Peptide Bond Geometry Studied by Solid-State NMR Spectroscopy THESIS Presented in Paritual Fulfillment of the Requirements for the Degree of Master of Science in the Graduate School of The Ohio State University By Chitrak Gupta, Graduate Program in Chemistry The Ohio State University 2013 Master’s examination committee: Professor Christopher Jaroniec, Advisor Professor Thomas Magliery Copyright by Chitrak Gupta 2013 Abstract Three-dimensional structure of proteins is an intrinsic property of every protein, and is directly related to its biological function. Studying protein structure is thus of immense importance to understand the mechanism by which the proteins perform such functions. In principle, the backbone structure of a protein can be completely described by a set of torsional angles. Indeed, a significant amount of structural studies of proteins have involved measuring the torsional angles defined by C’-N-Cα-C’ atoms (denoted φ) and N-Cα-C’-N atoms (denoted ψ). Little has been done, however, to study the third of the triplet, the ω-torsional angle, defined by the Cα-C’-N-Cα atoms, which is usually assumed to be planar, with the atoms arranged in a trans-like conformation. However, cis-like peptide bonds are known to exist, often at or near active sites, which make them biologically significant, although there are no reliable experimental methods apart from crystallographic studies to differentiate cis-peptide bonds form their trans- counterparts. This thesis is aimed at developing a new solid-state NMR experiment to study the peptide bond geometry. The primary objective is to differentiate cis- and trans-peptide bonds in polypeptides and proteins. Chapter 2 describes the synthesis of isotopically labeled model compounds with trans- and cis-peptide bond. The former, glycylglycine, was synthesized by solid phase peptide synthesis following standard protocols, with minor modifications to increase the yield. The latter, 2,5-diketopiperazine, was ii synthesized using microwave-assisted synthesis, a protocol recently published. Both the compounds could be obtained in high purity and integrity, as shown by their solution 1H and 13C NMR spectra. Chapter 3 describes the design of the NMR pulse sequence which can achieve this goal. The fundamental idea is to allow correlated evolution of two different anisotropic interactions under magic-angle spinning. Such evolutions are sensitive to the relative orientation of the two interactions. In this case, 13 α 15 C - N dipolar coupling and 13 C’-13Cα dipolar coupling have been correlated, to distinguish between the two geometries about the N-C’ bond. The possibility of using 13 13 C’ chemical shift anisotropy instead of 13 C’- Cα dipolar coupling has been discussed. Preliminary simulations have shown the detection of a cis-peptide bond to be possible, although, thus far, the experiment is not sensitive enough to measure deviation from planarity of a peptide bond. Chapter 4 shows the solid-state NMR spectra obtained from these model compounds. The 1D 13 C and 15 N CP-MAS spectra from both the compounds confirm their purity and integrity, and provide sufficient signal-to-noise to proceed with the proposed experiment. REDOR dephasing recorded on both the compounds are in good agreement with simulation. Chapter 5 discusses the future perspectives of this work, including extending it to bigger systems, and to use iii 15 N-1H dipolar coupling instead of 13 C’-13Cα dipolar coupling to increase sensitivity and enable measurement of the ω- torsional angle with reasonable accuracy. iv Dedication Dedicated to all hard workers awaiting recognition v Thank you Ma, Baba, and Ananya. vi Acknowledgments I shall take this opportunity to thank the variety of people without whose help this thesis would not have been a success. First of all, I would like to thank my advisor, Dr. Christopher Jaroniec, for his extremely helpful guidance throughout the course of my study in his research group. I would also like to thank all the former and current members of the group for the wonderful experience I have had working with them. I specially thank Judith Brown, Graduate Program Coordinator, and Jennifer Hambach, Graduate Admissions Coordinator, at the Department of Chemistry of The Ohio State University. Before and immediately after my arrival at OSU, I was full of confusions regarding the way the administrative processes works in this country, and Jennifer was always ready with help. Throughout my stay in the program, I have been immensely helped by the timely reminders from Judy, not to mention the multiple times I have been at her office with the strangest of problems. One of the most approachable people I have ever known, Judy has always had the answers to my questions. I would thank my parents for their unconditional love and support throughout my life, and for always encouraging me to pursue my dreams. I thank them for teaching me that life is bigger than success and failure, which has given me the strength to go on in difficult times. I would extend my gratitude to my high school teachers, speficially, Dr. vii Sekhar Pal, for getting me interested in Chemistry. I am grateful for having been taught by encouraging teachers, both at St. Stephen’s College and IIT Roorkee. My teachers not only helped me to grasp a fundamental understanding of Chemistry, but also inspired me to aim high. I have been extremely fortunate to have friends, both in Columbus and outside, who has made life livable. I would thank Souvagya, Chiranjit, Dwaipayan, Mithila, Arijit and Shiladitya for making my stay away from home a lot easier. I would forever cherish the potlucks and movie-sessions with Arijit and Shiladitya. I would never learn to cook and to make an apartment look decent, without Chiranjit. I am indebted to Mithila for teaching me to play squash and getting me to play table-tennis after fifteen years. No amount of gratitude is sufficient for Souvagya, who I could always bump into, for anything good and bad, but most importantly to discuss plans for long trips. My knowledge of history and digital electronics (and pretty much everything else) has increased significantly since meeting Dwaipayan. Some of the best times of my life have been those spent with each one of you at Buckeye Donuts, the coffee shop that by itself deserves special mention. I would thank Shreya and Joyeeta for always being just a phone call away, keeping me connected to the life I left behind in India. Last but not the least, I would express very special gratitude to Ananya, for being by my side and continuing giving me hope through thick and thin. I can never thank you enough for keeping me cheerful even during my darkest times. viii Vita April 22, 1986……………………….………………………………Born, Kolkata, India 2004………………………………………..Salt Lake School (Eng. Med.), Kolkata, India 2007………………………………..……………..B.Sc., University of Delhi, Delhi, India 2009………………..……M.Sc., Indian Institute of Technology Roorkee, Roorkee, India 2009-present………………………………..Graduate Teaching and Research assistant, Department of Chemistry, The Ohio State University Fields of study Major Field: Chemistry ix Table of contents Abstract…………………………………………………………………………ii Dedication……………………………………………………………………....v Acknowledgement……………………………………………………………...vii Vita………………………………………………………………………………ix List of tables……………………………………………………………………..xiii List of figures……………………………………………………………………xiv Abbreviations……………………………………………………………………xxi Chapter1:Introduction………………………………………………………………..1 1.1: Peptide bond geometry…..........................................................................1 1.1.1: Planarity of peptide bonds…………………….…………….…1 1.1.2: Unusual peptide bonds in proteins…………………………….2 1.1.3: Biological significance of non-proline cis-peptide bonds….…5 1.1.4: Proposed methodology: A solid-state NMR approach………..6 1.2: NMR spectroscopy……………………………………………………...8 1.2.1: Matter and spin………………………………………………..8 1.2.2: Zeeman splitting………………………………………………8 1.2.3: Fourier-transform (FT) NMR………………………………...11 x 1.2.4: Multidimensional NMR…………………………….………..12 1.2.5: Magnetic interactions……………………………….……….13 1.2.6: Solid-state NMR and magic-angle spinning………….…......20 1.2.7: Recoupling experiments……………………………….…….21 1.2.8: Measuring torsional angles………………………….………22 Chapter 2: Synthesis of model compounds……………………………………….24 2.1: Synthesis of model compound with trans-peptide…………………...25 2.1.1: Problems associated with solid phase peptide synthesis…...43 2.1.2: Synthesis of (U-13C2, 15N) glycylglycine…………………..44 2.2: Synthesis of model compound with cis-peptide……………………..47 2.2.1: Synthesis of 2,5-diketopiperazine with solid phase peptide synthesis…………………………………………………….……47 2.2.2: Synthesis of 2,5-diketopiperazine with microwave-assisted synthesis………………………………………………….………48 2.2.3: Synthesis of of (U-13C2, 15N) 2,5-diketopiperazine……….53 Chapter 3: Design of NMR experiment………………………………………….55 3.1: Recoupling experiments…………………………………………….55 xi 3.1.1: Heteronuclear dipolar recoupling…………………….....55 3.1.2: Homonuclear dipolar recoupling……………………….58 3.1.3: Recoupling of chemical shift anisotropy………………60 3.2: Cα-N-C-Cα experiment…………………………………………...60 3.2.1: Choice of C-C recoupling technique…………………..62 3.2.2: Choice of ratio of mixing times………………………..64 3.3: Utilizing carbonyl CSA for ω-angle measurement………………68 Chapter 4: Preliminary results……………………………………………….73 4.1: Solid-state 1D CP-MAS experiments…………………………...73 4.2: REDOR experiments……………………………………………77 Chapter 5: Conclusions and future work……………………………………80 References……………………………………………………………….….83 xii List of tables Table 1.1: Fraction of Xaa-Pro and Xaa-nonPro cis peptides as a function of structural resolution………………………………..……………………..…4 xiii List of figures Figure 1.1: Frequency distribution of ω-angle, energy difference from planar structure in peptides and proteins …………………..……………………………...…………1 Figure 1.2: Trans and cis peptide bond with and without proline…......................2 Figure 2.1: Glycylglycine…………..……………………………………………24 Figure 2.2: 2,5-diketopiperazine……….………………………………………..24 Figure 2.3: Proposed scheme for synthesis of glycylglycine…............................25 Figure 2.4: 1H NMR of glycylglycine (Bachem)…………….………………….26 Figure 2.5: 13C NMR of glycylglycine (Bachem)…………………………….…27 Figure 2.6: 1H NMR of product obtained by proposed scheme………..…….….29 Figure 2.7: 13C NMR of product obtained by proposed scheme...........................29 xiv Figure 2.8: 1H NMR of Fmoc-protected glycine (AAPPTec) in CDCl3 with TMS standard...............................................................................................................31 Figure 2.9: 1H NMR of Fmoc-protected glycine after attachment to resin and cleavage, dissolved in CDCl3 with TMS standard………………………….……………..31 Figure 2.10: 1H NMR of product obtained by treating resin-bound Fmoc-Gly-Gly with 50% TFA in DCM, dissolved in CDCl3 with TMS standard...………………….32 Figure 2.11: 1H NMR of product obtained after cleavage in DCM……………...34 Figure 2.12: ESI mass spectra of product obtained after cleavage in DCM……..35 Figure 2.13: 1H NMR of product obtained with modified scheme……………….36 Figure 2.14: 13C NMR of product obtained with modified scheme………………36 Figure 2.15: ESI mass spectra of product obtained with modified scheme………37 xv Figure 2.16: Overlaid 1H NMR of product obtained with modified scheme, with Fmoc removal done for 30 minutes, 4 hours and 24 hours…………………….……..39 Figure 2.17: 13 C NMR of product obtained with modified scheme, with Fmoc removal done for 24 hours................................................................................................40 Figure 2.18: The optimized synthesis scheme that was followed for the synthesis of (U13 C2, 15N) glycylglycine…………………………………………………….…41 Figure 2.19: RP-HPLC of glycylglycine at 220 nm……………………...…....42 Figure 2.20: Overlaid 1H NMR of glycylglycine (Bachem) and glycylglycine prepared by the optimized scheme………………………………………..…………….…42 Figure 2.21: HCl crystals of glycylglycine…………………………………...43 Figure 2.22: 1H NMR of (U-13C2, 15N) glycylglycine……………………….46 Figure 2.23: HCl crystal of (U-13C2, 15N) glycylglycine……………………..46 xvi Figure 2.24: Overlaid 13 C NMR of product obtained by microwave-assisted synthesis of 2,5-diketopiperazine, with hold times of 3, 6 and 9 minutes….……………...49 Figure 2.25: Overlaid 13 C NMR of product obtained by microwave-assisted synthesis of 2,5-diketopiperazine, with hold times of 15, 30 and 45 minutes…………….49 Figure 2.26: Overlaid 13C NMR of 2,5-diketopiperazine (Sigma) and 2,5-diketopiperazine synthesized by microwave-assisted synthesis……………………………….50 Figure 2.27: 1 H NMR of 2,5-diketopiperazine synthesized by microwave-assisted synthesis…………………………………………………………………...….50 Figure 2.28: Reaction profile for microwave-assisted synthesis of 2,5- diketopiperazine…………………………………………………………..……51 Figure 2.29: Overlaid HPLC at 220 nm of 2,5-diketopiperazine (Sigma) and 2,5diketopiperazine synthesized by microwave-assisted synthesis………..……..52 Figure 2.30: Overlaid synthesized by 1 NMR of 2,5-diketopiperazine (Sigma) and 2,5-diketopiperazine microwave-assisted synthesis after purification…………………………………………………………….………52 xvii HPLC Figure 2.31: Reaction profile for the microwave-assisted synthesis of (U-13C2, 15 N) 2,5- diketopiperazine………………………………………………….............………53 Figure 2.32: 13C NMR of (U-13C2, 15N) 2,5-diketopiperazine……………………54 Figure 3.1: Pulse sequence for Rotational Echo Double Resonance (REDOR) experiment………………………………………………………………………..56 Figure 3.2: Simulated 13C-15N REDOR dephasing………………………………57 Figure 3.3: 15N-1H dephasing pattern..…………………………..……………….57 Figure 3.4: Schematic representation of the pulse sequence for measuring relative orientation of αC-N and C’- αC internuclear vector………………………………61 Figure 3.5: Schematic representation for the same experiment with increased dimensionality……………………………………………………………………62 Figure 3.6: Simulations for the choice of 13 C-13C recoupling element. a. DRAWS b. MELODRAMA c. SPC-5 d. POST-C7…………………………………….......64 xviii Figure 3.7: Simulated dephasing spectra of REDOR and POST-C7…………..65 Figure 3.8: Simulations for the choice of ratio of REDOR and POST-C7 mixing time. a. REDOR mixing time = POST-C7 mixing time. b. REDOR mixing time = 2 * POST-C7 mixing time. c. REDOR mixing time = 0.5 * POST-C7 mixing time……………………………………………………………………………..67 Figure 3.9: Sensitivity of simulated dephasing to variation of ω-torsional angle…………………………………………………………………………….68 Figure 3.10: Schematic representation of the pulse sequence for measuring the relative orientation of the αC-N bond with carbonyl CSA……………………...............69 Figure 3.11: Simulated decay curve for the ROCSA experiment……………...71 Figure 3.12: Schematic representation of the pulse sequence for measuring the relative orientation of 15N-1H internuclear vector and carbonyl CSA…………………72 Figure 4.1: 1D 13 C CP-MAS spectrum of (U-13C2, 15 N) glycylglycine at magnetic field strength 500 MHz and 11.904 kHz magic-angle-spinning…………..………..74 xix Figure 4.2: 1D 15 N CP-MAS spectrum of (U-13C2, 15 N) glycylglycine at magnetic field strength 500 MHz and 11.904 kHz magic-angle-spinning…………..………75 Figure 4.3: 1D 13C CP-MAS spectrum of (U-13C2, 15N) 2,5-diketopiperazine at magnetic field strength 500 MHz and 11.904 kHz magic-angle-spinning…………….75 Figure 4.4: 1D 15N CP-MAS spectrum of (U-13C2, 15N) 2,5-diketopiperazine at magnetic field strength 500 MHz and 11.904 kHz magic-angle-spinning……………..76 Figure 4.5: 1D 13C spectrum of glycylglycine after CP transfers to 13 15 N and selectively to C (carbonyl)…………………………………..………………...................77 Figure 4.6: Experimental and simulated REDOR dephasing pattern of glycylglycine…………………………………………………………………78 Figure 4.7: Experimental and simulated REDOR dephasing pattern of 2,5diketopiperazine………………………………………………………………79 xx Abbreviations C’ Carbonyl carbon CP Cross Polarization CP-MAS Cross Polarization Magic Angle Spinning CSA Chemical Shift Anisotropy DIC N,N’-Diisopropylcarbodiimide DCM Dichloromethane DMAP 4-Dimethylaminopyrdine DMF Dimethylformamide DRAMA Dipolar Recovery At the Magic Angle DRAWS Dipolar Recovery with A Windowless Sequence ESI Electrospray Ionization Fmoc Fluorenylmethyloxycarbonyl FID Free Induction Decay FT Fourier Transform FT-NMR Fourier Transform Nuclear Magnetic Resonance HPLC High Performance Liquid Chromatography HOBt Hydroxybenzotriazole MAS Magic Angle Spinning xxi MELODRAMA MELding Of spin-locking and DRAMA NMR Nuclear Magnetic Resonance POST-C7 Permutationally Offset-STabilized C7 REDOR Rotational Echo Double Resonance RP-HPLC Reversed Phase High Performance Liquid Chromatography ROCSA Recoupling Of Chemical Shift Anisotropy SPC-5 Supercycled Permutationally Offset-STabilized C-5 TFA Trifluoroacetic Acid T-MREV Transverse MREV TMS Tetramethylsilane xxii Chapter 1 Introduction 1.1 Peptide bond geometry 1.1.1 Planarity of peptide bonds One of the most fundamental assumptions in protein structure is the assumption of peptide bonds being planar. This is due to partial double bond character of the N-C’ bond. This restricts peptide bonds to two possible conformations around N-C’ bond, where the ω torsional angle has value of 180º (trans) and 0º (cis) respectively. This was described by Corey and Pauling1 in1953 where they noted that (i) N-C’ bond has ~40% double bond character, and (ii) planarity of peptide bonds is a “sound structural principle”. However, in 1968, Ramachandran2 recognized the need for including the possibility of cis Figure 1.1: Figure taken from reference 6. Frequency distribution of ω angle (histogram), energy difference from planar structure for peptide (gray) and protein (black), from reference 5. Red points are the similar energies calculated from the distribution shown in histogram. 1 peptides in structural studies. The same group later showed3 that for peptides with bulky side chains, agreement between measured and calculated 3JHN-Hα can be improved by considering nonplanar peptide bonds. Much later, in 1991, Pople’s group used high level Hartree-Fock ab initio calculations to show4 that peptide bonds show significant flexibility about the N-C’ bond in the gas phase, with little energetic cost. Five years later, MacArthur and Thornton surveyed5 peptide and protein databases to show that experimentally derived structures show significant deviation from planar peptide bonds. They also calculated peptide bond energy wells around ω = 180º from ω angle distributions using MaxwellBoltzmann relationship, which were further updated by Edison6. Figure 1.1, taken from reference 6, shows the frequency distribution of ω angles along with original (reference 5) and updated values for energy of deviation of peptide bonds about planarity. 1.1.2 Unusual peptide bonds in proteins In addition to assumption of planarity, peptide bonds are usually assumed to be in the trans conformation. Indeed, most peptide bonds exist conformation. in This the can trans be explained qualitatively by the fact that cis-peptides have significantly higher energy thant their transFigure 1. 2: Figure taken from reference 7. Trans and cis peptide bonds with and without Proline 2 counterparts. Energy difference between trans- and cis-peptide bonds is ~ 10 kJ/mol for peptides in which C-terminal residue is anything other than Proline (Xaa-nonPro). This difference is ~ 2 kJ/mol if the C-teriminal residue is Proline (Xaa-Pro). This makes trans-peptides far more abundant than cis for both the categories. However, an attempt to quantify the above argument brings up a discrepancy. Based on these energy values, ~30% of C-terminus proline-containing (Xaa-Pro) peptides and ~1.5% of C-terminus non-proline (Xaa-nonPro) peptides should exist in the cis conformation. However, the fraction of peptides observed to exist in the cis conformation is significantly lower8, both for proline and non-proline peptides. Even more intriguing is the fact that the observed discrepancies are quite difference for proline and non-proline peptides. While observed cis-peptide bonds are lower than expected by a factor of ~ 6 for Xaa-Pro, the same is lower by a factor of ~ 50 for Xaa-nonPro. It is interesting to note that high resolution (< 2.0 Å) structures contain almost twice the number of cis Xaa-Pro bonds and four times the number of cis Xaa-nonPro bonds than medium and low resolution (> 2.5 Å) structures. This has been tabulated by Weiss and Hilgenfeld 8 (Table 1.1). This led the authors to believe that all the cis-peptides have not been identified. The authors also noted that most structure refinement programs allow for the possibility of a cis-peptide only in cases of Xaa-Pro unless specified explicitly. This immediately underlines the importance of direct measurement of cis-peptides, especially for XaanonPro, through experimental methods. Also to be noted is the observation made by Weiss and Hilgenfeld9 that in a structure where a cis-peptide is the correct structure, a 3 trans- conformation may be modeled with only the amide Nitrogen atom being displaced by 1.0 Å. In totality, there is a definite need to experimentally determine cis-peptide Table 1.1: Taken from reference 8. Fraction of Xaa-Pro and Xaa-nonPro cis peptides as a function of structure resolution < 2.0 Å 2.0 Å – 2.5 Å ≥ 2.5 Å of 571 291 184 96 of 153209 72576 52194 28448 All Number proteins Number peptide bonds Xaa-Pro 7413 3407 2566 1440 Xaa-nonPro 145796 69160 49628 27008 232 (0.32%) 140 (0.27%) 55 (0.19%) Number of cis 427 (0.28%) peptide bonds Xaa-Pro 386 (5.21%) 205 (6.02%) 129 (5.03%) 52 (3.61%) Xaa-nonPro 41 (0.028%) 27 (0.039%) 11 (0.022%) 3 (0.011%) bonds. While NMR spectroscopy can be used to distinguish Xaa-Pro trans- and cispeptides10-12, till date, no general methods exist for this purpose, apart from X-ray crystallography. The motivation, thus, is to develop a new experimental technique, that will reliably differentiate cis- and trans-peptide bonds, without depending on the presence of Proline at the C-terminus of the peptide. 4 1.1.3 Biological significance of non-proline cis-peptide bonds Although extremely low in frequency, non-proline cis-peptides often have an impact on the biological function of the protein they are a part of. Herzberg and Moult 13 examined 7 non-proline cis-peptide bonds, 6 of which were involved in ligand binding and/or catalytic activity. Later on, Weiss et. al. 14 reported a refined structure of cellular factor XIII, where they observed two non-proline cis peptides. As discussed by the authors, one of these contain the active site at its N-terminus, while the other is close to the dimerization interface. The authors also noted that the lack of any geometric need for a cis-peptide at either of these locations. They concluded that cis-trans isomerism of these peptides was the trigger for a conformational rearrangement, which provides a part of the energy for binding to the substrate. The same year, Stoddard and Pietrokovski15 reported the role of a non-proline cis-peptide in the self-splicing of DNA gyrase A. This and other instances of strained amino acid geometries in binding sites and catalytic locations instigated the authors to conclude that such geometries have “stored potential energy that may be used to drive biochemical reactions and other physical processes in the cell”. Non-proline cis-peptides have also been implicated in amyloid fibril formation, although a conclusive picture has not been reached as yet. Spencer et. al. 16 used solidstate NMR spectroscopy to measure 13C-13C distances in a 9-residue peptide representing the C-terminus of amyloid beta (Aβ), a major constituent of amyloid plaques that Alzheimer’s Disease and other related dieseases. From their results, they concluded that Gly37-Gly38 segment of β34-42 exists in cis-conformation. This result has been disputed later on17 by the same group by taking into account -inhomogeneous broadening effects that tend to make 13 C-13C distances longer than they actually are. However, even in this 5 work they could not confirm whether the said peptide bond is cis or trans, and only concluded that both possibilities were likely. Above examples highlight the need for an experimental technique capable of distinguishing trans- and cis- peptide bonds. The technique needs to be general, i.e., be able to pick up cis-peptides involving any two residues. Such a development, we hope, shall enable refinement of protein structure databases in terms of frequency of cispeptides, especially ones not containing a Proline residue at the C-terminus. Given the apparent role of such structurally-strained peptides in biological function, such a technique could potentially throw light on the mechanism of functioning of a variety of proteins. In addition, we tried to take this a step further. We wanted to see if, in addition to distinguishing between trans and cis peptide bonds, we could measure the deviation from planarity of a peptide (section 1.1.1). Depending on the sensitivity of this new experiment, it might be possible to obtain observable difference in the behavior of peptide bonds of varying ω angles. This will be discussed in Chapter 5. 1.1.4 Proposed methodology: A solid-state NMR approach With the problem defined and the importance outlined, proposed methodology for attacking the same shall be described here. In short, solid-state nuclear magnetic resonance (NMR) spectroscopy has been used with the goal of picking up cis peptides in proteins. NMR spectroscopy has a marked advantage over other techniques in that it allows site-specific resolution. Consequently, NMR spectroscopy is a popular method for studying structure and dynamics of proteins and other biological molecules. 6 Solution NMR has traditionally been the technique of choice for studying such large molecules. In solution, rapid molecular tumbling averages out anisotropic interactions, which results in sharp spectral lines, yielding good resolution. In solids, crystallites are frozen in space, which causes anisotropic interactions to dominate, giving broad spectra which are difficult to assign. However, with the advent of magic-angle spinning (MAS) solid-state NMR, such problems have been virtually overcome, and it is possible to obtain sharp, “solution-like” spectra in solid-state. With time, experimental approaches have been developed which allow reintroduction of anisotropic couplings during specified durations of a MAS solid-state NMR experiment. These interactions are extremely useful for the measurement of distances and torsional angles. In this thesis, a new solid-state NMR experiments shall be described which will be sensitive to the conformation of a peptide about the N-C’ bond. In the next section, the fundamental approaches of NMR spectroscopy shall be discussed. Chapter 2 shall discuss the preparation of model compounds with trans- and cis-peptide bonds. The details of the proposed experiment along with preliminary simulated results shall be described in Chapter 3. Some preliminary experimental results on the model compounds are shown in Chapter 4. Finally, future goals and possible modifications of the experiment will be discussed in Chapter 5. 7 1.2 NMR Spectroscopy 1.2.1 Matter and spin Matter is made up of atoms, which in turn are made up of electrons and nuclei. Each atomic nucleus has a unique number of protons and neutrons. The number of protons gives each nucleus their chemical identity, while the number of protons and neutrons together determine the nucleus’s mass and magnetic properties. The latter arises out of a special property, known as spin. Both proton and neutron has a spin of ½. The combined effect of the spins of each of these particles determine the ground state nuclear spin value for a given nucleus. It needs mention here that although nuclei can be excited to higher energy states, conditions required to achieve the same are prohibitive, and we restrict our discussion to nuclear ground states. Possible spin values for nuclear ground state are either integral (0, 1, 2 …) or half integral (1/2, 3/2, 5/2, ….). In this thesis, we shall only be considering nuclei whose nuclear spin has the value of ½. The nuclei most typically used in NMR of peptides and proteins are 1H, 13C, and 15 N. While 1H is the most abundant isotope of hydrogen, 13 C and 15 N are extremely rare in nature. Consequently, NMR signal cannot be detected for these nuclei in a natural abundance sample. Samples have to be isotopically labeled in order for these nuclei to be observed in an NMR experiment. 1.2.2 Zeeman splitting A nucleus with spin I (as described above) is (2I + 1)-fold degenerate. When such a nucleus is placed in a magnetic field, this degeneracy is lifted, and the (2I + 1) energy 8 levels are split. This is known as Zeeman splitting. For spin-1/2 (I = ½) nuclei, this generates two energy levels, which we call α and β. When a nucleus with non-zero nuclear spin I is placed in an external magnetic field of strength B0, it interacts with the magnetic field with interaction energy (E) given by: (1) where µ is the magnetic moment of the nucleus given by (2) where h is Planck’s constant (universal constant) and γ is the gyromagnetic ratio, characteristic of the nucleus. If the external field is assumed to be along the Z direction (in other words, if the direction of the external field is defined as the Z-axis), then the Hamiltonian for the Zeeman interaction can be written as: (3) where IZ is the spin angular momentum operator in the Z direction. From this point, the external field B0 will be assumed to be along the Z-axis. 9 In other words, energy difference between the two states α and β is given by equation (1). Thus, at a given temperature T and magnetic field strength B0, the relative population of the two states α and β (denoted nα and nβ respectively) are related by Boltzmann distribution as: (4) where kB is Boltzmann constant and E has been defined in equation (1). For magnetic field strengths typically used in NMR experiments and temperatures few fractions of a degrees K, the term inside parenthesis is extremely low, and a “hightemperature approximation” can be made wherein it can be written: which can be rearranged to obtain the fractional population of the higher energy state with respect to the ground state (5) At room temperatures and typical magnetic field strengths, this fraction is of the order of 10-5. This results in NMR being an insensitive technique. Nevertheless, this population difference is sufficient to detect signals with reasonably good signal-to-noise ratio. 10 1.2.3 Fourier-Transform (FT) NMR For spin-1/2 nuclei, equations (1) and (2) can be combined to write the difference in energy between states α and β as: (6) where ω0 is known as the larmor frequency of the nucleus, and is fixed for a given nucleus at a given magnetic field strength. Short radiofrequency pulses generated by RF coils around the sample, applied at a direction perpendicular to the external magnetic field (B0) will tilt the net magnetization away from B0. This induces an oscillating current in the RF coil, which can be detected as the signal. The signal is detected as a function of time (known as the free induction decay or FID), and then fourier transformed to generate the NMR spectrum. In the most general form, the FID can be written as: (7) where A is a constant depending on the experimental conditions, λ is damping constant, and Ω0 = ω0 – ωref, where ωref is a reference larmor frequency. This signal can then be fourier transformed as: (8) S(Ω) is the frequency-domain NMR spectrum. 11 It should be noted here that a pulse is described by a phase (Φ P), power (ωnut), frequency (ωP) and duration (τP). The angle βP by which the magnetization is tipped away from B0 given by: (9) where ωnut = |γBRF/2|, BRF being the oscillating RF field. It should also be noted here that the RF pulse interacts with the nucleus most strongly is the frequency ωP is exactly equal to the larmor frequency ω0. In this case, the pulse is said to be “on resonance” with the nucleus. Each nucleus is said to “resonate at” the larmor frequency. 1.2.4 Multidimensional NMR Although the problem of insensitivity of NMR spectroscopy (see section 1.2.2) can be overcome by using higher magnetic fields, the resolution is still not sufficient especially for large molecules like polypeptides and proteins. Such molecules typically have a large number of spins, and peaks from each of these make the spectrum extremely complicated and difficult to assign. To overcome this problem, different spins can be correlated so as to have additional dimensions in the spectrum. This was proposed by Jean Jeener 18 and demonstrated by Richard Ernst and corworkers19. Simplest example of a multidimensional NMR experiment is a 2D correlation experiment. In such experiments, transverse magnetization is created on one nucleus (say 15N) and is frequency-labeled for duration t1. This magnetization can then be transferred to another nucleus (say 13C). This transfer is usually through J-coupling or dipolar coupling (the following section discusses 12 these interactions briefly). Following the transfer, the signal is collected (for time t2). Thus, a two-dimensional FID is collected which can be double fourier transformed to generate a 2D 15N-13C correlation spectrum: In principle, higher dimensions are also achievable. Higher dimensions would increase resolution to aid spectral assigning, while being expensive in terms of time. Such multidimensional NMR experiment is extremely valuable for studying large biomolecules, as it allows for site-specific resolution. This is a feature unique to NMR spectroscopy, and gives it an advantage over other techniques commonly used for structural studies of biomolecules. 1.2.5 Magnetic interactions Having described the fundamentals of FT-NMR experiment, this section will describe the various magnetic interactions that are typically present in a sample of spin-1/2 nuclei placed in a magnetic field. 1.2.5.1 Chemical shift From what was discussed in the previous section, it would appear that all nuclei of a given type under a given magnetic field would have the exact same resonant frequency, as larmor frequenct depends only on the nucleus (gyromagnetic ratio) and magnetic field strength. However, this is not the case as the surrounding influences the resonant frequency of the nuclei. 13 The external magnetic field B0 induces a current in the electronic clouds surrounding the nucleus. This generates a microscopic magnetic field, Binduced, which can be expressed as (10) where δ is the chemical shift tensor, a rank two tensor. It is to be noted here that the symbol δ is used for the “deshielding” convention wherein resulting shift with respect to reference larmor frequency is +δ. A similar “shielding” convention can be used, where the symbol σ is used to describe a chemical shielding tensor, with opposite sign as the chemical shift tensor. In the shielding convention, Binduced is written as: In Cartesian coordinates, the chemical shift tensor can be written as: For each nuclear site, it is possible to choose an axis frame such that the chemical shift tensor becomes diagonal when expressed in the said axis frame. This axis frame is said to be the principal axis frame of the said chemical shift tensor. This interaction is small compared to the Zeeman interaction (see section 1.2.2), and hence, it can be approximated to only have the δ ZZ term (note that in the principal axis frame all off-diagonal terms are zero). Thus, the chemical shift Hamiltonian can be written as: 14 (11) This is known as the secular approximation. Here, is the ZZ element of the chemical shift tensor in the laboratory reference frame (the reference frame in which B0 field is along Z-axis), and θ defines the orientation of a particular crysyallite in a powder sample, with respect to this frame. Such orientation dependence means that all three principal components (δ XX, δYY, and δZZ) of the chemical shift tensor can contribute to the chemical shift Hamiltonian, depending on the oritentation of a given crystallite. The chemical shift interaction can be separated into an isotropic and an anisotropic component. As the name suggests, the isotropic component is independent of the crystallite orientation. These components, denoted δ iso and δaniso can be written as: The asymmetry parameter of the chemical shift, η, is given by: (12) In liquids, due to rapid molecular tumbling, the chemical shift is averaged to its isotropic value. This is because when the molecular tumbling is fast on the NMR timescale, each molecule can be assumed to have the same orientation with respect to B0. However, in most solids, due to the absence of such tumbling, the chemical shift has an orientational dependence. If the direction of B0 in the principal axis frame of the chemical shift tensor is b0 and is described by the polar angles (φ, θ), then it can be 15 shown that the effective resonant frequency (ωCS) for a nucleus of larmor frequency ω0 is given by: which for axially symmetric chemical shift tensors (η = 0) can be written as: (13) As can be seen from equation (13), in a powder, each crystallite has a different orientation with respect to B0, and consequently, experience slightly different chemical shifts. Each crystallite has its peak at a slightly different location, which superimpose to give a broad signal, known as “powder pattern”. This effect can be removed by magicangle spinning (see section 1.2.5). However, the oritentation-dependence of this anisotropy also makes it a useful probe for measuring torsional angles. The potential of chemical shift anisotropy (CSA) in measuring ω-angle in peptides and proteins will be discussed Chapter 5. 1.2.5.2 Spin-spin interactions A typical molecule contains more than one spins, each of which respond to the external magnetic field (and experience chemical shift) as described in the previous section. Additionally, each of these spins are coupled to each other. For spin-1/2 nuclei, the two possible mechanisms of spin-spin coupling are J-coupling and dipolar coupling. For both of these interactions, the Hamiltonian HC between nuclei i and j has the form: 16 (14) where I is the spin angular momentum (given by I2 = Ix2 + Iy2 + Iz2) and subscripts i and j denote the two nuclei, and C is a rank two tensor (similar to the chemical shift tensor described in the previous section) describing the said interaction. For each of these cases, random molecular tumbling in solution result in the tensors being averaged to their isotropic value, which is defined by the trace of the tensor. This is similar to the effect described for chemical shift, where molecular tumbling reduces it to its isotropic value. J-coupling J-coupling (also known as scalar coupling or “through-bond” coupling) is the coupling between two nuclei through bonding electrons. Their effect is strongest between two nuclei which are directly bonded, and die away usually within 3 bonds. In solution NMR, they cause splitting of NMR spectra. The J-coupling tensor has an anisotropic part which is extremely small and can be neglected. The resulting secular-approximated Hamiltonian is given by: If the chemical shift difference between the two species is significantly larger than the strength of the J-coupling, which is usually the case, then a further approximation (“weak coupling” approximation) can be made and the Hamiltonian further simplified as: While a common feature in solution NMR, they are usually dominated by stronger, dipolar couplings (described in the next section), in solids. J-couplings are not used in any of the experiments used in this thesis, and shall not be discussed any further. 17 Dipolar coupling Dipolar coupling (also known as “direct” coupling or “through-space” coupling) is the coupling between two nuclei without the involvement of electrons. The strength of the coupling is directly related to the distance between the two nuclei (falls off as inversecubed of the distance). The dipolar tensor D is of the form: (15) where x, y and z are the projections of the internuclear vector on the coordinate axes, and r is the length of the internuclear vector. Note that this is a traceless tensor, as the diagonal elements add up to zero. Using this in equation (13), it can be shown that the Hamiltonian for dipolar interaction between nuclei i and j is given by: (16) where I is the spin angular momentum and subscripts i and j denote the two nuclei, e ij is the unit vector parallel to the line joining the two nuclei, and bij is dipolar coupling constant given by: (17) 18 where µ0 is the permeability of free space, γi and γj the gyromagnetic ratios of the two nuclei, and rij the distance between the nuclei. Like chemical shift, the secular approximation can be used even for dipolar coupling to simplify the interaction Hamiltonian (equation 13). This simplified Hamiltonian can be written as: is the secular dipolar coupling and θij is the angle where between the vector eij and the external magnetic field. As for J-coupling, the weak coupling approximation can be made for dipolar coupling to further simplify the Hamiltonian as: (18) As mentioned earlier, in liquids, rapid molecular tumbling averages this interaction to its isotropic value, which is zero. Hence, dipolar coupling is usually not observed in solution. However, in solids, due to restricted motion, the anisotropic part is not averaged out. This is both a blessing and a nuisance. On one hand, this leads to “dipolar broadening” of solid-state NMR spectra, a problem that can be removed by magic-angle spinning (see section 1.5). On the other hand, dipolar coupling is a storehouse of information. The dependence of the dipolar coupling constant on internuclear distance (equation 16) makes dipolar coupling measurement a very useful tool for determining structural restraints. Additionally, angular dependence of the secular dipolar coupling constant makes it a useful probe for measuring torsional angles, a feature that has been utilized in this work. 19 1.2.6 Solid-state NMR and Magic-angle spinning The problems associated with anisotropic interactions, specifically, dipolar coupling and chemical shift anisotropy, have been described above. These interactions result in the spectral lines being too broad for meaningful assignment of the peaks. In order to obtain high-resolution NMR spectra of solids, it is necessary to remove these effects. This is achieved by magic-angle spinning. As can be seen from equations (13) and (18) and keeping in mind the orientational dependence of the term bij (equation 18), it can be seen that both of these terms have an orientational dependence of the form (3cos2θ – 1). Here, θ is the angle between B0 and orientation of the interaction tensor (chemical shift tensor and dipolar tensor respectively). In a powder sample, θ takes all possible values as all molecular orientations are possible. If we spin this sample about an axis inclined at an angle θ R with respect to B0, then θ varies with time as the molecule rotates. Under such circumstance, the average orientational dependence, <3cos2θ -1> (angular brackets denoting averaged) can be shown to be: (19) where β is the angle between the spinning axis and the Z axis of the principal axis frame of the tensor. θR, the angle between the sample and B0, is under the control of the experimenter. As can be seen from equation (19), if θR is set to 54.74º, the term on the left hand side 20 becomes 0. This angle is called the “magic-angle”, introduced by Andrew et. al20. and Lowe21. If the sample spinning is fast enough (so that θ is averaged rapidly compared to the dipolar coupling and the chemical shift anisotropy), the anisotropic components are averaged to zero, resulting in sharp spectral lines. The terms “rotor period” is used to refer to the time taken by the sample to complete one rotation. 1.2.7 Recoupling experiments As mentioned above, anisotropic interactions like chemical shift anisotropy and dipolar coupling, are both a blessing and a nuisance. While it is useful to remove their effect of line broadening from solid-state NMR spectra, it would also be profitable to be able to measure these quantities, as they contain useful structural information. For this reason, techniques have been developed to reintroduce these interactions under magic-angle spinning, for specific periods during the experiment. These are called “recoupling” experiments. Carefully designed sequence of pulses can be used to reintroduce chemical shift anisotropy, homonuclear dipolar coupling, and heteronuclear dipolar coupling. Some of these techniques have been used in this thesis, and will be described in Chapter 3. It is also possible to correlate two different anisotropic interactions to measure the relative orientation of the two tensors, which is the essence of measuring torsional angle using solid-state NMR. This is discussed in the next section. 21 1.2.8 Measuring torsional angles Measurement of torsional angles is a key step to structural studies of proteins and peptides. A polypeptide backbone has three different torsional angles, denoted φ, ψ and ω. Two of these, φ and ψ, are needed to completely describe the three-dimensional structure of the polypeptide chain. Consequently, more attention has been given to the measurement of these angles. A number of magic-angle spinning solid-state NMR experiments have been designed for measuring these torsional angles in proteins 22-31. This section shall explain the way these experiments correlate two different anisotropic interactions to give information about the relevant torsional angle. Consider a molecular fragment consisting of torsional angle about the 15 1 H-15N-13C-1H, in which the N-13C bond is to be measured31. In this experiment, 15 N transverse magnetization is generated, and dephased by reintroducing 1H-15N dipolar coupling, in a time-resolved fashion. This dephased magnetization is frequency-labeled (for site-specific resolution) and transferred to 13 C transverse magnetization. This magnetization is then dephased by reintroducing 1H-13C dipolar coupling in a timeresolved manner, and finally the magnetization is detected. The result is a series of 2D 15 N-13C correlation spectra in which each cross-peak is modulated by the relative orientation of the 1H-15N vector and the 1H-13C vector. The intensity of a given crosspeak is given by: where DN is the dephasing of the 15 coupling, DC is the dephasing of the N transverse magnetization due to 1H-15N dipolar 13 C transverse magnetization due to 1H-13C dipolar coupling, and T(τmix) is the magnetization transfer efficiency for the 15N to 13C transfer. 22 Each of these functions depend on the relative orientation of the molecular fragment with respect to B0. However, it is usually possible assume T(τmix) to be constant, and consider only the two other functions. The functional forms of these are: where ω is the orientation-dependent dipolar coupling being reintroduced ( 1H-15N amd 1 H-13C respectively). Hence, the intensity cross-peaks can be written as: where ωHN and ωHC are the orientation-dependent dipolar couplings being reintroduced for periods τN and τC respectively, and the angular brackets denote an average over all crystallites. The two dephasing periods τN and τC could be varied independently to generate effectively a 4D experiment. However, time required for such an experiment would be prohibitively long. Hence, the two dephasing periods are varied synchronously to generate a 3D experiment. This is the methodology that shall be used in this thesis for describing a new experiment for obtaining information on geometry about the N-C’ bond in polypeptides. In principle, different anisotropic interactions including dipolar couplings and chemical shift anisotropy can be correlated for this purpose. These shall be discussed in Chapter 3, and the most appropriate experiment shall be chosen from preliminary simulation results. Chapter 4 shall discuss some of the results from the experiments performed on the model compounds. Chapter 5 shall discuss some possible modifications to the experiments that can be made for better results. 23 Chapter 2 Synthesis of model compounds Linear diglycine (Figure 2.1) was chosen as a model compound with trans-peptide, and 2,5-diketopiperazine (Figure 2.2) was chosen as a model compound with cis-peptide. Both of these compounds are commercially available. However, for our experiments, we needed isotopically labeled (U-13C2, 15 N) versions of these compounds (see section 1.2.1). For this reason, both of these compounds were synthesized in-house. The commercially available versions were purchased and used for standardization purposes. This chapter describes the synthesis protocols used, optimazions required, and final results obtained for the synthesis of these two compounds. Figure 2.1: Glycylglycine Figure 2.2: 2,5-diketopiperazine All 1H NMR was taken at 400 MHz 1H larmor frequency and all 13C NMR was taken at 500 MHz 1H larmor frequency. All NMR spectra were recorded in D2O solvent unless specified otherwise. 24 2.1 Synthesis of model compound with trans-peptide Glycylglycine can be synthesized using solid-phase peptide synthesis using standard Fmoc-chemistry. The original scheme that was planned for this synthesis is shown in Figure 2.3. Resin used was Wang Resin and was obtained from AAPPTec. Fmoc- protected Glycine was also obtained from AAPPTec. DIC was obtained from Sigma Aldrich and HOBt.H2O was obtained from AnaSpec Inc. Figure 2.3: Proposed scheme for synthesis of glycylglycine 25 Glycylglycine was purchased from Bachem for standardization tests. 1H and 13C NMR of this compound is shown (Figures 2.4 and 2.5 respectively). Figure 2.4: 1H NMR of Glycylglycine (Bachem) The following protocol was followed for the synthesis scheme shown in Figure 2.3. ~ 2 g Wang Resin was washed with DCM for 30 minutes and 10% DMF in DCM for 30 minutes. 200 mg HOBt.H2O and 20 mg DMAP was added to Fmoc-protected Glycine and dissolved in minimum amount of DMF. 200 µL DIC was added and allowed to stand for 5 minutes. This mixture was then added to the resin and shaken for 14 hours. 0.1 mL Pyridine and 120 µL Acetic anhydride was added to it and shaken for further 30 minutes. 26 The liquid was drained and the beads washed alternately with DCM and DMF (four times each). After drying the beads were washed once more with DMF and 20% Piperidine in DMF was added. It was shaken for 30 minutes. Kaiser test was performed after draining the liquid and drying, to confirm the presence of free amine. The beads were again washed alternately with DCM and DMF (four times each). After drying, it was washed with 10% DMF in DCM. Figure 2.5: 13C NMR of Glycylglycine (Bachem) 27 200 mg HOBt.H2O and 20 mg DMAP was added to Fmoc-protected Glycine and dissolved in minimum amount of DMF. 200 µL DIC was added and the mixture was allowed to stand for 5 minutes. This was added to the beads and shaken for 8 hours. The liquid was washed alternately with DCM and DMF (four times each). 0.1 mL Pyridine and 120 µL Acetic anhydride was added to it and shaken for further 30 minutes. The liquid was drained and the beads washed alternately with DCM and DMF (four times each). After drying the beads were washed once more with DMF and 20% Piperidine in DMF was added. It was shaken for 30 minutes. Kaiser test was performed after draining the liquid and drying, to confirm the presence of free amine. The beads were again washed alternately with DCM and DMF (four times each). After drying, the beads were washed once more with DCM 50% TFA in DCM was added and shaken for 2 hours. The filtrate was collected. The beads were washed with small portions of TFA, and the was combined with the filtrate. DCM was evaporated at 35 ºC under vacuum, and the residue was dissolved in appropriate solvent (usually D2O) for taking 1H and 13C NMR. The product obtained was analyzed by taking 1H NMR (Fiugure 2.6) and 13 C NMR (Figure 2.7). Clearly, Glycylglycine was not the major product. The product(s) present could not be identified. Reason for not obtaining the product was not confirmed, but it was probable that since the peptide was so short, there was possible cyclization (or other side reactions) during cleavage from the resin. 28 Figure 2.6: 1H NMR of product obtained by proposed scheme Figure 2.7: 13C NMR of product obtianed by proposed scheme 29 If cyclization was indeed the problem, then cleavage of the peptide from the resin before the removal of the Fmoc group should solve the problem, as, in that case, there would be no free amine group during cleavage which could cause partial cyclization. What needed to be confirmed was whether cleavage could work in the presence of Fmoc (most protocols have Fmoc groups removed before cleavage from resin). To check this, Fmoc-protected Glycine was attached to the resin and then cleaved with 50% TFA in DCM. 1H NMR of Fmoc-protected Glycine is shown in Figure 2.8 and that of the product obtained is shown in Figure 2.9. These compounds are insoluble in D 2O, and these NMR spectra were taken in CDCl3 solvent, with TMS added as internal reference. Figure 2.9 is complicated as the compound is impure. However, presence of peak at 7.2 ppm confirms that the product contains Fmoc-group, which must have been cleaved from the resin. In other words, it was confirmed that deprotection works in the presence of Fmoc. The next thing that had to be checked is, whether 50% TFA in DCM can, apart from cleavage, also effect removal of the Fmoc group if treated for longer durations. This would make the working up of the reaction much simpler, as DCM solvent is a lot easier to remove than DMF, the solvent conventionally used for removal of Fmoc group. To check this, the synthesis scheme (Figure 2.3) was repeated until the last two steps. Then, the treatment with 50% TFA in DCM was continued for 4 hours, and the product was isolated by evaportating DCM. 1H of this compound is shown in Figure 2.10. 30 Figure 2.8: 1H NMR of Fmoc-protected Glycine (AAPPTec) in CDCl3 with TMS standard Figure 2.9: 1H NMR of Fmoc-protected Glycine after attachment to resin and cleavage, dissolved in CDCl3 with TMS standard 31 Figure 2.10: 1H NMR of product obtained by treating resin-bound Fmoc-Gly-Gly with 50% TFA in DCM, dissolved in CDCl3 with TMS standard The 1H NMR does show a 1:1 peak corresponding to the two methylene groups. However, comparison with 1H NMR of Glycylglycine (Figure 2.4) shows that the position of these peaks are shifted by ~ 1 ppm. This was clearly not the desired product. This clearly meant that this product had to be treated with piperidine in DMF for the removal of the Fmoc group. However, the work-up of that step was a problem that had to be solved. Normally, Fmoc removal is done while the peptide is bound to the resin beads. After the reaction, solvent (containing excess reactant and unwanted product) can be allowed to drain. However, if Fmoc removal was done after cleavage, purification of the product would be difficult. HPLC purification could be done if the product could be dissolved in water (Glycylglycine is soluble in water). However, that would require removal of DMF first, which is quite difficult. 32 If it was possible to do Fmoc removal in DCM instead of DMF, purification would be easier, as DCM could be easily removed. To check this, the synthesis was repeated, with cleavage from resin done with Fmoc still attached. This compound was treated with Piperidine in DCM. The 1H NMR (Figure 2.11) did not have the peaks corresponding to the product (Figure 2.4). Clearly, Fmoc removal had to be done in DMF. ESI Mass Spectra was taken for this compound (Figure 2.12). The product could not be identified even from its mass. At this point, a method had to be found to remove DMF after reaction was over. Two such methods were tried: 1. Adding Toluene to create azeotrope: Toluene was added to the solution in 1:5 ratio (i.e. 50 mL Toluene to be added to 10 ml DMF). This creates an azeotrope. Evaportation at 60 degrees was found to dry the solution in ~ 2 hrs. If it was still not dried, additional Toluene could be added and the process could be continued to yield a dry mixture. 2. Air-drying: The solution was taken in a 2-necked or 3-necked round-bottomed flask. One of the mouths was kept open and the rest (2 of them for 3-necked and 1 for 2-necked flasks) 33 were closed with septum. Through one of the septums (the only septum for 2-necked), air was inject by inserting a syringe which is connected to a tube attached to the air vent. Air flow was adjusted so that it is not high enough to spill the solution through the open mouth. Dry mixture was obtained overnight or in two days depending on the volume of the solvent. Figure 2.11: 1H NMR of product obtained after cleavage in DCM 34 Figure 2.12: ESI Mass Spectra of product obtained after cleavage in DCM Air-drying had the additional advantage of completely removing TFA. When Tolueneazeotrope was evaporated, the mixture still smelt of TFA. This is usually not reliable as TFA can interfere with other reactions. After air-drying, there was no smell of TFA. For this reason, this method of removing DMF was used henceforth. The synthesis was repeated with the modifications described above. 1H and 13 C NMR of the product is shown in Figures 2.13 and 2.14 respectively. As can be seen in the 1 H NMR (Figure 2.13), the 1:1 peaks are still shifted by ~ 1 ppm (compared to Figure 2.4). However, since it shows peaks corresponding to the Fmoc group, chances are that it is Fmoc-Gly-Gly-OH. To confirm this, ESI Mass Spectra was taken for this compound (Figure 2.15). 35 Figure 2.13: 1H NMR of product obtained with modified scheme Figure 2.14: 13C NMR of product obtained with modified scheme 36 Figure 2.15: ESI Mass Spectra of product obtained with modified scheme As can be seen, peak appears where it is expected for Fmoc-Gly-Gly-OH. This confirmed that this is Fmoc-Gly-Gly-OH. The only possible explanation for this is that, Fmoc removal becomes slower when the peptide is bound to the resin. This makes intuitive sense as, when the resin is bound, only the Fmoc group is oriented towards the outside, and it is easy for Piperidine molecules to “find” the Fmoc groups. Whereas, when not bound to resin, the Fmoc-GlyGly-OH molecules are randomly oriented, and the Piperidine molecule has to “find” the Fmoc group before the deprotection can take place. 37 With this contention, the process was repeated, this time the last deprotection done for 4 hours and 24 hours. Overlaid 1H NMR of the products are shown (Figure 2.16) As can be seen, 4 hour deprotection was still insufficient, while 24 hour deprotection shows 1:1 methylene peaks where they are expected for glycylglycine (see Figure 2.4). Also note that the product of 24-hour deprotection was insoluble in CDCl3 and hence was dissolved in D2O for 1H NMR, which gives the strong D2O peak in the last product. 13 C NMR of this product (Figure 2.17) also matches well with that of Glycylglycine (Figure 2.5). This method was used for the synthesis of (U-13C2, 15 N) Glycylglycine. The scheme is shown in Figure 2.18. The product, after air-drying, was dissolved in water and purified by RP-HPLC (Waters 1525) on a C18 column using water (with 0.01% TFA) and acetonitrile (with 0.01% TFA) as solvent. The concentration of H2O was varied from 95% to 70% in 20 minutes (8mL/min flow rate). Absorbance at 200 nm is shown (Figure 2.19). The peak at 10 minutes is confirmed to be Glycylglycine by 1H NMR obtained by lyophilizing the fraction and dissolving in D2O. Overlaid spectra of this compound with that of commercially available Glycylglycine (Bachem) is shown (Figure 2.20). 38 Figure 2.16: Overlaid 1H NMR of product obtained with modified scheme, with Fmoc removal done for 30 mins, 4 hours and 24hours. 39 Figure 2.17: 13 C NMR of product obtained with modified scheme, with Fmoc removal done for 24 hours 40 Figure 2.18: The optimized synthesis scheme that was followed for the synthesis of (U-13C2, Glycylglycine 41 15 N) Figure 2.19: RP-HPLC of Glycylglycine at 220 nm Figure 2.20: Overlaid 1H NMR of Glycylglycine (Bachem) and Glycylglycine prepared by the optimized scheme 42 The product was crystallized as its HCl salt. Needle-like crystals appeared in 2 hours (Figure 2.21). Figure 2.21: HCl crystal of Glycylglycine 2.1.1 Problems associated with solid phase peptide synthesis While being being reasonably simple, solid-phase peptide synthesis had one problem for our purposes. Usually, for such synthesis, the protected amino acids are not the most expensive reagent. Hence, most protocols can afford to follow a simple trick: in every step, 3-5 folds of excess amino acid is added, with respect to the resin loading capacity. This is to ensure that all the sites on the resin (or the peptide already attached) are 43 occupied by the amino acid, so as to maximize the yield. For our purpose, the isotopically labeled amino acid was indeed the most expensive reagent, and it was not possible to add too much excess. This meant that after the first addition, there would probably be a significant number of sites on the resin that would be unoccupied. Although Pyrdine and Acectic anyhydride was added to cap the unoccupied sites, it still meant that during the next attachment, some of the Fmoc-protected glycine would attach directly to the resin (unbound sites) rather than to the first Glycine. This, apart from reducing the yield, would cause the resultant compound to be impure, as there will be some Glycine contamination. This was a problem that could not be solved for the synthesis of (U-13C2, 15 N) Glycylglycine. 2.1.2 Synthesis of (U-13C2, 15N) Glycylglycine (U-13C2, 15N) Fmoc-Glycine was purchased from Cambridge Isotopes Ltd. 2.0021 g Wang Resin (AAPPTec) was washed with DCM for 30 mins and 10% DMF in DCM for 38 mins. 200.6 mg HOBt.H2O and 20.7 mg DMAP was added to (U-13C2, 15N) FmocGly-OH (Cambridge Isotopes Ltd.) and dissolved in minimum DMF. 200 μL DIC was added and allowed to stand for 5 mins. This mixture was added to the resin and shaken for 14 hrs. 0.1 mL pyridine and 120 μL Ac2O was added and it was shaken for further 30 mins. The liquid was drained and the beads were washed alternately with DCM and DMF four times. After drying, beads were washed once more with DMF, and 20% piperidine in DMF was added. It was shaken for 30 mins. Kaiser test was done after drying, and it showed presence of free amine. The beads were again washed alternately with DCM and DMF four times. After drying, it was ashed with 10% DMF in DCM. 44 212.4 mg HOBt.H2O and 13.9 g DMAP was added to (U-13C2, 15 N) Fmoc-Gly- OH (Cambridge Isotopes Ltd.) and dissolved in minimum DMF. 200 μL DIC was added and allowed to stand for 5 mins. This was added to the beads and shaken for 8 hrs. The liquid was drained and washed alternately with DCM and DMF four times. After drying, 50% TFA in DCM was added and shaken for 2 hrs. The filtrate was collected. The beads were washed with small volumes of TFA and the filtrate was combined. TFA was removed by evaporating under vacuum first at 60 ºC for 2 hrs, then at 35 ºC overnight. For complete removal of TFA, the mixture was air-dried for 24 hrs. 20 % piperidine in DMF was added to the residue and shaken for 24 hrs. The reaction mixture was air-dried, and the residue was dissolved in water and purified by RP-HPLC (Waters) with H2O-Acetonitrile solvent, with H2O concentration varying from 95% to 70% in 20 mins (8mL/min). The purified product was lyophilized and crystallized as its HCl salt. 1H NMR in D2O (Figure 2.22), apart from splitting due to labellings introduced, showed traces of impurity which is probably Glycine, introduced for reasons explained in the previous section. This might also be the reason why the crystal quality (Figure 2.23) was inferior compared to natural abundance Glycylglycine (Figure 2.21). 45 Figure 2.22: 1H NMR of (U-13C2, 15N) Glycylglycine Figure 2.23: HCl crystal of (U-13C2, 15N) Glycylglycine 46 2.2 Synthesis of model compound with cis-peptide 2.2.1 Synthesis of 2,5-diketopiperazine using solid phase peptide synthesis Cylcic peptides can synthesized using solid-phase peptide synthesis. On-resin cylcization, which involves cyclization of the peptide while being bound to the resin, has been reported in literature32, 33 . Taylor and coworkers34 have prepared three simple piperazinediones using Kaiser oxime resin, and Smith et. al35. have described the protocol for synthesizing cyclic dipeptides using on-resin cyclization on Kaiser oxime resin. However, Kaiser oxime resin is not compatible with Fmoc chemistry, and Boc chemistry has to be used. Boc chemistry has the following disadvantages: 1. Reactions are extremely slow. The attachment of the first peptide to the resin takes close to 24 hours (12 hours for Fmoc chemistry on Wang resin), and the subsequent peptides require 12 hours (4-5 hours for Fmoc chemistry). 2. Removal of Boc group is extremely difficult. It requires 95% TFA which is extremely dangerous. It is also a very slow reaction, requiring ~ 16 hours (Fmoc group can be removed in 2 hours using 50% TFA). Additionally, solid-phase peptide synthesis has the problems described in the previous section. For these reasons, 2,5-diketopiperazine was not synthesized using solid-phase peptide synthesis. Instead, a non-conventional approach was used, which is described in the following section. This method is extremely simple, very fast (2,5-diketopiperazine can be obtained from glycine in less than an hour), and the product extremely pure. 47 2.2.2 Synthesis of 2,5-diketopiperazine using microwave-asisted synthesis Microwave-assisted synthesis of 2,5-diketopiperazine has been described36. Briefly, 250 mg Glycine was taken in a 10 mL microwave reaction vessel, and 2 mL DMF was added. It was subjected to microwave reaction in CEM Discover SP microwave reactor. Set point temperature was set at 210 ºC and the hold time was varied to get maximum yield. The product can be distinguished from unreacted Glycine by 13 C NMR taken in D2O. While carbonyl of Glycine shows peak at ~172 ppm, 2,5-diketopiperazine shows peak at ~ 168 ppm. When the hold time is increased, the peak at 168 ppm is seen to increase in intensity at the cost of the peak at 172 ppm. Initial reactions were done with hold times of 3, 6, and 9 mins (Figure 2.24). Even at 9 mins, considerable amount of Glycine was uncreacted. Next, reactions were done with hold times of 15, 30 and 45 mins (Figure 2.25). It was seen that at 30 mins, all the glycine converted to 2,5-diketopiperazine. Comparison of this spectrum with that for control 2,5-diketopiperazine (obtained from Sigma Aldrich) confirmed the formation of the product (Figure 2.26). 1H NMR of the same compound is shown (Figure 2.27). Reaction profile for this reaction is shown (Figure 2.28). 48 Figure 2.24: Overlaid 13 C NMR of product obtained by microwave-assisted synthesis of 2,5- diketopiperazine, with hold times of 3, 6 and 9 minutes Figure 2.25: Overlaid 13C NMR spectra of product obtained by microwave-assisted synthesis of 2,5diketopiperazine with hold times of 15, 30 and 45 minutes 49 Figure 2.26: Overlaid 13C NMR of 2,5-diketopiperazine (Sigma) and 2,5-diketopiperazine synthesized by microwave-assisted synthesis Figure 2.27: 1H NMR of 2,5-diketopiperazine synthesized by microwave-assisted synthesis 50 Figure 2.28: Reaction profile for microwave-assisted synthesis of 2,5-diketopiperazine Although the NMR spectra looked clean, the product didn’t crystallize. So, it was passed through HPLC column under conditions similar to purification of Glycylglycine. As a control experiment, 2,5-diketopiperazine purchased from Sigma Aldrich was also passed through HPLC column in a similar fashion. Overlaid HPLC at 220 nm for control 2,5-diketopiperazine (obtained from Sigma Aldrich) and 2,5-diketopiperazine prepared by microwave-assisted synthesis is shown in Figure 2.29. This confirmed the product eluting at 10 mins was 2,5-diketopiperazine. It was further compared by 13C NMR of the product after lyophilizing the fractions (Figure 2.30). 51 Figure 2.29: Overlaid HPLC at 220 nm of 2,5-diketopiperazine (Sigma) and 2,5-diketopiperazine synthesized by microwave-assisted synthesis Figure 2.30: Overlaid 1H NMR of 2,5-diketopiperazine (Sigma) and 2,5-diketopiperazine synthesized by microwave-assisted synthesis after HPLC purification 52 2.2.3 Synthesis of (U-13C2. 15N) 2,5-diketopiperazine (U-13C2,15N) Glycine (purchased from Cambridge Isotopes Ltd.). 250 mg (U-13C2,15N) Glycine was taken in a 10 mL microwave reaction vessel, and 2 mL DMF was added. It was subjected to microwave reaction in CEM Discover microwave reactor. Set point temperature was set to 210 ºC. The hold time was kept at 30 mins. The reaction profile for this reaction is shown (Figure 2.31). The product was purified by HPLC as described above. The purified product showed splitting in 13C NMR (Figure 2.32) as expected due to the labeling introduced. Figure 2.31: Reaction profile for the microwave-assisted synthesis of (U-13C2, diketopiperazine 53 15 N) 2,5- Figure 2.32: 13C NMR of (U-13C2, 15N) 2,5-diketopiperazine 54 Chapter 3 Design of NMR experiment Section 1.2.8 described the essence of torsional angle measurement using solid-state NMR spectroscopy. Such experiments extensively use dipolar recoupling experiments (see section 1.2.7). The proposed experiments, too, shall make use of such recoupling techniques, and a more detailed discussion of these techniques shall be presented here. 3.1 3.1.1 Recoupling experiments Heteronuclear dipolar recoupling Techniques for measuring heteronuclear dipolar recoupling have a similar form. All of them are based on a spin echo experiment, which normally refocuses heteronuclear dipolar couplings. However, in these techniques, specially designed sequence of rf pulses are applied on one of the spins to prevent complete refocusing of the dipolar coupling. In this thesis, two different dipolar recoupling techniques shall be discussed: 13C-15N dipolar recoupling and 15N-1H dipolar recoupling. 13 C-15N dipolar recoupling The most common and robust technique for 13 C-15N dipolar recoupling is Rotational Echo Double Resonance (REDOR)37. The pulse sequence for this technique is shown in Figure 3.1. REDOR consists of two trains of pulses applied on the 15 N channel, which are rotor-synchronized. Pulses are applied every half rotor period on the 55 15 N channel. The two trains of pulses are separated by a period of two rotor periods, which contain, at the center, the echo pulse on the 13 C channel. The 13 C-15N dipolar coupling evolves only during the train of 15N pulses, known as the REDOR “mixing period”, denoted by t mix. The effect of the 13 C-15N dipolar coupling is to dephase the 13 C transverse magnetization. This in turn reduces the 13C peak intensity of the resultant NMR spectrum. Figure 3.1: Pulse sequence for Rotational Echo Double Reosnance (REDOR) experiment 37 In practice, a series of experiments are performed where t mix is incremented in steps of 2τR where τR is the rotor period. The 13 C peak intensity as a function of the REDOR mixing period is then plotted to give a REDOR dephasing pattern. Experimental REDOR dephasing patterns can be fitted to simulated patterns to extract the the 13 C-15N internuclear distance. A simulated REDOR experiment with 13C-15N distance of 1.5 Å is shown in Figure 3.2 56 Figure 3.2: Simulated 13C-15N REDOR dephasing. Simulation done using SPINEVOLUTION38 15 N-1H dipolar recoupling Most solid-state NMR spectroscopy experiments biomolecules involve of 1 H decoupling following initial 1H flip and CP transfer. This is because the large number of 1H’s typically present in a biomolecule are involved in strong dipolar Figure 3.3: 15N-1H dephasing pattern. Figure taken from reference 39 coupling network, which are not removed completely even after magic-angle spinning. Such decoupling techniques, apart from removing homonuclear 57 1 H-1H couplings, also remove heternuclear couplings of 1H with neighboring 15 13 C and/or N. An experiment like REDOR cannot be performed on 15 N-1H coupled systems for this reason, unless 1H-diluted proteins are prepared, which is beyond the scope of this thesis. The method employed for recoupling 15 N-1H dipolar interaction has been described39 in literature. It uses a special decoupling on 1H for specific duration of the experiment. Such special decoupling only removes homonuclear 1 H-1H dipolar interaction while allowing heteronuclear dipolar interactions to evolve for the specified duration. 15 N-1H dipolar coupling is stronger than 13 C-15N dipolar coupling. This can be easily concluded by looking at the expression for dipolar coupling constant (Equation 17) and noting that gyromagnetic ratio of 1H is much higher than that of 13C. For this reason, 15 N-1H dephasing curve decays much faster than 13C-15N decay, as shown by Figure 3.3, which is taken from reference 39. As can be seen, 0.25 ms, whereas 13 15 N-1H dephasing decays to zero in ~ C-15N dephasing decays to zero in ~ 2 ms (Figure 3.2). This feature has been utilized and is discussed in Chapter 5. 3.1.2 Homonuclear dipolar recoupling For the purposes of this thesis, discussion on homonuclear dipolar recoupling shall be restricted to recoupling of 13C-13C dipolar interaction. Unlike 13C-15N dipolar interaction, there is no one “best” technique for reintroducing 13 C-13C dipolar interaction. Since the first of its kind40 known as Dipolar Recovery At the Magic Angle (DRAMA), a series of 58 such methods have been published41-45 which are either entirely new or modification of a previously published method. From these choices, four methods were chosen as probably candidates for the purpose of the current work. Dipolar Recovery with A Windowless Sequence (DRAWS42), MELding Of spin-locking and DRAMA (MELODRAMA41), Permutationally Offset STabilized C7 (POST-C744) and Supercycled POST-C5 (SPC-545) experiments were chosen for their higher bandwidth and/or lower rf power requirements. The best candidate, to be ultimately used in the experiment, was chosen based on the sensitivity of the experiment, as obtained from simulations, described in section 3.2. The rf power requirement plays an important role. All of these techniques require the rf power on the 13C channel, in frequency units, to be a certain multiple of the sample spinning frequency. However, the rf power on the 1H channel for decoupling should be ~ 50 kHz higher than the 13 C channel power for effective decoupling. Taken together, a higher power requirement for the 13 C-13C recoupling technique would indirectly affect the power requirements on the 1H channel. Too high rf powers generates heat and are capable of damaging the sample and the NMR probe. Another point to be kept in mind is that these techniques also vary in terms of how well they compensate for chemical shift anisotropy. This is important as carbonyl carbon, a part of the spin system we are interested in, usually has high chemical shift anisotropy. 59 3.1.3 Recoupling of chemical shift anisotropy Like dipolar coupling, chemical shift anisotropy is also averaged to zero by magic angle sample spinning. However, this too can be reintroduced by specially designed pulse sequence46 . While high chemical shift anisotropy of carbonyl could be a nuisance when reintroducing 13 C-13C dipolar interaction, it itself could provide an alternative way of obtaining information about the geometry of the peptide bond. A change in the ω-angle would cause a change in the relative orientation of the 13 C-15N internuclear vector with respect to the chemical shift tensor of the carbonyl carbon. Thus, measuring the relative orientation of these two anisotropic interactions could also prove fruitful in measuring ωangle. The possibility of using this method is discussed in Chapter 5. 3.2 Cα-N-C-Cα experiment The ω-torsinal angle between residues n and (n+1) is defined by the nuclei αCn+1, Nn+1, C’n, and αCn. This angle can be thought of as the angle between the αCn+1- Nn+1 bond and the OCn- αCn bond, which can be measured by the correlated evolution of αCn+1- Nn+1 and C’n- αCn dipolar coupling. A schematic representation of the pulse sequence which can achieve this goal is shown in Figure 3.4. 60 Figure 3.4: Schematic representation of the pulse sequence for measuring relative orientation of αC-N and C’-αC internuclear vector Briefly, after initial 1H flip and CP transfer to 13C to create transverse 13C magnetization, 13 C-15N dipolar coupling is reintroduced by REDOR pulse sequence (see section 3.1.1). Following this, the magnetization is transferred to 15 N and then back to 13C. The last CP transfer is selective, so that only the carbonyl carbons are magnetized. This is ensured by adjusting the frequency offset of the CP spin-locking pulse and proper phase cycling. At this point, 13 C-13C dipolar coupling is reintroduced by an appropriate pulse sequence. This results in 13C magnetization which has been dephased by the correlated evolution of 13 C-15N and 13 C-13C magnetization. This magnetization is then detected as FID (see section 1.2.3) and fourier transformed to generate an 1D 13 C spectrum. A series of such spectra are generated where the duration of the recoupling periods are increased synchronously, as described in section 1.2.8. The intensity of the 61 13 C peak as a function of the recoupling period can be fitted to simulation to extract the relative oritentation of the said interactions. If this experiment is to be performed on a polypeptide, an additional dimension is necessary for spectral assignment. This can be achieved by introducing a chemical shift evolution period on 15 N (Figure 3.5). This frequency-labeled magnetization can then be transferred to carbonyl before 13 C-13C recoupling and detection. This signal after double fourier transform will result in a 2D 13 C-15N correlation spectra. A series of such spectra with synchronously varying recoupling periods can be collected, where the intensity variation of each peak can be fitted to simulations. Figure 3.5: Schematic representation for same experiment with increased dimensionality 3.2.1 Choice of C-C recoupling technique As mentioned in section 3.1.2, out of four homonuclear dipolar recoupling experiments, DRAWS, MELODRAMA, POST-C7 and SPC-5, the one best suited for this experiment was to be chosen. For this, each of these experiments was simulated using 62 SPINEVOLUTION34. Each simulation was performed on both trans and cis geometry. The geometry was defined by Cartesian coordinates of the four nuclei ( αCn+1, Nn+1, Cn’, and α Cn), obtained by generating the glycylglycine with trans and cis geometry, respectively, in PyMol47. The simulations were performed with magnetic field strength 500 MHz 1H larmor frequency and 11.904 kHz magic angle spinning (so that the rotor period was 84 µs). The results from the simulations are shown in Figure 3.6, (a)-(d) represent, respectively, DRAWS, MELODRAMA, POST-C7 and SPC-5 correlated evolution with REDOR. As can be seen, DRAWS (Figure 3.6a) yields little difference in the dephasing between trans and cis geometry. MELODRAMA (Figure 3.6b) yields some visible change, and the key feature of the expected difference in the trajectory is already present here. The trajectory of the trans geometry, after the initial decay, becomes flat during the region of ~ 1 ms – 1.5 ms of evolution time, whereas the same for cis decays steadily till ~ 3 ms. This feature is more dominant with POST-C7 and SPC-5 (Figures 3.6c and 3.6d respectively). Here, from the region of ~ 0.5 ms – 1 ms of evolution time, the trans trajectory flattens out while the cis trajectory shows a minima. While this is only a preliminary result on which more work needs to be done (described in section 3.2.2), this shows the direction in which to proceed. Although both POST-C7 and SPC-5 yielded trajectories that looked similar, POST-C7 was chosen as it had a higher recoupling efficiency compared to SPC-5. 63 3.2.2 Choice of ratio of mixing times The next factor that had to be fixed was the ratio of mixing times that was to be used. As was described in section 1.2.8, experiments for measuring torsional angles, in the most general form, are 4-dimensional experiments (three dimensional if we omit the chemical shift evolution dimension, as in this case). This would be the case if the two anistropic evolutions were allowed to evolve independently. However, to cut down on the time requirements, we vary the two dephasing periods synchronously. This means that the two mixing times can only be varied at a fixed ratio. This immediately raises a question as to what should this ratio be. Figure 3.6: Simulations for the choice of 13C-13C recoupling element. (a) DRAWS (b) MELODRAMA (c) SPC-5 (d) POST-C7 64 Such correlation experiments are most sensitive when maximum interference between the two evolutions can be achieved. This happens if both the interactions dephase to their respective minima at the same time. Figure 3.7 shows the simulated dephasing curve of both REDOR and POST-C7 performed on the same molecular fragment under similar conditions (αC-N-C’-αC fragment of trans glycylglycine, 500 MHz 1H larmor frequency, and 11.904 kHz magic angle spinning). Figure 3.7: Simulated dephasing of REDOR and SPINEVOLUTION38 65 POST-C7. Simulation done using As can be seen, POST-C7 decays to its minima much faster than REDOR. This is expected, as 13 C-13C dipolar coupling is much stronger compared to 13 C-15N dipolar coupling (which, once again, can be related to the gyromagnetic ratios of the two nuclei). Intuitively, it appears that allowing REDOR to evolve for twice as long as POSTC7 would result in the maximum interference. Indeed, in the plots shown in Figure 3.6, REDOR mixing period was twice as much as POST-C7 mixing period. To verify if this really is the most sensitive ratio, two more simulations were performed, one where REDOR mixing period was kept equal to POST-C7 mixing period, and another where REDOR mixing period was kept half of POST-C7 mixing. The simulated dephasing patterns are shown in Figure 3.8. It is to be noted that the X-axis has been labeled by POST-C7 mixing time in all the three plots shown. As can be seen, the most sensitive experiment is the one in which REDOR mixing period is kept twice that of POST-C7 mixing period. However, this is a parameter that is not too difficult to vary while running the actual experiments, and might be worthwhile to actually verify these predictions experimentally. At this point, we wanted to verify if we could, in addition to distinguishing cis and trans peptide bonds, also measure deviations from planarity of a peptide. To this end, similar experiment was simulated (with REODR mixing period kept at twice POST-C7 mixing period) on different geometries. Specifically, the αC-N-C’-αC torsional angle was varied from 180º (completely planar) to 150º (30º deviation from planarity), in steps of 10º. The simulated curves are shown in Figure 3.9 66 Figure 3.8: Simulation for the choice of ratio of REDOR and POST-C7 mixing time. (a) REDOR mixing time = POST-C7 mixing time (b) REDOR mixing time = 2 * POST-C7 mixing time (c) REDOR mixing time = 0.5 * POST-C7 mixing time 67 Figure 3.9: Sensitivity of simulated dephasing to variation of ω-torsional angle As can be seen from the simulations, sensitivity of the experiment to small deviations of the ω-torsinal angle is insignificant. This means, with the current experiment, it would be unreasonable to hope to measure deviations from planarity of the peptide bond. However, other modifications of this experiment are possible, and are discussed in the following section, which might improve sensitivity. 3.3 Utilizing carbonyl CSA for ω-angle measurement As discussed in section 1.2.5.1, chemical shift has an anisotropic component, called chemical shift anisotropy (CSA). In principle, this interaction should also be exploitable for torsional angle measurements. We shall focus on CSA of the carbonyl carbon as its 68 magnitude is much higher than any of the other nuclei relevant here. For our work, it can be seen that the relative orientation of the carbonyl CSA with respect to the αC-N bond is changed when the ω-torsinal angle changes. Hence, it should be possible, in theory, to correlate αC-N dipolar coupling evolution with carbonyl CSA to obtain information about the peptide bond geometry. Schematic representation of the pulse sequence for correlating αC-N dipolar coupling with carbonyl CSA is shown in Figure 3.10 Figure 3.10: Schematic representation of the pulse sequence for measuring the relative orientation of the αC-N bond with carbonyl CSA 69 This is similar to the pulse sequence shown in Figure 3.4 with the C-C recoupling period being replaced by the CSA recoupling period. Note that the magnetization prior to this second recoupling period has been transferred selectively to the carbonyl carbons. Chemical shift anisotropy is removed under the effect of magic-angle spinning. It can be reintroduced by effect of rf pulses, as described by Chan and Tycko 46. It has to be kept in mind that when simulating CSA, the CSA parameters, i.e. the anisotropy (δ aniso) and asymmetry (η) must be known a priori, along with the set of (α, β, γ) angles that define the orientation of the CSA with respect to the laboratory frame. For the present work, these values were taken from the work of Karlsson et. al48. They reported CSA parameters for six compounds. The compound they label as compound V, a tripeptide AGG, is similar to glycylglycine, the model compound that has been used in this work. They have reported the CSA parameters for the carbonyl of the Glycine at the center (denoted as ‘k’ in their paper). Using these values, the ROCSA dephasing was simulated in a similar manner. The simulated curve is shown in Figure 3.11. The feature that attracts attention here is the timescale of the decay. As can be seen, the magnetization decays in ~ 0.25 ms. This is almost an order of magnitude shorter than the timescale of decay of the REDOR experiment, ~ 1.5 ms (see Figure 3.2 and Figure 3.7). 70 Figure 3.11: Simulated decay curve for the ROCSA experiment46 As mentioned earlier, the experiment performs the best if the decay times of the two sequences are very similar. While in principle this technique could still be made to work by keeping the REDOR mixing time 10 times of the ROCSA mixing time, it should also be kept in mind that after 10 REDOR cycles most of the magnetization will decay due to relaxation processes which have not been considered in the simulations. We speculate that this problem would be circumvented if 15N-1H dipolar coupling was used instead of 13C-15N dipolar coupling. Relative orientation of 15N-1H internuclear vector with carbonyl CSA is affected in the same manner as the relative orientation of the C-N internuclear vector with carbonyl CSA. However, the additional advantage here is that 15 N-1H dipolar recoupling decays in the same timescale as carbonyl CSA, as can be 71 seen from Figure 3.3 (Figure 2 of reference 39). A schematic representation of the proposed experiment is shown in Figure 3.12. Figure 3.12: Schematic representation of the pulse sequence for measuring the relative orientation of 15 N-1H internuclear vector and carbonyl CSA 72 Chapter 4 Preliminary results This section shall describe some of the preliminary results obtained from solid-state NMR experiments on the model compounds. While the complete picture is has still proved elusive, results from some control experiment are worthwhile and give a clear direction for progress. All experiments described here were performed under magnetic field strength of 500 MHz 1H larmor frequency. Magic angle sample spinning rate was 11.904 kHz (rotor time of 84 µs). 4.1 Solid-state 1D CP-MAS experiments The isotopically labeled model compounds with trans- and cis-peptide bonds were synthesized, purified, and crystallized, as described in Chapter 2. Both of these crystals were packed into Varian T3 HXY 3.2 mm MAS probe by centrifugation. As a first step, 1D 13 C and 15 N solid-state CP-MAS spectra was obtained for both the samples. These spectra are shown in Figures 4.1-4.4. As can be seen, 1D 15 N spectra of Glycylglycine (Figure 4.2) shows an extra peak, probably arising from some impurity (which is presumably glycine, as explained in section 2.1.1). However, this does not pose a significant hindrance for the purposes of this experiment. This is because, all the pulse sequences proposed for this experiment involves CP transfers from 13Cα to 15N and then from 15N to 13C’. This is possible only in molecules containing a αC-N-C’ bond (peptide bond). Thus, only signal that would be seen in this experiment would arise out of compound(s) containing at least one peptide bond. Hence, small Glycine contamination is insignificant. 73 Also to be noted are the spinning sidebands of the carbonyl peak in 1D 13 C CP- MAS spectrum, for both Glycylglyine (Figure 4.1) and 2,5-diketopiperazine (Figure 4.2). This is due to high chemical shift anisotropy of carbonyl carbons, which is not completely removed at 11.904 magic-angle spinning. Spinning sidebands can be easily recognized as they are separated from the central peak (in frequency units), on either side, by an integral multiple of the spinning frequency. The spinning sidebands seen here are separated by 11.904 kHz (n=1 sideband). Although higher spinning speeds can eliminate such spinning sidebands, our experimental requirements do not allow us to spin our sample too high, as discussed in section 3.1.2. However, these sidebands are easy to detect and low in intensity, and hence does not cause too much of a hindrance. Figure 4.1: 1D 13C CP-MAS spectrum of (U-13C2, 15 MHz and 11.904 kHz magic-angle-spining 74 N) Glycylglycine at magnetic field strength 500 Figure 4.2: 1D 15N CP-MAS spectrum of (U-13C2, 15 N) Glycylglycine at magnetic field strength 500 MHz and 11.904 kHz magic-angle-spinning Figure 4.3: 1D 13C CP-MAS spectrum of (U-13C2, 15N) 2,5-diketopiperazine at magnetic field strength 500 MHz and 11.904 kHz magic-angle-spinning 75 Figure 4.4: 1D 15N CP-MAS spectrum of (U-13C2, 15N) 2,5-diketopiperazine at magnetic field strength 500 MHz and 11.904 kHz magic-angle-spinning The next step to proceed with the proposed experiment was to optimize the CP conditions for 13C to 15N and then from 15N selectively to 13C (carbonyl). The frequency offset of the spin-lock pulse for the selective CP and the phase cycling had to be set such that only carbonyl peak was seen in the 1D 13C spectrum. This was obtained for Glycylglycine, as shown in Figure 4.5. Note that the singal-to-noise is poorer (compared to Figure 4.1), as is expected, as there is some signal loss during CP transfer. Similar 1D 13C spectrum was obtained for 2,5-diketopiperazine (not shown). 76 Figure 4.5: 1D 13 C spectrum of Glycylglycine, after CP transfers to 15 N and selectively to 13 C (carbonyl) Intensity of this carbonyl peak can now be followed as a function of REDOR mixing time or POST-C7 mixing time. Once both the experiments have been calibrated, the two elements (REDOR and POST-C7 respectively) can be put together to run the complete sequence. 4.2 REDOR experiments The REDOR experiment was performed on both the model compounds, along with the two cross-polarization transfers, the conditions for which had already been optimized. The intensity of the carbonyl peak (Figure 4.5) was followed as a function of REDOR 77 mixing time. This dephasing pattern, along with simulated REDOR curve is shown in Figure 4.6 (Glycylglycine) and Figure 4.7 (2,5-diketopiperazine). For the simulation (performed using SPINEVOLUTION38, 13 C-15N distance was set to 1.35 Å for both the molecules for best fitting. Note that in these simulations, relaxation has not been included. Yet, the fitting with experimental data is quite good, as can be seen from the plots (Figures 4.6 and 4.7). Figure 4.6: Experimental (red dot) and simulated (blue line) REDOR dephasing pattern of Glycylglycine. Simulation done with 13C-15N distance set to 1.35 Å 78 Figure 4.7: Experimental (red dot) and simulated (blue line) REDOR dephasing pattern of 2,5diketopiperazine. Simulation done with 13C-15N distance set to 1.35 Å 79 Chapter 5 Conclusions and future work Accurate determination of protein structure is an important aspect of describing the biological function of a protein. The three dimensional structure of a protein backbone is completely described by the various torsional angles. This can be realized by considering the protein backbone as a chain of N atoms, which has 3N-6 degrees of freedom. These are N-1 bond lengths, N-2 bond angles, and N-3 torsional angles. While bond lengths and angles mostly tend to be fixed, torsional angles can vary widely, giving each protein its structure. For a peptide with M residues, there are 3(M-1) torsional angles, with three torsional angles (φ, ψ, and ψ) for every pair of residues. While φ and ψ torsional angles have been extensively studied, relatively less is known about the ω torsional angle, the one governing the geometry of the peptide bond. When making structural models, the peptide bond is assumed to be planar, specifically, in a trans-like conformation, unless known to be otherwise. With no general experimental method to distinguish cis- and trans-peptide bonds, this is a serious shortcoming of protein structure determination. This study is aimed at describing a solid-state NMR spectroscopic experiment which is sensitive to the conformation about the peptide bond. Section 1.2.8 describes the fundamental principle behind such a study, which involves correlating two anisotropic interactions under magic-angle spinning. Such correlations are known to be sensitive to the relative orientation of the two interactions. Chapter 3 describes the specific experiments chosen for this study, with simulation results showing that it would be possible for such experiments to differentiate between model compounds with trans and 80 cis peptide bonds. Chapter 2 has described the preparation of the model compounds with high purity. Trans-peptide bond, realized with linear diglycine, was synthesized by solidphase peptide synthesis using standard protocols. Cis-peptide bond, realized with 2,5diketopiperazine, was synthesized using non-conventional microwave-assisted synthesis. Both of these compounds yielded spectra with reliable signal-to-noise, and produced REDOR decay patterns that agree well with simulation, as shown in Chapter 4. ` Preliminary results from the simulations show that the primary objective, to distinguish cis-peptide bonds from their trans counterparts, should be possible with the experiment described here. However, more work is necessary to achieve the farther goal, that of measuring deviation of a peptide bond from planarity. Simulations show this experiment to be insensitive to such small changes in the ω-torsional angle. Modifications are required to make the experiment suitable for this purpose. The immediate next step would be to extend the applicability of this experiment to a full protein. Potential systems to be studied by this method could be Barstar or Acquaporin, each of which have a native cis-peptide bond, and are known to give good quality solid-state NMR spectrum. In such a scenario, the experiment has to be modified to include an additional chemical shift evolution dimension on 15N. This would result in a 2D 13 C-15N spectrum, where each cross-peak is modulated by information about the geometry about the peptide bond. A series of such 2D spectra can be collected, and the trajectory of each peak can be fit to simulations to extract the geometry information about each peptide plane. Additionally, in such a scenario, the experiment needs to be compensated for the effects of side-chains, which were not present in the model compounds. Specifically, the 81 POST-C7 sequence is known to be sensitive to presence of 13 C in the side chains. A potential solution could be introducing SEASHORE49 delays in between POST-C7 blocks. Another potential area of improvement is to increase the experiment sensitivity, so as to detect small changes in the ω-torsional angle. Primarily, sensitivity around ωangle of 180º is desired, as most peptide bonds exist in the trans-like conformation. One option might be the use of carbonyl CSA instead of 13C-13C dipolar coupling. Preliminary simulations and experiments (not shown) show that carbonyl CSA decays the magnetization much faster (~ 0.2 ms) than the 13 C-15N dipolar coupling evolves (~ 1.5 ms). This problem may potentially be solved by using 15N-1H dipolar coupling instead of 13 C-15N dipolar coupling. 15N-1H dipolar coupling, reintroduced by T-MREV39, is known to decay the magnetization in ~ 0.5 ms (Figure 3.3, taken from reference 39). Correlated evolution of these two interactions, in a ratio to be determined by simulations, could be expected to yield a more sensitive version of this experiment. 82 References 1. Corey, R.B.; Pauling L., Proc. Roy. Soc. ser B, 1953, 141, 10. 2. Ramachandran, G.N., Biopolymers, 1968, 6, 1494. 3. Ramachandran, G.N.; Lakshminarayanan, A.V.; Kolaskar, A.S., Biochim. Biophys. Acta, 1973, 303, 8. 4. Head-Gordon, T.; Head-Gordon, M.; Frisch, M.J.; Brooks, C.L., III; Pople, J.A., 1991, Journ. Am. Chem. Soc., 113, 5989. 5. MacArthur, M.W.; Thornton, J.M., Journ. Mol. Biol., 1996, 264, 1180. 6. Edison, A.S., Nature, 2001, 8, 201. 7. Pal, D.; Chakrabarti, P., Journ. Mol. Biol., 1999, 294, 271. 8. Weiss, M.S.; Jabs, A.; Hilgenfeld, R., Nat. Struct. Biol, 1998, 5, 676. 9. Weiss, M.S.; Hilgenfeld, R., Biopolymers, 1999, 50, 536. 10. Dorman, D.E.; Torchia, D.A.; Bovey, F.A., Macromolecules, 1973, 6, 80. 11. Droman, D.E.; Bovey, F.A., Journ. Org. Chem., 1973, 38, 2379. 12. Sarkar, S.K.; Young, P.E.; Sullivan, C.E.; Torchia, D.A., Proc. Nat. Acad. Sci., 1984, 81, 4800. 13. Herzberg, O.; Moult, J., Proteins, 1991, 11, 223. 14. Weiss, M.S.; Metzner, H.J.; Hilgenfeld, R., FEBS Lett, 1998, 423, 291. 15. Stoddard, B.L.; Pietrokovski, S., Nat. Struct. Biol., 1998, 5, 3. 16. Spencer, R.G.S.; Halverson, K.J.; Auger, M.; McDermott, A.E.; Griffin, R.G.; Lansbury, P.T. Jr., Biochemistry, 1991, 30, 10382. 83 17. Lansbury, P.T. Jt,; Costa, P.R.; Griffiths, J.M.; Simon, E.J.; Auger, M.; Halverson, K.J.; Kocisko, D.A.; Hendsch, Z.S.; Ashburn, T.T.; Spencer, R.G.S.; Tidor, B.; Griffin, R.G., Nat. Struct. Biol., 1995, 2, 990. 18. Jeener, J. Lecture at International Ampere Summer School, Basko Polje, Yugoslavia. In 1971. 19. Aue, W.P.; Bartholdi, E.; Ernst, R.R., Journ. Chem. Phys., 1976, 64, 2229. 20. Andrew, E.R.; Bradbury, A.; Eades, R.G., Nature, 1958, 182, 1659. 21. Lowe, I.J., Phys. Rev. Lett., 1959, 2, 285. 22. Feng, X.; Lee, Y.K.; Sandstrom, D.; Eden, M; Maisel, H.; Sebald, A.; Levitt, M.H., Chem. Phys. Lett, 1996, 257, 314. 23. Schmidt-Rohr, K.; Journ. Am. Chem. Soc., 1996, 118, 7601. 24. Feng, X.; Verdegem, P.J.E.; Lee, Y.K.; Sandstrom, D.; Eden, M; Bovee-Geurts, P.; de Grip, W.J.; Lugtenberg, J.; de Groot, H.J.M.; Levitt, M.H., Journ. Am. Chem. Soc., 1997, 119, 6853. 25. Hong, M.; Gross, J.D.; Griffin, R.G., Journ. Phys. Chem. B, 1997, 101, 5869. 26. Costa, P.R.; Gross, J.D.; Hong, M.; Griffin, R.G., Chem. Phys. Lett., 1997, 280, 95. 27. Feng, X.; Eden, M.; Brinkmann, A.; Luthman, H.; Eriksson, L.; Graslund, A.; Antzutkin, O.N.; Levitt, M.H., Journ. Am. Chem. Soc., 1997, 119, 12006. 28. Hong, M.; Gross, J.D.; Hu, W.; Griffin, R.G., Journ. Magn. Reson., 1998, 135, 159. 29. Reif, B.; Hohwy, M.; Jaroniec, C.P.; Rienstra, C.M.; Griffin, R.G., Journ. Magn. Reson., 2000, 145, 132. 30. Ladizhansky, V.; Veshtort, M.; Griffin, R.G., Journ. Magn. Reson., 2002, 154, 317. 84 31. Riesntra, C.M.; Hohwey, M.; Mueller, L.J.; Jaroniec, C.P.; Reif, B.; Griffin, R.G., Journ. Am. Chem. Soc., 2002, 124, 11908. 32. Goodfellow, V.S.; Laudeman, C.P.; Gerrity, J.I.; Burkard, M.; Strobel, E.; Zuzack, J.S.; McLeod, D.A., Molecular Diversity, 1996, 2, 97. 33. Szardenings, A.K.; Burkoth, T.S., Tetrahedron, 1997, 53, 6573. 34. Osapay, G.; Bouvier, M.; Taylor, J.W., Techniques in Protein Chemistry II, Academic: San Diego, 1991, pp 221-231. 35. Smith, R.A.; Bobko, M.A.; Lee, W., Bioorg. Med. Chem. Lett., 1998, 8, 2369. 36. Nonappa, K.A.; Lahtinen, M.; Kolehmainen, E., Green Chemistry, 2011, 13, 1203. 37. Gullion, T.; Schaefer, J., Journ. Magn. Reson., 1989, 81, 196. 38. Veshtort, M.; Griffin, R.G.; Journ. Magn. Reson., 2006, 178, 248. 39. Hohwy, M.; Jaroniec, C.P.; Reif, B.; Rienstra, C.M.; Griffin, R.G., Journ. Am. Chem. Soc., 2000, 122, 3218. 40. Tycko, R.; Dabbagh, G., Chem. Phys. Lett., 1990, 173, 461. 41. Sun, B.Q.; Costa, P.R.; Kocisko, D.; Lansbury, P.T.; Griffin, R.G., Journ. Chem. Phys., 1995, 102, 702. 42. Gregory, D.M., Mol. Phys., 1996, 89, 1835. 43. Lee, Y.K.; Kurur, N.D.; Helmle, M.; Johannessen, O.G.; Nielsen, N.C.; Levitt, M.H., Chem. Phys. Lett., 1995, 242, 304. 44. Hohwy, M.; Jakobsen, H.J.; Eden, M; Levitt, M.H.; Nielsen, N.C., Journ. Chem. Phys., 1998, 108, 2686. 45. Hohwy, M.; Rienstra, C.M.; Jaroniec, C.P.; Griffin, R.G., Journ. Chem. Phys., 1999, 110, 7983. 85 46. Chan, J.C.C.; Tycko, R., Journ. Chem. Phys., 2003, 118, 8378. 47. The PyMOL Molecular Graphics System, Version 1.5.0.4 Schrödinger, LLC. 48. Karlsson, T.; Popham, J.M.; Long, J.R.; Oyler, N.; Drobny, G.P., Journ. Am. Chem. Soc., 2003, 125, 7394. 49. Paravastu, A.K.; Tycko, R., Journ. Chem. Phys., 2006, 124, 194303. 50. Helmus, J., 2011, Electronic Thesis or Dissertation. Retrieved from https://etd.ohiolink.edu/ 86
© Copyright 2025 Paperzz