.0 (16.2) 10.1 (24.5) 6.9 (10.3) 11.0 (17.9) 10.5 (15.5) Charged, negative NA3 (NA) 59.6 (13.7) 8.1 (11.3) NA (25.6) NA (18.3) 14.6 (37.3) 20.3 (22.3) 19.4 (21.6) Neutral, polar 1.2 (1.0) 8.0 (4.6) 46.6 (15.0) 10.1 (11.5) 12.5 (18.6) 8.3 (13.3) 72.1 (17.9) 11.4 (14.6) Hydrophobic 16.8 (2.3) 19.6 (16.2) 18.0 (9.1) 18.6 (25.9) 25.8 (30.5) 18.9 (28.9) 14.6 (95.6) 18.3 (24.4) 1 Amino acids were divided into 4 categories: charged positive (K, R, H), charged negative (D, E), neutral polar (N, Q, S, T, C, M, F, W, Y), hydrophobic (G, A, V, I, L, P). 2 The apparent dissociation constants Kd computed from the changes of 1H and 15N chemical shifts upon salt titrations are averaged for each residue type: note that only residues for which binding was detected and quantified are considered. Results based on 15N chemical shifts are given in parenthesis. 3 NA: not available. 3.6 Discussion In our research, we systematically studied the effects of fourteen salts with eight different anions on the conformation of buffer-insoluble protein, ephrin-B2 cytoplasmic domain. For the first time, we demonstrated that this cytoplasmic domain could be solubilized in unsalted water, with the N-terminal fragment highly unstructured while the C-terminal 33 residues adopt a similar conformation as the isolated ephrinB-33 whose NMR structure in buffer has been studied previously (Song et al. 2002). According to the results of CD and NMR studies, the addition of salts 51 even up to 100 mM could not induce significant conformational change to the protein, suggesting that being unstructured of ephrin-B2 cytoplasmic domain is an intrinsic property. Surprisingly, in our studies we found that by eliminating the interference of background buffer ions, eight anions of fourteen different salts could directly bind to the protein with high selectivity and saturation at physiological relevant salt concentrations. More importantly, the binding of the anions to the residues are not uniform along the sequence, but shows preference. Most of residues with relatively large chemical shift changes upon the binding of ions are located on the highly disordered N-terminal of the protein. One reasonable explanation is that the side chains and amide groups are highly accessible at this region. However, water accessibility could not be the main reason because the main binding region of eight anions is Arg14-His29, which constitutes a continuous positively charged surface. This indicates that besides the conformations, the electrostatic property of the residues is also responsible in mediating the specific protein-ion binding. This result is consistent with previous idea that at low salt concentrations (< 100 mM) electrostatic interactions play dominant roles in protein-ion interactions, not directly related to Hofmeister effects. On the other hand, we also observed that ions could bind to negatively charged and non-polar residues, and interestingly the binding affinities of these two kinds of residues show no big difference, suggesting that the ions seem to bind to the main chain of the residues not the side chains. From the results, we also noticed that NaF appears to be quite different from 52 other salts as it could bind to all types of amino acids, even the relatively folded last 33 residues. One possible explanation is that the size of F- is quite small and could access the buried amide protons of the last 33 residues of the cytoplasmic domain. In this regard, the study of the effects of salts on well-structured protein using high resolution NMR is quite necessary. Thus in the following studies, we choose a small well-folded protein, ww4, as a model to study salts’ effects. Our results reveal an extremely complex scenario that the binding of salts to proteins are not only salt specific, but also residue specific at physiologically-relevant salt concentrations, which might have significant implications in decrypting human diseases related to protein aggregation. Many reports have revealed the important role of ions in mediating diseases caused by protein aggregation (Raman et al. 2005; Calamai et al. 2006; Jain et al. 2010; Bellotti and Udgaonkar, 2008), but it remains elusive to which degree ions specifically induce the protein aggregation. As all cellular processes occur in buffers, we believe that ions have important roles in modulating protein functions, and many specific ion effects on proteins at low, physiologically relevant concentrations remains to be discovered yet. 53 Chapter 4 Studies of Interactions between Anions and Well-Folded Protein WW4 54 Result 4.1 Structural characterization of WW4 The solution NMR structure of WW4 in phosphate buffer and its binding with Nogo-A peptide has already been determined (Qin et al. 2008). So in my project, I mainly study the structural properties of WW4 in unsalted water and the effects of different salts to its structure. The lyophilized WW4 powder after purification with HPLC was solubilized in unsalted water and the pH of the protein sample was 4.0 at a concentration of 25 μM. The far-UV CD spectrum of the sample at pH 4.0 indicates that it is a β-turn rich protein (Figure 4.1a). To study the structural properties at physiological relevant pH, we adjusted the pH of the protein sample with microliter amounts of quite diluted sodium hydroxide to 6.4. The CD spectrum of the sample at pH 6.4 is quite similar as the spectrum at pH 4.0 (Figure 4.1a), implying that WW4 has very similar secondary structures at these two different pHs. Subsequently we added Na2SO4 and NaCl into the protein samples to a final concentration of 20mM at a protein concentration of 25 μM and got similar far-UV CD spectra at both pHs (Figure 4.1b,c), indicating that the addition of these two salts could not induce significant secondary structure changes. We also added NaSCN into the sample to 20mM and acquired the far-UV CD spectrum. However, the spectrum is of low quality because of high noise level and the result is less conclusive. 55 Figure 4.1 CD characterization of WW4. (a) Far UV CD spectra of the WW4 domain (25 μM) in water at pH 6.4 (blue) and 4.0 (red). (b) Far UV CD spectra of the WW4 domain (25 μM) in water at pH 6.4 (blue), and in the presence of 20 mM Na2SO4 (red) and 20 mM NaCl (green). (c) Far UV CD spectra of the WW4 domain (25μM) in water at pH 4.0 (blue), and in the presence of 20 mM Na2SO4 (red) and 20 mM NaCl (green). The HSQC spectrum of WW4 has large spectral dispersions in both 1H (~2.9 ppm) and 15N (~22 ppm), suggesting that it is well-folded. To study the effects of salts, we titrated the protein samples with Na2SO4, NaCl and NaSCN under three solution conditions (in water at pH 4.0 and pH 6.4, in 20 mM phosphate buffer at pH 6.4) (Figure 4.2). From the results we could see that even with addition of salts up to 200 mM, there are no significant changes in the HSQC spectral dispersions, but only HSQC peaks shifted. This result is consistent with the result of far-UV CD, suggesting that salts could not induce dramatic secondary and tertiary structure changes to the protein. 56 Na2SO4 NaCl NaSCN Figure 4.2 NMR HSQC titrations with Na2SO4, NaCl and NaSCN. Superimposition of two-dimensional 1H-15N NMR HSQC spectra at a WW4 concentration of 250 μM at 25 ºC in water at pH 6.4 (a), in the buffer at pH 6.4 (b), and in water at pH 4.0 (c); in the absence (black) and in the presence of different salts at 20 mM (green); 150 mM (red) and 200 mM (blue). Peaks with significant chemical shift changes are labeled: red for residues with amide proton H/D exchange rate (Kex) 0.03 ppm) are labeled. of amide protons of each residue by tracking the shifts of every HSQC peak with the increase of salt concentrations (Figure 4.3). 58 From the figures we could observe that different salts could induce different patterns of chemical shift difference. But for the same salt, the overall patterns are highly similar in unsalted water at pH 4.0 and 6.4. In water at pH 6.4, Na2SO4 induces dramatic shifts (>0.03 ppm) of three residues (Phe31, Lys32 and Asn36), NaSCN induces nine (Trp9, Glu10, Glu16, Gly17, Asp23, Arg27, Lys32, Arg35 and Asn36), while NaCl only induces two (Arg35 and Asn36). Besides, we also titrated the protein sample in water at pH 6.4 with sodium phosphate up to 200 mM and it seems to perturb only two residues (Glu16 and Asn36) significantly (Figure 4.4). Figure 4.4 NMR HSQC titrations with Na2HPO4. (a) Residue-specific chemical shift differences of amide protons (1H) of a WW4 domain upon addition of Na2HPO4 at 150 mM (blue bars) and 200 mM (red circles). (b) Residue-specific apparent dissociation constants (Kd) for Glu16 and Lys32. Experimental (dots) and fitted (lines) values are shown for the 1H chemical shift changes induced by gradual addition of Na2HPO4. 59 Strikingly, from all the results above we observed that with the pre-existence of 20 mM sodium phosphate, the HSQC peak shift patterns are different with those samples in water. Some of the residues that are not affected either by sodium phosphate or Na2SO4, NaCl and NaSCN separately, could suddenly be perturbed significantly when titrated with the three salts in buffer (20 mM sodium phosphate). For example, Trp9, Asp23 and Asn25 are not significantly perturbed by Na2SO4 in water either at pH 6.4 or 4.0; but they are largely perturbed by Na2SO4 with the pre-existence of phosphate buffer (Figure 4.2 and 4.3). One explanation for the change of shift pattern is the non-specific electrostatic screening imposed by the presence of 20 mM sodium phosphate. To evaluate whether this is the only reason, we titrated the protein WW4 with Na2SO4 in the pre-existence of 150 mM NaCl, which has larger ionic strength than 20 mM sodium phosphate and is similar as the physiological concentration in blood. Interestingly we found that the presence of 150 mM NaCl did not significantly change the shift pattern, but only attenuate the shift amplitude (Figure 4.5). Besides, the binding affinity of Na2SO4 to the residues Phe31 and Lys32 seem to reduce about three fold with the presence of 150 mM NaCl (Figure 4.5) as compared to binding affinity of single Na2SO4 titration (Table 2). To determine the contributions of cations to the perturbation, we further titrated the protein WW4 with MgSO4, KCl and KSCN in water at pH 6.4. With the addition of salts, the HSQC shift patterns are almost the same as those sodium salts with same anions (Figure 4.6-4.8), indicating that the observed salts effects are mostly induced by the binding of anions to the protein, not cations. 60 Figure 4.5 NMR HSQC titrations by Na2SO4 in the pre-existence of 150 mM NaCl. (a) Residue-specific chemical shift differences of amide protons (1H) upon addition of Na2SO4 at 200 mM to a WW4 sample in water at pH 6.4 (red bars) and in water with the pre-existence of 150 mM NaCl at pH 6.4 (blue bars). (b) Residue-specific apparent dissociation constants (Kd) for Phe31 and Lys32. Experimental (dots) and fitted (lines) values are shown for the 1H chemical shift changes induced by gradual addition of Na2SO4. 61 a 114 120 126 108 b 114 120 126 108 c 114 120 126 9.4 Figure 4.6 NMR HSQC titrations with MgSO4.Superimposition of two-dimensional 1H-15N NMR HSQC spectra at a WW4 concentration of 250 μM at 25 ºC in water at pH 6.4 in the presence of Na2SO4 (black) and MgSO4 (red) at 20 mM (a); 100 mM (b) and 200 mM (c). 108 9.0 8.6 8.2 7.8 7.4 7.0 6.6 62 a 114 120 126 108 b 114 120 126 108 c 114 120 126 9.4 Figure 4.7 NMR HSQC titrations with KCl. Superimposition of two-dimensional 1H-15N NMR HSQC spectra at a WW4 concentration of 250 μM at 25 ºC in water at pH 6.4 in the presence of NaCl (black) and KCl (red) at 20 mM (a); 100 mM (b) and 200 mM (c). 108 9.0 8.6 8.2 7.8 7.4 7.0 6.6 63 a 114 120 126 108 b 114 120 126 108 c 114 120 126 9.4 Figure 4.8 NMR HSQC titrations with KSCN. Superimposition of two-dimensional 1H-15N NMR HSQC spectra at a WW4 concentration of 250 μM at 25 ºC in water at pH 6.4 in the presence of NaSCN (black) and KSCN (red) at 20 mM (a); 100 mM (b) and 200 mM (c) 108 9.0 8.6 8.2 7.8 7.4 7.0 6.6 64 4.3 Quantitative assessment of binding affinity To observe the binding affinity between the protein and ions more clearly, we fitted the apparent dissociation constants (Kd) of all residues with 1H chemical shift difference > 0.03 ppm by monitoring the shifts of HSQC peaks upon titrating and subsequently fitting the shift tracings as shown in figure 4.9. We fitted both backbone and sidechain amide protons and summarized the Kd values in table 2. Figure 4.9 Apparent dissociation constants for representative amide protons. (a-b) Residue-specific apparent dissociation constants (Kd) for backbone amide proton of Phe31 and sidechain amide proton of Asn36 titrated by Na2SO4 respectively under different conditions. (c-d) Residue-specific apparent dissociation constants (Kd) for backbone amide proton of Glu16 and sidechain amide proton of Asn25 titrated by NaSCN respectively under different conditions. Experimental (dots) and fitted (lines) values are shown for the 1H chemical shift changes induced by gradual addition of two salts (Na2SO4 and NaSCN). Red is for the data in water (pH 6.4), green for those in water (pH 4.0), and blue for those in 20 mM sodium phosphate buffer (pH 6.4). 65 66 a “Buffered” refers to “in 20 mM sodium phosphate (pH 6.4)”. bIn calculating the average values, Kd values > 250 mM are not included. Table 2. Apparent Dissociation Constants (Kd) for Binding of Salts to WW4 Domain Interestingly we observed that although the residues perturbed by Na2SO4 are much less than NaSCN, the binding affinity of Na2SO4 is much stronger, with average Kd values of 15.7, 32.0 and 86.3 mM for the backbone amide protons in water at pH 4.0, 6.4 and in buffer at pH 6.4. The binding affinities of NaSCN, NaCl and Na2HPO4 are much lower and quite similar, with average values of about 100 mM. For the sidechain amide protons, they seem to prefer to interact with Na2SO4, especially at pH 4.0. Besides, the Kd values for the sidechain amide protons are about 3-4 fold larger than those of backbones under the same condition, suggesting that the interactions of anions with backbone and sidechain amide protons are independent. Surprisingly, we found that with the pre-existence of 20 mM sodium phosphate, the binding affinities are significantly reduced compared to those of single salt titration for Na2SO4, NaSCN and NaCl as shown in Figure 4.9 and Table 2. The average apparent dissociation constant of Na2SO4 to backbone amide protons increased about 3 times with the presence of buffer, but we could still observe saturation from the titration curve (Figure 4.9). However, for NaCl and NaSCN, the titration curves seem to become liner with the presence of buffer and the apparent dissociation constant cannot be fitted. 4.4 Properties of binding surfaces for different anions According to previous results in section 4.3, different salts could perturb different residues. To study the distribution properties of these residues on the structure of the protein, we mapped the binding sites onto the NMR structure of WW4 (Qin et al. 67 2008) (Figure 4.10). Interestingly, most of the residues seem to locate on the relatively exposed area. Figure 4.10 Binding sites on WW4. (a-c) WW4 NMR structures with significantly perturbed residues colored in green, by 200 mM Na2SO4 in water (pH 6.4), the buffer (pH 6.4) and water (pH 4.0) respectively. (d-f) by 200 mM NaSCN in water (pH 6.4), the buffer (pH 6.4) and water (pH 4.0) respectively. (g-h) by 200 mM NaCl in water (pH 6.4) and the buffer (pH 6.4) respectively. (i) by 200 mM Na2HPO4 in water (pH 6.4). Red is used to label residues with significant changes only in the presence of the buffer (pH 6.4). Sticks are used to indicate the side chain amide protons with significant 1H chemical shift changes (>0.03 ppm). 68 To see the accessibility of these residues, we obtained the exposure degree of amide protons to solvent though H/D exchange experiments (Englander et al. 1997) and fitted the H/D exchange rates (Kex) (Figure 4.11). The exchange rates of the backbone amide protons are summarized in Table 3. Figure 4.11 H/D exchange experiments of the WW4 domain (a) HSQC spectrum of WW4 acquired in 2 min after the lyophilized sample was dissolved in D2O. The red letter is used to label residues with their peaks significantly perturbed while the black for residues not significantly perturbed by salts. (b-c) Experimental (dots) and fitted (lines) values are shown for the HSQC peak intensities in H/D exchange experiments for WW4 residues whose peaks are significantly (b) and not significantly (c) perturbed by salts. (d) WW4 NMR structure colored with H/D exchange rates (Kex): blue for residues with HSQC peaks disappeared in 2 min after the lyophilized sample was dissolved in D2O; green for residues with Kex > 5 h-1 and red for residues with Kex 5 h-1 (Figure 4.11 d). However, for NaSCN, some of the binding residues, such as Trp9, Glu10 and Asp23, are well protected (with Kex < 5 h-1), locating on two central β-strands (Figure 4.11 d). The residue Glu10 is even one of the two most protected residues (another is Ile11), with a Kex of 0.97 h-1 (Table3). To further study the relationship between specific binding and electrostatic potential, we calculated the electrostatic potential surfaces of WW4 using APBS (Figure 4.12). Surprisingly, we found that the amide protons which could interact with Na2SO4, NaCl and Na2HPO4 are mostly located in the areas that are relatively positively charged, while NaSCN is even able to bind to amide protons of Trp9, Glu10, Val22 and Asp23 that are located on negatively charged regions (Figure 4.12). This suggests that the binding of NaSCN to amide protons is relatively independent on the electrostatic potential while for Na2SO4, NaCl and Na2HPO4, the binding is highly 70 electrostatically dependent. This might be consistent with the phenomenon that the binding affinities of NaSCN are almost the same at two different pHs, pH 4.0 and 6.4, while for Na2SO4, the binding affinities at pH 4.0 have about two-fold increase compared to those at pH 6.4 (Table 2). Figure 4.12 The electrostatic potential of WW4 at pH 6.4 and 4.0 respectively, visualized at the level of the accessible surface of the protein, with blue and red corresponding to positive and negative potential values. In this project, we studied the interactions between anions and well-structured protein WW4 and found that sulfate, chloride and thiocyanate could bind to the well-folded protein at specific residues and affinities and sulfate turned out to be the strongest binder. Previous research and reviews have proposed that chloride and sulfate ions have high charge-density and are highly hydrated and thus could only bind well exposed amide protons driven mostly by electrostatic interactions (Mason and et al., 2011; Parsons and et al., 2011). Maybe this could explain why sulfate turned out to be the strongest binder to both unstructured protein cytoplasmic domain of ephrinB2 and well-folded protein WW4 in our research. 71 Besides, we also interestingly found that with the existence of 20 mM sodium phosphate, the binding patterns of these three salts are totally changed and the binding affinities are largely reduced, while this phenomenon is not observed during titrations with the pre-existence of 150 mM sodium chloride. The reduction of binding affinities may be because of electrostatic screening as the pre-existence of 150 mM NaCl could also reduce the binding affinities of sulfate to WW4. Similar results were also got by Jolivalt (1998) when they study the interaction between bovine pancreatic trypsin inhibitor (BPTI) and thiocyanate using NMR. They found that in the presence of 40 molar equivalents of KCl, the chemical shift variations of two protons induced by the addition of thiocyanate were reduced (Jolivalt and et al., 1998), indicating that screening on the electrostatic field have some effects on the chemical shift perturbation. However, the electrostatic screening could not be the only reason because the presence of 20 mM phosphate buffer changed the chemical shift perturbation patterns while the presence of 150 mM NaCl could not. One doubt for this result is that the pH might be slightly changed during titration without the existence of buffer while in the presence of 20 mM phosphate buffer, the pH is stable. So we monitored the pH of the samples during titration and found that the pH kept to the same because the pH of salt solutions is the same as the protein samples and could not perturb the pH during titration. Thus we suspect that maybe there are non-additive interactions between phosphate and sulfate anions. Previously it was commonly believed that the effects of anions are independent and additive. To further demonstrate this hypothesis, the titration experiments still need to be repeated. 72 As described in the introduction, Nogo-A peptide could bind to the well-folded WW4 domain over the Thr28-Thr29-Thr30 sequence on the third β-strand (Qin, 2008). In our studies, we interestingly found that the binding regions of sulfate and thiocyanate on WW4 and the binding region of Nogo-A have significant overlap. As proteins are exposed to various inorganic salts under physiological conditions and most cellular processed happen in buffers, we hypothesize that the specific ion-binding may be related to many protein functions. A recent review proposed that the protein-ion interactions seem to be more important than previously expected in mediating various aspects of proteins under the crowded conditions where electrostatic properties of the proteins become dominant (Laue and Demeler, 2011). Thus a deep understanding of the anion-protein interactions is quite important for the study of protein functions in cells (Laue and Demeler, 2011). 73 REFERENCES Adams R.H., Klein R. (2000) Eph receptors and ephrin ligands: essential mediators of vascular development. Trens Cardiovasc Med.10, 183-188. Baker NA, Sept D, Joseph S, Holst MJ, McCammon JA. (2001) Electrostatics of nanosystems: application to microtubules and the ribosome. Proc. Natl. Acad. Sci. USA. 98, 10037-10041. Bellotti V and Chiti F. (2008) Amyloidogenesis in its biological environment: challenging a fundamental issue in protein misfolding diseases. Curr. Opin. Struct. Biol. 18, 771-779. Bénas P, Legrand L, Riès-Kautt M. (2002) Strong and specific effects of cations on lysozyme chloride solubility. Acta Cryst., D58, 1582-1587. Bian H, Wen X, Li J, Chen H, Han S, Sun X, Song J, Zhuang W, Zheng J. (2011) Ion clustering in aqueous solutions probed with vibrational energy transfer. Proc. Natl. Acad. Sci. USA. 108, 4737-4742. Boström M, Williams DRM, Ninham BW. (2001) Specific ion effects: why DLVO theory fails for biology and colloid systems. Phys. Rev. Lett. 87, 168103. Boström M, Williams DRM, Ninham BW. (2002) The influence of ionic dispersion potentials on counterion condensation on polyelectrolytes. J. Phys. Chem. B, 106, 7908-7912. Boström M, Williams DRM, Ninham BW. (2003) Special ion effects: why the properties of lysozyme in salt solutions follow a Hofmeister series. Biophys. J. 85, 686-694. Brückner K., Pablo Labrador J., Scheiffele P., Herb A., Seeburq P.H., Klein R. (1999) EphrinB ligands recruit GRIP family PDZ adaptor proteins into raft membrane microdomains. Neuron. 22, 511-524. Brünger AT, Adams PD, Clore M, Deland WL, Gros P, Grosse-Kunstleve RW, Jiang J, Kuszewski J, Nilges M, Pannu NS, Read RJ, Rice LM, Simonson T, Warren GL. (1998) Crystallography & NMR system: a new software suite for macromolecular structure determination. Acta Cryst. D54, 905-921. Calamai M, Kumita JR, Mifsud J, Pattini C, Ramazzotti M, Ramponi G., Taddei N, Chiti F, Dobson CM. (2006) Nature and significance of the interactions between amyloid fibrils and biological polyelectrolytes. Biochemistry, 45, 12806-12815. 74 Cao X, Xue L, Han L, Ma L, Chen T, Tong T. (2011) WW Domain-containing E3 ubiquitin protein ligase 1 (WWP1) Delays cellular senescence by promoting p27Kip1 degradation in human diploid fibroblasts. J. Biol. Chem. 286, 33447– 33456. Carbonnaux C, Ries-Kautt M, Ducruix A. (1995) Relative effectiveness of various anions on the solubility of acidic Hypoderma lineatum collagenase at pH 7.2. Protein Sci., 4, 2123-2128. Chakrabarti P. (1993) Anion binding sites in protein structures. J. Mol. Biol. 234, 463-482. Chen C, Sun X, Guo P, Dong XY, Sethi P, Cheng X, Zhou J, Ling J, Simons JW, Lingrel JB, Dong JT. (2005) Human Kruppel-like factor 5 is a target of the E3 ubiquitin ligase WWP1 for proteolysis in epithelial cells. J. Biol. Chem. 280, 41553–41561 Chen C, Matesic LE. (2007) The Nedd4-like family of E3 ubiquitin ligases and cancer. Cancer Metastasis Rev. 26, 587-604. Cho Y, Zhang Y, Christensen T, Sagle LB, Chilkoti A, Cremer PS. (2008) Effects of Hofmeister anions on the phase transition temperature of elastin-like polypeptides. J. Phys. Chem. B, 112, 13765-13771. Clevers H, and Batlle E. (2006) EphB/EphrinB receptors and Wnt signaling in colorectal cancer. Cancer Res. 66, 2–5. Collins KD. (1995) Sticky ions in biological systems. Proc. Natl. Acad. Sci. USA, 92, 5553-5557. Collins KD. (1997) Charge density-dependent strength of hydration and biological structure. Biophys J. 72, 65-76. Collins KD. (2004) Ions from the Hofmeister series and osmolytes: effects on proteins in solution and in the crystallization process. Methods, 34, 300-311. Collins KD. (2006) Ion hydration: implications for cellular function, polyelectrolytes, and protein crystallization. Biophys Chem. 119, 271-281. Collins KD, Washabaugh MW. (1985) The Hofmeister effect and the behavior of water at interfaces. Q. Rev. Biophys. 18, 323–422. Cowan C.A., Henkemeyer M. (2001) The SH2/SH3 adaptor Grb4 transduces B-ephrin reverse signals. Nature, 413, 174-179. 75 Delaglio F, Grzesiek S, Vuister GW, Zhu G, Pfeifer J, Bax A. (1995) NMRpipe: a multidimentional spectral processing system based on UNIX pipes. J. Biomol. NMR, 6, 277-293. Dolinsky TJ, Nielsen JE, McCammon JA, Baker NA. (2004) PDB2PQR: an automated pipeline for the setup, execution, and analysis of Poisson-Boltzmann electrostatics calculations. Nucleic Acids Research. 32, 665-667. Englander SW, Mayne L, Bai Y, Sosnick TR. (1997) Hydrogen exchange: the modern legacy of Linderstrøm-Lang. Protein Sci. 6, 1101-9. Fedoroff OY, Townson SA, Golovanov AP, Baron M, Avis JM. (2004) The structure and dynamics of tandem WW domains in a negative regulator of notch signaling, Suppressor of deltex. J. Biol. Chem. 279, 34991-35000. Flanagan JG, Vanderhaeghen P. (1998) The ephrins and Eph receptors in neural development. Annu. Rev. Neurosci. 21, 309-345. Foord RL, Leatherbarrow RJ. (1998) Effect of osmolytes on the exchange rates of backbone amide protons in proteins. Biochemistry, 37, 2969-2978. Galinier R, Gout E, Lortat-Jacob H, Wood J, Chroboczek J. (2002) Adenovirus protein involved in virus internalization recruits ubiquitin-protein ligases. Biochemistry. 41, 14299–14305 Gilliland GL. (1988) A biological macromolecule crystallization database: A basis for a crystallization strategy. J. Crystal Growth, 90, 51-59. Gokarn YR, Fesinmeyer RM, Saluja A, Cao S, Dankberg J, Goetze A, Remmele RL. Jr., Narhi LO, Brems DN. (2009) Ion-specific modulation of protein interactions: anion-induced, reversible oligomerization of a fusion protein. Protein Sci., 18(1), 169-179. Gokarn YR, Fesinmeyer RM, Saluja A, Razinkove V, Chase SF, Laue TM, Brems DN. (2011) Effective charge measurements reveal selective and preferential accumulation of anions, but not cations, at the protein surface in dilute salt solutions. Protein Sci. 20, 580-587. Güntert P. (2004) Automated NMR structure calculation with CYANA. Methods Mol. Biol. 278, 353-378. Hall DL, Darke PL. (1995) Activation of the herpes simplex virus type 1 protease. J. Biol. Chem., 270(39), 22697-22700. 76 Harvey KF, Kumar S. (1999) Nedd4-like proteins: an emerging family of ubiquitinprotein ligases implicated in diverse cellular functions. Trends Cell Biol. 9, 166– 169. Hochachka PW, Somero GN. (1984) Biochemical Adaption. Princeton, NJ: Princeton University Press. Hofmeister F. (1888) Zur Lehre von der Wirkung der Salze. II. Arch. exp. Pathol. Pharmakol. 24, 247-260. Holder N, Klein R. (1999) Eph receptors and ephrins: effectors of morphogenesis. Development, 126, 2033-2044. Huang H, Melacini G. (2006) High-resolution protein hydration NMR experiments: probing how protein surfaces interact with water and other non-covalent ligands. Anal. Chim. Acta. 564, 1-9. Jain S, Udqaonkar JB. (2010) Salt-induced modulation of the pathway of amyloid fibril formation by the mouse prion protein. Biochemistry, 49, 7615-7624. Jaravine VA, Rathgeb-Szabo K, Alexandrescu AT. (2000) Microscopic stability of cold shock protein A examined by NMR native state hydrogen exchange as a function of urea and trimethylamine N-oxide. Protein Sci. 9, 290-301. Johnson BA, Blevins RA. (1994) NMRView: a computer program for the visualization and analysis of NMR data. J. Biomol. NMR, 4, 603-614. Jolivalt C, Böckmann A, Riès-Kautt M, Ducruix A, Guittet E. (1998) Characterization of the interaction between bovine pancreatic trypsin inhibitor and thiocyanate by NMR. Biophys. Chem. 71, 221-234. Kunz W. (2009) Specific ion effects. World Scientific Publishing. Kunz W. (2010) Specific ion effects in colloidal and biological systems. Curr. Opion. Coll. Inter. Sci. 15, 34-39. Kunz W, Henle J, Ninham BW. (2004) ‘Zur Lehre von der Wirkung der Salze’ (about the science of the effect of salts): Franz Hofmeister's historical papers. Curr. Opin. Coll. Inter. Sci. 9, 19-37. Laue T, Demeler B. (2011) A postreductionist framework for protein biochemistry. Nat. Chem. Biol. 7, 331-334. 77 Lee B, Richards FM. (1971) The interpretation of protein structures: estimation of static accessibility. J. Mol. Biol. 55, 379-400. Li M, Liu J, Ran X, Fang M, Shi J, Qin H, Goh JM, Song J. (2006) Resurrecting abandoned proteins with pure water: CD and NMR studies of protein fragments solubilized in salt-free water. Biophys J. 91, 4201-4209. Lu Q, Sun E.E., Klein R.S., Flanagan J.G. (2001) Ephrin-B reverse signaling is mediated by a novel PDZ-RGS protein and selectively inhibits G protein-coupled chemoattraction. Cell. 105, 69-79. Macomber RS. (1992) An introduction to NMR titration for studying rapid reversible complexation. J. Chem. Educ. 69, 375-378. Mason PE, Neilson GW, Dempsey CE, Barnes AC, Cruickshank JM. (2003) The hydration structure of guanidinium and thiocyanate ions: implications for protein stability in aqueous solution. Proc Natl Acad Sci U.S.A. 100, 4557-61. Moren A, Imamura T, Miyazono K, Heldin CH, Moustakas A. (2005) Degradation of the tumor suppressor Smad4 by WW and HECT domain ubiquitin ligases. J. Biol. Chem. 280, 22115–22123. Munishkina LA, Henriques J, Uversky VN, Fink A. (2004) Role of protein-water interactions and electrostatics in alpha-synuclein fibril formation. Biochemistry, 43, 3289-3300. Nguyen Huu NS, Ryder WD, Zeps N, Flasza M, Chiu M, Hanby AM, Poulsom R, Clarke RB, Baron M. (2008) Tumour-promoting activity of altered WWP1 expression in breast cancer and its utility as a prognostic indicator. J. Pathol. 216, 93–102. Ninham BW, Yaminsky V. (1997) Ion binding and ion specificity: the Hofmeister effect and Onsager and Lishitz theoties. Langmuir, 13, 2097-2108. O’Brien R, DeDecker B, Fleming KG, Sigler PB, Ladbury JE. (1998) The effects of salt on the TATA binding protein-DNA interaction from a hyperthermophilic archaeon. J. Mol. Biol. 279, 117-125. O’Leary DD, Wilkinson DG. (1999) Eph receptors and ephrins in neural development. Curr. Opin. Neurobiol. 9, 65-73. Omta AW, Kropman MF, Woutersen S, Bakker H. (2003) Negligible effect of ions on the hydrogen-bond structure in liquid water. Science, 301, 347-349. 78 Pace CN, Vajdos F, Fee L, Grimsley G, Gray T. (1995) How to measure and predict the molar absorption coefficient of a protein. Protein Sci. 4, 2411-2423. Parsons DF, Boström M, Lo Nostro P, Ninham BW. (2011) Hofmeister effects: interplay of hydration, nonelectrostatic potentials, and ion size. Phys Chem Chem Phys. 13, 12352-67. Pasquale EB. (2005) Eph receptor signalling casts a wide net on cell behaviour. Nat. Rev. Mol. Cell. Biol. 6, 462-475. Pasquale EB. (2010) Eph receptors and ephrins in cancer: bidirectional signaling and beyond. Nat. Rev. Cancer. 10, 165-180. Pedersen JS, Flink JM, Dikov D, Otzen DE. (2006) Sulfates dramatically stabilize a salt-dependent type of glucagon fibrils. Biophys. J. 90, 4181-4194. Plant PJ, Yeger H, Staub O, Howard P, Rotin D. (1997) The C2 domain of the ubiquitin protein ligase Nedd4 mediates Ca2+-dependent plasma membrane localization. J. Biol. Chem. 272, 32329–32336. Qin H, Pu HX, Li M, Ahmed S, Song J. (2008) Identification and structural mechanism for a novel interaction between a ubiquitin ligase WWP1 and Nogo-A, a key inhibitor for central nervous system regeneration. Biochemistry. 47, 13647-13658. Raman B, Chatani E, Kihara M, Ban T, Sakai M, Haseqawa K, Naiki H, Rao ChM, Goto Y. (2005) Critical balance of electrostatic and hydrophobic interactions is required for beta 2-microglobulin amyloid fibril growth and stability. Biochemistry, 44, 1288-1299. Ramos CH, Baldwin RL. (2002) Sulfate anion stabilization of native ribonuclease A both by anion binding and by the Hofmeister effect. Protein Sci. 11, 1771-1778. Ries-Kautt M, Ducruix A. (1989) Relative effectiveness of various ions on the solubility and crystal growth of lysozyme. J. Biol. Chem. 264, 745-748. Rode BM, Fitz D, Jakschitz T. (2007) The first steps of chemical evolution towards the origin of life. Chem Biodivers. 4, 2674-702. Salis A, Bilanicova D, Ninham BW, Minduzzi M. (2007) Hofmeister effects in enzymatic activity: weak and strong electrolyte influences on the activity of Candida rugosa lipase. J. Phys. Chem. B. 111, 1149-1156. 79 Saluja A, Crampton S, Kras E, Fesinmeyer RM, Remmele RL. Jr., Narhi LO, Brems DN, Gokarn YR. (2009) Anion binding mediated precipitation of a peptibody. Pharm. Res. 26(1), 152-160. Schmucker D, Zipursky SL. (2001) Signaling downstream of Eph receptors and ephrin ligands. Cell, 105, 701-704. Scicolone G., Ortalli A.L., Carri N.G. (2009) Key roles of Ephs and ephrins in retinotectal topographic map formation. Brain Res. Bull. 79, 227-247. Seo SR, Lallemand F, Ferrand N, Pessah M, L’Hoste S, Camonis J, Atfi A (2004) The novel E3 ubiquitin ligase Tiul1 associates with TGIF to target Smad2 for degradation. EMBO J. 23, 3780-3792. Shuker SB, Hajduk PJ, Meadows RP, Fesik SW. (1996) Discovering high-affinity ligands for proteins: SAR by NMR. Science. 274, 1531-1534. Song J. (2009) Insight into "insoluble proteins" with pure water. FEBS Lett. 583, 953-959. Song J, Jamin N, Gilquin B, Vita C, Ménez A. (1999) A gradual disruption of tight side-chain packing: 2D 1H-NMR characterization of acid-induced unfolding of CHABII. Nat Struct Biol. 6, 129-34. Song J, Vranken W, Xu P, Gingras R, Noyce RS, Yu Z, Shen SH, Ni F. (2002) Solution structure and backbone dynamics of the functional cytoplasmic subdomain of human ephrin B2, a cell-surface ligand with bidirectional signaling properties. Biochemistry. 41, 10942-10949. Sudol M. (1996) Structure and function of the WW domain. Prog. Biophys. Mol. Biol. 65, 113-132. Sudol M, Chen HI, Bougeret C, Einbond A, Bork P. (1995) Characterization of a novel protein-binding module–the WW domain. FEBS Lett. 369, 67–71. Tadeo X, Pons M, Millet O. (2007) Influence of the Hofmeister anions on protein stability as studied by thermal denaturation and chemical shift perturbation. Biochemistry. 46(3), 917-923. Wang J, Peng Q, Lin Q, Childress C, Carey D, Yang W. (2010) Calcium activates Nedd4 E3 ubiquitin ligases by releasing the C2 domain-mediated auto-inhibition. J. Biol. Chem. 285, 12279–12288. 80 Warren JC, Cheatum SG. (1966) Effect of neutral salts on enzyme activity and structure. Biochemistry. 5, 1702-1707. Williamson MP, Havel TF, Wüthrich K. (1985) Solution conformation of proteinase inhibitor IIA from bull seminal plasma by 1H nuclear magnetic resonance and distance geometry. J. Mol. Biol. 182, 295–315. Wishart DS, Bigam CG, Yao J, Abildgaard F, Dyson HJ, Oldfield E, Markley JL, and Sykes BD. (1995) “1H, 13C and 15N chemical shift referencing in biomolecular NMR” J. Biomol. NMR, 6, 135-140. Wishart DS, Sykes BD, Richards FM. (1991) Relationship between nuclear magnetic resonance chemical shift and protein secondary structure. J. Mol. Biol. 222, 311-333. Wishart DS, Sykes BD. (1994) Chemical shifts as a tool for structure determination. Methods Enzymol. 239, 363-392. Wüthrich K. (1986). NMR of Proteins and Nucleic Acids. New York (NY), USA: Wiley-Interscience. Xu Q, Mellitzer G, Robinson V, Winkinson DG. (1999) In vivo cell sorting in complementary segmental domains mediated by Eph receptors and ephrins. Nature. 399, 267-271. Zhang X, Srinivasan SV, Lingrel JB (2004) WWP1-dependent ubiquitination and degradation of the lung Kruppel-like factor, KLF2. Biochem. Biophys. Res. Commun. 316, 139–148. Zhang Y, Cremer PS. (2006) Interactions between macromolecules and ions the Hofmeister series. Curr. Opin. Chem. Biol. 10, 658-663. Zhang Y, Cremer PS. (2009) The inverse and direct Hofmeister series for lysozyme. Proc. Natl. Acad. Sci. USA. 106, 15249-15253. Zhang Y, Cremer PS. (2010) Chemistry of Hofmeister anions and osmolytes. Annu. Rev. Phys. Chem. 61, 63-83. Zhao C., Irie N., Takada Y., Shimoda K., Miyamoto T., Nishiwaki T., Suda T., and Matsuo K. (2006) Bidirectional ephrinB2-EphB4 signaling controls bone homeostasis. Cell Metab. 4, 111–121. Zhi X, Chen C. (2012) WWP1: a versatile ubiquitin E3 ligase in signaling and diseases. Cell. Mol. Life Sci. 69, 1425-1434. 81 Appendix Table: Dissociation Constants (Kd) of different salts to ephrinB2 cytoplasmic domain Na2SO4 H N L11 G15 K17 H22 H25 S26 H29 T37 S64 M87 V99 Kd 0.39 1.07 1.12 1.56 1.04 0.76 1.04 1.31 1.3 1.59 49.05 error 0.08 0.15 0.44 1.31 0.16 0.18 0.33 0.15 0.25 0.69 10.34 R19 H22 R23 K24 H25 S26 Q28 H29 L38 Kd 1.3 0.64 0.28 0.33 0.89 0.46 1.43 1.04 2.31 error 0.19 0.14 0.05 0.04 0.13 0.04 0.16 0.15 0.59 NaH2PO4 L11 V12 G15 R20 H22 K24 H29 D63 S64 I83 V99 I55 D76 I83 V99 2.93 2.51 4.81 5.45 0.81 0.85 13.75 59.59 8.03 62.25 25.5 17.38 13.66 21.11 24.36 0.41 0.79 0.48 0.9 0.17 0.43 2.34 13.14 1.67 21.04 3.04 5.44 4.21 7.94 2.09 S8 V12 R19 R21 R23 K24 H25 S26 Q28 H29 T30 1.27 1.95 9.65 7.1 6.77 4.07 7.62 4.05 7.88 5.95 5.38 0.15 0.33 1.8 1.73 1.66 0.79 1.53 0.61 1.36 1.39 1.09 NaF V12 G15 S26 H29 T30 E51 28.05 22.27 67.15 73.39 26.05 9.38 3.96 3.53 22.26 16.8 6.4 0.94 S8 V12 S26 Q28 T30 T32 14.4 18.26 14.68 22.44 12.46 10.95 2.52 1.91 2.1 5.49 2.71 2.51 82 D63 E71 D76 V99 D76 V99 5.36 14.62 3.13 3.74 6.18 3.49 0.45 1.51 0.46 0.18 0.71 0.25 D54 I55 E71 K72 7.47 5.42 20.39 13.36 0.54 0.3 5.14 2.08 NaCl V12 G15 Y18 R21 H22 R23 K24 H25 S26 H29 T40 D63 21.9 15.24 6.53 9.22 22.71 16.38 19.84 13.79 16.78 26.14 7.08 25.63 2.32 6.61 0.49 1.48 1.3 1.6 2.72 2.32 2.06 3.89 0.64 6.64 V12 Y18 R19 R20 R21 H22 H25 S26 H29 T30 T32 25.85 8.46 16.91 12.41 11.6 21.44 12.01 11.35 22.72 11.35 15.03 3.56 1.2 2.58 2.83 1.86 3.16 2.23 1.31 6.26 1.99 3.28 NaBr V12 Y18 R20 R21 H22 R23 K24 H25 S26 H29 S50 M87 D63 25.81 8.88 6.01 9.82 11.56 8.4 8.45 11.51 22.27 14.89 3.97 14.87 18.34 2.43 0.96 0.76 1.92 1.33 0.89 0.82 1.59 4.5 2.88 1.73 3.42 7.01 V12 Y18 R19 R20 R21 H22 R23 H25 S26 H29 T30 T32 30.51 17.45 13.78 20.59 18.8 33.07 26.21 26.18 13.93 32.95 14.92 28.12 2.62 2.24 2.75 4.58 3.39 6.27 6.31 6.19 4.02 7.17 4.88 8.32 NaNO3 V12 Y18 R21 H22 R23 K24 15.73 5.22 1.42 4.44 4.4 7.03 1.99 0.6 0.19 0.26 0.33 1 V12 R19 H22 R23 K24 H25 20.09 12.26 6.51 10.43 11.83 8.74 2.6 1.14 1.98 0.74 1.33 1.21 83 H25 S26 H29 L59 E71 M87 8.49 14.44 15.81 22.07 14.56 5.25 1.56 2.59 3.04 4.63 3.89 0.41 S26 H29 T30 L59 E71 9.23 11.84 17.35 37.81 37.3 1.62 2.01 2.93 7.86 5.48 NaI K17 Y18 H22 R23 K24 H25 S26 H29 E51 L59 Y70 Y82 9.05 7.26 10.74 5 9.3 11.22 17.37 20.43 20.32 14.62 239.71 24.08 1.63 1.06 1.94 0.24 1.78 2.21 2.73 4.67 2.84 2.48 51.92 4.4 Y18 R19 H22 R23 K24 H25 S26 T30 E51 L59 Y82 9.69 14.68 18.54 6.37 35.63 14.09 15.36 21.01 22.26 95.4 25.37 1.46 2.49 2.93 0.42 5.02 3.78 3.35 6.08 5.56 26.85 5.02 NaSCN V12 R14 Y18 H22 R23 K24 H25 S26 E51 L59 M87 21.41 17.69 6.56 7.83 3.31 9.79 13.78 19.65 19.39 15.25 7.88 3.34 2.44 0.64 0.42 0.41 1.29 1.94 2.88 1.76 1.32 0.95 V12 Y18 R19 H22 R23 K24 H25 S26 T30 E51 L59 25.71 8.1 11.91 26.25 6.18 19.92 13.06 12.26 23.34 21.61 23.28 2.11 1.37 1.37 4.46 1.24 3.16 2.18 2.41 3.49 2.79 3.6 84 PUBLICATION Miao L, Qin H, Koehl P, Song J. (2011) Selective and specific ion binding on proteins at physiologically-relevant concentrations. FEBS Lett. 585, 3126-3132. 85 [...]... the effects of cations in promoting the precipitation of proteins are not as pronounced as anions However, recent years more and more studies showed that the effect of cations on crystallization of proteins is comparable to anions, especially at low concentrations (Carbonnaux et al 1995) A reasonable interpretation given (Kunz, 2009) was that both cations and anions can interact with proteins with specificities... while at high salt concentrations (0.8-1.5 M) it follows a direct Hofmeister series: SCN- < I- < ClO4- < Br- < NO3- < Cl- (Zhang and Cremer, 2009) In most previous studies, people focus their attention on moderate to high (> 0.1 M) salt concentrations as it is commonly believed that non-specific, electrostatic interactions are dominant at low salt concentrations (< 0.1 M), while at higher salt concentrations, ... 1.1.4 Specific protein-ion binding The study of interactions between ions and proteins has last for decades and in most studies, the interaction turned to be specific However, these specific interactions seem to be highly dependent on salt concentrations A typical example is the study of cloud-point temperature by Zhang and Cremer (2009) At low salt concentrations (< 9 0.5 M), the cloud-point of lysozyme... that ions could interact directly with glucagon fibrils like structural ligands and thus assist the formation of fibrils 4 O’Brien et al (1998) studied salts’ effects on the interactions of the TATA binding protein (TBP) with DNA They found that only at high salt concentrations the protein-DNA interaction could be detected and ions appeared to influence the binding A compelling explanation was that... protein interactions Ions have strong effects on the solution interactions of proteins, not only protein-protein interactions, but also DNA-protein interactions Gokarn et al (2009) reported that the reversible oligomerization of a fusion protein peptibody A (PbA) is modulated by specific anion-protein interactions Pedersen et al (2006) also investigated the effects of different salts on fibrillation... hydrated cations, by contrast, could interact more efficiently with the negatively charged and strongly hydrated side-chain carboxylates and backbone carbonyls Collins also proposed that besides the charged side chains, the weakly hydrated anions could also interact with other polar and non-polar groups, which are considered weakly hydrated (Collins, 2004) Paul Cremer and his colleagues (Zhang and. .. found that even at low salt concentrations (< 0.1 M), anions could selectively and preferentially bind to protein surface Chakrabarti (1993) also systematically analyzed the binding of 52 phosphate and sulfate ions in 34 different protein crystal structures In his review, he listed detailedly which amino acids could bind to the anions and even gave the possible geometric parameters for protein-anion interactions... ephrinB2 cytoplasmic domain in unsalted water; 2 To study the effects of different salts to the intrinsic unstructured protein; 3 To study the characteristics of the interactions between anions and unstructured proteins; 4 To study the characteristics of the interactions between anions and well-folded protein WW4 23 Chapter 2 Materials and Methods 24 2.1 DNA manipulation 2.1.1 Polymerase chain reaction... denaturation and destabilization by chaotropes The solubility of a basic protein, Peptibody A (PbA), was also found to be affected by ions Besides, the solubility of PbA turned to be more affected by anions than by cations (Saluja et al 2009) Enzymatic activity might be also mediated by the conformational changes induced by ions Kosmotropes, which are commonly believed to stabilize the native conformation,... perturbation of amide protein chemical shift Moreover, hydrogen exchange is also widely used to study the interactions between salts, osmolytes and proteins (Jolivart et al., 1998; Foord et al., 1998; Jaravine et al., 2000), as the NH resonances will exchange to ND in D2O and the NH signal decays So in this thesis, I employed NMR spectroscopy to study the interactions between ions and proteins at low ... moderate to high (> 0.1 M) salt concentrations as it is commonly believed that non-specific, electrostatic interactions are dominant at low salt concentrations (< 0.1 M), while at higher salt concentrations, ... comparable to anions, especially at low concentrations (Carbonnaux et al 1995) A reasonable interpretation given (Kunz, 2009) was that both cations and anions can interact with proteins with specificities... of cations in promoting the precipitation of proteins are not as pronounced as anions However, recent years more and more studies showed that the effect of cations on crystallization of proteins

Ngày đăng: 02/10/2015, 17:13

Từ khóa liên quan

Tài liệu cùng người dùng

  • Đang cập nhật ...

Tài liệu liên quan