SINGLE-MOLECULE AND IMAGING STUDIES OF -UNFOLDING CONFORMATIONAL DYNAMICS: THE MULTIPLE-STATE AND MULTIPLE-CHANNEL LANDSCAPE

Zijian Wang

A Dissertation

Submitted to the Graduate College of Bowling Green State University in partial fulfillment of the requirements for the degree of

DOCTOR OF PHILOSOPHY

May 2016

Committee:

H. Peter Lu, Advisor

Gabriela Bidart-Bouzat Graduate Faculty Representative

Ksenija D. Glusac

George Bullerjahn

© 2016

Zijian Wang

All Rights Reserved iii ABSTRACT

H. Peter Lu, Advisor

Protein conformational dynamics often plays a critical role in protein functions. We have characterized the spontaneous folding-unfolding conformational fluctuation dynamics of calmodulin (CaM) at thermodynamic equilibrium conditions by using single-molecule resonance energy transfer (FRET) spectroscopy. We studied protein folding

dynamics under simulated biological conditions to gain a deep, mechanistic understanding of

this important biological process. We have identified multiple folding transition pathways and

characterized the underlying of the single-molecule protein conformational

fluctuation trajectories. Our results suggest that the folding dynamics of CaM molecules

involves a complex multiple-pathway multiple-state energy landscape, rather than an energy

landscape of two-state dynamical process. Our probing single-molecule FRET fluctuation

experiments demonstrate a new approach of studying spontaneous protein folding-unfolding

conformational dynamics at the equilibrium that features recording long time single-molecule

conformational fluctuation trajectories. This technique yields rich statistical and dynamical

information far beyond traditional ensemble-averaged measurements.

We characterize the conformational dynamics of single CaM interacting with C28W. The single CaM molecules are partially unfolded by GdmCl, and the folded and unfolded CaM molecules are approximately equally populated. Under this condition, the majority of the single protein CaM undergoes spontaneous folding-unfolding conformational fluctuations. Using single molecule FRET spectroscopy, we study each of the single protein’s conformational dynamics in iv the presence of C28W-CaM interactions. The results show an interesting folding-upon-binding

dynamic process, and a conformational selection mechanism is further confirmed.

The effect of molecular crowding on protein folding process is a key issue in the understanding of protein folding dynamics in living cells. Due to the complexity and interplay between various interactions existing in an equally favored environment of protein folding and unfolding conformational dynamics, such simple reduced entropic enhancement model do not suffice in describing protein folding conformational dynamics. We observe, at higher concentration of crowding reagent Ficoll 70, single protein molecules spontaneously denature into unfolded proteins which involves a combined process of polymer-polymer interaction, entropic effects and solvation thermodynamics and dynamics. Such heterogeneous unfolding process can serve as a first step to a mechanistic understanding of living cell disease as a result of molecular crowding effect, protein aggregates and fibril formation.

v

To My Mom

Whose Insights Influence Me the Most vi ACKNOWLEDGMENTS

First and foremost I want to thank my advisor H. Peter Lu. It has been an honor to be one of

his graduate students. He has taught me, both consciously and unconsciously, how good

experiments are done. I appreciate all his contributions of time, ideas and funding to make my

Ph. D. experience productive. I am also grateful for the excellent example he has provided as a

successful chemist and professor.

I am thankful to all my dissertation committee members: Dr. Ksenija D. Glusac, Dr. George

Bullerjahn and Dr. Gabriela Bidart-Bouzat for their precious time. I also want to acknowledge all

the group members, both current members and past members from Dr. H. Peter Lu’s group, for

setting a competitive, productive and hard-working environment. I also want to thank Dr.

Desheng Zheng, Dr. Yufan He, Dr. Yuanmin Wang, Dr. Jin Cao and Dr. Qing Guo for their

friendship and help. Especially, I’d like to thank Dr. Takashige Fujiwara for training me on tuning Lasers and technical support.

I gratefully acknowledge Delta Electronics Inc. for generously providing me the Delta

Electronics Research fellowship. The Delta Electronics Research fellowship supported me for the 2012-2013 academic year.

I would also like to thank many faculty and staff members at the Center for Photochemical

Sciences and the Department of Chemistry: Nora Cassidy, Alita Frater, Charles Codding, Doug

Martin, and Hilda Miranda, for their help.

Lastly, I would like to thank my family for all their love and encouragement, for my parents who raised me with a love of science and supported me in all my pursuits. Without their understanding and love, I would never be able to finish this journey.

vii

TABLE OF CONTENTS

Page

CHAPTER I. INTRODUCTION……………………… ...... 1

1.1 Introduction to Single-Molecule Fluorescence Spectroscopy...... 1

1.2 Single-Molecule Studies of Protein Conformational Dynamics ...... 3

1.3 An Overview of Single-Molecule Studies of Protein Folding ...... 4

1.4 Our Sample Protein Calmodulin (CaM) ...... 7

1.5 Research Objective, Specific Aims and Dissertation Overview ...... 9

1.6 Reference ...... 11

CHAPTER II. EXPERIMENTAL SECTION ...... 18

2.1 Principles of Experimental Techniques ...... 18

2.1.1 Principles of Confocal Microscopy ...... 18

2.1.2 Principles of Fluorescence Resonance Energy Transfer ...... 20

2.1.3 Signal Detection Techniques: Introduction to APD ...... 25

2.2 Experimental Details ...... 26

2.2.1 Experimental Setup of Single-Molecule FRET combined with Confocal

Microscope...... 26

2.2.2 Materials and Sample Preparation……………………………… ...... 29

2.2.3 Statistical Analyses of Single-Molecule Intensity Trajectories ...... 32

2.3 Protein Folding Dynamics in Condensed Phases...... 35

2.4 Theory of Polymer Solutions ...... 37

2.5 References ...... 38 viii

CHAPTER III. PROBING SINGLE- MOLECULE PROTEIN FOLDING

CONFORMATIONAL DYNAMICS USING SINGLE-MOLECULE FRET

SPECTROSCOPY………………………...... 41

3.1 Introduction ...... 41

3.2 Materials and Methods ...... 43

3.2.1 Sample Preparation and Characterization…………………………… . 43

3.2.2 Single-Molecule Imaging and FRET Measurements………………… 45

3.3 Results and Discussion ...... 47

3.3.1 Single-Molecule FRET Trajectories Monitored Unfolding of Single CaM

molecules…………………………… ...... 47

3.3.2 Autocorrelation Analyses of CaM Folding-Unfolding Conformational

Fluctuation Dynamics …………………………… ...... 50

3.3.3 Non-Exponential Distribution of Folding Waiting Time Indicates Multiple

Folding Intermediates …………………………… ...... 58

3.3.4 Model Analyses of Conformational Dynamics and Energy Landscape of

Single-Molecule CaM Folding …………………………… ...... 61

3.4 Conclusions ...... 69

3.5 References ...... 70

CHAPTER IV. PROBING SINGLE-MOLECULE PROTEIN FOLDING-UNPON-BINDING

CONFORMATIONAL DYNAMICS USING SINGLE-MOLECULE FRET SPECTROSCOPY.

...... ………………………...... 80

4.1 Introduction ...... 80

4.2 Materials and Methods ...... 83 ix

4.2.1 Sample Preparation and Characterization…………………………… . 83

4.2.2 Single-Molecule Imaging and FRET Measurement ...... 84

4.3 Results and Discussion ...... 85

4.4 Conclusions ...... 92

4.5 References ...... 93

CHAPTER V. PROBING SINGLE-MOLECULE PROTEIN FOLDING CONFORMATIONAL

DYNAMICS IN CROWDED EVIRONMENT USING SINGLE-MOLECULE FRET

SPECTROSCOPY ………………………… ...... ………………………………… 98

5.1 Introduction ...... 98

5.2 Materials and Methods ...... 99

5.3 Results and Discussion ...... 102

5.4 Conclusions ...... 114

5.5 Reference ...... 115

x

LIST OF FIGURES

Figure Page

1.1 Conceptual Figure of FRET ...... 3

1.2 Flexible Conformations of a Protein ...... 4

1.3 Protein Folding Energy Landscape ...... 7

1.4 3D Crystal Structure of CaM ...... 9

2.1 Conceptual Figure of Confocal Microscope ...... 20

2.2 FRET Efficiency vs Inter-Dye Distance Curve ...... 21

2.3 Conceptual Figure of Protein Folding FRET ...... 24

2.4 Conceptual Figure of FRET ...... 25

2.5 A Schematic Representation of APD ...... 26

2.6 Our Single-Molecule Setup...... 28

2.7 A Cartoon of Our Sample ...... 32

2.8 TCAD Map in a Typical Single-Molecule Experiment ...... 34

2.9 A Schematic Representation of Molecular Crowding ...... 38

3.1 Our Single-Molecule Imaging System...... 46

3.2 Typical Trajectories and Data ...... 48

3.3 Correlation Analyses of Single-Molecule Data ...... 52

3.4 Single-Molecule Conformational Fluctuation Dynamics ...... 54

3.5 Waiting-Time Distribution of Single-Molecule Data ...... 60

3.6 Protein Folding Pathway Distribution ...... 66

4.1 Protein Folding Binding Process ...... 82

4.2 Single-Molecule Imaging and Correlation Analysis...... 86 xi

4.3 Protein Folding Binding and EFRET ...... 87

4.4 Waiting Time Distribution Analysis of Protein Folding ...... 89

4.5 Folding Pathway Distribution in Protein Folding Binding ...... 91

5.1 Single-Molecule Imaging and Analysis ...... 103

5.2 EFRET Distributions ...... 104

5.3 Correlation Analyses ...... 106

5.4 Folding Pathway Distribution in Protein Folding ...... 112

5.5 A Cartoon Showing Folding Process ...... 113

xii

LIST OF TABLES

Table Page

2.1 List of concentrations used in our experiments ...... 30

2.2 List of concentrations used in our folding-binding experiments ...... 31

4.1 List of concentrations used in our folding-binding experiments ...... 84

1

CHAPTER I. INTRODUCTION

This chapter is dedicated to the introduction of single-molecule studies of protein

conformational dynamics and a brief introduction of single-molecule protein folding.

1.1 Introduction to Single-Molecule Fluorescence Spectroscopy

Single-molecule spectroscopy developed in the 1990s is by far the most important technique

to study one molecule at a time.1-13 The dynamics and heterogeneity revealed by single-molecule

experiments proved to be invaluable, when researchers try to understand biological processes at

the molecular level. Single-molecule fluorescence spectroscopy is powerful for studying

complex biological processes, such as enzymatic reactions14, 15, protein folding dynamics,16-18

and other dynamic processes.19 Single-molecule fluorescence imaging is a powerful tool to

probe details in living cells and nanostructures.20-22 Single-molecule fluorescence resonance

energy transfer (smFRET) spectroscopy probes conformational changes of dye labeled individual biomolecules with a sensitivity down to about 1 nm to 8 nm spatial range and sub-millisecond temporal resolution,23, 24 being an ideal approach to analyze protein conformational dynamics.

Single-molecule FRET spectroscopy can monitor conformational changes of macromolecules

containing fluorophores. Depending on the photophysics properties of the fluorophores,

single-molecule FRET spectroscopy typically can reveal detailed molecular scale dynamics and the conformational fluctuation information of biomolecules.25 New developments in

single-molecule spectroscopy revolutionized the study of traditional biophysics. Recently, new

developments in this technique also made it possible to probe various interaction pathways in

which real biological processes occur.26 The experimental output of a single-molecule FRET 2

experiment is typically a photon trajectory containing rich fluctuation dynamic information which cannot be straightforwardly obtained by ensemble-averaged experiments. By removal of

ensemble average, one can analyze detailed molecular scale dynamics from such experiments. Of particular relevance to my research is the observation of single-molecule trajectories which record dynamic events of individual systems in condensed phases. Motivated by the experimental developments, theorists have analyzed intermittency, interconversion between two conformational states, and photon statistics.

Single-molecule fluorescence resonance energy transfer (smFRET), which applies FRET at the single-molecule level, is a measurable technique for studying real-time structural dynamics of individual biomolecules. It measures distance typically in 10-80 Å range. Generally speaking, a pair of donor and acceptor is attached to two specific sites of one or two target molecules, and energy transfer from donor to acceptor occurs through non-radiative induced dipole-dipole interaction energy transfer mechanism. The energy transfer efficiency (EFRET) between the two fluorophores depends on the inter-distance between the donor fluorophore and acceptor fluorophore attached to the same molecule:

1 EFRET  6 1+rR / 0 

where r is the separate distance between two fluorophores, the FRET donor and acceptor,

respectively. R0, the Förster radius, is a function of dipole orientations of two fluorophores,

refractive index of the medium, spectral overlap integral between donor emission and acceptor

absorbance, and quantum yield of the donor. 3

Figure 1.1. Conceptual Figure of FRET. Förster resonance energy transfer or fluorescence resonance energy transfer (FRET) Jablonski diagram.

1.2 Single-Molecule Studies of Protein Conformational Dynamics

Proteins participate in various processes inside living cells such as metabolism, gene

expression, cell signaling, and protein-protein interactions. A mechanistic understanding of protein conformation, structure and function is critical in biophysics studies. The structure and function relationship of proteins confused generations of scientists. Such structure and function relationship is critical to understand biochemical reactions and biological processes involved in living cells. Ensemble-averaged techniques such as X-ray diffraction and nuclear magnetic resonance studies only reveal averaged blurred physical pictures of the true protein motions.

Only single-molecule techniques can reveal the temporal and spatial heterogeneity of protein conformational motion and dynamics by removal of ensemble average.

One type of protein which is of critical importance is the flexible protein involved in various signaling processes inside living cells.24, 26 These proteins play the role of messengers delivering messages from side to side inside living cells by engaging themselves in extensive 4

protein-protein interactions. The nature of such protein-protein interactions is still a mystery and under extensive studies. Based on limited experimental data, various models have been proposed to understand the nature of such interaction processes such as induced fit and conformational

selection models. One of the best ways to resolve this puzzle is by removal of ensemble average

using single-molecule techniques. The rich information obtainable in detailed fluctuation

dynamic trajectories of single-molecule techniques can provide detailed mechanistic

understanding of critical questions of this kind.

Figure 1.2. Flexible Conformations of a Protein. A cartoon showing the flexible confirmations a

protein can take in real living cell environments where extensive protein-protein interactions

exist. The conformational fluctuation dynamic model is adapted to account for folding reaction

and binding flexibility.

1.3 An Overview of Single-Molecule Studies of Protein Folding

Over the last 50 years, there have been intensive studies on protein folding mechanisms and dynamics.27-30 A widely accepted perspective suggests that a protein folding process is essentially 5

a navigation on a funnel-shaped energy landscape towards a global energy minimum.31-33 This multi-dimensional energy landscape is typically rugged and complex involving multiple folding routes and metastable states. Single-molecule FRET is a powerful approach to reveal selectively the folded and unfolded protein conformational subpopulations from otherwise hidden in ensemble-averaged overall population distribution.34-36 Using smFRET, there are significant advances in analyzing the protein unfolded state dynamics,34, 35 and protein folding

dynamics.36 Generally, a two-state model with two distinct folded and unfolded conformations is sufficient to describe the protein folding dynamic processes. However, for large proteins with

more than 100 amino acids, this simple two-state folding dynamic scheme is often insufficient or

inapplicable.37-40

Since the question about the spontaneity of proteins folding into their via sampling of an enormous number of possible conformations first appeared 50 years ago, the

protein folding problem, although with significant progress, still remains to be a fascinating topic.

Theoretically, the protein folding process is navigation on a funnel-shaped energy landscape towards a global energy minimum. This multi-dimensional energy landscape can be rugged and extremely complex involving multiple folding routes and metastable states. Using smFRET, there are advances in characterization of protein unfolded state dynamics, protein collapse (from unfolded state to a compact yet folded state) dynamics, protein folding dynamics19 and ‘one-state’

downhill folding process.41-44 Usually, a two-state model with two distinct folded and unfolded conformations is sufficient to describe the dynamic processes. However, for large proteins with

more than 100 amino acids, this simple two-state folding dynamic scheme is not applicable. A 6

recent example is a demonstration of existence of six distinct folding intermediates of a 214

protein adenylate kinase.45

After gene expression, single proteins start to fold into a unique functional structure without

external help. However, inside the living cells, this process normally happens in a crowded

environment which involves various kinds of protein-protein interactions. Little is known about

the folding dynamics of single CaM molecules in the presence of extensive interaction with other peptide molecules. The motivation of our study is to probe the folding-unfolding conformational

dynamics, which is the most important conformational dynamics of single proteins in this

“crowded” environment in the presence of polymer molecules. The outcome of this study helps us to understand the critical role played by protein-protein interactions in the single protein folding process, which can provide insight into the future study of protein misfolding and protein aggregates related to human diseases (e.g., amyloid diseases) inside living cells.

Single-molecule techniques are powerful tools to probe molecular scale dynamic processes without ensemble averaging. The molecular details of single protein folding are readily

observed one molecule at a time in single-molecule experiments. As a key advantage,

single-molecule FRET allows a clear separation of folded and unfolded subpopulations, and thus

enables a further quantitative analysis of the properties of protein conformational fluctuation

dynamics without interference from undesired ensemble averaged signals. Since

single-molecule FRET can provide distance and protein conformational fluctuation dynamic

information without ensemble-averaging, it is promising to observe intramolecular

conformational dynamics at equilibrium. 7

Figure 1.3. Protein Folding Energy Landscape. A cartoon showing the existence of possible multiple pathways involved in the single-molecule protein folding process. It is consistent with the multidimensional funnel-shaped folding energy landscape physical picture which was theoretically proposed.

1.4 Our Sample Protein Calmodulin (CaM)

CaM, a 148-residue protein responsible for intracellular calcium-sensing, plays a crucial role

in a number of biological processes including cell signaling, muscle contraction, and energy

metabolism. CaM has two globular domains, and the conformational dynamics of the domains

has been under extensive studies.46-48 Traditional methods, such as nuclear magnetic resonance

and X-ray crystallography,49, 50 have provided detailed insights into the mechanisms and 8

dynamics of CaM conformational changes. However, the complexity of the protein dynamics, especially, the conformational fluctuations involved in CaM biological functions are still not fully characterized by using ensemble-averaged dynamic measurements or by static structural

analyses alone. CaM is capable of regulating activities of many proteins via interacting with Ca2+

ions. When the intracellular calcium level is high, four Ca2+ ions bind to the CaM in order to

reduce Ca2+ level. Calmodulin has four EF-hand motifs that change its conformation upon

binding Ca2+ ions. Each EF-hand motif contains two α-helices connected by a 12-residue peptide loop. The calcium ion changes the conformation of CaM by binding to the loop region and changes the relative positions of the α-helices. Upon binding Ca2+, CaM undergoes large

conformational changes which can be directly measured by single-molecule fluorescence

techniques in real-time. The crystal structure of Ca2+-loaded CaM exhibits a dumbbell shape

structure which forms an excellent prototype for single-molecule studies.

Single-molecule spectroscopy is capable of dissecting protein conformational fluctuations

under physiological conditions in real time. As a result of the flexibility and the dumbbell shape

of the protein CaM, it is possible to monitor the conformational dynamics of CaM with

single-molecule FRET spectroscopy. The typical time-resolution of single-molecule FRET spectroscopy is milliseconds, which are the characteristic time scales of protein motion. Our

work on probing the folding-unfolding conformational dynamics of CaM under denature

conditions of guanidinium chloride (GdmCl) by using single-molecule FRET spectroscopy is a significant step towards a mechanistic molecular scale understanding of CaM folding process. 9

Figure 1.4. 3D Crystal Structure of CaM. Three dimensional crystal structure of calmodulin

from protein data bank. The blue balls indicate the calcium ion. As we can see from the picture, calmodulin has two globular domains each consisting of two calcium ion binding sites.

C28W is a short peptide chain consisting of 28 amino acid residues. C28W is an oligomer

acting as the effective binding domain of the plasma membrane Ca-ATPase. The binding of

C28W to CaM can have significant effect on CaM protein conformations. A mechanistic

understanding of the binding induced conformational dynamic process is under extensive study.

In general, a two-state dynamic model involving loosely binding and tightly binding terminals is sufficient to describe such dynamic process. A main focus of our study is to further probe the role played by C28W-CaM protein-protein interactions in the CaM protein folding process.

1.5. Research Objective and Specific Aims, and Dissertation Overview

Our work is to probe the folding-unfolding conformational fluctuation dynamics of

CaM under denature conditions by using smFRET spectroscopy. The protein folding-unfolding

conformational dynamics is the essential part of this thesis study. We have identified the critical 10

concentration of GdmCl, at which single-molecule CaM undergoes spontaneous folding-unfolding conformational fluctuation with about equal probability of dwelling on the

folded and unfolded conformational states. Thermodynamically, such a condition is ideal for

studying equilibrium spontaneous fluctuations without external driving force. Recording

single-molecule conformational fluctuation trajectories and analyzing equilibrium fluctuation

dynamics, we are able to identify the nature of the CaM folding dynamics which involves

multiple pathways and multiple states. Using a dynamic model analysis, we have further

identified the distribution of the folding transition pathways and the energetic features of the

folding energy landscape of CaM.

We use single-molecule FRET spectroscopy to probe the folding-unfolding conformational

dynamics of CaM interacting with C28W under the mild unfolding concentration of denaturant

solvent guanidinium chloride (GdmCl). The critical concentration of denaturant GdmCl, at which

single protein undergoes folding-unfolding conformational fluctuation, has been characterized by

our previous study. Since the interaction characterized by previous studies revealed an induced

binding picture in the native state of CaM, we see a more complicated dynamic picture of CaM interacting with C28W under a partially denatured condition at which single CaM molecules

have larger conformational space to explore.51

In chapter five of this thesis, we observe, at higher concentration of crowding reagent Ficoll

70, single protein molecules spontaneously denature into unfolded proteins which are a combined process of polymer-polymer interactions, entropic effects and solvation thermodynamics. Such a heterogeneous unfolding process can serve as a first step to a 11

mechanistic understanding of living cell disease as a result of molecular crowding effects, protein aggregates and fibril formations.

1.6 Reference:

1 Nie, S.; Zare R. N. Optical Detection of Single Molecules. Annu. Rev. Biophys. Biomol.

Struct. 1997, 26, 567-596.

2 Moerner, W. E.; Orrit, M. Illuminating Single Molecules in Condensed Matter. Science 1999,

283 (5480), 1670-1676.

3 Betzig, E.; Chichester, R.J. Single Molecules Observed by Near-Field Scanning Optical

Microscopy. Science 1993, 262, 1422-1425.

4 Macklin, J. J.; Trautman, J. K.; Harris, T. D.; Brus, L. E. Imaging and Time-Resolved

Spectroscopy of Single Molecules at an Interface. Science 1996, 272, 255-258.

5 Xie, X. S.; Trautman, J. K. Optical Studies of Single Molecules at Room Temperature. Annu.

Rev. Phys. Chem. 1998, 49, 441-480.

6 Ishijima, A.; Yanagida, T. Single Molecule Nanobioscience. Trends Biochem. Sci. 2001, 26

(7), 438-444.

7 Weiss, S. Fluorescence Spectroscopy of Single Biomolecules. Science 1999, 283 (5408),

1676-1683.

8 Moerner, W. E. New Directions in Single-Molecule Imaging and Analysis. Proc. Natl. Acad.

Sci. U.S.A. 2007, 104 (31), 12596-12602.

9 Orrit, M. The Motions of an Enzyme Soloist. Science 2003, 302 (5643), 239-240. 12

10 Betzig, E.; Patterson, G.H.; Sougrat, R.; Lindwasser, O.W.; Olenych, S.; Bonifacino, J.S.;

Davidson, M.W.; Lippincott-Schwartz, J.; Hess, H.F. Imaging Intracellular Fluorescent

Proteins at Nanometer Resolution. Science, 2006, 313 ( 5793) 1642-1645.

11 Hell, S.W.; Wichmann, J. Breaking the diffraction resolution limit by stimulated emission:

stimulated-emission-depletion fluorescence microscopy. Opt. Lett. 1994, 19(11) 780-782.

12 Rust, M.J.; Bates, M.; Zhuang, X. Stochastic optical reconstruction microscopy (STORM)

provides sub-diffraction-limit image resolution. Nat. Methods. 2006, 3(10): 793–795.

13 Lu, H. P. Probing Single-Molecule Protein Conformational Dynamics. Acc. Chem. Res. 2005,

38 (7), 557-565.

14 Lu, H. P.; Xun, L. Y.; Xie, X. S. Single-Molecule Enzymatic Dynamics. Science 1998, 282

(5395), 1877-1882.

15 English, B. P.; Min, W.; van Oijen, A. M.; Lee, K. T.; Luo, G. B.; Sun, H. Y.; Cherayil, B. J.;

Kou, S. C.; Xie, X. S. Ever-Fluctuating Single Enzyme Molecules: Michaelis-Menten

Equation Revisited. Nat. Chem. Biol. 2006, 2 (2), 87-94.

16 Schuler, B.; Lipman, E. A.; Eaton, W. A. Probing The Free-Energy Surface for Protein

Folding with Single-Molecule Fluorescence Spectroscopy. Nature 2002, 419(6908), 743-747.

17 Rief, M.; Gautel, M.; Oesterhelt, F.; Fernandez, J.M.; Gaub, H.E. Reversible unfolding of

individual titin immunoglobulin domains by AFM. Science. 1997 276(5315):1109-12.

18 Rief, M.; Oesterhelt, F.; Heymann, B.; Gaub, H.E. Single Molecule Force Spectroscopy on

Polysaccharides by Atomic Force Microscopy. Science. 1997 275(5304): 1295-1297. 13

19 Lu, H. P.; Iakoucheva, L. M.; Ackerman, E. J. Single-Molecule Conformational Dynamics of

Fluctuating Noncovalent DNA-Protein Interactions in DNA Damage Recognition. J. Am.

Chem. Soc. 2001, 123 (37), 9184-9185.

20 Tan, X.; Nalbant, P.; Toutchkine, A.; Hu, D. H.; Vorpagel, E. R.; Hahn, K. M.; Lu, H. P.

Single-Molecule Study of Protein-Protein Interaction Dynamics in a Cell Signaling System. J.

Phys. Chem. B. 2004, 108 (2), 737-744.

21 Liu, S.; Bokinsky, G.; Walter, N. G.; Zhuang, X. Dissecting the Multistep Reaction Pathway

of an RNA Enzyme by Single-Molecule Kinetic "Fingerprinting". Proc. Natl. Acad. Sci.

U.S.A. 2007, 104 (31), 12634-12639.

22 Visscher, K.; Schnitzer, M. J.; Block, S. M. Single Kinesin Molecules Studied with a

Molecular Force Clamp. Nature 1999, 400 (6740), 184-189.

23 Roy, R.; Hohng, S.; Ha, T. A Practical Guide to Single-Molecule FRET. Nat. Methods 2008,

5 (6), 507-516.

24 Selvin, P. R.; Ha, T. Single-Molecule Techniques: A Laboratory Manual; Cold Spring Harbor

Laboratory Press: Cold Spring Harbor, NY, 2008.

25 Ha, T. J.; Ting, A. Y.; Liang, J.; Caldwell, W. B.; Deniz, A. A.; Chemla, D. S.; Schultz, P. G.;

Weiss, S. Single-Molecule Fluorescence Spectroscopy of Enzyme Conformational Dynamics

and Cleavage Mechanism. Proc. Natl. Acad. Sci. U.S.A. 1999, 96 (3), 893-898.

26 Chen, Y.; Hu, D. H.; Vorpagel, E. R.; Lu, H. P. Probing Single-Molecule T4 Lysozyme

Conformational Dynamics by Intramolecular Fluorescence Energy Transfer. J. Phys. Chem.

B 2003, 107 (31), 7947-7956. 14

27 Anfinsen, C. B.; Haber, E.; Sela, M.; White, F. H., Jr. The Kinetics of Formation of Native

Ribonuclease During Oxidation of the Reduced Polypeptide Chain. Proc. Natl. Acad. Sci.

USA 1961, 47, 1309-1314.

28 Schuler, B.; Eaton, W. A. Protein Folding Studied by Single-Molecule FRET. Curr. Opin.

Struc. Biol. 2008, 18 (1), 16-26.

29 Schuler, B.; Hofmann, H. Single-Molecule Spectroscopy of Protein Folding

Dynamics-Expanding Scope and Timescales. Curr. Opin. Struc. Biol. 2013, 23 (1), 36-47.

30 Zoldak, G.; Rief, M. Force as a Single Molecule Probe of Multidimensional Protein Energy

Landscapes. Curr. Opin. Struc. Biol. 2013, 23 (1), 48-57.

31 Onuchic, J. N.; Luthey-Schulten, Z.; Wolynes, P. G. Theory of Protein Folding: the Energy

Landscape Perspective. Annu. Rev. Phys. Chem. 1997, 48, 545-600.

32 Dill, K. A.; Chan, H. S. From Levinthal to Pathways to Funnels. Nat. Struct. Biol. 1997, 4 (1),

10-19.

33 Thirumalai, D.; Klimov, D. K. Deciphering the Timescales and Mechanisms of Protein

Folding Using Minimal Off-Lattice Models. Curr. Opin. Struc. Biol. 1999, 9 (2), 197-207.

34 Deniz, A. A.; Laurence, T. A.; Beligere, G. S.; Dahan, M.; Martin, A. B.; Chemla, D. S.;

Dawson, P. E.; Schultz, P. G.; Weiss, S. Single-Molecule Protein Folding: Diffusion

Fluorescence Resonance Energy Transfer Studies of the Denaturation of Chymotrypsin

Inhibitor 2. Proc. Natl. Acad. Sci. USA 2000, 97 (10), 5179-5184.

35 Nettels, D.; Gopich, I. V.; Hoffmann, A.; Schuler, B. Ultrafast Dynamics of Protein Collapse

from Single-Molecule Photon Statistics. Proc. Natl. Acad. Sci. USA 2007, 104 (8), 15

2655-2660.

36 Nettels, D.; Hoffmann, A.; Schuler, B. Unfolded Protein and Peptide Dynamics Investigated

with Single-Molecule FRET and Correlation Spectroscopy from Picoseconds to Seconds. J.

Phys. Chem. B 2008, 112 (19), 6137-6146.

37 Haran G. How, When and Why Proteins Collapse: the Relation to Folding. Curr. Opin. Struc.

Biol. 2012, 22 (1), 14-20.

38 Sherman, E.; Haran, G. Coil-Globule Transition in the Denatured State of a Small Protein.

Proc. Natl. Acad. Sci. USA 2006, 103 (31), 11539-11543.

39 Aznauryan, M.; Nettels, D.; Holla, A.; Hofmann, H.; Schuler, B. Single-Molecule

Spectroscopy of Cold Denaturation and the Temperature-Induced Collapse of Unfolded

Proteins. J. Am. Chem. Soc., 2013, 135 (38), 14040-14043.

40 Hoffmann, A.; Kane, A.; Nettels, D.; Hertzog, D. E.; Baumgartel, P.; Lengefeld, J.; Reichardt,

G.; Horsley, D. A.; Seckler, R.; Bakajin, O.; et al. Mapping Protein Collapse with

Single-Molecule Fluorescence and Kinetic Synchrotron Radiation

Spectroscopy. Proc. Natl. Acad. Sci. USA 2007, 104 (1), 105-110.

41 Chung, H. S.; Louis, J. M.; Eaton, W. A. Experimental Determination of Upper Bound for

Transition Path Times in Protein Folding from Single-Molecule Photon-by-Photon

Trajectories. Proc. Natl. Acad. Sci. USA 2009, 106 (29), 11837-11844.

42 Chung, H. S.; McHale, K.; Louis, J. M.; Eaton, W. A. Single-Molecule Fluorescence

Experiments Determine Protein Folding Transition Path Times. Science 2012, 335 (6071),

981-984. 16

43 Shaw, D. E.; Maragakis, P.; Lindorff-Larsen, K.; Piana, S.; Dror, R. O.; Eastwood, M. P.;

Bank, J. A.; Jumper, J. M.; Salmon, J. K.; Shan, Y.; et al. Atomic-Level Characterization of

the Structural Dynamics of Proteins. Science 2010, 330 (6002), 341-346.

44 Garcia-Mira, M. M.; Sadqi, M.; Fischer, N.; Sanchez-Ruiz, J. M.; Munoz, V. Experimental

Identification of Downhill Protein Folding. Science 2002, 298 (5601), 2191-2195.

45 Pirchi, M.; Ziv, G.; Riven, I.; Cohen, S. S.; Zohar, N.; Barak, Y.; Haran, G. Single-Molecule

Fluorescence Spectroscopy Maps the Folding Landscape of a Large Protein. Nat. Commun.

2011, 2, 493.

46 Chin, D.; Means, A. R. Calmodulin: A Prototypical Calcium Sensor. Trends Cell Biol. 2000,

10 (8), 322-328.

47 James, P.; Vorherr, T.; Carafoli, E. Calmodulin-Binding Domains - Just 2-Faced or

Multifaceted. Trends Biochem. Sci. 1995, 20 (1), 38-42.

48 Liu, R.; Hu, D.; Tan, X.; Lu, H. P. Revealing Two-State Protein-Protein Interactions of

Calmodulin by Single-Molecule Spectroscopy. J. Am. Chem. Soc. 2006, 128 (31),

10034-10042.

49 Babu, Y. S.; Sack, J. S.; Greenhough, T. J.; Bugg, C. E.; Means, A. R.; Cook, W. J.

Three-Dimensional Structure of Calmodulin. Nature 1985, 315 (6014), 37-40.

50 Chattopadhyaya, R.; Meador, W. E.; Means, A. R.; Quiocho, F. A., Calmodulin Structure

Refined at 1.7 Angstrom Resolution. J. Mol. Biol. 1992, 228 (4), 1177-1192.

51 Slaughter, B. D.; Unruh, J. R.; Price, E. S.; Huynh, J. L.; Bieber Urbauer, R. J.; Johnson, C.

K. Sampling Unfolding Intermediates in Calmodulin by Single-Molecule Spectroscopy. J. 17

Am. Chem. Soc. 2005, 127 (34), 12107-12114.

18

CHAPTER II. EXPERIMENTAL SECTION

This chapter is dedicated to the description of experimental techniques in our single-molecule protein folding studies and the sample preparation procedures used in our experiments.

2.1 Principles of Experimental Techniques

2.1.1 Principles of Confocal Microscopy

Developed in 1940, confocal microscopy is one of the most commonly used microscope techniques in biological and biophysical research. It is first developed by Marvin Minsky, then at

Harvard. It has the advantage in resolution over traditional microscopes in that it has the ability

to control depth of field, elimination or reduction of background signals away from the focal

plane which are out-of-focused lights. The key element to the confocal microscope approach is

the use of spatial filtering techniques to eliminate and reduce out-of-focus light signals. The

confocal microscopy is specifically designed for biological research in living organisms to

improve spatial resolution which is powerful and revolutionary. By 1971, lasers were introduced

as the light sources for confocal microscopy. In the mid-1980s, confocal microscopes have been evolved into their modern form similar to the ones in modern biological research labs used on a daily basis. A conceptual figure of confocal microscopy is shown in figure 2.1. The central idea of confocal microscopes is to use two pinhole apertures at both ends of the optical path: one aperture is put in front of the laser light source, and another aperture is set in front of the photon detector. Once these two pinhole apertures form optical conjugate planes, fluorescence signals from unwanted parts of the specimen which are from out of focus plane are blocked.1-6 19

The principle in laser scanning microscopy is presented in figure 2.1. Coherent light emitted

by the laser light source system passes through a pinhole aperture in a conjugate plane (confocal)

with a scanning point on the specimen and a second pinhole aperture positioned in front of the

detector (APD). As the laser is reflected by a dichromatic mirror, we scanned across the sample in a defined focal plane, secondary fluorescence signals emitted from points on the sample (in the same focal plane) reflect back through the dichromatic mirror and are focused as a confocal point at the detector pinhole aperture with undesired signals rejected. The unwanted

fluorescence signal is blocked by the pinhole and a diffraction limited image with typically a resolution of the excitation laser wavelength is obtained with significantly higher resolution.

We further combine this technique with a piezo controlled scanner. By carefully controlling the

step size of approximately 200 nanometers, we can easily achieve diffraction limited images of

single-molecules. By focusing the laser spot onto single molecules, we continue to record

single-molecule trajectories which contain dynamic information recording single-molecule

dynamic properties. Such combined techniques help us to remove ensemble-averaged signals

to reveal heterogeneity. By studying detailed fluctuation dynamics, we obtain valuable

information about biological systems at the single-molecule level.

In traditional wide field microscopy, the entire specimen is subjected to light illumination from an incoherent mercury lamp, and the resulting image of secondary fluorescence emission

(blurring the image) can be viewed directly in the eyepieces or projected onto the surface of a photo-detector (APD or CCD). In contrast to this simple concept, the mechanism of image formation in a confocal microscope is fundamentally different. It helps us to achieve higher 20

spatial resolution using pinhole systems. As discussed above, the confocal fluorescence

microscope consists of a laser excitation source, a scanning component with optical and

electronic components, electronic detectors, and a computer for acquisition of molecular images.

The scanning component is at the heart of the confocal system and is the part we used for raster

scanning the sample, as well as collecting the photon signals from the sample. A typical scanning

component contains inputs from the external laser sources, fluorescence filter sets and

dichromatic mirrors and a piezo raster scanning system.

Figure 2.1. Conceptual Figure of Confocal Microscope. A scheme showing a conceptual basis of

a modern confocal microscope which consists of laser light sources, detectors, sample stage

filters, and pinholes.

2.1.2 Principles of Fluorescence Resonance Energy Transfer

Fluorescence resonance energy transfer between two dyes, donor and acceptor, has proven to be a powerful spectroscopic technique for measuring distances within molecular systems.

Excitation energy of the donors is transferred to the acceptor via an induced dipole-dipole 21

non-radiative interaction.7-13 FRET efficiency measures the relative populations in an induced

dipole interaction via energy transfer. The efficiency of energy transfer can be determined by

spectroscopic approach. The transfer efficiency, E, is given by energy transfer results in a decrease in fluorescence measurable parameters. However, to study kinetics in ensemble

measurements using fluorescence intensity and excited-state lifetime measurements of the donor

and FRET, the reactions have to be synchronized to be at the initial step. To probe dynamic

information of a system, we subsequently measure the decay of the synchronized signal as a

function of time. To quantify the ensemble measurements FRET signal, the molecules have to be

prepared in one state before transferring energy to the acceptor molecules.

1.0

0.8

R =5.4nm 0.6 0 FRET E 0.4

0.2

0 2 4 6 8 10 Distance (nm) Figure 2.2. FRET Efficiency vs Inter-Dye Distance Curve. A calculated sinusoidal shaped

function curve of FRET Efficiency versus distance between two dye molecules in question is

shown. The R0 which corresponds to a 50% of FRET efficiency of Cy3-Cy5 dye pair is 5.4nm. 22

Fluorescence resonance energy transfer is a form of electronic energy transfer between

fluorescent, typically organic, molecules. The energy transfer process between the donor and

the acceptor molecules is via the non-radiative energy transfer mechanism involving a virtue photon. The excited-state lifetime of the donor molecule is reduced in such energy transfer dynamics which is different from an emission reabsorption mechanism. The nature of such energy transfer processes between the donor and the acceptor molecules is not via transmission of photons. Historically, FRET is referred to as an energy transfer by inducing electronic resonance between the donor and acceptor molecules. Effective energetic coupling of FRET between the donor and acceptor molecules also requires a emission and absorption spectral overlap between the donor and acceptor molecules. Theoretically, such a spectral overlap bundles a lot of complicated information into a simple functional form which is firstly derived by

Forster. The spectral overlap includes nuclear overlap factors, which are separated from the

electronic coupling term by the Born-Oppenheimer approximation, since the photon absorption is considered instantaneous, which turns the spectral overlap into the form of quantum mechanical coupling factors. Energy conservation and nuclear overlap factors, separated from the electronic coupling by the Franck-Condon principle, are shown to relate emission and absorption events of the donor and acceptor molecules.7-13 The energy transfer rate can be written

as:

23

if the electronic coupling is independent of energy, then we can write,

which describes the rate of the energy transfer process.

The electronic coupling between the donor and the acceptor molecules can be described by a coulombic contribution part and a short range contribution part. It is assumed that the coulombic part plays a major role in the energetic coupling between the donor and acceptor molecules.

In addition to the dipole-dipole interaction, an orientation factor is also taken into account. This orientation factor describes the rotation of the dipoles of our dye molecules used in our FRET measurements.

However, in our single-molecule experiments the orientation factor is fixed to be 2/3. Such fact is due to the experimental condition of our single-molecule experiments. The typical

time-resolution of single-molecule experiments is at the millisecond time scale, yet single dye orientation dynamics is at the picosecond time scale. To our first approximation we only need to use the averaged orientational factor.

Single-molecule fluorescence resonance energy transfer (single molecule FRET), which applies FRET at single molecule level, is a measurable technique for studying real-time 24

structural dynamics of individual biomolecules, with measuring distance typically in 30-80 Å

range. Generally, a donor-acceptor pair is attached to two specific sites of one or two target molecules, and the energy transfer from donor to acceptor occurs through non-radiative induced

dipole-dipole interactions. Practically, we use the distance dependent formula for distance

conversion in our single-molecule FRET experiments. The energy transfer efficiency (EFRET)

depends on the inter-distance between the donor fluorophore and acceptor fluorophore:

1 EFRET  6 1+rR / 0 

where r is the separate distance between two fluorophores, the FRET donor and acceptor, respectively. R0, as the Förster radius, is a function of dipole orientations of two fluorophores, refractive index of the medium, spectral overlap integral between donor emission and acceptor absorbance, and quantum yield of the donor as discussed in details above.

Figure 2.3. Conceptual Figure of Protein Folding FRET. A cartoon representing using smFRET to monitor single-molecule protein, DNA and various biomolecule conformational motions. By monitoring emission intensities from donor and acceptor organic dye molecules, conformational 25

information can be obtained from experimental data.

Figure 2.4. Conceptual Figure of FRET. Energy levels involved in typical FRET experiments.

The laser light source is used to first excite the donor molecule and then the donor molecule

transfers its excitation energy to the acceptor molecule via non-radiative dipole-dipole interaction.

The energy transfer efficiency depends on the distance between the donor and acceptor

molecules making FRET highly sensitive to a distance change.

2.1.3 Signal Detection techniques: Introduction to APD

Avalanche photodiode (APD) is used in our single-molecule experiments due to its low dark count and high photon to electron transition efficiency. It is a highly sensitive tool to probe molecular scale dynamics by recording molecular scale signal photon by photon. The time-resolution in our experiment is typically in milliseconds, roughly the same scale as protein conformational dynamics. An avalanche photodiode is a semiconductor-based photodetector (photodiode) which is operated with a relatively high reverse 26

voltage. The carrier generation process amplifies the low signal; subsequently we reach single

photon counting sensitivity in our single-molecule experiments. The avalanche process effectively

amplifies the photocurrent which effectively turns low signals to measurable signals in our

experiments. Therefore, avalanche photodiodes can be used for sensitive detections.

Figure 2.5. A Schematic Representation of APD. A schematic representation of the working

mechanism of the typical avalanche photodiode.

2.2 Experimental Details

2.2.1 Experimental Setup of Single-Molecule FRET Combined with Confocal

Microscope

We use single-molecule photon-stamping spectroscopic approach to record FRET

trajectories of CaM at different concentrations of denaturant solvents. Using this approach, we

are able to record the emission photon time trajectories from both the donor and acceptor with

specific detection time for each detected photon. The experimental setup is an inverted confocal microscope (Axiovert 200, Zeiss) that uses a crystal laser (532nm CW) as the light source for excitation. The laser beam focuses through a 100× oil immersion objective lens (1.3 27

NA, 100×, Zeiss) onto the upper surface of cover slip after the excitation light is reflected up by a dichroic beam splitter (z532rdc, Chroma Technology). To obtain confocal microscopy image,

we use an x-y closed-loop piezo position scanning stage for raster-scan of the sample sandwich

(Figure 2.6A). The fluorescence is collected through the same objective, and the FRET photon signal is split by a dichroic beam splitter (640dcxr, Chroma Technology) into two different

wavelengths. We use two Si avalanche photodiode single photon counting modules

(SPCM-AQR-16, Perkin Elmer Optoelectronics): 570nm for the donor channel and 670nm for the acceptor channel.14, 15

We use the same single-molecule photon stamping approach to record FRET trajectories at

different concentration of C28W. This approach records photon emissions from donor and acceptor channels one by one with arrival times, and the intensity trajectory can be constructed

as a function of time. Förster resonance energy transfer or fluorescence resonance energy transfer

(FRET) is a certain type of energy transfer realized by non-radiative dipole-dipole

interaction. The experimental setup is an inverted confocal microscope (Axiovert 200, Zeiss)

which uses a crystal laser (532nm CW) as the light source for excitation (Figure 2.6). The laser

beam focuses through a 100× oil immersion objective lens (1.3 NA, 100×, Zeiss) onto the upper

surface of the cover slip after reflected up by a dichroic beam splitter (z532rdc, Chroma

Technology). To obtain confocal microscopy images, we use an x-y closed-loop piezo position

scanning stage for raster-scan of the sample sandwich. The fluorescence is collected through the

same objective and the signal is split by a dichroic beam splitter (640dcxr) into two different

wavelengths: 570nm and 670nm which are the emission wavelengths of Cy3 and Cy5 28

respectively. To detect the signal from the two channels, two Si avalanche photodiode single photon counting modules (SPCM-AQR-16, Perkin Elmer Optoelectronics) are used for recording

the photons from the donor and acceptor.

Figure 2.6. Our Single-Molecule Setup. (A) Single-molecule fluorescent experimental setup. It

is an inverted confocal microscope (Axiovert 200, Zeiss) which uses a laser (532nm CW) as the

excitation light source.35, 40 Fluorescence photons from donor and acceptor are both directed onto

avalanche photodiodes to acquire emission images and FRET intensity trajectories. (B) Image

obtained from confocal microscope of single CaM molecules. The left-hand side is the image

obtained from the Cy3 donor channel and the right-hand side is the image obtained from the Cy5

acceptor channel. The bright spots are single molecules with diffraction limited (~300nm in 29

diameter) image. Each image is obtained by laser focus raster-scanning and collecting fluorescence of Cy3-Cy5 D-A labeled single-molecule CaM.

2.2.2 Materials and sample preparation

Sample Preparation and Characterization. The CaM is mutated with cysteine residues on

N-terminal domain at residue 34 and C-terminal domain at residue 110, and a FRET dye pair

Cy3/Cy5 as donor/acceptor is covalently tethered onto the protein via thiolation reactions. In

our experiment, the change of FRET efficiency (EFRET) ranges from 0.6 to 0.2 corresponding to

the donor-acceptor (D-A) distance change of about 5.0 to 6.7 nm, and the Forster radius R0 of

Cy3-Cy5 pair is ~5.4 nm. The samples for single-molecule conformational folding-unfolding

dynamics measurements are prepared inside a 1% agarose gel with 99% of buffer solution (Type

VII, Sigma). We make samples of CaM into different concentrations of denaturant GdmCl in

the mixture of 1 nM CaM, 1.25μL and oxygen scavenger with Trolox solution to obtain a 10μL mixture of enzyme and denaturant solvent. The Trolox solution is prepared previously by

dissolving about 1 mM 6-hydroxy-2,5,7,8-tetramethylchroman-2-carboxylic acid (Trolox) in 1

mg/mL glucose oxidase, 0.8% D-glucose and 0.04 mg/mL catalase in order to protect fluorescent

dyes from photobleaching or blinking as a result of triplet state oxygen quenching as well as

other photophysical processes. We then heat the 10μL 1% agarose gel just above its

gel-transition temperature (26°C) and quickly mix the above enzyme solution with denaturant

solution and the gel between two clean cover glasses to form a sandwiched sample. All

solutions are prepared with 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid (HEPES buffer) 30

at pH 7.4. To probe conformational dynamics of single-molecule enzymes at different

concentration of denaturant solvents, we carry out concentration dependent experiments with

different ratio of mixture in the sample.

Trolox+Oxygen CaM GdmCl Agarose Gel Scavenger ~1 nM CaM+0M 0.2μL 200nM 0μL 8M 9.8μL 10μL Gdmcl ~1 nM CaM+1M 0.2μL 200nM 2.5μL 8M 7.3μL 10μL Gdmcl ~1 nM CaM+2M 0.2μL 200nM 5μL 8M 4.8μL 10μL Gdmcl

Table 2.1. List of concentrations used in our experiments.

In our protein-protein interaction experiments, the single CaM molecule has mutation on

N-terminal domain at residue 34 and C-terminal domain at residue 110, and fluorescent dye pair

Cy3/Cy5 was tethered onto the protein via thiolation. These two dyes serve as a spectroscopic ruler for the measurement of conformational fluctuation of the protein molecule interacting with

C28W. In our experiment, the change of FRET efficiency corresponds to the distance change of about 2 nm. On the other hand, the Forster radius R0 of Cy3-Cy5 pair is ~5.4 nm, so the distance

change of the two dyes monitoring the conformational change falls into measureable range. In

our experiment, the samples for single-molecule conformational folding-unfolding dynamics measurements are prepared inside the agarose gel (1% by weight, Type VII, Sigma). For example, we make a sample of CaM into 2 M of denaturant solvent GdmCl by first mixing 0.2μL 200 nM

CaM, 2.5μL 8 M GdmCl, 0.1mM CaCl2 and 8.55μL oxygen scavenger with Trolox solution to obtain a 10μL mixture of protein and denaturant solvent with 100nM C28W. The Trolox solution 31

is made previously by dissolving about 1 mM

6-hydroxy-2,5,7,8-tetramethylchroman-2-carboxylic acid (Trolox) in 1 mg/mL glucose oxidase,

0.8% D-glucose and 0.04 mg/mL catalase in order to protect fluorescent dye from

photobleaching or blinking as a result of triplet state formation or other irreversible

photophysical processes. We then heat the 10μL 1% agarose gel just above its gel-transition

temperature (26°C) and quickly mix the above protein with denaturant solvent solution with the

gel between two clean cover glasses to form a sandwich. All solutions are prepared with

Phosphate buffered saline (PBS buffer) at pH 7.4. Since we want to probe conformational

dynamics of single protein molecule at different concentration of C28W peptide, we carry out concentration dependent experiments with different ratio of mixture in the sandwich as listed

below.

Trolox+Oxygen Agarose CaM GdmCl Scavenger Gel ~1 nM CaM+2M 0.2μL 200nM 2.5μL 8M 9.8μL 10μL Gdmcl+100 nM C28W ~1 nM CaM+2M 0.2μL 200nM 2.5μL 8M 8.55μL 10μL Gdmcl+500 nM C28W ~1 nM CaM+2M Gdmcl+1000 nM 0.2μL 200nM 2.5μL 8M 7.3μL 10μL C28W

Table 2.2. List of concentrations used in our folding-binding experiments.

In our single-molecule crowding experiment, all solutions are prepared with HEPES buffer at pH

7.4. Since we want to probe conformational dynamics of single enzyme molecule at different

concentration of Ficoll 70, we carry out concentration dependent experiments with different ratio

of mixture in the sandwich. 32

Figure 2.7. A Cartoon of Our Sample. We prepare our samples in agarose gel 1% by weight.

The agarose gel formed pores with diameter about 200 nm. The single protein molecules have

the size of approximately 6 nm. The protein molecules which are much smaller in dimensions

compared to the pore size can freely rotate inside the pore.

2.2.3 Statistical Analyses of Single-Molecule Intensity Trajectories.

To analyze the FRET signal fluctuation trajectories in order to obtain the CaM conformational folding-unfolding conformational dynamics, we have applied a newly developed

2D regional correlation mapping analysis to analyze our single-molecule photon-stamping trajectories. The 2D regional correlation mapping analysis calculates a two-dimensional

cross-correlation amplitude distribution and acts as a guide to find the anti-correlated portion of a selected trajectory. In such analysis, each of the single-molecule FRET trajectories are scanned

with different starting and ending time, and the cross-correlation amplitude of each time segment

with distinct starting and ending time are calculated. Color bar is used to indicate the amplitude

of cross-correlation in order to locate the anti-correlated portion with negative value of

cross-correlation amplitudes. 33

stop tstop t Ccross(,:)()()()() tt start stop  ItItdtAD   ItItAD  tstart  tstart

Where IA and ID are the photon count intensities of donor and acceptor, and tstart and tstop give the

scanning window width. The cross-correlation functions are calculated with different tstart and tstop along a pair of single-molecule FRET trajectories {IA(t)} and {ID(t)}. The detailed description of the analysis algorithm can be found elsewhere. Applying this method, we are able to identify the time segments along a specific trajectory with a strong anti-correlated cross-correlation indicated by a negative value of cross-correlation amplitudes (cold color).

Typically, single-molecule FRET trajectories are dominated by shot noises or averaged intensity drifts as a result of local environment fluctuations. Our two-dimensional cross-correlation

amplitude distribution analysis enables us to locate true anti-correlated segments from single-molecule FRET trajectories for further dynamics analyses.

We apply auto-correlation and cross-correlation calculations to analyze our single-molecule

photon-stamping trajectories after the identification of each anti-correlated single-molecule time segments. The cross-correlation and auto-correlation function are defined as:

IAD0  I  t  IA 0   IADD  I t  I  Ctcross    IIAD00    IIIIA 00  AD   D 

IAA0  I  t  IA 0   IAAA  I t  I  auto  Ct  22 IA 0  IIAA0   

where IA(t) and ID(t) representing acceptor and donor intensitie, and and are the

means of the intensity trajectories respectively. 34

Figure 2.8. TCAD Map in a Typical Single-Molecule Experiment. (Upper panel) The TCAD

map and cross correlation functions calculated from a simulated two-band fluctuation trajectory

that consists of three sections: cross-correlated (I: 1–500 data points), non-correlated (II: 501–

1000 data points) and anti-correlated (III: 1001–1500 data points). (A) 2D regional correlation

analysis by TCAD mapping. The hot color represents the positive amplitude and the cold color

represents the negative amplitude. (B) Correlation functions calculated from the three sections of

correlated, non-correlated, and anti-correlated fluctuation data corresponding to sections I, II, III, and the whole data trajectory (I + II + III). It is evident that a conventional correlation calculation from the whole data trajectory gives no correlation amplitude, whereas the 2D regional 35

correlation analysis gives definitive analysis of the correlation behavior for each specific

fluctuation in the long trajectory. (Lower panel) An example of 2D regional correlation analysis

of single-molecule FRET fluctuation data. The experimental FRET two-band (D – A) fluorescence intensity fluctuation trajectory is measured from a D–A labeled kinase enzyme

protein molecule showing a conformational change fluctuation in a buffer solution. (C) A TCAD

map calculated from a two-band (D–A) FRET fluorescence fluctuation trajectory. The red and

black trajectories are the donor and acceptor signals, respectively. (D) Cross correlation functions

calculated from different sections of the trajectory. It is clear that the dynamics can be averaged

out if only a whole-trajectory calculation is carried out. The anti-correlated FRET fluctuations can only dominate the fluorescence intensity trajectories in fraction of time periods but not all the time due to non-correlated and correlated thermal fluctuation background noises.

2.3 Protein Folding Dynamics in Condensed Phases

The new single-molecule techniques have allowed the investigation of the mechanism of

formation of basic structural elements of proteins. Since the basic structure elements of

proteins fold at a faster time scale than typical measurement time scales of single-molecule

FRET spectroscopy, which is at the millisecond time scale, it cannot be directly probed. However,

the folding dynamics speed limit can be calculated and simulated. The measured protein folding

dynamics time scale is often much longer than the theoretical limit. Such elongation of time

scales typically implies the existence of a folding barrier and multiple intermediate states of

protein folding. In other words, by directly measuring protein folding dynamics using

single-molecule techniques, we can directly measure the height and energetic features of the 36

protein folding dynamics. We employ condensed phase dynamic models to model protein

folding dynamic processes in condensed phases. The simplest model describing chemical

dynamics in the condensed phase is Kramer’s theory. Another fact, which is not so obvious, is

that sometimes for a simple protein involving fewer than 100 amino acids the activation barrier

of the protein folding process may disappear. The measured folding time of these proteins serves

as a benchmark to further understand larger and more complex proteins involving more amino acids. Such is the hypothesis we use, when we try to understand the energetic features of

protein folding dynamics in condensed phases.

Direct applications of Kramer’s theory of unimolecular reaction rates in condensed phases

will give us an upper bond estimation of the height of the activation barrier. Such theory assumes

that the activation barrier crossing dynamics can be described by a one-dimensional diffusion

along a reaction coordinate. The time of the folding dynamic process can be expressed by

τfolding = 2πτlimit exp(ΔG/kT) where τfolding is the folding time of single-molecule CaM measured in our experiment, τlimit is the theoretical upper bound of single-molecule CaM folding time, and ΔG is the free energy

barrier and k is the Boltzmann’s constant T is the temperature.

The vast majority of measurements yield the formation time of a loop is less than 0.1 μs, and

α-helices formation is approximately 0.5 μs. The formation time of β hairpins is greater than

0.5 μs. From these data we understand the theoretical folding time limit should be at the

microsecond scale. By using single-molecule FRET spectroscopy we further measure the

distance changes in protein folding dynamic processes. We also can measure the characteristic 37

timescale of protein folding dynamic processes. From these two, we further can characterize

the dynamic behavior of protein folding in condensed phases. The conformational diffusion

coefficients calculated from the distance changes and folding time can give us a quantitative understanding of the energetic features of condensed phase protein folding dynamic processes.

2.4 Theory of Polymer Solutions

Proteins are heteropolymer made of 20 different monomer types (amino acids). They self-assemble into well-defined three dimensional structures to perform biological functions. In principle, protein solutions can be understood as polymer solutions. Protein dynamics is also a form of polymer dynamics. The denatured state of proteins has been described for many years as

being similar to a random coil-like polymer. By tuning the chemical property of the polymer solution, we performed our single-molecule folding-unfolding experiments by changing the

solvent properties. The protein molecules subsequently undergo conformational changes in

different solvent environments. Protein molecules change from the unfolded state to the folded

state. This is known in polymer science as the coil-globule transition. In good

0.55 solvents, the radius of gyration is Rg = 0.345N nm. This formula is confirmed by small-angle

X-ray scattering. The hydrodynamic radius is 1.5 times smaller than the radius of gyration.

Since linear scaling theory holds for small degrees of polymerization, using the linear length scaling suggested by the homopolymer collapse theory, theoretical upper bound of single-molecule CaM folding time can be estimated by a simple diffusion model.

Synthetic polymers such as PEG, Ficoll are commonly used as a means to simulate

molecular crowding in living cells. Polymers of different sizes were used in the past to study 38

protein-protein associations under crowded environments. Such studies have a significant

contribution to the understanding of diseases resulting from over-expression of proteins inside

living cells. However, the whole polymer solution and matrix is highly dynamic instead of

static. So our single-molecule study proves to be a powerful tool to study such complex dynamic properties.

Figure 2.9. A Schematic Representation of Molecular Crowding. Schematic representation of

proteins in crowded polymer solutions. As we change the concentration of the matrix polymer solution, the protein-protein interaction pattern changes.

2.5 References

1 Nie, S.; Zare R. N. Optical Detection of Single Molecules. Annu. Rev. Biophys. Biomol.

Struct. 1997, 26, 567-596.

2 Minsky, M. Microscopy Apparatus: US 3,013,467.

3 Davidovits, P.; Egger, M. D. Scanning Laser Microscope. Nature 1969, 223 (5208): 831. 39

4 Egger, M. D. (1971). Scanning Laser Microscope for Biological Investigations. Applied

optics. 1971, 10 (7): 1615–1619

5 Amos, W.B.; White, J.G.: How the Confocal Laser Scanning Microscope Entered Biological

Research. In: Biology of the Cell / under the Auspices of the European Cell Biology

Organization. Band 95, Nummer 6, September 2003, S. 335–342

6 Prasad, V.; Semwogerere, D.; Weeks, E.R. Confocal Microscopy of Colloids. J. Phys.: Cond.

Mat. 2007 19, 113102.

7 Förster, T. Zwischenmolekulare Energiewanderung und Fluoreszenz. Annalen der Physik

1948, 437, 55-75.

8 Förster T. Fluorescence of Organic Compounds Gettingen: Vandenhoeck & Ruprecht:

1951:312.

9 Stryer, L.; Haugland, R. P. Energy Transfer: A Spectroscopic Ruler. Proc. Natl. Acad. Sci.

USA 1967, 58, 719-730.

10 Clegg, R.M. Fluorescence Resonance Energy Transfer. Curr. Opin. in Biotech. 1995, 6:103-l

10.

11 Ha, T.; Enderle, T.H.; Ogletree, D.F.; Chemla, D.S.; Selvein, P.R.; Weiss, S. Probing the

Interaction Between Two Single Molecules: Fluorescence Resonance Energy Transfer

Between a Single Donor and a Single Acceptor. Proc. Natl. Acad. Sci. USA 1996, 93,

6264-6268.

12 Deniz, A.A.; Dahan, M.; Grunwell, J.R.; Ha, T.; Faulhaber, A.E.; Chemla, D.S.; Weiss, S.;

Schultz, P.G. Single-Pair Fluorescence Resonance Energy Transfer on Freely Diffusing 40

Molecules: Observation of Förster Distance Dependence and Subpopulations Proc. Natl.

Acad. Sci. USA 1999, 96, 3670–3675.

13 Ambrose, E.J. A Surface Contact Microscope for the Study of Cell Movements. Nature 1956,

178 (4543): 1194.

14 Yanagida, T.; Sako, Y.; Minoghchi, S. Single-Molecule Imaging of EGFR Signalling on the

Surface of Living Cells. Nature Cell Biology 2000, 2 (3): 168–172.

15 He, Y.; Lu, M.; Lu, H.P. Single-Molecule Photon Stamping FRET Spectroscopy Study of

Enzymatic Conformational Dynamics. Phys. Chem. Chem. Phys., 2013, 15, 770-775.

41

CHAPTER III. PROBING SINGLE-MOLECULE PROTEIN FOLDING

CONFORMATIONAL DYNAMICS USING SINGLE-MOLECULE FRET

SPECTROSCOPY

This chapter is dedicated to the study of single-molecule protein folding using single-molecule FRET.

3.1 Introduction

Single-molecule fluorescence spectroscopy is powerful to study complex biological processes one molecule at a time.1 Sub populations involved in enzymatic reactions2,3, protein folding dynamics4 and other dynamic processes5,6 can be well characterized without ensemble

averaging. Single-molecule fluorescence resonance energy transfer (smFRET) spectroscopy monitors conformational changes of dye labeled single biomolecules with sensitivity down to a 2

nm to 10 nm.7 Such distance sensitivity and millisecond temporal resolution makes smFRET an

ideal tool to probe protein conformational dynamics.

Since Anfinsen first posed the question about the spontaneity of proteins folding into their native state through an enormous number of possible conformations 50 years ago, the protein

folding problem, although with significant progress, still remains to be a fascinating topic.8

Theoretically, the protein folding process can be understood as a navigation on a funnel-shaped

energy landscape towards a global energy minimum, the folded state.9-11 This multi-dimensional

energy landscape can be rugged and extremely complex involving multiple folding routes and

metastable states. Experimentally, mainly two types are involved: force-probe techniques using 42

atomic force microscopy (AFM) or optical tweezers,12 and fluorescent spectroscopy especially smFRET. Single-molecule FRET is previously applied to reveal folded and unfolded protein subpopulations from otherwise ensemble average population distribution.4, 13 Subsequently, using smFRET, there are advances in characterization of protein unfolded state dynamics,13-15 protein collapse (from unfolded state to a compact yet folded state) dynamics16-18 and protein folding dynamics.19 Most recently, based on previous work,20 a remarkable application of smFRET to determine the transition path times of protein folding is achieved.21 The average transition path times are reproduced well in an all-atom explicit solvent simulation on a special-purpose super computer ANTON, which recently become available.22 On the other hand, distinct from two-state folders, protein with ‘one-state’ downhill folding process is also investigated via smFRET, and its dynamics well specified.23 For the studies mentioned above, two-state folders with two distinct folded and unfolded conformations serve the purpose well. However, when it comes to large protein with more than 100 amino acids, this simple two-state folding dynamic scheme does not apply. A recent example is a study of a 214 amino acid protein adenylate kinase, six distinct folding intermediates may exist.24

Calmodulin (CaM) is a 148-residue protein responsible for intracellular calcium-sensing.

It is crucial in many biological processes including muscle contraction and energy metabolism.25,

26 CaM has two globular domains, and the conformational dynamics of the two domains is under

extensive studies.27 Traditional methods involving nuclear magnetic resonance and X-ray

crystallography have provided detailed insights into the mechanisms and dynamics of CaM

conformational motions.28-30 As a result of the flexibility and the dumbbell shape of the protein, 43

with dye labeling at the two domains, it is possible to monitor the conformational dynamics of

CaM with single-molecule spectroscopy, such as smFRET.27 The typical time-resolution can be milliseconds which is the characteristic time scale of protein motions. Similarly, free diffused types of experiments have been carried out on single CaM in different concentrations of denaturant urea. However, dynamic information is limited in such experiments due to the limitation of the length of single-molecule trajectories (typically microsecond scale).31-39

In this chapter, we use single-molecule FRET spectroscopy to probe the folding-unfolding conformational dynamics of CaM under the mild denatured condition of denaturant guanidinium

chloride (GdmCl). The critical concentration of GdmCl, at which single proteins undergo folding-unfolding conformational fluctuation, has been achieved. Subpopulation of proteins with folded and unfolded states equally populated is obtained. These results allow us to address the following intriguing questions about single-molecule protein folding-unfolding dynamics. What exactly are the sequences of microscopic events that ultimately take single CaM molecules from the unfolded conformation to the folded native states? Is it a simple two-state folding dynamic scheme or more complex process involved multiple states? Is there a distribution of such sequences or folding transition paths, and what will be the underlying energy landscape nature of such a folding channel distribution? Is it static disorder or dynamic disorder associated with such an energy landscape or is it a highly dynamic energy landscape with certain distinct “ruggedness” present along each folding routes?

3.2 Materials and Methods

3.2.1. Sample Preparation and Characterization. 44

The CaM is mutated with cysteine residues on N-terminal domain at residue 34 and

C-terminal domain at residue 110, and a FRET dye pair Cy3/Cy5 as donor-acceptor (D-A) is

covalently tethered onto the protein on mutated cysteine residue 34 at N-terminal domain and on mutated cysteine residue 110 at C-terminal domain via thiolation reactions.40 In our experiment,

the change of FRET efficiency (EFRET) ranges from 0.6 to 0.2 corresponding to the D-A distance

change of about 5.0 nm to 6.7 nm in single-molecule protein conformation, and the Forster radius R0 of Cy3-Cy5 pair is ~5.4 nm at which EFRET of Cy3-Cy5 is 0.5. The samples for

single-molecule conformational folding-unfolding dynamics measurements are prepared with a 1%

agarose gel containing 99% of buffer solution (Type VII, Sigma). Single CaM molecules can rotate freely to perform biological functions and chemicals such as electrolytes and denaturant

GdmCl can diffuse uninterruptedly.6, 40, 41 We make samples of CaM into different concentrations of denaturant GdmCl in the mixture of 1 nM CaM, 1.25μL and oxygen scavenger with Trolox solution to obtain a 10μL mixture of protein and denaturant solvent GdmCl solution. The

Trolox solution is prepared by dissolving about 1 mM

6-hydroxy-2,5,7,8-tetramethylchroman-2-carboxylic acid (Trolox) in 1 mg/mL glucose oxidase,

0.8% D-glucose and 0.04 mg/mL catalase in order to protect fluorescent dye from

photobleaching or blinking due to the triplet state oxygen quenching as well as other

photophysical processes.40 We then heat the 10μL 1% agarose gel just above its gel-transition

temperature (26°C) and quickly mix the above protein solution with denaturant solution GdmCl

and the gel between two clean cover glasses to form a sandwiched sample. All solutions are

prepared with 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid (HEPES buffer) at pH 7.4 45

with 1mM EGTA.38 To probe conformational dynamics of single-molecule protein at different

concentration of denaturant solvent, we carry out concentration dependent experiments with different ratio of mixture in the sample.40

3.2.2. Single-Molecule Imaging and FRET Measurements.

We use single-molecule photon-stamping spectroscopic approach to record smFRET D-A trajectories of CaM at different concentrations of denaturant solvent GdmCl. Using this approach, we are able to record the emission photon time trajectories from both donor and acceptor with specific detection time for each detected photons. The experimental setup is an inverted confocal microscope (Axiovert 200, Zeiss) that uses a laser (532nm, CW) as the light source for excitation.40 The Laser beam focuses through a 100× oil immersion objective lens (1.3

NA, 100×, Zeiss) onto the upper surface of cover slip after the excitation light is reflected up by a dichroic beam splitter (z532rdc, Chroma Technology). To obtain confocal microscopy image,

we use an x-y closed-loop piezo position scanning stage for raster-scan of the sample sandwich

(Figure 3.1A). The fluorescence is collected through the same objective, and the FRET photon signal is split by a dichroic beam splitter (640dcxr, Chroma Technology) into two different

wavelength 570nm for the donor channel and 670nm for the acceptor channel with two Si

avalanche photodiode single photon counting modules (SPCM-AQR-16, Perkin Elmer

Optoelectronics). A detailed description of the experimental setup is reported previously. 35, 40 46

Figure 3.1. Our Single-Molecule Imaging System. (A) Single-molecule fluorescent experimental setup. It is an inverted confocal microscope (Axiovert 200, Zeiss) which uses a laser (532nm CW) as the excitation light source.35, 40 Fluorescence photons from donor and acceptor are both

directed onto avalanche photodiodes to acquire emission images and FRET intensity trajectories.

(B) Image obtained from Confocal Microscope of single CaM molecule. The left-hand side is

image obtained from the Cy3 donor channel and the right-hand side is image obtained from the

Cy5 acceptor channel. The bright spots are single molecules with diffraction limited (~300nm

in diameter) image. Each image is obtained by laser focus raster-scanning and collecting

fluorescence of Cy3-Cy5 D-A labeled single-molecule CaM.

47

3.3 Results and Discussion

3.3.1 Single-Molecule FRET Trajectories Monitored Unfolding of Single CaM

Molecules.

Figure 3.1B shows typical images from our smFRET imaging microscopy using an inverted

confocal microscope. By raster-scanning the sample, a 20μm×20μm sample image yields

bright spots with 300nm resolution of single molecules. We attribute that the imaging spot features are due to single CaM molecules confined in agarose gel. Single CaM molecules embedded in agarose gel can rotate freely to perform its biological functioning and chemicals such as electrolytes and denaturant GdmCl can diffuse uninterruptedly.40, 42 From each specific single-molecule CaM imaging spot, we are able to obtain continuous D-A fluorescence intensity trajectories (Figure 3.2A). Typically, we gather a number of single-molecule fluorescence intensity trajectories at different concentrations of denaturant GdmCl, and we calculate the EFRET of each of the single-molecule fluorescence intensity trajectories measured using this equation.

ItA() EtFRET () AA IAD()() t I t DD

Where  A and  D are the emission quantum yields of acceptor and donor dyes, respectively, and  A and  D are the acceptor and donor detection efficiencies, respectively. Here the

AA correction factor is ~1 in our experiment conditions. DD 48

Figure 3.2. Typical Trajectories and Data. (A) Typical D-A signals and EFRET trajectories. Green and red lines indicate the donor and acceptor channel respectively. (B) Single-molecule EFRET trajectory calculated following formula. (C) The histogram distribution from EFRET trajectory gives the average EFRET of a certain single-molecule. (D-F) Distribution of EFRET of different samples at different concentration of GdmCl, we carry out concentration dependent experiment of GdmCl of 0M 1M and 2M. The decrease in EFRET mean values characterizes the unfolding of CaM molecule at higher concentration of denaturant solvent.

To identify the specific condition that facilitates the CaM to have a measurable probability of staying either folded or unfolded states, we have characterized the folding and unfolding state 49

distributions by single-molecule fluorescence intensity trajectories measurements of EFRET under

different denaturant conditions (Figure 3.2D-2F). The decrease in EFRET mean values of the single-molecule CaM EFRET distributions are due to the increased unfolding probability of the single CaM molecules under the increased denaturant GdmCl concentrations from zero to 2 M.

The EFRET distribution consists of both the folded subpopulation with EFRET 0.6 and unfolded

subpopulation with EFRET 0.2. Single-molecule FRET spectroscopy allows folded and unfolded

molecules to be distinguished on the basis of the significant distance dependence of energy

transfer between smFRET donor and acceptor dyes. More importantly, single-molecule

sub-populations of partially folded with equal amount of time dwelling on folded and unfolded

conformational states can be thoroughly examined by studying detailed fluctuation dynamics of

each single-molecule fluorescence intensity trajectory. We locate and record singe-molecule trajectories from individual molecules undergoing spontaneous folding-unfolding conformational fluctuation under the condition providing a roughly 50%-50% folding-unfolding conformational state probability, i.e. the transition midpoint. At the titration midpoint, approximately half of

the population of CaM is in folded conformational state whereas another half of the population is

in unfolded conformational state. Our ensemble-averaged titration experiment of Cy3-Cy5 D-A labeled CaM molecules in solution with different concentration of GdmCl yields 2M concentration of denaturant GdmCl to be the titration midpoint where the folded and unfolded conformational states of CaM molecules are equally populated (Supporting Information). In term of single-molecule experiments, such titration midpoint condition is ideal for studying spontaneous folding-unfolding conformational fluctuation dynamics, since individual molecule 50

undergoes spontaneous conformational fluctuations without external driving force. The EFRET

distribution measured under no GdmCl condition, at which CaM molecules are in folded native

states, has a mean of 0.55 ± 0.07 that corresponds to 5.2 ± 0.2 nm D-A distance, a folded

conformation state; whereas the distribution of EFRET measured under 2M GdmCl has a mean of

0.36 ± 0.17 that corresponds to 6.1 ± 0.8 nm D-A distance, an unfolded conformation state. On

average, a decrease in EFRET corresponds to distance change between the D-A Cy3-Cy5 dye pair

which further indicates a conformational distance change along the FRET coordinate.

Therefore, under the unfolding effect of denaturant GdmCl, we measured that the D-A Cy3-Cy5

dye pair distance change of 0.9 ± 1.0 nm for the folded and unfolded states of CaM. In our

experiments, single-molecule CaM is embedded in 1% agarose gel in which a single CaM

molecule can rotate freely to perform its biological functions, and chemicals, such as electrolyte and denaturant GdmCl, can diffuse uninterruptedly.40, 42

3.3.2. Autocorrelation Analyses of CaM Folding-Unfolding Conformational Fluctuation

Dynamics.

We analyze the fluctuation dynamics of our single-molecule FRET trajectories by

calculating the autocorrelation functions of selected segments using our 2D regional correlation

mapping analysis.39, 40 By applying the 2D regional correlation mapping analysis, we identify

specific anti-correlated segments along a trajectory and zoom into that segment to study detailed

dynamic fluctuation information by calculating autocorrelation and cross-correlation functions.

Fitting the autocorrelation functions with single exponential functions, we are able to

characterize the rate of protein conformational fluctuation (Figure 3.3).39, 40 The correlation rate 51

of protein conformational fluctuation measured by single-molecule D-A fluorescence intensity fluctuation trajectories gives a broad distribution due to the local environment heterogeneity,

thermal fluctuations, dynamic disorder and static disorder.2-4 At low concentration of GdmCl, autocorrelation function analysis of single-molecule trajectories of CaM molecules shows a typical autocorrelation rate that corresponds to millisecond time scale, the characteristic

timescale of native state protein motions.2-4, 39, 40 Nevertheless, Figure 3.4B shows a significant

distinction of distributions of autocorrelation rates between different individual CaM molecules

under different concentration of denaturant GdmCl. The CaM molecules, in native states, show

more significant autocorrelation rates at millisecond time scale. However, the significant

correlation rates diminishes for unfolded CaM single-molecules, which is due to the unfolding of

CaM molecules turning to random coils under the denaturant GdmCl condition; and the correlation fluctuation rates of single-molecule protein regulated conformational motion becomes slower. The two-dimensional regional cross-correlation analyses help us to identify time segments along each single-molecule intensity trajectory with significant anti-correlations

(Figure 3.3F). We zoom in to study each of the anti-correlated time segments by calculating the

cross-correlation and autocorrelation respectively.

52

Figure 3.3. Correlation Analyses of Single-Molecule Data (A) A part of single-molecule intensity-time trajectories of the donor and acceptor from the long trajectories shown in Figure

2A. (B) Autocorrelation functions of the donor (green) and the acceptor (red) calculated from

single-molecule fluorescence intensity trajectory shown in (A). The time interval, we calculate

the autocorrelation functions and cross-correlation function is indicated by the blue frame. The

-1 fitted single exponentials (green and red) yield fluctuation rate kdonor = 2.2 ± 0.7 s and kacceptor=

-1 3.1 ± 0.6 s . (C) A part of calculated EFRET trajectory by using formula. (D) Cross-correlation

function from the single-molecule intensity-time trajectories of the donor and acceptor shown in 53

-1 (A). The fitted single exponentials (black) yield fluctuation rate kcross =2.0 ± 0.2 s which is

similar to the fluctuation rates captured by autocorrelation functions of the donor and the

acceptor. (E) Distribution of EFRET measured in (C) yields a mean of EFRET 0.42 ± 0.04. (F)

The result of analysis on the single-molecule donor-acceptor fluorescence trajectories shown in

(A). The cold color represents that the D-A intensity fluctuation is anti-correlated, whereas the warm color represents that the D-A is correlated.

The fluctuation dynamics of correlation function analyses yields fluctuation rates of conformational dynamics of single protein molecules.3, 4 Since the fitted exponential function of autocorrelation function has a similar fluctuation rate of donor and acceptor channel which indicates the autocorrelation come from the same source, the single-molecule protein

conformational fluctuations (Figure 3.3B and 3.3D). Fitted exponential function of

cross-correlation function yielding similar fluctuation rate further confirms the anti-correlated

cross-correlation function between the donor and acceptor fluctuation intensity trajectories with

the essentially same correlation rate within the error bar (Figure 3.3D), a typical anti-correlation

FRET D-A trajectory fluctuation dynamics showing a protein conformational motion measured by anti-correlated smFRET D-A intensity trajectories. Such protein conformational motion

captured by smFRET gives anti-correlated two-band D-A fluorescence intensity trajectories.

Both autocorrelation functions from the donor and acceptor signal trajectories and the anti-correlated cross correlation function between the donor and acceptor signal fluctuation trajectories have essentially the same fluctuation rate, which strongly indicates that the 54

fluctuations are dominated by the protein folding-unfolding conformational fluctuation probed

by the smFRET D-A signal fluctuation trajectory measurement (Figure 3). We focus on the autocorrelations from the donor channel only, and we plot the fitted exponential autocorrelation function rates at various concentrations of denaturants GdmCl in which single-molecule protein

CaM undergoes different conformational fluctuation dynamics as a result of different local environment. The time interval at which the autocorrelation functions are calculated, we also compute the EFRET (Figure 3.3A, 3.3C and 3.3E), and these two parameters, autocorrelation function fluctuation rate and EFRET, serve as a characteristic value to represent a local fluctuation dynamics. The two-dimensional contour plot of EFRET vs fitted autocorrelation function correlation rate is shown in Figure 4A. As the single CaM molecules get unfolded, not only the

EFRET decreases, but also the fluctuation rate of the autocorrelation function decreases.

Figure 3.4. Single-Molecule Conformational Fluctuation Dynamics. (A) Contour plot of EFRET

vs fitted autocorrelation function correlation rate. We see as the proteins become gradually

unfolded, the autocorrelation fluctuation rate of the protein dynamics become slower by a factor

of 100. We construct this contour plot by calculating autocorrelation function at a specific time

interval along a trajectory and compute the EFRET at this time interval (Figure 3.3A, 3.3C and 55

3.3E). These two parameters, autocorrelation function fluctuation rate and EFRET, correspond to one point on the contour plot, serving as a characteristic value to represent a local fluctuation dynamic behavior. (B) Distribution of fluctuation rate of correlation functions at various concentration of denaturant solvent. The autocorrelation fluctuation rates are 18 ± 10 s-1, 9 ±

11 s-1 , and 3 ± 3 s-1 at 0M GdmCl (blue), 1M GdmCl (green), and 2M GdmCl (red), respectively.

Quantitatively, the fitted exponential functions obtained from fitting the single-molecule

FRET intensity trajectory donor autocorrelation functions give the fluctuation rates from different single-molecule trajectories and distinct individual molecules at different concentrations of denaturant GdmCl. Single molecules experience different local environments and undergo distinct conformational fluctuation dynamics probed by autocorrelation function analyses.

Figure 4B, the distribution of these various autocorrelation fluctuation rates, shows a significant shift of the distribution as the concentration of GdmCl increases resulting in denaturing of single-molecule CaM which turns single-molecule CaM into random coils. From

autocorrelation fluctuation conformational dynamic analyses, the mean autocorrelation

fluctuation rates are 18 ± 10 s-1 , 9 ± 11 s-1 , and 3 ± 3 s-1 at 0M GdmCl, 1M GdmCl and 2M

GdmCl, respectively. We attribute the significantly different fluctuation dynamics to the CaM folding-unfolding conformational states and fluctuation dynamics under the different denature

GdmCl conditions, which create heterogeneous local environments. In 2M concentration of

GdmCl, CaM molecules denature into random coils without ordered secondary and tertiary structures, and the conformational dynamics probed by autocorrelation function reveals slow fluctuation dynamics at rate of 3 ± 3 s-1 which is a result of loss of regulated protein motion at the 56

timescale of millisecond as is expected for native state protein. The distribution of fluctuation

rates also become narrower as the single-molecule CaM unfolds in 2M denaturant GdmCl. The narrower distribution of conformational fluctuation rates is related to protein conformational space sampling speed which is an essential part of various recognition processes involved in biological function.2-4, 13, 35

To further understanding the folding-unfolding conformational fluctuation dynamics probed

by autocorrelation function analyses of single-molecule D-A fluorescence intensity trajectories

measured at the folding-unfolding titration equilibrium conditions, we use a kinetic model to account for the single-molecule protein CaM folding-unfolding conformational fluctuation

dynamics.3, 43, 44 Conventional kinetics experiments measure the relaxation of concentration of a

large ensemble of molecules after a perturbation (such as fast mixing or a temperature jump).

In contrast, single-molecule experiments measure the probability at specific states of an individual molecule conformational fluctuation, and the dynamic information can be extracted by measuring single-molecule spontaneous fluctuation dynamics at equilibrium; specified by the

Onsager’s regression hypothesis and the fluctuation dissipation theorem.44-46 The fluctuation dissipation theorem dictates that the relaxation of macroscopic non-equilibrium disturbances is governed by the same laws as the regression of spontaneous microscopic fluctuations in an equilibrium system.44-46 For a two-state spontaneous fluctuation dynamic model, the autocorrelation function directly probes the fluctuation dynamic process under detailed balance by reflecting the summation of forward and backward reaction rate kf + kb as the fluctuation rate of fitted exponential function of autocorrelation. We generalize this argument by expressing a 57

multiple-step fluctuation dynamic process by separately model the forward and backward reaction rate kf and kb by dividing the whole dynamic process into forward and backward half

reactions.43-46 For a three-state kinetic scheme, the mean first passage time of the rate process, or

reaction, is calculated by the flux method under detailed balance rate processes.43 We express

the observed forward and backward reaction rate by a polynomial of all reaction rates involved.

Where k1 and k-1 are the forward and backward reaction rates of step-one reaction, and k2 and k-2 are the forward and backward reaction rates of step-two reaction. In such a case, by assuming k-2 to be small, we derive the expression of the mean first passage time of the forward reaction:

kk 12 1 t . By further assuming k-1 equals zero, we get the mean first passage time of k1 k 2 k 2 11 the forward half reaction t . The reciprocal of t gives us the observed forward kk12 kk12 reaction rate kf  . By the same argument, the observed backward reaction rate is kk12 kk12 kb  . For a two-state dynamic scheme, the autocorrelation function is kk12

kfb k  t C t   e .3 For a generalized three-state (or easily generalized n-state) dynamic scheme, the

43-46 observed kf and kb is a polynomial of rate constant of each steps. It is clear that the

autocorrelation function analysis is a capable approach to probe the conformational fluctuations

of single proteins under spontaneous detailed balanced folding-unfolding conformational

fluctuations. By probing autocorrelation function fluctuation rates from single-molecule D-A

intensity fluctuation dynamic trajectories, we are able to directly probe and monitor protein

conformational fluctuation dynamics rates.47-52 58

3.3.3. Non-Exponential Distribution of Folding Waiting Time Indicates Multiple

Folding Intermediates.

Subtle conformational dynamic signatures are straightforwardly analyzed by the distribution

of on-times and off-times of mFRET D-A intensity fluctuation trajectories. 2, 3, 13, 53 The on-time

and off-time are the “waiting time” for the CaM folding and unfolding conformational dynamics, respectively. Such on-time and off-time or “waiting time” correspond to the dwelling time of

protein folding-unfolding conformational states. The nature of single-molecule CaM folding dynamic processes lies in the unique feature of such “waiting time” distributions. To characterize such detailed dynamic behavior, we further bin our data in the smFRET intensity trajectories to one millisecond bin (Figure 3.5A). The histogram of donor intensity trajectory

yields a distinct two-peaked occurrence distribution. The peak corresponding to the high

photon counts is unfolded state of single-molecule CaM, while the low photon count peak represents folded state of single-molecule CaM. To obtain a folding waiting time distribution,

we set up a threshold, at which the waiting times of the folding and unfolding states are

separately identified and read out. For the trajectory shown in Figure 3.5, we choose the value

10 as the threshold value, and subsequent folding waiting time distribution is shown in Figure

3.5C.3 Those short unfolding events that last less than three binning times of 3 ms total cannot be reliably differentiated from the measurement photon counting shot noise and are not counted as unfolding events.13, 53

Noticeably, the folding waiting time distribution is non-exponential resulting from a non-two-state folding-unfolding dynamic model.2, 3, 13 The distribution is also distinct from a 59

Gaussian distribution and has a broad coverage of time scales.3 The broad folding waiting time

distribution, reflecting the heterogeneity, is typically associated with complex protein

conformational dynamics involving multiple states and multiple pathways. This distinction clearly rules out the possibility of two-state folding-unfolding dynamics for single-molecule

CaM. The gamma distribution shaped folding waiting time distribution indicates a more complex dynamic mechanism involving a convolution of multiple Poisson processes of the folding-unfolding conformational fluctuations associated with multiple intermediates and multiple steps, which suggests multiple-state and multiple-pathway funnel-shaped folding energy landscape involving transition states, metastable states, and misfolded states.53, 63 Detailed spontaneous conformational fluctuation dynamics measurements yield such distinct statistics of folding waiting time distribution which, from dynamic perspective, is likely associated with a multiple-step Markovian dynamic process.53 Our detailed analysis of the folding waiting time distribution confirms a multiple folding pathways with multiple intermediates. Compared to a random search for the folded native state, our single-molecule spectroscopic analysis reveals that the CaM folding process is much more complex with multiple folding pathways running parallel with one another. Opposite to one-dimension reaction coordinate, single-molecule CaM folding process involving a folding network without rate-limiting step as considered in most two-state chemical reaction modeling of protein folding.53 60

Figure 3.5. Waiting-Time Distribution of Single-Molecule Data. (A) Donor intensity trajectory obtained from our single-molecule confocal microscopy with 1 ms binning time. The black dotted line indicates the threshold criterion separating the on- and off-time. (B) A histogram of such trajectory yields two peaks separating the “on” and “off” time which in our experiment correspond to folding and unfolding waiting time. Subsequently, a distribution of such “on” and “off” can be constructed. (C) The on-time distribution is a gamma shaped distribution representing a multiple step dynamic scheme. It is different in shape from both exponential distribution resulting from two-state dynamic scheme and Gaussian distribution. Simulated data (red line) from a multiple step Markovian dynamic process which facilitates a one dimensional random walk type of conformation diffusional process.

61

3.3.4. Model Analyses of Conformational Dynamics and Energy Landscape of

Single-Molecule CaM Folding.

We attribute the non-exponential folding waiting time distribution to the existence of multiple folding intermediates. 3, 13, 53, 54 To further characterize this multiple intermediate state dynamics, we exploit a one dimensional random walk model, which has been used successfully in the modeling of multiple step single-molecule enzymatic reaction.13, 53 The basic approach is

as the following: without prior knowledge of the shape of the energy landscape, we assume a

uniform rate k to each step of the overall rate process involved in single-molecule CaM

folding-unfolding conformational dynamics. Each step of one dimensional random walk model

is considered to be a Poisson process modeling single-molecule protein folding intermediate conversion process. The convolution of several Poisson process with uniform rate k gives

gamma function shaped distribution (Eqn. 2-4) to model non-exponential non-Gaussian shaped

folding waiting time distribution of single-molecule CaM folding-unfolding conformational

dynamics. This distribution reproduces the mean and standard deviation of the original folding

waiting time distribution of single-molecule CaM folding-unfolding conformational dynamics.

The number of Poisson process steps modeling single-molecule protein folding intermediate

conversion process involved in this convolution calculation gives an estimation of how many

Poisson rate processes are present in the overall rate process of single-molecule CaM folding-unfolding conformational dynamics. This number is the lower bound estimation of the number of intermediate involved in the dynamics process of protein folding. Quantitatively, the mean value of the original non-exponential folding waiting time distribution is 13.2 ± 9.0 ms. 62

The simulation data yields mean value 13.0 ± 7.2 ms which reproduces the original distribution well. The simulated distribution involves two Poisson rate processes which indicates a 2-step three-state dynamic process (Figure 3.5D).

P t  Aexp  t /  where P(t) is the probability distribution of the Poisson rate process step times, τ is the averaged Poisson rate process step time, and A is the distribution weight constant.

The Poisson rate process step time is the duration between two adjacent states, and it is different from the formation time of the intermediate states or folded state of single-molecule CaM folding-unfolding conformational dynamics. In our model the folding waiting time distritution is the convolution distribution of Poisson rate process step times. To calculate the convolution of function f(t) and g(t), the integration equation is

t f g  t   f  v  g  t  v  dv 0 Based on equation 6, a consecutive intermediate steps involved in a two-state dynamic model of

folding-unfolding conformational fluctuation is expressed as a convolution of two consecutive

exponential waiting-time distribution. The general probability function involving arbitrary

number of folding intermediates is deduced to be

n1   n ttexp /  PATn   n 1!  

Where n (1, 2, 3,…, N) is the index of the intermediate steps;  is the mean formation time of a

folding intermediate through a single-step process and A is the normalizing factor of this

probability distribution.

To further specify the shape of the energy landscape, we carry out a more detailed dynamic 63

analysis by calculating the conformational diffusion coefficients of single-molecule CaM

folding-unfolding conformational dynamics. Via a dynamic modeling, the folding waiting time

distribution gives the conformational diffusion coefficients.3 Briefly, we model the

folding-unfolding conformational dynamics of single-molecule CaM as a classical particle one-dimensional multiple-step random walk in the presence of a force field, and the position

distribution density function can be calculated by the Master equation.13 Following derivation of our previous work, the conformational diffusion coefficient of single-molecule CaM folding-unfolding dynamic process is

2 t2 X t  unfold N    D  3 2 tunfold

t where D is the conformational diffusion coefficient. The mean unfolding time, unfold , and the t 2 standard deviation of the unfolding time distribtuion, unfold , are directly measured in our

experiment. The total drifting distance of the folding-unfolding conformational motion,

XtN   , is associated to the folding-unfolding conformational distance change. From EFRET

distribution data, we already obtain the mean conformational drift distance XtN   to be 0.9 ±

1.0 nm. Using equation 8,13 we calculate the diffusion coefficient D of this two-step dynamic

process to be 1.4×10-13 cm2/s. The diffusion coefficient is directly related to the shape of the

underlying energy landscape, since it reflects the roughness of the potential energy surface of

single-molecule CaM folding-unfolding conformational dynamic process.

Theoretically, we estimate the size of single-molecule CaM in the presence of denaturant

GdmCl. Although there are different models involving complex intra-chain molecular 64

interactions, we choose Gaussian chain model because it is the most simplified model and

55-58 catches the essential properties of an unfolded protein. The unfolded radius of gyration Rg of

0.55 the single-molecule CaM protein in GdmCl, Rg = 0.345N nm is 5.4 nm, which is consistent

with experimental value of 6.1 ± 0.8 nm. N is the number of monomers in the polypeptide

56 chain. Such scaling relation of Rg is strongly supported by small-angle X-ray scattering. The

folding speed limits of single domain proteins are provided experimentally59, and the

single-molecule single domain protein folding time is well characterized under Gaussian chain

assumption.56, 59 The vast majority of measurements yield the formation time of loop is less than

0.1 μs, and α–helix formation is approximately 0.5 μs. The formation time of β hairpin is greater than 0.5 μs.59 Since linear scaling theory holds for small degree of polymerization, using the linear length scaling suggested by the homopolymer collapse theory,60 theoretical upper

bound of single-molecule CaM folding time τlimit of 148 amino acid CaM is 3.0μs. Since CaM

polypeptide chain is more like of a heteropolymer than a homopolymer there are additional

factors to be taken into account while describing protein collapse. The estimation of theoretical

upper bound of single-molecule CaM folding time τlimit based on homopolymer collapse theory is

essentially accurate.61 We note that the folding time τ = /(3D) with D as the diffusion

2 2 1.1 2 59 coefficient and = 6Rg = 0.7N nm is the mean square end-to-end distance of a polymer.

-7 2 We estimated Dlimit to be 1.9 × 10 cm /s as the theoretical upper bound of single-molecule CaM

conformational diffusion.59 The roughness of the free energy barrier involved in the

2 2 folding-unfolding dynamic process is determined by D = Dlimit exp(-β ε ), ε is a measurement of the roughness of the single-molecule CaM protein folding free energy barrier.62-64 For CaM 65

folding-unfolding conformational diffusion, the diffusion coefficient D is 1.4×10-13 cm2/s from our experimental measurement discussed above, which gives the roughness of the single-molecule CaM protein folding free energy barrier ε of 3.8±0.8kT; k is the Boltzmann

constant and T is the temperature. Compared with the hydrogen bonding (~2-12 kT),74, 75 this value of free energy barrier roughness is at the scale of hydrogen bonding interaction energy in a protein folding process associating with breaking and forming of a number of hydrogen bonds.

On the other hand, the height of the free energy barrier is estimated by using Kramer’s

barrier-crossing theory.59 We first estimate the actual folding time of CaM by using τ = /(3D)

2 0.5 to be 0.89s in which () is experimental value measured by EFRET 6.1 ± 0.8 nm and D is experimental value measured by the time bunching effect of single-molecule CaM folding

process. In our experiment the single-molecule CaM folding-unfolding conformational

diffusion coefficient D is 1.4×10-13 cm2/s. We use Kramer’s theory to estimate the empirical free energy barrier of single-molecule protein folding process. Kramer’s theory assumes that the dynamic process of protein folding can be described as a one-dimensional diffusion along a reaction coordinate, and the minimum and maximum of the free energy surface are parabolic.59

From Kramer’s theory,

τfolding = 2πτlimit exp(ΔG/kT) where τfolding is the folding time of single-molecule CaM measured in our experiment, τlimit is the theoretical upper bound of single-molecule CaM folding time, and ΔG Is the free energy

barrier and k is the Boltzmann’s constant T is the temperature. The height of the free energy

barrier ΔG is estimated to be 11kT, which is consistent with previous optical tweezers 66

measurement for CaM.63 Our measurement yields an free energy barrier slightly lower than

optical tweezers measurements, this difference is likely due to the distinct structure involved in the measurement assays using different denature methods resulting in slightly different energetic feature of energy landscape.65 Nevertheless, our analysis reveals that instead of a two-state dynamic scheme our study shows a more complex and higher dimensional dynamic process on a multiple-pathway multiple-state energy landscape. By recording detailed fluctuation dynamics at the thermodynamic equilibrium under the condition of 2M GdmCl, we characterize such unique dynamic feature via a conformational diffusional modeling of single-molecule CaM folding-unfolding conformational dynamic process.

Figure 3.6. Protein Folding Pathway Distribution. (A) We postulate a schematic representation

of protein folding process along one specific pathway based on our single-molecule spectroscopy

measurements. The nuclear coordinate Q is a projected one dimensional coordinate of a three 67

dimensional funnel shaped energy landscape. We estimate the roughness of the potential

energy surface and the free energy barrier height based on dynamic modeling. The roughness

of the potential surface 3.8±0.8kT is likely to be entropic traps which are the thermodynamic features of multiple folding intermediates observed in the folding-unfolding dynamic process of

CaM. The 11kT obtained from Kramer’s theory of barrier crossing is likely to be the empirical

overall free energy barrier height. (B) The distribution we constructed with number of steps involved in each of the dynamic process vs. conformational diffusion coefficient. The vertical axis is occurrence. (C) We extract the distribution of folding pathways of single CaM molecule by measurements of spontaneous conformational fluctuations at folding-unfolding

thermodynamic titration equilibrium. Different single-molecule CaM protein folding pathways

are labeled with different color with distinct probability. Two-step folding dynamic process

involving one intermediate labeled with red has a probability of 23% among all single-molecule

protein folding-unfolding conformational fluctuation dynamic analyses. Different folding

dynamic process involving multiple intermediate labeled with black, purple and blue which have

a probability of 17%, 33% and 21%. (D) A conceptual representation of single-molecule CaM

protein folding-unfolding conformational dynamic process measured by our single-molecule

FRET. In general a multiple-state folding dynamic process involving more than one

intermediate is observed.

To study the heterogeneity of the CaM folding process, we have analyzed 52 single-molecule protein folding-unfolding conformational fluctuation trajectories to acquire a distribution of conformational diffusion coefficient versus number of steps involving Poisson 68

rate processes (Figure 3.6B). For each of the single-molecule FRET folding-unfolding signal fluctuation trajectories, we simulate the number of steps involved in the single-molecule CaM folding-unfolding conformational dynamic process, and we also calculate the single-molecule

CaM folding-unfolding conformational diffusion coefficient. A two dimensional distribution is

constructed with conformational diffusion coefficient vs. number of steps (Figure 3.6B). This

distribution is clearly a manifestation of a multiple-pathway multiple-state energy landscape with distinct folding pathways on which single-molecule CaM navigate. Moreover, these various folding pathways have a distribution with distinct probabilities. We count the total number of

occurrence of each single-molecule CaM folding-unfolding fluctuations dynamics with same number of steps involving Poisson rate processes, and calculate the ratio of each protein folding pathway among all the single-molecule folding-unfolding fluctuation dynamics trajectories we measure. Our results suggest that there are at least four different folding pathways of CaM molecule folding process (Figure 3.6B and 3.6C) involving multiple intermediate states have different probabilities ranged from 17% to 33%, based on all single-molecule protein folding-unfolding conformational fluctuation trajectories analyzed (Figure 3.6C and 3.6D).

Furthermore, we have identified that the different protein folding pathways involve different

protein folding conformational diffusion coefficients, a significant indication of

multiple-pathway and multiple-states protein folding energy.18, 19 The comparable energy scale of

the roughness 3.8±0.8kT and hydrogen bonding is a further manifestation of a protein folding energy landscape with many local minima corresponds to multiple folding intermediates (Figure

3.6D). Presumably, the folding intermediates is likely to have partially folded domains of 69

single-molecule CaM.65

3.4 Conclusions

Using more sensitive single-molecule spectroscopic measurements, we have achieved the

specification of the underlying protein folding pathways by monitoring the protein

folding-unfolding dynamics at equilibrium. We have characterized the multiple-pathway

multiple-state folding dynamics and energy landscape of single CaM molecules under conditions

of various denaturant GdmCl by using single-molecule FRET spectroscopy measurements and

correlated model analysis.66-70 We utilized the protein folding-unfolding at the denaturant

titration midpoint condition in our single-molecule CaM folding-unfolding experiments to probe

single-molecule CaM undergo spontaneous folding-unfolding conformational fluctuations, identifying the correlated conformational fluctuation rates and single-molecule EFRET distributions. Folded state of single-molecule CaM corresponds to a high EFRET and a fast fluctuation rate at millisecond time scale, whereas unfolded state of single-molecule CaM

71-78 corresponds to a low EFRET and a slow fluctuation rate at second time scale. We characterized detailed single-molecule CaM folding-unfolding fluctuation dynamics by analyzing folding waiting time distribution that indicates the existence of multiple intermediate states.

Furthermore, we have identified the specific folding routes with distinct probability distributions.

The free energy barriers and roughness of free energy barrier calculated from the dynamic model based on our experimental results provide the primary energetic features of CaM folding-unfolding conformational dynamics. The comparable energy scales of roughness and free energy barrier suggest that the multiple intermediate states serve as entropic traps on the 70

folding energy landscape. More interestingly, our dynamic model gives an estimation of the

number of folding intermediates along each of the folding pathways. Overall, the approach of

our dynamics measurements under equilibrium dynamic is demonstrated to be effective and

powerful to explore the energy landscape and to verify the theoretical model of

multiple-pathway and multiple-states folding energy landscape based on fluctuation dissipation

theorem. Presumably, a fluctuating folding energy landscape with multiple-state

multiple-pathway is likely to be more energetically efficient and kinetically effective for a

consecutive protein folding dynamical process in living cells. 3.5 Reference:

1 For a review, see Nie, S.; Zare R. N. Optical detection of single molecules. Annu. Rev.

Biophys. Biomol. Struct. 1997, 26, 567-596.

2 Lu, H. P. Probing Single-molecule protein conformational dynamics. Acc. Chem. Res. 2005,

38 (7), 557–565.

3 Lu, H. P.; Xun, L. Y.; Xie, X. S. Single-molecule enzymatic dynamics. Science 1998, 282

(5395), 1877–1882.

4 English, B. P.; Min, W.; van Oijen, A. M.; Lee, K. T.; Luo, G. B.; Sun, H. Y.; Cherayil, B. J.;

Kou, S. C.; Xie, X. S. Ever-fluctuating single enzyme molecules: Michaelis-Menten equation

revisited. Nat. Chem. Biol. 2006, 2 (2), 87–94

5 Schuler, B.; Lipman, E. A.; Eaton, W. A. Probing the free-energy surface for protein folding

with single-molecule fluorescence spectroscopy. Nature 2002, 419, 743-747.

6 Lu, H. P.; Iakoucheva, L. M.; Ackerman, E. J. Single-molecule conformational dynamics of 71

fluctuating noncovalent DNA-Protein interactions in DNA damage recognition. J. Am. Chem.

Soc. 2001, 123 (37), 9184-9185.

7 Tan, X.; Nalbant, P.; Toutchkine, A.; Hu, D.; Vorpagel, E. R.; Hahn, K.M.; Lu, H. P.

Single-molecule study of protein−protein interaction dynamics in a cell signaling system. J.

Phys. Chem. B. 2004, 108 (2), 737-744.

8 Liu, S. X.; Bokinsky, G.; Walter, N. G.; Zhuang, X. W. Dissecting the multistep reaction

pathway of an RNA enzyme by single-molecule kinetic “fingerprinting”. Proc. Natl. Acad.

Sci. U.S.A. 2007, 104 (31), 12634–12639.

9 Visscher, K.; Schnitzer, M. J.; Block, S. M. Single kinesin molecules studied with a

molecular force clamp. Nature 1999, 400, 184–189.

10 Roy, R.; Hohng, S.; Ha, T. A practical guide to single-molecule FRET. Nat. Methods 2008, 5,

507–516.

11 Selvin, P. R.; Ha, T. Single-Molecule Techniques: A Laboratory Manual; Cold Spring Harbor

Laboratory Press: Cold Spring Harbor, NY, 2008.

12 Ha, T. J.; Ting, A. Y.; Liang, J.; Caldwell, W. B.; Deniz, A. A.; Chemla, D. S.; Schultz, P. G.;

Weiss, S. Single-molecule fluorescence spectroscopy of enzyme conformational dynamics

and cleavage mechanism. Proc. Natl. Acad. Sci. U.S.A. 1999, 96 (3), 893–898.

13 Chen, Y.; Hu, D. H.; Vorpagel, E. R.; Lu, H. P. Probing single-molecule T4 lysozyme

conformational dynamics by intramolecular fluorescence energy transfer. J. Phys. Chem. B

2003, 107 (31), 7947–7956.

14 Anfinsen, C.; Haber, E.; Sela, M.; White, F. H. The kinetics of formation of native 72

ribonuclease during oxidation of the reduced polypeptide chain. Proc. Natl. Acad. Sci. USA

1961, 47 (9), 1309–1314.

15 For a review, see Eaton, W. A.; Schuler, B. Protein folding studied by single molecule FRET.

Curr. Opin. Struct. Biol. 2008, 18 (1), 16-26.

16 For a review, see Schuler, B.; Hofmann, H. Single-molecule spectroscopy of protein folding

dynamics—expanding scope and timescales. Curr. Opin. Struct. Biol. 2013, 23 (1), 36-47.

17 For a review, see Rief, M. Force as a single molecule probe of multidimensional protein

energy landscapes. Curr. Opin. Struct. Biol. 2013, 23 (1), 48-57.

18 Onuchic, J. N.; Luthey-Schulten, Z.; Wolynes, P. G. Theory of protein folding: the energy

landscape perspective. Annu. Rev. Phys. Chem. 1997, 48, 545–600.

19 Dill, K. A.; Chan, H. S. From Levinthal to pathways to funnels. Nat. Struct. Biol. 1997, 4,

10–19.

20 Thirumalai, D.; Klimov, D. K. Deciphering the timescales and mechanisms of protein folding

using minimal off-lattice models. Curr. Opin. Struct. Biol. 1999, 9 (2), 197–207.

21 Deniz A. A.; Laurence T. A.; Beligere G. S.; Dahan M.; Martin A. B.; Chemla D. S.; Dawson

P. E.; Schultz P. G.; Weiss S. Single-molecule protein folding: Diffusion fluorescence

resonance energy transfer studies of the denaturation of chymotrypsin inhibitor 2. Proc. Natl.

Acad. Sci. USA 2000, 97 (10), 5179-5184.

22 Nettels D.; Gopich I. V.; Hoffmann A.; Schuler B. Ultrafast dynamics of protein collapse

from single-molecule photon statistics. Proc. Natl. Acad. Sci. USA 2007, 104 (8), 2655-2660.

23 Nettels D.; Hoffmann A.; Schuler B. Unfolded Protein and Peptide Dynamics Investigated 73

with Single-Molecule FRET and Correlation Spectroscopy from Picoseconds to Seconds. J.

Phys. Chem. B 2008, 112 (19), 6137-6146.

24 Haran G. How, when and why proteins collapse: the relation to folding. Curr. Opin. Struct.

Biol. 2012, 22 (1), 14-20.

25 Sherman E.; Haran G. Coil–globule transition in the denatured state of a small protein. Proc.

Natl. Acad. Sci. USA 2006, 103 (31), 11539-11543.

26 Ziv G.; Thirumalai D.; Haran G. Collapse transition in proteins. Phys. Chem. Chem. Phys.

2009, 11, 83-93.

27 Hoffmann A.; Kane A.; Nettels D.; Hertzog D. E.; Baumgärtel P.; Lengefeld J.; Reichardt G.;

Horsley D. A.; Seckler R.; Bakajin O.; Schuler B. Mapping protein collapse with

single-molecule fluorescence and kinetic synchrotron radiation circular dichroism

spectroscopy. Proc. Natl. Acad. Sci. USA 2007, 104 (1), 105-110.

28 Chung H. S.; Louis J. M.; Eaton W. A. Experimental determination of upper bound for

transition path times in protein folding from single-molecule photon-by-photon trajectories.

Proc. Natl. Acad. Sci. USA 2009, 106 (29), 11837-11844.

29 Chung H. S.; McHale K.; Louis J. M.; Eaton W. A. Single-molecule fluorescence

experiments determine protein folding transition path times. Science 2012, 335 (6071),

981-984.

30 Shaw D. E.; Maragakis P.; Lindorff-Larsen K.; Piana S.; Dror R. O. Atomic-level

characterization of the structural dynamics of proteins. Science 2010, 330 (6002), 341-346.

31 Garcia-Mira M. M.; Sadqi M.; Fischer N.; Sanchez-Ruiz J. M.; Munoz V. Experimental 74

identification of downhill protein folding. Science 2002, 298 (5601), 2191-2195.

32 Pirchi M.; Ziv G.; Riven I.; Cohen S. S.; Zohar N.; Barak Y.; Haran G. Single-molecule

fluorescence maps the folding landscape of a large protein. Nat. Commun. 2011, 2, 493.

33 Chin, D.; Means, A. R. Calmodulin: a prototypical calcium sensor. Trends in Cell Biology

2000, 10 (8), 322-328.

34 James, P.; Vorherr, T.; Carafoli, E. Calmodulin-binding domains: just two faced or

multi-faceted? Trends Biochem. Sci. 1995, 20 (1), 38-42.

35 Liu, R. C.; Hu, D. H.; Tan, X.; Lu, H. P. Revealing two-state protein-protein interactions of

calmodulin by single-molecule spectroscopy. J. Am. Chem. Soc. 2006, 128 (31), 10034–

10042.

36 Babu, Y. S.; Sack, J. S.; Greenhough, T. J.; Bugg, C. E.; Means, A. R.; Cook, W. J.

Three-dimensional structure of calmodulin. Nature 1985, 315, 37-40.

37 Chattopadhyaya, R.; Meadorl, W. E.; Means, A. R.; Quiocho, F. A. Calmodulin structure

refined at 1.7 A resolution. J. Mol. Biol. 1992, 228 (4), 1177-1192.

38 Slaughter B. D.; Unruh J. R.; Price E. S.; Huynh J. L.; Bieber Urbauer R. J.; Johnson C. K.

Sampling unfolding intermediates in calmodulin by single-molecule spectroscopy. J. Am.

Chem. Soc. 2005, 127 (34), 12107-12114.

39 Wang, X.; Lu, H. P. 2D Regional correlation analysis of single-molecule time trajectories. J.

Phys. Chem. B. 2008, 112 (47), 14920-14926.

40 He, Y.; Li, Y.; Mukherjee, S.; Wu, Y.; Yan, H.; Lu, H. P. Probing single-molecule enzyme

active-site conformational state intermittent coherence. J. Am. Chem. Soc. 2011, 133 (36), 75

14389-14395.

41 Nie, S.; Chiu, D. T.; Zare R. N. Probing individual molecules with confocal fluorescence

microscopy. Science 1994, 266 (5187), 1018–1021.

42 We have obtained the SEM imaging of agarose gel by super-critical point sample preparation

at different concentration of agarose gel from 0.2% to 2%. The imaged pore size formed by

the network of agarose gel averagely rage from 300 nm to 50 nm. Single-molecule CaM

freely rotate and perform biological function in agarose gel matrix and the translational

motion of individual CaM does not extend beyond the laser diffraction limited focus spot.

43 Cao, J. S. Michaelis−Menten equation and detailed balance in enzymatic networks. J. Phys.

Chem. B 2011, 115 (18), 5493-5498.

44 Chandler, D. Introduction to Modern ; Oxford University Press: New

York, 1987.

45 Oppenheim, I.; Shuler, K. E.; Weiss, G. H. Stochastic Processes in Physics and Chemistry;

MIT Press: Cambridge, MA, 1977.

46 Huang, K. Lectures on Statistical Physics and Protein Folding; World Scientific Publishing:

Singapore, 2005.

47 Barkai, E.; Jung, Y.; Silbey, R. Time-dependent fluctuations in single molecule spectroscopy:

a generalized Wiener-Khintchine approach.Phys. Rev. Lett. 2001, 87 (20), 207403

48 Barkai, E.; Silbey, R.; Zumofen, G. Transition from simple to complex behavior of single

molecule line shapes in disordered condensed phase. J. Chem. Phys. 2000, 113, 5853–5867.

49 Flomenbom, O.; Klafter, J.; Szabo, A. What Can One Learn from Two-State Single-Molecule 76

Trajectories? Biophys. J. 2005, 88 (6), 3780–3783.

50 Flomenbom, O.; Velonia, K.; Loos, D.; Masuo, S.; Cotlet, M.; Engelborghs, Y.; Hofkens, J.;

Rowan, A. E.; Nolte, R. J. M.; Van der Auweraer, M.; de Schryver, F. C.; Klafter, J. Stretched

exponential decay and correlations in the catalytic activity of fluctuating single lipase

molecules. Proc. Natl. Acad. Sci. U.S.A. 2005, 102 (31), 2368–2372.

51 He, Y.; Barkai, E. Super- and sub-Poissonian photon statistics for single molecule

spectroscopy. J. Chem. Phys. 2005, 122 (18), 184703.

52 Zwanzig, R. Nonequilibrium Statistical Mechanics; Oxford University Press: New York,

2001.

53 Wang, Y.; Lu, H. P. Bunching effect in single-molecule T4 lysozyme nonequilibrium

conformational dynamics under enzymatic reactions. J. Phys. Chem. B 2010, 114 (19),

6669-6674.

54 Xie, X. S. Single-molecule approach to enzymology. Single Mol. 2001, 4, 229-236.

55 Uversky, V. N. Natively unfolded proteins: A point where biology waits for physics. Protein

Sci. 2002, 11 (4), 739–756.

56 Ziv, G.; Thirumalai, D.; Haran, G. Collapse transition in proteins. Phy. Chem. Chem. Phys.

2009, 11, 83-93.

57 de Gennes, P.-G. Scaling concepts in polymer physics; Cornell University Press: Ithaca,

1979.

58 Des Cloizeaux, J.; Jannink, G. Polymers in solution: their modeling and structure; Clarendon

Press: Oxford, 1990. 77

59 Kubelka, J. J.; Hofrichter, J. J.; Eaton, W. A. The protein folding 'speed limit'. Curr. Opin.

Struct. Biol. 2004, 14, 76-88.

60 Pitard E. Influence of hydrodynamics on the dynamics of a homopolymer. Eur. Phys. J. B

1999, 7, 665-673.

61 The more sophisticated heteropolymer model considering the complexity of specific amino

acids interactions in the polypeptide chain may be more accurate. Such discussion requires

detailed simulation study which is beyond the scope of our current paper.

62 Socci N. D.; Onuchic J. N.; Wolynes P. Diffusive dynamics of the reaction coordinate for

protein folding funnels. J. Chem. Phys 1996, 104 (15), 5860-5868.

63 Stigler, J.; Ziegler, F.; Gieseke, A.; Gebhardt, J. C.; Rief, M. The complex Folding network of

single calmodulin molecules. Science 2011, 334 (6055), 512-516.

64 Zwanzig, R. Diffusion in a rough potential. Proc. Natl. Acad. Sci. U.S.A. 1988, 85,

2029-2030.

65 Stirnemann, G.; Kang, S.; Zhou, R.; Berne, B. J. How force unfolding differs from chemical

denaturation. Proc. Natl. Acad. Sci. U.S.A .2014, 111 (9), 3413-3418

66 Wang, J.; Oliveira, R. J.; Chu, X.K.; Whitford, P. C.; Chahine, J.; Wei, H.; Wang, E.K.;

Onuchic, J. N.; Leite, V. B. P. Topography of funneled landscapes determines the

thermodynamics and kinetics of protein folding. Proc. Natl. Acad. Sci. USA, 2012, 109 (39),

15763–8.

67 Dutta, P.; Sen, P.; Halder, A.; Mukherjee, S.; Sen, S.; Bhattacharyya, K. Solvation dynamics

in a protein–surfactant complex. Chemical Physics Letters 2003, 377 (1), 229-235. Pal, N.; 78

Verma, S. D.; Sen, S. Probe position dependence of DNA dynamics: comparison of the

time-resolved stokes shift of groove-bound to base-stacked probes. J. Am. Chem. Soc. 2010,

132 (27), 9277-9279.

68 Sen, S.; Paraggio, N. A.; Gearheart, L. A.; Connor, E. E.; Issa, A.; Coleman, R. S.; Wyatt, M.

D.; Berg M. A. Effect of protein binding on ultrafast DNA dynamics: characterization of a

DNA: APE1 complex. Biophysical Journal 2005, 89 (6), 4129-4138

69 Biju, V.; Anas, A.; Akita, H.; Shibu, E. S.; Itoh, T.; Harashima, H.; Ishikawa M. FRET from

quantum dots to photodecompose undesired acceptors and report the condensation and

decondensation of plasmid DNA. ACS nano 2012, 6 (5), 3776-3788

70 Zheng, D.; Kaldaras, L.; Lu, H. P. Total internal reflection fluorescence microscopy

imaging-guided confocal single-molecule fluorescence spectroscopy. Rev. Sci. Instrum. 2012,

83 (1), 013110−013110−5.

71 Zheng, D.; Lu, H. P. Single-molecule enzymatic conformational dynamics: spilling out the

product molecules. J. Phys. Chem. B. 2014, 118 (31), 9128-9140.

72 Lu, M.; Lu, H. P. Probing protein multidimensional conformational fluctuations by

single-molecule multiparameter photon stamping spectroscopy. J. Phys. Chem. B. 2014, 118

(41), 11943-11955.

73 Bartko, A. P.; Dickson R. M. Imaging three-dimensional single molecule orientations. J. Phys.

Chem. B. 1999, 103 (51), 11237–11241

74 Sticke, D. F.; Presta, L. G.; Dill, K. A.; Rose, G. D. Hydrogen bonding in globular proteins. J.

Mol. Biol. 1992, 226, 1143-1159. 79

75 Rose, G. D.; Wolfenden, R. Hydrogen-bonding, hydrophobicity, packing, and protein-folding.

Annu. Rev. Bioph. Biom. 1993, 22, 381−415.

76 Bates, M.; Huang, B.; Dempsey, G. T.; Zhuang, X. W. Multicolor super-resolution imaging

with photo-switchable fluorescent probes. Science 2007, 317, 1749-1753.

77 Hohng, S.; Joo, C.; Ha, T. Single-molecule three-color FRET. Biophys. J. 2004, 87,

1328-1337.

78 Lee, N. K.; Kapanidis, A. N.; Koh, H. R.; Korlann, Y.; Ho, S. O.; Kim, Y.; Gassman, N.; Kim,

S. K.; Weiss, S. Three-color alternating-laser excitation of single molecules: Monitoring

multiple interactions and distances. Biophys. J. 2007, 92, 303-312.

80

CHAPTER IV. PROBING SINGLE-MOLECULE PROTEIN

FOLDING-UNPON-BINDING CONFORMATIONAL DYNAMICS USING

SINGLE-MOLECULE FRET SPECTROSCOPY

This chapter is dedicated to the study of single molecule protein folding study in the presence of protein-protein interaction.

4.1 Introduction

Single-molecule spectroscopy developed in the 1990s is by far the most important technique to study one molecule at a time. The dynamics and heterogeneity revealed by single-molecule experiments prove to be invaluable when researchers try to understand biological processes at

the molecular level, such as cell signaling.1-6 Recently, new developments in this technique also

make it possible to probe various interaction pathways in which real biological processes

occur.7-10 For example, one study involving investigation of protein-protein interactions of

calmodulin (CaM) with peptide shows these interactions can significantly alter the normal

conformational dynamics of the target protein.11

After gene expressions, single proteins start to fold into a unique functional structure without external help. However, inside the living cells, this process normally happens in a crowded

environment which involves various kinds of protein-protein interaction. Little is known about

the folding dynamics of single CaM molecules in the presence of extensive interactions with

other peptide molecules. The motivation of this study is to probe the folding-unfolding

conformational dynamics, which is the most important conformational dynamics of single 81

proteins, in this “crowded” environment. The outcome of this study will help us to understand the critical role played by protein-protein interactions in the single protein folding process, which

can provide insight into the future study of protein misfolding and protein aggregates related to

human diseases (e.g., amyloid diseases).12

Calmodulin (CaM) is a 148-residue protein responsible for intracellular calcium-sensing. It

is crucial in many biological processes including muscle contraction and energy metabolism.6,7

CaM has two globular domains and the binding mechanism between calcium ions and the two domains of CaM are under extensive studies.8,9 Traditional methods involving nuclear magnetic resonance and X-ray crystallography have provided detailed insights into the mechanisms and dynamics of this reaction.7-9 As a result of the flexibility and the dumbbell shape of the protein,

with dye labeling at the two domains, it is possible to monitor the dynamics of CaM with

single-molecule spectroscopy, such as fluorescence resonance energy transfer (FRET).12 The

typical time-resolution can be milliseconds which is the characteristic time scale of protein

motions.

Cyanine is a commonly used dye molecule, which can be attached onto the protein with

thiolation. The CaM has mutations at residue 34 and residue 110 on the N- and C-terminal

respectively, and fluorescent dye pair Cy3/Cy5 was tethered onto the protein. These two dyes

served as a spectroscopic ruler for the measurement of conformational fluctuations of the protein

molecules, since the measurement efficiency of energy transfer between the two dye molecules

will reveal distance information of the protein conformation: 82

Where R0 is the distance at which energy transfer is 50% efficient. For our specific dye pair

Cy3-Cy5, the R0 value is measured to be 5.4 nm.

Figure 4.1. Protein Folding Binding Process. A cartoon of protein folding-binding process.

C28W is an oligomer of 28 amino acid residues which can interact with CaM through

binding to one of the two terminals of the two domains. Previous research proved a two-state

interaction picture for the CaM/C28W and figured out the general induced bending motion of the

CaM through interaction with C28W.11

In this chapter, we use single-molecule FRET spectroscopy to probe the folding-unfolding

conformational dynamics of CaM interacting with C28W under the mild unfolding concentration

of denaturant solvent guanidinium chloride (GdmCl) (Figure 2). The critical concentration of

denaturant of GdmCl, at which single proteins undergo folding-unfolding conformational

fluctuations, has been characterized by our previous study. Since the interaction characterized by

previous studies revealed an induced binding picture in native states of CaM, we observe a more

complicated dynamic picture of CaM interacting with C28W under partially denatured

conditions at which single CaM molecules have large conformational space to explore.

83

4.2 Materials and Methods

4.2.1 Sample Preparation and Characterization

The single CaM molecule has mutation on N-terminal domain at residue 34 and C-terminal

domain at residue 110, and fluorescent dye pair Cy3/Cy5 was tethered onto the protein via

thiolation. These two dyes served as a spectroscopic ruler for the measurement of conformational

fluctuation of the protein molecule interacting with C28W. In our experiment, the change of

FRET efficiency corresponds to the distance change of about 2 nm. On the other hand, the

Forster radius R0 of Cy3-Cy5 pair is ~5 nm, so the distance change of the two dyes monitoring the conformational change falls into measureable range. In our experiment, the samples for single-molecule conformational folding-unfolding dynamics measurements are prepared inside the agarose gel (1% by weight, Type VII, Sigma). For example, we make a sample of CaM into 2

M of denaturant solvent GdmCl by first mixing 0.2μL 200 nM CaM, 2.5μL 8 M GdmCl, 0.1mM

CaCl2 and 8.55μL oxygen scavenger with Trolox solution to obtain a 10μL mixture of protein and denaturant solvent with 100nM C28W. The Trolox solution is made previously by dissolving about 1 mM 6-hydroxy-2,5,7,8-tetramethylchroman-2-carboxylic acid (Trolox) in 1 mg/mL

glucose oxidase, 0.8% D-glucose and 0.04 mg/mL catalase in order to protect fluorescent dye from photobleaching or blinking as a result of triplet state formation or other irreversible

photophysical processes. We then heat the 10μL 1% agarose gel just above its gel-transition

temperature (26°C) and quickly mix the above protein with denaturant solvent solution with the

gel between two clean cover glasses to form a sandwich. All solutions are prepared with

Phosphate buffered saline (PBS buffer) at pH 7.4. Since we want to probe conformational 84

dynamics of single protein molecule at different concentration of C28W peptide, we carry out concentration dependent experiments with different ratio of mixture in the sandwich as listed below.10

Trolox+Oxygen CaM GdmCl Agarose Gel Scavenger ~1 nM CaM+2M 0.2μL 200nM 2.5μL 8M 9.8μL 10μL Gdmcl+100 nM C28W ~1 nM CaM+2M 0.2μL 200nM 2.5μL 8M 8.55μL 10μL Gdmcl+500 nM C28W ~1 nM CaM+2M 0.2μL 200nM 2.5μL 8M 7.3μL 10μL Gdmcl+1000 nM C28W

Table 4.1. List of concentrations used in our folding-binding experiments.

4.2.2 Single-Molecule Imaging and FRET Measurement

Single-Molecule photon stamping approach is used to record FRET trajectories at different concentration of C28W. This approach records emission of photons from donor and acceptor

channel one by one with arrival time, and the intensity trajectory can be constructed as a function of time. Förster resonance energy transfer or fluorescence resonance energy transfer (FRET) is a certain type of energy transfer realized by nonradiative dipole-dipole interaction (Figure

4.4). The experimental setup is an inverted confocal microscope (Axiovert 200, Zeiss) which uses a crystal laser (532nm CW) as the light source for excitation (Figure 5). The Laser beam focused through a 100× oil immersion objective lens (1.3 NA, 100×, Zeiss) onto the upper surface of cover slip after reflected up by a dichroic beam splitter (z532rdc, Chroma Technology).

To obtain confocal microscopy image, we use an x-y closed-loop piezo position scanning stage for raster-scan of the sample sandwich. The fluorescence is collected through the same objective 85

and the signal is split by a dichroic beam splitter (640dcxr) into two different wavelength 570nm

and 670nm which are the emission wavelength of Cy3 and Cy5 respectively. To detect the signal

from the two channels, two Si avalanche photodiode single photon counting modules

(SPCM-AQR-16, Perkin Elmer Optoelectronics) were used for recording the photons from donor

and acceptor. A more detailed description of the experimental setup is in the literature.10,11

4.3 Results and Discussion

Figure 4.2 shows typical image obtained by our inverted confocal microscopy. By raster-scan the sample, a 20μm × 20μm sample image gives bright spots of single CaM molecules confined in the network formed by agarose gel (~200nm in diameter). The spots indicating single protein molecules are diffraction limited (~300nm). We then focus our laser

beam on each specific sample spot to gain continuous donor-acceptor (D-A) fluorescence

intensity trajectories shown in figure 2. Typically, we gather approximately seventy data points at

different concentration of C28W peptide, and the FRET efficiency of each of the sample

measured is calculated.

12-15 Here the correction factor (φA×ηA)/(φD×ηD) is ~1 in our experiment conditions. 86

Figure 4.2. Single-Molecule Imaging and Correlation Analysis. Typical Donor/Acceptor signals

and FRET efficiency trajectories. Green and red lines indicate the donor and acceptor channel respectively. The histogram distribution from FRET Efficiency trajectory gives the FRET efficiency of a certain single-molecule.

We observe shifts in the peak values of the distribution of all efficiencies at different concentration of C28W interacting with folding-unfolding fluctuating CaM resulting in either decrease or increase of FRET efficiency (Figure 3). These changes in FRET efficiency clearly evidence the conformational change of single CaM molecules interacting with different concentration of C28W. The samples with FRET efficiency at around 0.6 are assigned to be in the native state and those with FRET efficiency at 0.3 are assigned to be the unfolded ones. In this study, we only focus on those fluctuating CaM molecules which interact with different concentrations of C28W, so to obtain different distribution at various concentration of denaturant 87

solvent is unnecessary. When the folded and unfolded single CaM molecules are approximately

equally populated, those molecules with FRET efficiency at 0.4 are those fluctuating proteins

interacting with C28W peptide.

Figure 4.3. Protein Folding Binding and EFRET. Distribution of FRET efficiency of different

samples at different concentration of GdmCl. The shift in peak values characterizes the unfolding

of CaM molecule at higher concentration of denaturant solvent.

We analyze the dynamical behaviors of our data sets by calculating the autocorrelation functions. By fitting the autocorrelation functions with exponential decays, we are able to 88

characterize the time scale of protein motion (Figure 1). Each fitted autocorrelation function gives us correlation lifetime of FRET efficiency, and the lifetimes have a broad distribution as a result of environmental heterogeneity. We also observe that at high concentration of C28W, CaM has a typical autocorrelation function which has a sharp decay as a result of our induced folding hypothetical picture.15-21 The shift of FRET efficiency distribution reveals the interesting induced folding or induced unfolding phenomenon.

The exponential decays obtained from fitting the autocorrelation functions give decay

lifetime from different sample points at different concentration of GdmCl. We plot distribution of

these various fluctuation lifetimes and an interesting shift of the distribution as the concentration

of GdmCl increased emerged (Figure 1). In our previous work, at 2M of denaturant solvent, there

are more big taus compared to the lower concentration and blank samples. We attribute this change to denature of the CaM molecule. At high concentration of GdmCl, most of the CaM molecules are denatured into random coil, and the conformational dynamics reveal weak correlation (slow protein motion at time scale of ~10s).22-28 89

Figure 4.4. Waiting Time Distribution Analysis of Protein Folding. Single-molecule

folding-unfolding fluctuation trajectory is shown on the upper channel. The distinct high-low

fluctuation pattern indicates a protein folding-unfolding conformational dynamics. We analyzed such folding-unfolding process by the bunching effect and simulated the number of possible intermediate states.

The conformational diffusion coefficients can also be calculated from the folding waiting time distribution. From FRET efficiency distribution data, we already obtained the mean

29-35 conformational drift distance XtN   to be 1.22nm. Using this formula, we calculate the

diffusion coefficient of this 2-step dynamic process to be D~2.92×10-13 cm2/s. The diffusion coefficient is directly related to the underlying energy landscape, since it represents the local 90

mean square fluctuation of the activation barrier of the folding-unfolding dynamic process.

2 t2 X t  unfold N    D  3 2 tunfold

Theoretically, using polymer theory,36-39 the unfolded radius of gyration of the protein in

0.55 GdmCl can be calculated Rg = 0.345N nm which is 5.39 nm (experiment value 6.70 nm). The folding speed limits of single domain proteins are provided experimentally40 and the folding time well characterized with a Gaussian chain assumption.40 Using the linear length scaling suggested

41 by the homopolymer collapse theory, theoretical upper bound folding timeτlimit of 148 amino acid CaM is 3 μs. Notice folding time τ = /(3D) with D as the diffusion coefficient and

2 2 1.1 2 -8 2 = 6Rg = 0.7N nm . So Dlimit can be estimated to be 18.97 × 10 cm /s. The local mean square fluctuation of the activation barrier involved in the folding-unfolding dynamic process

2 2 -13 can be determined by D = Dlimitexp(-β ΔE ). In our case, diffusion coefficient D~2.92×10

cm2/s which givesΔE is about 3.7 kT. Compared with hydrogen bonding (2 – 12kT), this value of activation barrier fluctuation is at reasonable scale. The net increase of such “energy landscape

roughness” is about 20%, a significant change which could be due to a recognition process

between the ligand and the target protein. On the other hand, the height of the activation barrier

can be estimated using Kramer’s dynamics.40 The actual folding time of CaM can be calculated to be 0.26s. From Kramer’s dynamics, τfolding = 2πτlimit exp(ΔG/kT). So the height of the

activation barrierΔG = 10.2 kT, consistent with previous optical tweezers measurement.42 91

Figure 4.5. Folding Pathway Distribution in Protein Folding Binding. The single-molecule protein folding pathway distribution features a multiple-channel and multiple-state

conformational dynamics. The convergence of such a distribution clearly revealed a

conformational-selection mechanism of folding-upon-binding dynamical process.

We analyze 35 trajectories to acquire a distribution of such dynamic processes. We characterized each single protein folding-unfolding conformational dynamics by calculating the

diffusion coefficients and the number of intermediate states. Folding-unfolding trajectories with

same number of steps and similar conformational gathered into different domains which is a

manifestation of the underlying multi-dimensional energy landscape. This cluster is a clear proof

of the existence of folding pathway distribution. Moreover, these various folding pathways have

a distribution with different probabilities assigned to each of them. To better see this, we

constructed a two dimensional color plot, with the color bar indicating occurrence. Four different

folding pathways started to emerge, and the probability of each of these pathways are calculated 92

(Figure 4.5).

4.4 Conclusion

In this study, we characterize the conformational dynamics of single CaM under interaction

with C28W. We denature single CaM molecules by adding in 2M denaturant GdmCl, and the

ratio between folded and unfolded CaM molecules is approximately one to one. Under this

condition, the majority of the single protein undergoes folding-unfolding fluctuation which has

been studied by our previous research. Our hypothesis is that the addition of C28W peptide,

which can bind to the two heads of single CaM molecules, will significantly change the

conformation of the single protein and further enhance the folding process of single proteins. We

probe this folding process by measuring single-molecule FRET which can reveal conformational

dynamic information.43

The limitation of our present study will lie in the complexity of the effect on proteins of

denaturant, the chemical GdmCl. It is well studied that the effect of the chemical denaturant

GdmCl on protein will be the destruction of the secondary and tertiary structure. The unfolding

process is a result of loss of hydrophobic interaction between amino acids. However, the binding

process of C28W to CaM is also a result of hydrophobic amino acid interactions between these

two. Whether the denaturant will also have any effect on the C28W-CaM interaction remains

unknown. This effect could intrinsically make our experiment more complex. However, if we

can observe a substantial change in the measured distribution of FRET efficiency, we know the

C28W-CaM interaction is active and we can carry out our following dynamic analyses. 93

4.5 Reference:

1 Moerner, W. E.; Orrit, M. Illuminating Single Molecules in Condensed Matter Science 1999,

283 (5408), 1670-1676.

2 Nie, S.; Zare, R. N. Optical Detection of Single Molecules. Annual Review of Biophysics and

Biomolecular Structure 1997, 26, 567-596.

3 Weiss, S. Fluorescence Spectroscopy of Single Biomolecules. Science 1999, 283 (5408),

1676-1683.

4 For reviews, see: Acc. Chem. Res. (special issue on single molecule spectroscopy) 2005, 38

(7).

5 For a review, see: Protein-Ligand Interactions;Nienhaus, G. U., Ed.; Humana Press: New

Jersey, 2005.

6 Chin, D.; Means, A. R. Trends in Cell Biology 2000, 10 (8), 322-328.

7 James, P.; Vorherr, T.; Carafoli, E. Calmodulin: a prototypical calcium sensor. Trends

Biochem. Sci. 1995, 20 (1), 38-42.

8 Babu, Y. S.; Sack, J. S.; Greenhough, T. J.; Bugg, C. E.; Means, A. R.; Cook, W. J.

Three-dimensional structure of calmodulin. Nature 1985, 315, 37-40.

9 Chattopadhyaya, R.; Meadorl, W. E.; Means, A. R.; Quiocho, F. A. Calmodulin structure

refined at 1.7 Å resolution. J. Mol. Biol. 1992, 228 (4), 1177-1192.

10 Chen, Y.; Hu, D. H.; Vorpagel, E. R.; Lu, H. P. Probing Single-Molecule T4 Lysozyme

Conformational Dynamics by Intramolecular Fluorescence Energy Transfer. J. Phys. Chem.

B. 2003, 107, 7947–7956. 94

11 Liu, R. C.; Hu, D. H.; Tan, X.; Lu, H. P. Revealing Two-State Protein−Protein Interactions of

Calmodulin by Single-Molecule Spectroscopy. J. Am. Chem. Soc. 2006, 128, 10034–10042.

12 Dobson, C. M. Protein chemistry. In the footsteps of alchemists. Science 2004, 304,

1259-1262

13 Deniz AA, Laurence TA, Beligere GS, Dahan M, Martin AB, Chemla DS, Dawson PE,

Schultz PG, Weiss S: Protein folding from a highly disordered denatured state: The folding

pathway of chymotrypsin inhibitor 2 at atomic resolution. Proc Natl Acad Sci USA 2000, 97,

5179-5184.

14 Nettels D, Gopich IV, Hoffmann A, Schuler B. Ultrafast dynamics of protein collapse from

single-molecule photon statistics. Proc Natl Acad Sci USA 2007, 104, 2655-2660.

15 Nettels D, Hoffmann A, Schuler B. Unfolded Protein and Peptide Dynamics Investigated

with Single-Molecule FRET and Correlation Spectroscopy from Picoseconds to Seconds. J

Phys Chem B 2008, 112, 6137-6146.

16 Haran G. How, when and why proteins collapse: the relation to folding. Curr Opin Struct

Biol 2012, 22, 14-20.

17 Sherman E, Haran G. Coil–globule transition in the denatured state of a small protein. Proc

Natl Acad Sci USA 2006, 103:11539-11543.

18 Ziv G, Thirumalai D, Haran G. Collapse transition in proteins. Phys Chem Chem Phys 2009,

11:83-93.

19 Hoffmann A, Kane A, Nettels D, Hertzog DE, Baumga¨ rtel P, Lengefeld J, Reichardt G, 95

Horsley DA, Seckler R, Bakajin O et al.. Mapping protein collapse with single-molecule

fluorescence and kinetic synchrotron radiation circular dichroism spectroscopy. Proc Natl

Acad Sci USA 2007, 104:105-110.

20 Chung HS, Louis JM, Eaton WA. Experimental determination of upper bound for transition

path times in protein folding from single-molecule photon-by-photon trajectories. Proc Natl

Acad Sci USA 2009, 106:11837-11844.

21 Chung HS, McHale K, Louis JM, Eaton WA. Single-Molecule Fluorescence Experiments

Determine Protein Folding Transition Path Times. Science 2012, 335:981-984.

22 Shaw DE, Maragakis P, Lindorff-Larsen K, Piana S, Dror RO. Atomic-level characterization

of the structural dynamics of proteins. Science 2010, 330:341-346.

23 Garcia-Mira MM, Sadqi M, Fischer N, Sanchez-Ruiz JM, Munoz V. Experimental

Identification of Downhill Protein Folding. Science 2002, 298:2191-2195.

24 Pirchi M, Ziv G, Riven I, Cohen SS, Zohar N, Barak Y, Haran G. Single-molecule

fluorescence spectroscopy maps the folding landscape of a large protein. Nat Commun 2011,

2, 493.

25 Chin, D.; Means, A. R. Calmodulin: a prototypical calcium sensor. Trends in Cell Biology

2000, 10 (8), 322-328.

26 James, P.; Vorherr, T.; Carafoli, E. Calmodulin-binding domains: just two faced or

multi-faceted? Trends Biochem. Sci. 1995, 20 (1), 38-42.

27 Liu, R. C.; Hu, D. H.; Tan, X.; Lu, H. P. Revealing Two-State Protein−Protein Interactions of

Calmodulin by Single-Molecule Spectroscopy. J. Am. Chem. Soc. 2006, 128, 10034–10042. 96

28 James, P.; Vorherr, T.; Carafoli, E. Calmodulin-binding domains: just two faced or

multi-faceted? Trends Biochem. Sci. 1995, 20 (1), 38-42.

29 Babu, Y. S.; Sack, J. S.; Greenhough, T. J.; Bugg, C. E.; Means, A. R.; Cook, W. J.

Three-dimensional structure of calmodulin. Nature 1985, 315, 37-40.

30 Chattopadhyaya, R.; Meadorl, W. E.; Means, A. R.; Quiocho, F. A. Calmodulin structure

refined at 1.7 Å resolution. J. Mol. Biol. 1992, 228 (4), 1177-1192.

31 Slaughter B. D., Unruh J. R., Price E. S., Huynh J. L., Bieber Urbauer R. J., Johnson C. K.

Sampling Unfolding Intermediates in Calmodulin by Single-Molecule Spectroscopy. J Am

Chem Soc 2005, 127, 12107-12114.

32 Wang, X.; Lu, H. P. 2D Regional Correlation Analysis of Single-Molecule Time

Trajectories. J. Phys. Chem. B. 2008, 112, 14920-14926.

33 He, Y.; Li, Y.; Mukherjee, S.; Wu, Y.; Yan, H.; Lu, H. P. Probing Single-Molecule Enzyme

Active-Site Conformational State Intermittent Coherence. J. Am. Chem. Soc. 2011, 133,

14389-14395.

34 Wang, Y.; Lu, H. P. Bunching Effect in Single-Molecule T4 Lysozyme Nonequilibrium

Conformational Dynamics under Enzymatic Reactions. J. Phys. Chem. B 2010, 114,

6669-6674.

35 Xie, X. S. Single-Molecule Approach to Enzymology. Single Mol. 2001, 4, 229-236.

36 Uversky, V. N. Natively unfolded proteins: a point where biology waits for physics. Protein

Sci. 2002, 11, 739–756.

37 Ziv, G.; Thirumalai, D.; Haran, G. Collapse transition in proteins. Phy. Chem. Chem. Phys. 97

2009, 11(1), 83-93.

38 de Gennes, P.-G. Scaling concepts in polymer physics; Cornell University Press: Ithaca,

1979.

39 Des Cloizeaux, J.; Jannink, G. Polymers in solution: their modeling and structure; Clarendon

Press: Oxford, 1990.

40 Kubelka, J. J.; Hofrichter, J. J.; Eaton, W. A. The protein folding 'speed limit'. Curr. Opin.

Struct. Biol. 2004, 14(1), 76-78.

41 Pitard E. Influence of hydrodynamics on the dynamics of a homopolymer. Eur. Phys. J. B

1999, 7, 665-673.

42 Socci N. D.; Onuchic J. N.; Wolynes P. Diffusive dynamics of the reaction coordinate for

protein folding funnels. J. Chem. Phys 1996, 104, 5860-5868.

43 Stigler, J.; Ziegler, F.; Gieseke, A.; Gebhardt, J. C.; Rief, M. The complex folding network of

single calmodulin molecules. Science 2011, 334, 512-516.

98

CHAPTER V. PROBING SINGLE-MOLECULE PROTEIN FOLDING

CONFORMATIONAL DYNAMICS IN CROWDED EVIRONMENT USING

SINGLE-MOLECULE FRET SPECTROSCOPY

This chapter is dedicated to the study of single-molecule protein folding dynamics in the presence of molecular crowding effects.

5.1. Introduction

The interior of living cells contains large numbers of macromolecules such as proteins, DNA and RNA molecules1-5. Volume fraction of occupied macromolecular agents is as large as 40%6,

7. Existence of such macromolecular networks has significant effects on the behavior of protein

conformational dynamics inside such networks6-8. For example, enzymatic reactions and

amyloid formations are profoundly affected in real crowded biological environments7. Protein

folding processes in the presence of molecular crowding is crucial to understand protein folding processes inside living cells. Various molecular scale microenvironments are readily emulated by molecular crowding effects such as extensive protein-macromolecule interactions and molecular confinement effects.7

Single-molecule techniques are powerful tools to probe molecular scale dynamic processes without ensemble averaging9-15. Molecular details of single protein folding are readily observed one molecule at a time12. As a key advantage, single-molecule FRET allows a clear separation of the folded and unfolded subpopulations, and thus enables a further quantitative

analysis of the properties of single-molecule protein conformational fluctuation dynamics 99

without interference from undesired ensemble averaged signals10, 12. Since single-molecule

FRET can provide distance and protein conformational fluctuation dynamic information without ensemble-averaging, it is promising to show intramolecular conformational dynamics to be observed at the thermal equilibrium10.

5.2 Materials and Methods

The single CaM molecule is mutated on N-terminal domain at residue 34 and C-terminal

domain at residue 110, and fluorescent dye pair Cy3/Cy5 are tethered onto the protein via

thiolation. These two dyes serve as a spectroscopic ruler for the measurement of

conformational fluctuation of the protein molecules in different concentration of crowding

reagent Ficoll 70. On the other hand, the Forster radius R0 of Cy3-Cy5 pair is ~5.4 nm, so the distance change of the two dyes monitoring the conformational change falls into measureable

range. In our experiment, we prepare the samples for single-molecule conformational

folding-unfolding dynamics measurements in the presence of crowding reagent Ficoll 70 inside the agarose gel (1% by weight, Type VII, Sigma). We make samples of CaM into different concentrations of denaturant solvent GdmCl in the mixture of 200 nM CaM, 1.25μL and oxygen scavenger with Trolox solution to obtain a 10μL mixture of enzyme and denaturant solvent.

The Trolox solution is made previously by dissolving about 1 mM

6-hydroxy-2,5,7,8-tetramethylchroman-2-carboxylic acid (Trolox) in 1 mg/mL glucose oxidase,

0.8% D-glucose and 0.04 mg/mL catalase in order to protect fluorescent dye from

photobleaching or blinking as a result of triplet state formation and other photophysical

processes. We then heat the 10μL 1% agarose gel just above its gel-transition temperature 100

(26°C) and quickly mix the above enzyme solution with denaturant solvent solution and the gel between two clean cover glasses to form a sandwich. All solutions are prepared with HEPES buffer at pH 7.4. Since we want to probe conformational dynamics of single enzyme molecule at different concentration of Ficoll 70, we carry out concentration dependent experiments with different ratio of mixture in the sandwich (Supporting Information).

Single-molecule photon stamping approach is used to record FRET trajectories at different

concentration of crowding reagent. This approach records emission of photons from donor and

acceptor channel one by one with arrival time. We construct the intensity trajectory as a function

of time with desired time resolution from recorded raw data. The experimental setup is an

inverted confocal microscope (Axiovert 200, Zeiss) which used a crystal laser (532nm CW)

delivering excitation. The Laser beam focus through a 100× oil immersion objective lens (1.3

NA, 100×, Zeiss) onto the upper surface of cover slip after reflected up by a dichroic beam

splitter (z532rdc, Chroma Technology). To obtain confocal microscopy image, we use an x-y

closed-loop piezo position scanning stage for raster-scanning of the sample sandwich.

Fluorescence from single molecules is collected through the same objective and the signal was

split by a dichroic beam splitter (640dcxr) into two different wavelengths 570nm for the donor

channel and 670nm for the acceptor channel. To detect the signal from two channels, two Si avalanche photodiode single photon counting modules (SPCM-AQR-16, Perkin Elmer

Optoelectronics) are used to record photons from donor and acceptor. A more detailed description of the experimental setup is in the literature.16-20

We first apply a 2D regional correlation mapping analysis to our single-molecule photon 101

stamping data. This 2D regional correlation mapping analysis calculates a two-dimensional

cross-correlation amplitude distribution (TCAD). In this analysis, each of the trajectories is

scanned with different starting and ending time. The cross-correlation amplitude of each time

segment with distinct starting and ending time are calculated. We use color bar to indicate the

stop tstop t amplitude of cross-correlation. Ccross(,:)()()()() tt start stop  ItItdtAD   ItItAD  . tstart  tstart

Where IA and ID are the two-band photon count intensities signal of donor and acceptor, and tstart and tstop give the scanning window width. The cross-correlation functions are calculated with different tstart and tstop along a pair of smFRET trajectories {IA(t)} and {ID(t)}. The cross-correlation functions are calculated with different starting and ending time. Via this method, we identify time segments along a specific trajectory with strong cross-correlation indicated by cold color blue. Since typically, single-molecule FRET trajectories are dominated by shot noises or average intensity drifts as a result of undesired protein motion. This method enabled us to locate true anti-correlated segments from single molecule trajectories for our further dynamics analyses.

We apply auto-correlation and cross-correlation analyses to our single-molecule

photo-stamping data after the identification of each anti-correlated single-molecule time

segments. The cross-correlation and auto-correlation function are defined to be

IAD0  I  t  IA 0   IADD  I t  I  Ctcross    IIAD00    IIIIA 00  AD   D  102

IAA0  I  t  IA 0   IAAA  I t  I  auto  Ct  22 IA 0  IIAA0   

where IA(t) and ID(t) representing acceptor and donor intensitie, and and are the

means of the intensity trajectories respectively. Ccross(t) and Cauto(t) are cross-correlation and

autocorrelation functions.21-23

5.3 Results and Discussion

Figure 5.1 shows typical image from our inverted confocal microscopy. By raster-scan the

sample, a 20μm × 20μm sample image yields bright spots of single CaM molecules confined

in the network formed by agarose gel (~200nm in diameter). The spots indicating single

protein molecule are diffraction limited with approximately 300nm in diameter. We pin-point

each specific sample spot to gain continuous donor-acceptor (D-A) fluorescence intensity

trajectories shown in Figure 5.1. Typically, we gather more than seventy data points which consists of single-molecule trajectories of 40 seconds long at different concentration of

denaturant GdmCl, and we calculate the EFRET of each of the sample measured using the formula.

ItA() EtFRET () AA IAD()() t I t DD

Where  A and  D are the emission quantum yields of acceptor and donor dye molecules,

respectively, and  A and  D are the acceptor and donor detection efficiencies, respectively.

AA Here the correction factor is ~1 in our experiment conditions. DD 103

Figure 5.1. Single-Molecule Imaging and Analysis. Image obtained from Confocal Microscope of single CaM molecule. The bright spots are single molecules with diffraction limited (~300nm

in diameter) image. Typical Donor/Acceptor signals and EFRET trajectories. Green and red lines indicate the donor and acceptor channel respectively. The histogram distribution from EFRET trajectory gives the EFRET of a certain single-molecule. 104

Figure 5.2. EFRET Distributions. Distribution of FRET efficiency of different samples at different concentration of Ficoll 70. The shift in peak values characterizes the unfolding of CaM molecule

at higher concentration of crowding reagent.

The shifts in peak values of the distribution of all FRET efficiencies at different 105

concentration of Ficoll 70 mixed with CaM are the results of change of FRET efficiency (Figure

5.2) which are due to the unfolding and refolding of the single protein molecules. These changes in EFRET clearly evidence the conformational change of single CaM molecules under different concentrations of Ficoll 70. The EFRET distribution is consisting of both the folded

subpopulation with EFRET 0.6 and unfoled subpopulation with EFRET 0.3. In this study, we only

focus on the folding-unfolding fluctuating CaM molecules with EFRET 0.45. An ensemble average measurements and single-molecule measurements yield similar results that the 2M denature condition is the critical point of such titration. The FRET efficiency distribution at

50g/L Ficoll 70 is 0.36 and at 100g/L Ficoll 70 is 0.41, at which CaM molecules start to refold.

This conformational change measured by FRET is strong evidence of single CaM molecules enhanced stability by crowding reagent Ficoll 70. 24-30 As we increase Ficoll 70 concentration close to living cell macromolecule concentration ~300g/L, a remarkable decrease and broadening of EFRET distribution is observed. Such heterogeneous unfolding process as a result of crowding

effect cannot be resolved by conventional ensemble average measurements.31-39 We resolve

subpopulations of unfold and fold protein conformations and study detailed dynamic fluctuations

at the equilibrium. Mechanistic understanding of the single molecule protein folding energy

landscape can be extracted from this analysis.

106

Figure 5.3. Correlation Analyses. Autocorrelation function calculated from FRET efficiency

trajectory. In different concentrations of crowding reagent, the CaM molecule in its native state

show a correlation decay. Auto-correlations from intensity trajectory. Donor intensity trajectory

obtained from our single-molecule confocal microscopy (1 ms). The black line indicates the

threshold criterion separating the on- and off-time. Brown lines are the identified successive on- and off-time. The on-time distribution is a gamma shaped distribution representing a multiple step dynamic scheme. The simulated data is also shown.

We analyze the dynamical behaviors of our data sets by calculating the autocorrelation functions. By fitting the autocorrelation functions with exponential decays, we are capable of 107

characterizing the time scale of protein conformational fluctuation (Figure 5.3). Qualitatively,

each exponential function gives correlation lifetime of donor channel, and the broad distribution

of the lifetimes is a result of local environment heterogeneity. At low concentration of Ficoll 70,

CaM has a typical autocorrelation function with a sharp decay as a result of native protein

motion with typical time scale of millisecond. A strong contrast is shown in Figure 5.3. In

higher concentration of Ficoll 70, the CaM molecules, in its denatured states, show weak auto-correlation decay yielding second scale dynamics. Second scale dynamics is considerably slower than millisecond scale dynamics. This slow conformational fluctuation dynamics is due to the unfolding of CaM molecule in crowdedness which turns the proteins into random coils; the

correlation of protein’s regular motion becomes weaker. The two-dimensional regional

cross-correlation analyses help us to identify the time segments with strong anti-correlation of each single-molecule intensity trajectory. We zoom in to study each of the anti-correlated time segments by calculating the cross-correlation and auto-correlation respectively.

Subtle conformational dynamic signatures can be straightforwardly analyzed by the distribution of on- and off-times. The on-time and off-time are the “waiting time” for the CaM

folding and unfolding. To characterize such detailed dynamic behavior, we further bin our

trajectories to be one millisecond as shown (Figure 5.3). To obtain a folding waiting time distribution, we set up a threshold, at which the folding and unfolding states are separated. For the trajectory shown in Figure 5.3, we choose the value 7 as the threshold value, and subsequent folding waiting time distribution is shown.

Noticeably, the folding waiting time distribution is non-exponential resulting from a 108

multiple-state folding-unfolding dynamic scheme. Such distribution is also distinct from a

Gaussian distribution. This distinction clearly rules out the possibility of two-state

folding-unfolding dynamics for single CaM molecules. The gamma shaped folding waiting

time distribution indicates a more complex dynamic scheme involving multiple intermediates

and multiple steps. We attribute the non-exponential folding waiting time distribution to the

existence of multiple folding intermediates. To further characterize this multiple-intermediate

state dynamics, we exploit a one dimensional random walk model, which has been applied successfully in the modeling of multiple step single-molecule enzymatic reaction. We assume a

uniform rate constant k to each step, since we don’t have prior knowledge of the shape of the underling energy surface. The convolution of several Poisson process with uniform k gives gamma function shaped distribution to reproduce the mean and standard deviation of the original folding waiting time distribution with certain accuracy. The error of this recipe will be reflected in the difference between the simulated and measured mean. The number of Poisson steps involved in this convolution calculation gives an estimation of the lower bound of how many

Poisson rate processes are present. Since we assigned each Poisson step to a folding intermediate,

the simulated step number will be our best estimation of folding intermediate state number.

P t  Aexp  t /  where P(t) is the probability distribution of the Poisson rate process step times, τ is the averaged

Poisson rate process step time, and A is the distribution weight constant. To calculate the

convolution of function f(t) and g(t), the integration equation is 109

t f g  t   f  v  g  t  v  dv 0 n1   n ttexp /  PATn   n 1! 

Where n (1, 2, 3,…, N) is the index of the intermediate steps;  is the mean formation time of a folding intermediate through a single-step process and A is the normalizing factor of this probability distribution. The simulated distribution shown in the figure involves two Poisson rate processes which indicated a 2-step three state dynamics.40-42

Presumably, macromolecular crowding effect plays an enhancing role in protein folding process. However from our single-molecule experiment, we deduce polymer crowding might have enhancing and diminishing the protein folding stability in that over-crowdedness could potentially reduce the stability of protein molecules in the matrix. Such is a consequence of a balance of hydrophobic, hydrophilic and solvation thermodynamics and dynamics. Using polymer solution theory, we straightforwardly understand such counterintuitive phenomenon by calculating Helmholtz free energy and force exerted by the polymer matrix. In concentrated polymer solution of Ficoll 70, the very existence of another polymer will automatically exert a force on the target polymer. Size of the polymer R=N0.5φ-0.35b, where b is the monomer size,

φ the volume fraction and N degrees of polymerization. Blob size g=(Nb2)/(φ2R2). In our current analyses, our target polymer is the single-molecule CaM protein molecules. Since we increase rather than decrease the artificial crowder molecule Ficoll 70, the Ficoll 70 molecule applies a force on the CaM protein molecules. The exerted force is at molecular scale force typically at pico-newton scale used to unfolding or manipulates protein molecules. Since 110

polymer force is in principle entropic force originates from thermal fluctuation, it only makes sense to consider such subtle force and time scale in condensed phase protein folding dynamics.

To further quantitatively understand the force feature of our current experiment, we calculate

thermodynamic quantities from equilibrium polymer solution theory. The entropic nature of the

forces applied in single-molecule protein folding experiment can be straightforwardly calculated

from statistical mechanics. Since statistical mechanics results are textbook standards, we don’t

have to start from classical partition function, rather we directly use equipartition theorem. From

equipartition theorem, to every degree of freedom, there is thermal energy kT associated to that degree of freedom solely as a result of and thermal fluctuation energy. The Helmholtz free energy F = (kTN)/g = (kTφ2R2)/b2 is essentially the amount of kT thermal energy stored in

an extended polymer. The physical picture of polymer solution chemistry directly implies an

extended fibril like equilibrium solution dynamics of single protein molecules. Basically, it means when we increase the concentration of polymer solution, due to monomer-monomer

interactions, the polymers automatically reorganize themselves to form the concentrated polymer solution. To estimate the forces in this macroscopic dynamic polymer solution, we differentiate the Helmholtz free energy with respect to R to get the force exerted on target polymer f =

(2kTN0.5φ1.65)/b. For Ficoll 70, the specific volumeφ is 0.67ml/g. In our experiment, the

crowded region has a polymer concentration of 300g/L, the volume fraction of Ficoll 70 is 20%.

In this crowded region, we observe protein unfolding resolved from single-molecule data. The

monomer size of polymer Ficoll 70 is 3.9 Å. So the calculated force is 18.8 pN. From AFM and optical tweezer experiments, such force is strong enough to denature protein into unfolded 111

states. More recently, an experiment demonstrates that such force can also crush single proteins

into unfolded states.

We also estimate the conformational diffusion coefficient from the folding waiting time

distribution. From our previous experiments, we already have the mean conformational drift

distance XtN   to be 1.22nm. Using formula , we calculate the diffusion coefficient D of this

2-step dynamic process to be 13.2×10-13 cm2/s. The diffusion coefficient is directly related to

the shape of the underlying energy landscape, since it reflects the local mean square fluctuation

of the activation barrier of the folding-unfolding dynamic process.

2 t2 X t  unfold N    D  3 2 tunfold

t where D is the conformational diffusion coefficient. The mean unfolding time, unfold , and the

t 2 standard deviation of the unfolding time distribtuion, unfold , are directly measured in our

experiment. The total drifting distance of the folding-unfolding conformational motion,

Xt  N , is associated to the folding-unfolding conformational distance change. The activation barrier is as high as 12 kT, a 10% increase from simple protein folding experiments. However the increase in the local mean square fluctuation is approximately 20%. This extraordinary fact lies in the molecular scale interaction between the polymer Ficoll 70 and our target protein CaM. We compare this number to standard Hydrogen bonding, which is typically 2-10kT. The existence of

crowding reagent increases the fluctuation dynamics of hydrogen bonding within CaM protein

molecule by making the folding energy landscape “rougher”. Such subtle molecular scale

conformational interaction energy might play a key role in sub-cellular protein-protein 112 interactions.

Figure 5.4. Folding Pathway Distribution in Protein Folding. The scatter plot we constructed with number of steps involved in each of the dynamic process vs. conformational diffusion coefficient. The folding-unfolding dynamics pattern naturally clustered into separated domains.

Subsequently, a CaM folding energy landscape can be extracted, and a distribution of the folding paths with distinct probability assigned to each of them can be achieved.

We have analyzed approximately 30 trajectories under different crowding conditions to acquire a distribution of such dynamic pattern. For each of them, we simulate the number of steps involved in the dynamic process, and we also calculate the folding-unfolding 113 conformational diffusion coefficient. A two dimensional scatter plot is constructed with conformational diffusion coefficient vs. number of steps (Figure 5.4). Clearly, a pattern emerged. Folding-unfolding trajectories with same number of steps and similar conformational diffusion coefficient cluster together. This cluster behavior is clearly a manifestation of an energy landscape with distinct folding pathways. Moreover, these various folding pathways have a distribution with distinct probabilities. To better see this, we convert the two-dimensional scatter plot to be a two dimensional energy landscape plot with color arrow indicating different folding pathways. Our results suggest four different folding pathways of

CaM molecule folding process both in the crowding enhancing and diminishing regions.

Figure 5.5. A Cartoon Showing Folding Process. A cartoon showing a combination of molecular scale effects lead to protein unfolding under buffer condition. Such molecular crowding effects leading to a protein unfolding can be seen as a concerted effect of solvation and macromolecular crowding.

114

5.4 Conclusions

We demonstrate single-molecule protein folding-unfolding conformational dynamics in the

presence of molecular crowding effect provided by Ficoll 70. In our single-molecule FRET experiment, we observe large heterogeneity of EFRET which is clear evidence of single-molecule protein refolding and unfolding processes. Such remarkable detailed information of

conformational dynamics eludes ensemble-averaged experiments due to averaging. The enhanced folding process is entropic driven process which stabilizes single protein molecules and refolds proteins into their native states. However, at higher concentrations of crowding reagent Ficoll 70, we observe the unfolding of single protein molecules which are a combined process of polymer-polymer interactions, entropic and solvation mechanism. Utilizing the polymer solution theory, we show such unfolding processes can be understood as an isotropic force exerted on single-molecule protein solely due to the existence of host polymers.

Subtle conformational fluctuation dynamics can be readily observed and studied by single-molecule techniques. The longer folding conformational fluctuation time indicates an enthalpic trap provided by the presence of molecular crowding reagents. In other words, extensive polymer-polymer interactions slow down folding conformational fluctuations by favoring the unfolded states of single-molecule proteins. By our dynamic model, we observe a converged folding pathway due to normal crowding reagents as a result of entropic effects.

Simply, the existence of polymer provides excluded volume which decreases the sampling conformations available for protein molecules. However, in concentrated regime of polymer solutions, such entropic effects are out-favored by polymer-polymer interactions and solvation 115

effects which destabilize single-protein molecules. The unfolding process due to molecular

crowding is highly heterogeneous, which is a manifestation of complex, cooperative biological

mechanisms.

5.5 Reference:

1 Lu, H. P.; Xun, L. Y.; Xie, X. S. Single-Molecule Enzymatic Dynamics. Science 1998, 282,

1877–1882.

2 Chen, Y.; Hu, D. H.; Vorpagel, E. R.; Lu, H. P. Probing Single-Molecule T4 Lysozyme

Conformational Dynamics by Intramolecular Fluorescence Energy Transfer. J. Phys. Chem.

B 2003, 107, 7947–7956.

3 Schuler, B.; Lipman, E. A.; Eaton, W. A. Probing The Free-Energy Surface for Protein

Folding with Single-Molecule Fluorescence Spectroscopy. Nature 2002, 419, 743-747.

4 Lu, H. P.; Iakoucheva, L. M.; Ackerman, E. J. Single-Molecule Conformational Dynamics of

Fluctuating Noncovalent DNA-Protein Interactions in DNA Damage Recognition. J. Am.

Chem. Soc. 2001, 123, 9184-9185.

5 Tan, X.; Nalbant, P.; Toutchkine, A.; Hu, D.; Vorpagel, E. R.; Hahn, K.M.; Lu, H. P.

Single-Molecule Study of Protein-Protein Interaction Dynamics in a Cell Signaling System. J.

Phys. Chem. B. 2004 108, 737-744.

6 Ha, T. J.; Ting, A. Y.; Liang, J.; Caldwell, W. B.; Deniz, A. A.; Chemla, D. S.; Schultz, P. G.;

Weiss, S. Single-Molecule Fluorescence Spectroscopy of Enzyme Conformational Dynamics

and Cleavage Mechanism. Proc. Natl. Acad. Sci. U.S.A. 1999, 96, 893–898.

7 Anfinsen, C. B.; Haber, E.; Sela, M.; White, F. H., Jr. The Kinetics of Formation of Native 116

Ribonuclease During Oxidation of the Reduced Polypeptide Chain. Proc. Natl. Acad. Sci.

USA 1961, 47, 1309-1314.

8 Onuchic, J. N.; Luthey-Schulten, Z.; Wolynes, P. G. Theory of Protein Folding: the Energy

Landscape Perspective. Annu. Rev. Phys. Chem. 1997, 48, 545-600.

9 Dill, K. A.; Chan, H. S. From Levinthal to Pathways to Funnels. Nat. Struct. Biol. 1997, 4 (1),

10-19.

10 Thirumalai, D.; Klimov, D. K. Deciphering the Timescales and Mechanisms of Protein

Folding Using Minimal Off-Lattice Models. Curr. Opin. Struc. Biol. 1999, 9 (2), 197-207.

11 Zoldak, G.; Rief, M. Force as a Single Molecule Probe of Multidimensional Protein Energy

Landscapes. Curr. Opin. Struc. Biol. 2013, 23 (1), 48-57.

12 Deniz, A. A.; Laurence, T. A.; Beligere, G. S.; Dahan, M.; Martin, A. B.; Chemla, D. S.;

Dawson, P. E.; Schultz, P. G.; Weiss, S. Single-Molecule Protein Folding: Diffusion

Fluorescence Resonance Energy Transfer Studies of the Denaturation of Chymotrypsin

Inhibitor 2. Proc. Natl. Acad. Sci. USA 2000, 97 (10), 5179-5184.

13 Nettels, D.; Gopich, I. V.; Hoffmann, A.; Schuler, B. Ultrafast Dynamics of Protein Collapse

from Single-Molecule Photon Statistics. Proc. Natl. Acad. Sci. USA 2007, 104 (8),

2655-2660.

14 Nettels, D.; Hoffmann, A.; Schuler, B. Unfolded Protein and Peptide Dynamics Investigated

with Single-Molecule FRET and Correlation Spectroscopy from Picoseconds to Seconds. J.

Phys. Chem. B 2008, 112 (19), 6137-6146.

15 Haran G. How, When and Why Proteins Collapse: the Relation to Folding. Curr. Opin. Struc. 117

Biol. 2012, 22 (1), 14-20.

16 Sherman, E.; Haran, G. Coil-Globule Transition in the Denatured State of a Small Protein.

Proc. Natl. Acad. Sci. USA 2006, 103 (31), 11539-11543.

17 Ziv, G.; Thirumalai, D.; Haran, G. Collapse Transition in Proteins. Phy. Chem. Chem. Phys.

2009, 11 (1), 83-93.

18 Hoffmann, A.; Kane, A.; Nettels, D.; Hertzog, D. E.; Baumgartel, P.; Lengefeld, J.; Reichardt,

G.; Horsley, D. A.; Seckler, R.; Bakajin, O.; et al. Mapping Protein Collapse with

Single-Molecule Fluorescence and Kinetic Synchrotron Radiation Circular Dichroism

Spectroscopy. Proc. Natl. Acad. Sci. USA 2007, 104 (1), 105-110.

19 Chung, H. S.; Louis, J. M.; Eaton, W. A. Experimental Determination of Upper Bound for

Transition Path Times in Protein Folding from Single-Molecule Photon-by-Photon

Trajectories. Proc. Natl. Acad. Sci. USA 2009, 106 (29), 11837-11844.

20 Chung, H. S.; McHale, K.; Louis, J. M.; Eaton, W. A. Single-Molecule Fluorescence

Experiments Determine Protein Folding Transition Path Times. Science 2012, 335 (6071),

981-984.

21 Shaw, D. E.; Maragakis, P.; Lindorff-Larsen, K.; Piana, S.; Dror, R. O.; Eastwood, M. P.;

Bank, J. A.; Jumper, J. M.; Salmon, J. K.; Shan, Y.; et al. Atomic-Level Characterization of

the Structural Dynamics of Proteins. Science 2010, 330 (6002), 341-346.

22 Garcia-Mira, M. M.; Sadqi, M.; Fischer, N.; Sanchez-Ruiz, J. M.; Munoz, V. Experimental

Identification of Downhill Protein Folding. Science 2002, 298 (5601), 2191-2195Pirchi M.;

Ziv G.; Riven I.; Cohen S. S.; Zohar N.; Barak Y.; Haran G. Nat. Commun. 2011, 2, 493-450. 118

23 Chin, D.; Means, A. R. Calmodulin: A Prototypical Calcium Sensor. Trends Cell Biol. 2000,

10 (8), 322-328.

24 James, P.; Vorherr, T.; Carafoli, E. Calmodulin-Binding Domains - Just 2-Faced or

Multifaceted. Trends Biochem. Sci. 1995, 20 (1), 38-42.

25 Liu, R.; Hu, D.; Tan, X.; Lu, H. P. Revealing Two-State Protein-Protein Interactions of

Calmodulin by Single-Molecule Spectroscopy. J. Am. Chem. Soc. 2006, 128 (31),

10034-10042.

26 James, P.; Vorherr, T.; Carafoli, E. Calmodulin-Binding Domains - Just 2-Faced or

Multifaceted. Trends Biochem. Sci. 1995, 20 (1), 38-42.

27 Babu, Y. S.; Sack, J. S.; Greenhough, T. J.; Bugg, C. E.; Means, A. R.; Cook, W. J.

Three-Dimensional Structure of Calmodulin. Nature 1985, 315 (6014), 37-40.

28 Chattopadhyaya, R.; Meador, W. E.; Means, A. R.; Quiocho, F. A., Calmodulin Structure

Refined at 1.7 Angstrom Resolution. J. Mol. Biol. 1992, 228 (4), 1177-1192.

29 Slaughter, B. D.; Unruh, J. R.; Price, E. S.; Huynh, J. L.; Bieber Urbauer, R. J.; Johnson, C.

K. Sampling Unfolding Intermediates in Calmodulin by Single-Molecule Spectroscopy. J.

Am. Chem. Soc. 2005, 127 (34), 12107-12114.

30 Wang, X.; Lu, H. P. 2D Regional Correlation Analysis of Single-Molecule Time Trajectories.

J. Phys. Chem. B. 2008, 112 (47), 14920-14926.

31 He, Y.; Li, Y.; Mukherjee, S.; Wu, Y.; Yan, H.; Lu, H. P. Probing Single-Molecule Enzyme

Active-Site Conformational State Intermittent Coherence. J. Am. Chem. Soc. 2011, 133 (36),

14389-14395. 119

32 Wang, Y.; Lu, H. P. Bunching Effect in Single-Molecule T4 Lysozyme Nonequilibrium

Conformational Dynamics Under Enzymatic Reactions. J. Phys. Chem. B 2010, 114 (19),

6669-6674.

33 Xie, X. S. Single-Molecule Approach to Enzymology. Single Mol. 2001, 2 (4), 229-236.

34 Uversky, V. N. Natively Unfolded Proteins: A Point Where Biology Waits for Physics.

Protein Sci. 2002, 11 (4), 739-756.

35 Ziv, G.; Thirumalai, D.; Haran, G. Collapse Transition in Proteins. Phy. Chem. Chem. Phys.

2009, 11 (1), 83-93.

36 de Gennes, P.-G. Scaling Concepts in Polymer Physics; Cornell University Press: Ithaca,

1979.

37 Des Cloizeaux, J.; Jannink, G. Polymers in Solution: Their Modeling and Structure;

Clarendon Press: Oxford, 1990.

38 Kubelka, J.; Hofrichter, J.; Eaton, W. A. The Protein Folding 'Speed Limit'. Curr. Opin. Struc.

Biol. 2004, 14 (1), 76-88.

39 Pitard E. Influence of Hydrodynamics on the Dynamics of a Homopolymer. Eur. Phys. J. B

1999, 7 (4), 665-673.

40 Socci, N. D.; Onuchic, J. N.; Wolynes, P. G. Diffusive Dynamics of the Reaction Coordinate

for Protein Folding Funnels. J. Chem. Phys 1996, 104 (15), 5860-5868.

41 Stigler, J.; Ziegler, F.; Gieseke, A.; Gebhardt, J. C. M.; Rief, M. The Complex Folding

Network of Single Calmodulin Molecules. Science 2011, 334 (6055), 512-516.

42 Zwanzig, R. Diffusion in a Rough Potential. Proc. Natl. Acad. Sci. U.S.A. 1988, 85 (7), 120

2029-2030.