172
Protein Structure Dynamics and Function at Micelle and Particle Interfaces Michael Chin Submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in the Graduate School of Arts and Sciences COLUMBIA UNIVERSITY 2015

Protein Structure Dynamics and Function at Micelle and

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Protein Structure Dynamics and Function at Micelle and

Protein Structure Dynamics and Function at Micelle and Particle Interfaces

Michael Chin

Submitted in partial fulfillment of the requirements

for the degree of Doctor of Philosophy

in the Graduate School of Arts and Sciences

COLUMBIA UNIVERSITY

2015

Page 2: Protein Structure Dynamics and Function at Micelle and

© 2015

Michael Chin

All rights reserved

Page 3: Protein Structure Dynamics and Function at Micelle and

ABSTRACT

Protein Structure Dynamics and Function at Micelle and Particle Interfaces

Michael Chin

This work is an investigation into how protein function and structural dynamics are

affected by their interactions with detergents and particles. Recent studies have reported

the addition of certain non-ionic surfactants increase the catalytic rates of enzymes. This

is a departure from the established understanding of surfactant/protein interactions

wherein charged-surfactants tend to denature the macro-molecular structure, while other

detergents can be used to isolate and preserve protein structures. The mechanism of

denaturation by ionic surfactants is well understood, as is the preservation and

crystallization of protein structures with detergents. But why should non-ionic surfactants

increase observed activity? This question remains unanswered. Furthermore, the

explanations provided with previous studies where enzyme activity enhancement via

surfactant interactions are inconsistent with each other, and are highly specific to the

system in which they are studied. These contradicting conclusions highlight the need to

propose a general theory of protein-surfactant interactions which can explain these

reported changes in enzyme activity that can be applied to any of these reported systems.

The theory proposed in this thesis is as follows: Unlike ionic surfactants, non-ionic

surfactants minimally bind to the protein structure. Instead, bound water regions

surrounding non-ionic surfactant aggregates (micelles) interact with the surface hydration

Page 4: Protein Structure Dynamics and Function at Micelle and

of adjacent proteins, thus affecting the free energy of solvent forces at the protein surface.

This, in turn, affects the small scale fluctuations of the proteins structures which

influenced properties important to enzyme function, such as flexibility and substrate

binding energy. Aside from enzyme interactions with surfactant, the implications of this

proposed theory extend to any system in which functional proteins readily interact with a

surface. The two parallel examples also reported in this dissertation are the structural

dynamics of carrier proteins adsorbed onto metal phosphate adjuvant particles and a

mathematical model describing changes to FRET efficiency due to the fluctuating

motions of a light active protein

Three major objectives were established in order to test this hypothesis: 1) Establish

model system of enzyme and surfactants, and a means to test enzyme efficacy. 2) Study

the differences in surfactants and surfactant aggregates that would attribute to changes in

enzyme activity and 3) Test the flexibility or the enzyme structures as a function of

surfactant structure and concentration to see if there is, indeed, a connection between the

surfactant and the enzyme activity through a modification of the protein’s structural

dynamics.

Subtilisin proteinase and horseradish peroxidase were ultimately chosen as model

enzymes to study. Initial activity assays confirmed previously reported results in which

charged surfactants such as sodium dodecyl sulfonates had little effect or negatively

impacted enzyme activity. Longer exposures to these anionic surfactants result in a loss

Page 5: Protein Structure Dynamics and Function at Micelle and

of activity which can be attributed to enzyme denaturation. Enzyme activity in non-ionic

surfactants such as Polyoxyethylene (4) lauryl ether (Brij-30) and alkyl polyglucosides

(APG), increased. Curiously, dodecyl-β-D-maltoside (DM), a sugar-based non-ionic

surfactant of similar head group structure to APG, did not induce enhanced enzyme

activity, giving us two similar surfactants with which we can compare to determine what

factors, if not surfactant molecular structure, affect enzyme activity.

Surface tension, dynamic light scattering, electron spin resonance and NMR were used to

elucidate the structural differences between alkyl polyglucoside (APG) and dodecyl

maltoside micelles (DM), and which of these characteristics are attributing to the

differences in activity. It was found that APG transitions between two micelle structure

conformations as a function of concentration. After surpassing an initial critical micelle

concentration, APG forms loosely packed aggregates that are attracted to each other to

form even larger super-aggregates. Passing a second transitional concentration, the

micelles form dense, rod-like structures. It should be noted that enzyme activity peaks in

the presence of the loosely-packed super-aggregates and drops back to normal values

once the micelles transition into their rod-like configuration. Conversely, dodecyl

maltoside, which did not induce greater enzyme activity, does not exhibit this dual phase

transition. Instead, the surfactants in DM solutions exceeding the critical micelle

immediately form tightly packed, single layered spherical micelles. From these

experiments we conclude that the densely packed micelles like those formed by DM and

APG at higher concentrations do not affect enzyme structure the same way as those

loosely packed ones at lower APG concentrations.

Page 6: Protein Structure Dynamics and Function at Micelle and

Life-time fluorescence techniques were employed to fulfill the third objective and

measure the structural dynamics of the model proteins. In both subtilisin and horseradish

peroxidase, we observed a similar pattern of enzyme activity changes as a function of

APG and Brij-30 concentration. Anionic SDS surfactant did not affect this flexibility

parameter. Furthermore, we demonstrated that the flexibility of HRP is not affected by

dodecyl maltoside much in the same way dodecyl maltoside did not affect the enzyme

activity in any significant way.

The connection between micelle structure and enzyme flexibility can be attributed to the

structure and dynamics of the solvent between the micelle and enzyme. The final extent

of influence of solvent dynamics on enzyme function is still contentious; however the

experimental results from this work lend support this theory. Building off models

presented by Viparelli on the influence of micelle bound water on enzyme activity, and a

model describing the rate of “solvent-slaved” protein functions as a function of solvent

relaxation times, a new model which relates changes in enzyme activity as a function of

surfactant concentration is developed and presented here in this body of work.

Page 7: Protein Structure Dynamics and Function at Micelle and

i

Contents

List of Figures ................................................................................................................................... iii

1 Introduction ............................................................................................................................. 1

1.1 Initial findings and what is in the literature .................................................................... 2

1.2 Gaps in knowledge .......................................................................................................... 7

1.3 Central goal of dissertation: .......................................................................................... 10

2 Hypothesis and rationale ....................................................................................................... 11

2.1 Rationale for Hypotheses: ............................................................................................. 12

3 Research plan and objectives ................................................................................................ 18

3.1 Sub-objectives: .............................................................................................................. 18

4 Methods and Materials ......................................................................................................... 21

4.1 Objective 1 - Find suitable model system to test enzyme hyperactivity via interaction

with surfactants ......................................................................................................................... 21

4.1.1 Enzyme selection ................................................................................................... 22

4.1.2 Surfactants ............................................................................................................. 24

4.1.3 Subtilisin assay: ...................................................................................................... 26

4.1.4 Subtilisin initial rate activities and kinetics ........................................................... 27

4.1.5 HRP Kinetics: .......................................................................................................... 28

4.1.6 Subtilisin assay side note: Behavior of Fluorescein Probe (optimizing subtilisin

assay method) ....................................................................................................................... 29

4.2 Objective 2: Characterizing surfactant / particle properties in protein solutions ........ 40

4.2.1 DLS ......................................................................................................................... 40

4.2.2 Surface tension ...................................................................................................... 42

4.2.3 Electron paramagnetic resonance ......................................................................... 43

4.3 Objective 3: Measuring structural dynamics of model proteins ................................... 46

4.3.1 Life time fluorescence anisotropy ......................................................................... 46

4.3.2 Life time fluorescence decay ................................................................................. 49

4.4 Objective 3a: Dynamics and function changes at inorganic particle surfaces ............. 54

4.4.1 UV-Vis: ................................................................................................................... 56

4.4.2 Zeta potential measurement ................................................................................. 57

Page 8: Protein Structure Dynamics and Function at Micelle and

ii

4.4.3 FTIR ........................................................................................................................ 57

5 Results and discussion ........................................................................................................... 59

5.1 Objective 1: Establishing a model system ..................................................................... 59

5.1.1 5.1.1 Enzyme product yield assays. ....................................................................... 60

5.1.2 Initial rate assays – Subtilisin protease ................................................................. 63

5.1.3 HRP activity in Brij-30, APG and DM ...................................................................... 73

5.2 Discussion of subtilisin and HRP assay and activity results ........................................... 78

5.2.1 Sub section: kinetics parameters Km and Vmax ................................................... 82

5.3 Objective 2: Aggregation / micelle properties .............................................................. 95

5.3.1 Surface tension: ..................................................................................................... 95

5.3.2 Micelle size: ........................................................................................................... 97

5.3.3 ESR investigation of micelle structure: APG vs DM ............................................... 99

5.3.4 NMR: (Cooperative work with C. Totland of University of Bergen) .................... 101

5.3.5 IR investigation of Bio-Surfactant degradation via enzyme hydrolysis: .............. 106

5.4 Objective 3: Protein structure dynamics: .................................................................... 114

5.4.1 Life-time fluorescence: Anisotropy ..................................................................... 114

5.4.2 Time resolved fluorescence to measure flexibility of horseradish peroxidase: .. 119

5.5 Objective 3a: Protein flexibility and orientation at particle surfaces:......................... 127

5.6 Objective 3b: Modeling enzyme activity as a function of surfactant interaction and

concentration. ......................................................................................................................... 132

5.7 Results and discussion conclusions ............................................................................. 138

6 Conclusions and Next Steps ................................................................................................ 149

6.1 Next steps: ................................................................................................................... 152

7 References ........................................................................................................................... 154

Page 9: Protein Structure Dynamics and Function at Micelle and

iii

List of Figures

Figure 1-1 - Initial subtilisin activities in anionic and non-ionic surfactants ................................... 2

Figure 1-2 - Increased hydrolytic rate seen in Bacillus amyloliquefaciens α-amylase (BAA) as a

function of Brij-35 concentration as reported by Hosino et al. ...................................................... 3

Figure 1-3 - Three phases that enzyme hydrolysis can occur. Bulk (Sf), Bound water interface (Sb)

and within the micelle as described by Viparelli et al. .................................................................... 6

Figure 1-4 - Top down view of three protein chains which compose of a bacteriorhodopsin unit.

One is shaded blue with a white outline indicating the location of the retinol chromophore. ...... 9

Figure 2-1 - Range of time scales associated with protein structure dynamics. This study with

focus on elastic fluctuations occurring in the ps range. ................................................................ 13

Figure 2-2 - General schematic and description of induced fit model. Active site conformation

changes play an important role in minimizing energy barrier to product formation [33] ............ 14

Figure 4-1 - Dodecyl maltoside ...................................................................................................... 25

Figure 4-2 - Example of colormetric activity assay ........................................................................ 27

Figure 4-3 - Schematic illustrating liberation of free fluoresein probe from labled casein .......... 28

Figure 4-4 - Fluoresence spectroscopy of fluorescein fluorophore dilution series ....................... 30

Figure 4-5 - Fluorescein peak wavelength vs. concentration ........................................................ 31

Figure 4-6 - Fluorescein peak intensity vs. Concentration ............................................................ 32

Figure 4-7 - Divergence from intensity / concentration linarity starting at 0.1 mM of flourescein

....................................................................................................................................................... 33

Figure 4-8 - Reaction profiles of subtilisin catalyzed proteolysis (0.001 to 0.1 mM substrate

concentrations) ............................................................................................................................. 34

Figure 4-9 - End of reaction fluoresence intensity vs. substrate concentration (.0001 to 0.1 mM)

....................................................................................................................................................... 35

Figure 4-10 - Casein hydrolysis reaction profile (0.0005 to 0.025 mM substrate concentration) 36

Figure 4-11 - End of reaction fluoresence intensity (low substrate concentration 0.0005 to 0.025

mM) ............................................................................................................................................... 37

Figure 4-12 - Set of initial rates plotted against substrate concentration drawn with error

distribution. ................................................................................................................................... 38

Figure 4-13 - Schematic of Zeeman Effect energy splitting ........................................................... 43

Figure 4-14 - Typical nitroxyl group ESR spectrum and peak parameters used to calculate

rotational correlation times .......................................................................................................... 44

Figure 4-15 - Dansyl fluoride structure.......................................................................................... 46

Figure 4-16 - example of parallel (VV) and perpendicular(VH) fluorescence decays with

instrument response function (IRF)............................................................................................... 47

Figure 4-17 - anisotropy (r(t)) derived from VV(t) and VH(t). [62] ............................................... 48

Figure 4-18 - Structure of horseradish peroxidase. Trp residue and heme group are highlighted.

....................................................................................................................................................... 50

Figure 4-19 - Structure of CRM 197 carrier protein used in vaccine formulations ....................... 55

Figure 4-20 - Concentration standards of BSA measured by UV-Vis at 280nm ............................ 56

Page 10: Protein Structure Dynamics and Function at Micelle and

iv

Figure 5-1 - Subtilisin activity with and without SDS anionic surfactant as a function of exposure

time ............................................................................................................................................... 60

Figure 5-2 - Subtilisin activity with and without SDS anionic surfactant as a function of detergent

concentration ................................................................................................................................ 62

Figure 5-3 - Enzyme activity assays of subtilisin in various surfactants (relative to enzyme in

buffer) ............................................................................................................................................ 63

Figure 5-4 - Fluorescence spectrum of casein substrate proteolysis as a function of time. Peak

intensity is found at 519 nm. ......................................................................................................... 65

Figure 5-5 - Fluorescence spectrum of casein substrate proteolysis as a function of time. Peak

intensity is found at 519 nm. ......................................................................................................... 65

Figure 5-6 - Initial subtilisin rate profile experiments confirm faster initial proteolysis rates in

alkyl polyglucoside solutions. ........................................................................................................ 67

Figure 5-7 - Subtilisin catalyzed proteolysis reaction profile in APG solutions ............................. 67

Figure 5-8 - Normalized values of the initial rates derived from reaction profiles shown in Fig. 5.7

....................................................................................................................................................... 68

Figure 5-9 - Subtilisin catalyzed proteolysis reaction profile in Brij-30 surfactant ....................... 69

Figure 5-10 - Initial reaction rates of subtilisin mediated proteolysis in Brij-30 non-ionic

surfactants ..................................................................................................................................... 69

Figure 5-11 - Proteolysis reaction profile in Brij-35 surfactant ..................................................... 70

Figure 5-12 - Initial reaction rates of subtilisin catalyzed proteolysis in Brij-35 surfactant .......... 71

Figure 5-13 - Normalized reaction rates of Subtilisin in low concentrations. APG CMC is

approximately 0.2 mM. ................................................................................................................. 72

Figure 5-14 - Subtilisin activity in Brij-30 solutions ....................................................................... 72

Figure 5-15 - HRP initial reaction rates in APG surfactant ............................................................ 74

Figure 5-16 - - Initial reaction rates of HRP in APG (columns) and surface tension measurements

of APG (diamonds) ......................................................................................................................... 75

Figure 5-17 - Normalized initial rate values of HRP activity in dodecyl maltoside concentrations

....................................................................................................................................................... 76

Figure 5-18 - Normalized initial reaction rates of subtilisin (red) compared to those of HRP

enzyme (black) ............................................................................................................................... 77

Figure 5-19 - Figure 5.17. Comparison of normalized initial rates of HRP enzyme in APG (black

squares) vs DM (red triangles) ...................................................................................................... 77

Figure 5-20 - Subtilisin mediated proteolysis reaction profiles as a function of substrate

concentration ................................................................................................................................ 84

Figure 5-21 - Non-linear regression fitting of reaction rates of subtilisin in buffer vs. casein

substrate concentration done in Origins 8.0 software ................................................................. 84

Figure 5-22 - Initial rates of subtilisin catalyzed proteolysis vs substrate concentration in 10mM

of APG ............................................................................................................................................ 85

Figure 5-23 - Initial rates of subtilisin catalyzed proteolysis vs substrate concentration in 20mM

of APG ............................................................................................................................................ 86

Figure 5-24 - Subtilisin Vmax as a function of APG concentration ................................................ 87

Figure 5-25 - Subtilisin Vmax as a function of DM concentration ................................................. 88

Page 11: Protein Structure Dynamics and Function at Micelle and

v

Figure 5-26 - Subtilisin Vmax as a function of Brij-30 concentration ............................................ 88

Figure 5-27 - Subtilisin Vmax as a function of Brij-35 concentration ............................................ 89

Figure 5-28 - Subtilisin Km as a function of APG concentration .................................................... 90

Figure 5-29 - Subtilisin Km as a function of DM concentration ..................................................... 91

Figure 5-30 - Subtilisin Km as a function of Brij-35 concentration ................................................ 91

Figure 5-31 - Subtilisin Km as a function of Brij-30 concentration ................................................ 92

Figure 5-32 - Surface tension vs APG concentration ..................................................................... 96

Figure 5-33 - Surface tension comparisons of dodecyl maltoside (single surfactant) to alkyl

polyglucosides, a surfactant mixture of similar composition. ....................................................... 97

Figure 5-34 - APG micellesize vs concentration ............................................................................ 98

Figure 5-35 - DM micelle size vs concentration ............................................................................ 98

Figure 5-36 - Rotational correlation times (measure of local viscosity) as a function of surfactant

(APG and DM) concentration in buffer only. ................................................................................. 99

Figure 5-37 - Rotational correlation times (measure of local viscosity) as a function of surfactant

(APG and DM) concentration in buffer with 0.01 mg/ml HRP. ................................................... 100

Figure 5-38 - Comparison of the line widths of the main APG CH2 peak (2-12) for the 0.15 and 12

mM APG sample in 0.01 mg/ml HRP in 100mM potassium phosphate buffer ........................... 102

Figure 5-39 - The line width here is the line width of the main CH2 peak (2-12) in the NMR

spectrum. The line widths are measured by using the deconvolution tool in the Topspin software

(version 1.3). ................................................................................................................................ 103

Figure 5-40 - Overlay of the aggregate mobility data presented in figure 5.31 with the activity of

HRP in APG (triangles) ................................................................................................................. 104

Figure 5-41 - NOESY spectra with water signal suppression using excitation sculpting with

gradients, at different concentrations of APG at constant HRP concentration. A possible

interpretation in terms of APG aggregate structure is presented to the left. ............................ 105

Figure 5-42 - Sodium Dodecyl Sulfate ......................................................................................... 107

Figure 5-43 - n-Dodecyl β,D Maltoside ........................................................................................ 107

Figure 5-44 - IR spectrum of SDS and Dodecyl Maltoside between 1350 and 850 cm-1

wavenumber range. .................................................................................................................... 108

Figure 5-45 - IR spectrum of SDS and Dodecyl Maltoside between 1160 and 960 cm-1

wavenumber range. .................................................................................................................... 109

Figure 5-46 - Schematic of cellulase hydrolysis mechanism ....................................................... 110

Figure 5-47 - FTIR of Dodecyl maltoside with a-Cellulase indicating possible loss of C-O-C bond

..................................................................................................................................................... 111

Figure 5-48 - Proposed reaction of surfactant catalysis .............................................................. 111

Figure 5-49 - FTIR from 1800 to 1500 cm-1 possibly showing the carboxyl group located at the

cellulase active site. ..................................................................................................................... 112

Figure 5-50- Rotational correlation time (blue diamonds) and subtilisin activity (orange squares)

as a function of Brij-30 concentration. ........................................................................................ 117

Figure 5-51 - Decay correlation times of dansyl probe attached to subtilisin as a function of APG

detergent. .................................................................................................................................... 118

Page 12: Protein Structure Dynamics and Function at Micelle and

vi

Figure 5-52 -Top. Raw data (Blue) and fitting of sum of 3 exponential decay functions (orange) of

hrp in buffer. Bottom: residual of fitted line vs raw decay data. ................................................ 119

Figure 5-53 - Decay data and multiple exponential fitting (including residuals below) of HRP

enzyme in 0.05 mM APG. ............................................................................................................ 122

Figure 5-54 - Decay data and multiple exponential fitting (including residuals below) of HRP

enzyme in 2.0 mM APG. .............................................................................................................. 123

Figure 5-55 - Decay data and multiple exponential fitting (including residuals below) of HRP

enzyme in 10.0 mM APG. ............................................................................................................ 124

Figure 5-56 - Fluorescence decay times as a function of surfactant (APG) concentration ......... 125

Figure 5-57 - Fluorescence decay times as a function of surfactant (dodecyl maltoside)

concentration .............................................................................................................................. 126

Figure 5-58 - Amount of CRM-197 adsorbed onto AlPO4 particles at pH 4 and 5 as a function of

initial protein concentration. ...................................................................................................... 127

Figure 5-59 - CRM-197 structure with Trp residues highlighted ................................................. 128

Figure 5-60 - Fluorescence decay times resolved from CRM in pH 7 and 5.8 buffer and adsorbed

onto AlPO4 at pH 5.8 ................................................................................................................... 129

Figure 5-61 - The decay contribution of each decay time. .......................................................... 130

Figure 5-62 - a) (left) hydration sphere of densely packed micellese such as that of dodecyl

maltoside and high conc. APG consists of tightly bound water. b) (right) loosely packed micelles

result in more mobile hydration sphere ..................................................................................... 137

Figure 5-63 - Minimization of loosely bound hydration layers leading to enzyme/micelle

interaction ................................................................................................................................... 137

Figure 5-64 - Schematic differentiating micell formation in DM (top) vs APG (bottom) ............ 144

Page 13: Protein Structure Dynamics and Function at Micelle and

vii

Acknowledgements

I would like to extend my deepest thanks to all those who have supported this

undertaking and made this work possible: Firstly, to my advisor, Professor P.

Somasundaran who has guided my efforts throughout the years with wisdom, knowledge,

and patience.

I would also like to acknowledge my committee members, Professor’s Arthur Palmer III,

John Hunt, Xi Chen, and Raymond Farinato. Their advice and insight throughout the

course of this research have been invaluable. Similarly, the input from Dr. Nagaraj,

Professor Moudgil, my fellow co-workers, and all of the participating members of the

Columbia Center for Particulate and Surfactant Systems (CPASS), has helped to refine

this work into what it is today.

Finally I wish to thank my family and friends: My parents and sister for their ever-present

emotional and financial support, and my best-friend, who convinced me to pursue this

unique opportunity during a time when I might have decided otherwise.

Page 14: Protein Structure Dynamics and Function at Micelle and

1

1 Introduction

Enzymes are transitioning from niche technologies to a prominent tool in industrial

processing, energy production, and pharmaceuticals and medicine. One common use for

enzymes is in commercial laundry applications, where they are incorporated into

detergents to break down sugars, lipids, and proteins, so that stains ingrained into fabrics

are removed with less thermal and physical agitation [1], thereby conserving water and

energy. Improving upon the function, stability and reliability of enzymes and other

functional proteins is an increasingly important goal to fields of alternative energy as

well. In the case of bio-fuels, enzymes are a prominent strategy towards efficiently

liberating free sugars from cellulosic bio-feed stocks under low-energy conditions with

minimal impact to the environment. Other emerging concepts include bio-fuel cells [2,

3], in which enzymes catalyze sugar into free electrons [4], and bio-solar[5], in which

light-harvesting proteins are used to capture light and transfer energy to reaction centers

in photovoltaic cells. Control of macromolecules behavior also plays a major role in

medical technology - such as the interactions between anti-gen and adjuvant particles in

vaccine formulation. Understanding the interactions of enzymes and other proteins with

detergents and at particle is becoming an increasingly important goal. This particular

work will focus on elucidating the interactions between enzymes and detergents,

particularly, the mechanism by which certain surfactants have been reported to increase

enzyme activity.

Page 15: Protein Structure Dynamics and Function at Micelle and

2

1.11.11.11.1 Initial Initial Initial Initial findingsfindingsfindingsfindings and what is in the literatureand what is in the literatureand what is in the literatureand what is in the literature

Surfactant dependent enzyme activity was observed by our own initial experiments when

testing the activity of a subtilisin protease in solutions of detergents typically used in

detergent applications, including: anionic sodium dodecyl sulfonates and bio-based, non-

ionic, alkyl polyglucosides. As expected, ionic surfactants slightly diminished the activity

of the enzyme relative to a buffer only solution, presumably from denaturation of the

enzyme structure[6] (Figure 1.1) the two non-ionic surfactants tested, including the

sugar-based APG, increased enzyme activity, however.

SDS LDS Brij-30 APG

-10

0

10

20

30

40

50

60

70

Activity r

ela

tive

to

su

btilis

in in

bu

ffe

r

Sufactant

Figure 1-1 - Initial subtilisin activities in anionic and non-ionic surfactants

Our finding is not the first. In the past 20 years there have been several reports of

elevated enzymatic activity in the presence of non-ionic surfactants. This phenomenon

has been observed in wide range of enzyme types and target applications, ranging from

Page 16: Protein Structure Dynamics and Function at Micelle and

3

amylases and proteases for digesting sugars and proteins suspended in bulk solutions and

from fabric surfaces[7-10] to cellulases used to liberate free sugars from crystalline

polysaccharides (figure 1.2) for use in bio-fuel production [11-13]. In our own

preliminary experiments, we have observed enhanced proteolysis in the presence of sugar

based non-ionic alkypolyglucoside (APG) as well as a polyoxyethylene lauryl ethers (Brij

series surfactants). This enzymatic hyperactivity has also been observed in reverse

micelle systems[14, 15]. Regardless of the enzyme, substrate, or solvent environment, it

is particular non-ionic surfactants that are consistently observed promoting this

enhancement. Ionic surfactants, on the other hand, denature protein structures [6, 16-18]

reducing the proteins function.

Figure 1-2 - Increased hydrolytic rate seen in Bacillus amyloliquefaciens α-amylase (BAA) as a function of Brij-35

concentration as reported by Hosino et al.

While there is an apparent and consistent connection between non-ionic surfactants and

increased enzyme activity in the literature, the underlying mechanism that explains how

these surfactants are inducing this phenomenon is not clear. Shome et al [14] conclude

Page 17: Protein Structure Dynamics and Function at Micelle and

4

that additions of non-ionic surfactants to the cationic surfactant mixtures reduce the

surface charge at the head-group / protein interface, limiting the chance of electro-static

inhibition of an active site. Kaar et al [13] suggests that the non-ionic Tween surfactants

helped to maintain cellulase structure as it digested cellulosic bio feed stock into free

sugars; Erikson and colleges disagree with this assessment [11], and show results

supporting their own hypothesis that the non-ionic surfactants prevent non-specific and

permanent cellulase adsorption onto the cellulose surface, rendering the enzyme inactive.

Alternatively, Hoshino et al [7] attribute their observations of enhanced amylase activity

in Brij-35 surfactants to both the enzyme and substrate having an affinity for the micelle

surface, concentrating both elements at micelle interface thereby increasing maximum

reaction rate. While these empirically derived conclusions explain the enzyme activity

enhancement in each system unique to their respective study, a universal theory that

might explain why non-ionic surfactants should increase activity in *ALL* of these

studies has not yet been produced.

Theoretical models have also been formulated to mathematically describe, patterns of

enzyme hyperactivity and activity loss as a function of non-ionic surfactant

concentrations, starting with theoretical modeling of enzymes in reverse phase systems.

In 1988, Kabanov and colleagues [10] pointed out that enzyme activity in reverse micelle

systems are maximized at specific water to surfactant ratios. They postulated that reverse-

micelles in specific solvent concentrations are just large enough to house the enzyme,

thereby maximizing the interactions between micelle headgroups and the enzyme surface.

Page 18: Protein Structure Dynamics and Function at Micelle and

5

Their proposed model predicts enzyme activity as a function of total surface area of

enzyme and micelle interaction.

In parallel to Kabanov’s work, Bru, Sanchez-Ferrer and Garcia-Carmona also outlined a

theoretical model explaining enzyme superactivity in reverse micelle systems [19],

emphasizing the role of three possible pseudo-phases: hydrophobic regions of surfactant

tails, bound water phase at the surfactant heads, and free water which occurs inside the

inverse micelles as they grow larger. By taking into account the kinetics of micelle

aggregation and enzyme / substrate partitioning throughout the three pseudo-phases Bru

and colleagues developed a model to express the concentration of enzyme and substrate

in any of the three phases. This knowledge was coupled with an assumption that catalytic

rates differ in each phase, their final simulation accurately match previous experimental

observations of super-activity and inhibition of enzymes in reverse micelle systems.

Viparelli et al built upon Bru’s work [20], applying the “pseudo-phase approach” to

standard micelles in aqueous solutions. They modeled the steady state concentrations of

enzymes within bound water boundaries at the micelle interface versus the free water

phases in the bulk over a range of surfactant concentrations, and enzyme and substrate

affinities. Also like Kabanov and Bru etal, Viparelli concluded that maximizing enzyme

concentration in the bound water pseudo-phase of the micelle surface plays a key role in

enzyme super-activity.

Page 19: Protein Structure Dynamics and Function at Micelle and

6

Figure 1-3 - Three phases that enzyme hydrolysis can occur. Bulk (Sf), Bound water interface (Sb) and within the

micelle as described by Viparelli et al.

These theoretical studies attempt to predict enzyme activity in the presence of non-ionic

surfactants and offer a general model that can be applied universally to different enzyme,

substrates and surfactants. However, all the studies mentioned here only provide cursory

physical mechanisms to explain the faster specificity and catalytic constants of the

enzymes within these bound water regions. One of the papers highlights this:

“Of course, there may be other "molecular mechanisms" that can explain the enzyme

catalytic activity as a bell-shaped function of micelle dimensions. Nevertheless, despite

the possible variety and complexity of these mechanisms, for a formal kinetic description

of observed regularities it is sufficient to consider only three types of micelles: "small",

"large" and "optimal". The catalytic activities are different for enzymes solubilized in

micelles of each type: the kcat value is maximal in optimal micelles”

– Kabanov et al. 1988

Page 20: Protein Structure Dynamics and Function at Micelle and

7

1.21.21.21.2 Gaps in knowledgeGaps in knowledgeGaps in knowledgeGaps in knowledge

In section 1.1 we mentioned a considerable body of work extending into the experimental

and theoretical aspects of this enhancement phenomenon. However, when comparing

them all side-by-side, a considerable gap in knowledge becomes evident. As previously

stated, there is still no consistent explanation for how a seemingly inert colloid surface,

such as at a non-ionic micelle surface, can influence the activity of an enzyme. And while

the theoretical studies offer models that predict how enzyme activity may respond to

surfactants as a function of surfactant concentration, the physical explanation as to why

such an enhancement phenomenon should occur are lacking. This work attempts to fill in

this gap in understanding by studying one aspect of enzyme function that has not been

studied in the aforementioned works: structural dynamics, and how they might be

influenced by interactions with particular surfactants.

The protein structures and structural dynamics play an integral role in the function of

many proteins. This can be clearly seen in enzymes, where conformational changes in the

protein structure are a perquisite to the substrate binding, transition state formation and

energy minimization steps required for enzymatic catalysis. Other proteins that rely on

structure dynamics include: proton pumps, ion transport channels and photosystems.

These functional proteins are becoming increasingly prominent in emerging alternative

energy technologies such as bio-fuel production, bio-solar and enzymatic fuel cells; as

such, they will be exposed to surfaces and systems outside the native cellular

Page 21: Protein Structure Dynamics and Function at Micelle and

8

environment. Understanding the influence of these various ex vivo conditions on the

structure dynamics and functionality of proteins also becomes an increasingly prominent

priority.

Aside from enzymes, many other proteins rely on conformational changes at various

scales to function. This project will aim to extend the mechanisms elucidated from these

enzyme/surfactant systems, to other protein/colloid interfaces. To do this, we will

construct a model to describe how changes to the structural dynamics of a light

harvesting protein such as bacteriorhodopsin can affect its overall energy transfer

mechanism. Bacteriorhodopsin is a membrane protein found in halophilic bacteria. They

use sunlight to pump protons across the cell membranes to maintain proper chemical

gradient, which is in turn used to create ATP through ATP synthase[21, 22]. The

structure is composed of three protein chains that are embedded in the bacteria

membranes. In each of these chains is the light sensitive chromophore retinol which

undergoes a Cis-Trans conformational change when excited by a photon. This

conformation change induces further structural shifts along the protein chain which

ultimately move protons in a particular direction. This retinal switch occurs in the time

scale of 200-500 femto-seconds while the subsequent structure changes related to proton

pumping occur within picoseconds[23, 24]. X-ray diffraction and NMR demonstrate that

the collective motions during these pumping cycles have an average deviation of 3Å.

Clearly, these mechanisms are fast and precise. It is not inconceivable that attaching these

proteins to inorganic surfaces for use in bio-solar applications, may alter with these

dynamics and hence their electron transfer properties.

Page 22: Protein Structure Dynamics and Function at Micelle and

9

Figure 1-4 - Top down view of three protein chains which compose of a bacteriorhodopsin unit. One is shaded blue

with a white outline indicating the location of the retinol chromophore.

In addition, changes in protein structure, structural dynamics, and inter-particular forces

are paramount to the stability of many drug formulations, such as vaccine carrier proteins

adsorbed onto a solid, metal phosphate, particle surfaces[25]. There is limited knowledge

on the protein structure, and dynamics of protein-particle, and protein-protein interactions

when absorbed onto a particle surface such as aluminum. Gradual changes in protein

structure as well as its structural dynamics on particle surface may result in shifts in

protein electrostatics, hydrodynamics and steric interactions may lead to protein-protein

attraction. The structural dynamics of these vaccine carrier proteins as well as those of

bacteriorhodopsin at particle surfaces are not well understood, study parallel to studying

protein interactions at micelle interfaces may give us greater insights into protein

behavior at a solid-liquid interface and how forces that arise at these interties affect the

structure, structural dynamics of macro-molecules.

Page 23: Protein Structure Dynamics and Function at Micelle and

10

1.31.31.31.3 Central gCentral gCentral gCentral goal of oal of oal of oal of dissertationdissertationdissertationdissertation::::

Our goal is to propose one such “molecular mechanism” that would explain this broadly

observed enzyme hyper-activity phenomenon in the presence of particular surfaces

observed in previous literature as well as through our own experiments. In addition, we

aim to study proteins such as CRM-197 diphtheria toxin mutant on the surfaces of hard

colloids such as metal oxide particles. If successful, we believe the lessons learned from

the comparisons of proteins at soft micelle surfaces vs. hard particles can be applied to

any system where functional proteins are utilized in close packed environments as it is in

the fields of renewable energy, waste management, home-personal care and medicine.

Explanations in the works cited above are varied and inconsistent, proposing non-specific

binding, structure preservation and enzyme substrate concentration at the bound water

interface, to explain enzyme hyperactivity while theoretical modeling infers faster Kcat

of the enzyme itself. What has not yet been considered in the realm of surfactant /

enzyme interactions is the role of structural dynamics in enzyme function and how they

can be affected by the bound water at these micelle interfaces. Furthermore, small scale

structure dynamics play a major role in the charge transfer mechanisms of proton

pumping proteins. How these motions are affected by colloid surfaces will also be

investigated.

Page 24: Protein Structure Dynamics and Function at Micelle and

11

2 Hypothesis and rationale

We have proposed and tested the following hypotheses:

1) The surfactant enhances enzyme activity by improving flexibility of the enzyme

without denaturing or distorting the native structure

2) These flexibility improvements are influenced by micelles that extend the range

of bound water regions around the protein structure which impart less kinetic

energy into the protein structure via thermal fluctuations and shifts in electro-

static bonding constants, minimizing the amplitude and frequency of atomic

collisions within the structure, thereby increasing rate and range of induced

conformation changes.

3) The implications of these outcomes extend to other applications where functional

proteins are used in dense and restricted environments, often coupled directly to a

surface. I.E bio-solar, cellulosic bio-fuels and vaccine applications

Page 25: Protein Structure Dynamics and Function at Micelle and

12

2.12.12.12.1 Rationale Rationale Rationale Rationale forforforfor HHHHypotheseypotheseypotheseypotheses:s:s:s:

It is well known that proteins are not static structures [26-29] but rather dynamic bodies

that exhibit motions over a wide range of time and length scales (figure 2.1). These

motions fall under three major categories [30]:

1) Fluctuations (i.e. atomic vibrations): These motions are random, very fast and

cover distances less than 0.5Å. The energy for these motions comes from kinetic

energy inherent in the system as a function of temperature. Spatial displacement

range between 0.01 – 1 Å, temporal scale 10-15 to 10-11s

2) Collective motions: This category includes the motions of groups of atoms that

are covalently linked in such a way the group moves as a unit. These collective

motions are further divided into two subcategories: those that occur quickly, but

infrequently, like the flipping of tyrosine ring, and those that occur slowly. The

energy for these transitions is also derived from the thermal energy of the protein.

Spatial displacement between 0.01 to 5.0 Å, temporal scale 10-12 to 10-3 s

3) Induced motions: The third category includes motions that are triggered

conformational changes. These are the motions of individual side-chains or whole

protein sections, which occur as a response to a specific stimulus. The energy for

triggered conformational change comes from specific interactions. 0.5 to >10Å,

temporal scale 10-9 to 103s

Page 26: Protein Structure Dynamics and Function at Micelle and

13

Figure 2-1 - Range of time scales associated with protein structure dynamics. This study with focus on elastic

fluctuations occurring in the ps range.

The roles of these dynamics are a subject of great interest for elucidating the many

aspects of protein function [31]. Enzyme function is closely linked to the dynamic nature

of its structure. In the late 19th century, it was thought that the highly specific nature of

enzyme / substrate binding predicated that the active site be shaped exactly to the

cleavage site, as a key hole is perfectly shaped for one specific key, otherwise known as

the “lock-and-key” model. However, as characterization techniques improved, this static

view of enzyme mechanic had to be reconsidered. And so in the 50 some years since the

introduction of the lock-and-key model, conceptualization of enzymes and proteins had

shifted away from those of static structures; to instead, as soft assemblies that are

constantly sampling random conformations about an mean structure position [32, 33].

In the 1950’s, the static “lock-and-key” view of enzyme/substrate interactions was

modified into the “induced fit” model, wherein enzyme structures undergo a

conformational change as the active site is reshaped during its interactions with the

Page 27: Protein Structure Dynamics and Function at Micelle and

14

substrate [27, 34]. This dynamic model of enzyme/substrate interaction opened up

pathways to the elucidation of the various mechanisms employed by enzymes to lower

the barrier energy of chemical reaction: Inducing bond strain, bringing reactive groups

into close proximity or by creating a covalently bonded intermediate complex are several

examples of energy barrier reduction pathways induced by conformation changes [33].

Because of this prominent role that structure dynamics has in substrate binding and

energy barrier minimization, investigation of the relationship between the enzymes’

structural dynamic properties, and their catalytic effectiveness has become an intensely

studied field of research [28, 29, 35-38].

Figure 2-2 - General schematic and description of induced fit model. Active site conformation changes play an

important role in minimizing energy barrier to product formation [33]

Since it has been established that flexibility plays a major role in many protein functions,

this project will focus on studying factors that affect flexibility, particularly those that

stem from interactions with interfaces within the proteins’ local environment. Studying

the link between a protein’s local environment and flexibility has been investigated in the

Page 28: Protein Structure Dynamics and Function at Micelle and

15

past. Using methods that measure local flexibility about the enzyme active site, coupled

with comparative enzyme kinetics studies, several groups have demonstrated that

solvents play a substantial role in dictating enzyme flexibility [39-42]. These studies

consistently cite solvent interaction with structurally bound water as the root cause of

flexibility change. In parallel, many other studies point to hydration as an integral part of

protein function [43-45] [46], affecting bond distances [47], stabilizing structure,

facilitating protein folding [48] and acting as a lubricant, facilitating conformation

transitions [49], whether or not these same effects can be induced or influenced by

detergent micelles or particles has yet to be considered.

However, many studies have demonstrated that inorganic particles and surfactant

aggregates manipulate bound water dynamics at their interfaces. Recent work done by

Fayer et al utilizes ultra-fast infrared spectroscopy to observe orientation dynamics of

water at the interface of reverse AOT micelles [50, 51]. They found two distinct

reorientation correlation times; slow restructuring is attributed to bound water near the

micelle interface while a significantly faster transitioning phase is assigned to bulk-like

water in the micelle core. By decreasing the size of these micelles from 10 to 1 nm

diameter, Fayer et al. observed a transition between the amounts of this slow interfacial

water relative to fast moving bulk. In micelles less than 3nm in diameter, the interfacial

water reorientation characteristics are dominant. Comparisons were also made between

ionic (AOT) and non-ionic (Igepal) and it was found that the restructuring times are more

influenced by total surface area in the system rather than the chemical composition of the

hydrophilic groups. Given the connection between protein structure dynamics and

Page 29: Protein Structure Dynamics and Function at Micelle and

16

hydrating water, it is reasonable to hypothesize that those colloid interfaces which do not

adsorb directly to protein structure, can still influence the dynamics of the structure

through bound water interactions at the micelle surface.

While solvent conditions may be considered an “external” effecter on protein structure

dynamics, the protein itself provides its own “internal” mechanism that regulates

flexibility and conformational change. For some time the effects of smaller, pico-second

scale fluctuations of local atomic groups on the range of movement of global structures

has been studied [52, 53]. These thermodynamically driven fluctuations have been shown

to dictate the range and rate of conformation changes [53-55] resulting in a hierarchy of

dynamics. In some cases, these fluctuations are the driving force facilitating the moment

of larger and slower movements, such as the hinge bending seen in enzymes like

adenylate kinase [30, 54]. Conversely, frequent internal collisions of adjacent groups are

responsible for limiting net displacement [56] Hindering overall flexibility of the overall

structure [57, 58]. On the importance of dynamics on function [59]:

“Understanding the role of enzyme dynamics in the chemical step is both important and

complex. (Garcia-Viloca etal) describes possible equilibrium and nonequilibrium

contributions and discusses motions of the enzyme that can lower the free energy of

activation of enzyme reactions. Crossing of the transition state region is usually very fast

(in the fs to ps range) in enzyme reactions, and the reaction seems slow only because the

activation barrier makes it improbable to reach the transition state. Thus, fs to ps

motions must play a role, although whether they contribute to the faster rate of the

reaction in the enzyme versus that in solution has yet to be determined.”

- Karplus (2010)

Page 30: Protein Structure Dynamics and Function at Micelle and

17

Our initial experiments with subtilisin protease have demonstrated greater enzyme

activity in solution with akyl polyglucoside surfactant, a phenomenon that has been seen

in multiple studies in the past decade. Surfactants already play a major role in protein

stability, crystallization and solubilization[60, 61], but the possibility of surfactants

improving enzymatic activity through modification of structural dynamics has not yet

been considered. This study intends to elucidate this connection. However it will not be

enough to only demonstrate a correlation between flexibility and activity in the presence

of surfactants when the underlying mechanism of surfactant / enzyme structure remains

unknown. Various studies provide clues to the nature of these interactions. One likely

mechanism is through the protein structure hydration, which affects the overall free

energy of the structure[62], lubricating internal motions and directing folding pathways.

Measurements of protein structural dynamics in different solvent environments also

indicate changes occur because of bound water being displaced from adsorption sites -

this denotes the importance of bound water to structure dynamics [39, 47]. As mentioned

above, the surfaces of micelles play a large role in influencing the volume of bound water

in the system[42, 50], as well as the rate at which hydrogen bonds are exchanged[47].

This project intends to study these interfacial bound water dynamics as the primary

mechanism by which non-ionic surfactant interact with protein structures, thereby

influencing structural dynamics and dependent functions, such as enzyme activity. Since

ceramic and oxide nanoparticles are also known to attract bound water or affect water

structure at interfaces [63, 64], the structure dynamics of various functional proteins,

including bacteriorhodopsin, subtilisin enzyme and BSA have also been studied at hard

particle surfaces in parallel with the surfactant studies.

Page 31: Protein Structure Dynamics and Function at Micelle and

18

3 Research plan and objectives

Overall Goal: To demonstrate that proteins at a non-ionic interface are not completely

free of influence from its local environment, and that these interactions play can still play

a major role in affecting the steady-state dynamics of the proteins structure and thereby

it’s function. In addition this project aims to propose a mechanism by which non-ionic

surfaces may affect protein structural dynamics.

3.13.13.13.1 SubSubSubSub----objectives: objectives: objectives: objectives:

1. Select a system where enzyme activity / kinetics parameters are enhanced or retarded

by surfactant interactions

2. Determine the link between changes in local and/or global structure flexibility and

enzyme activity relative to surfactant concentration

2.1. Study the possible link of micelle structure vs. surfactant concentration

2.2. Correlate changes in protein structural flexibility to changes to bound water

dynamics with increasing surfactant concentration

3. Find parallel changes in function and structure dynamics of proteins

3.1. Demonstrate that changes in steady state conformation fluctuations and overall

structural mobility can be influenced by solid particle surfaces as well

4. Propose mechanism by which colloid surfaces can affect conformation fluctuations

without denaturing protein structure

Page 32: Protein Structure Dynamics and Function at Micelle and

19

4.1. Propose model which describes FRET efficacy as a function of protein flexibility

and fluctuation for use in bio-solar applications

Our initial goal was to elucidate the mechanism dictating the “hyperactivity”

phenomenon seen in enzymes in non-ionic surfactant solutions. We hypothesized that

changes in the enzyme structural flexibility are induced by bound water interactions

found at the micelle/water interface. From there, we extend the findings from this

particular surfactant / enzyme system to develop a holistic understanding of the

interactions between colloid surfaces and protein structural dynamics. To test our

hypotheses the project was divided into a series of four objectives. The first was to

determine a model system with which to test enzyme/surfactant interactions. This was

done primarily through enzyme activity and kinetics assays. The second objective was to

measure the structural mobility of the enzyme and observe how these changes with

respect to surfactant structure and concentration. A correlation was found between these

enzyme flexibilities and enzyme kinetics properties observed from the activity

measurements found while fulfilling our first objective. Since this correlation between

flexibility and surfactant concentration was found, we proceeded to our final objective

which was to describe a general mechanism by which colloid surfaces influence protein

structural dynamics. Our hypothesis, based on theoretical work of Bru, Viparelli and

Kabanov and their respective groups cited that the bound water region at the micelle

interface is instrumental in this hyperactivity phenomenon. Work to further study the

effects of bound water structure on enzyme activity and flexibility are now on going,

however, preliminary tests of enzyme activity in Hofmeister series of ions confirm the

Page 33: Protein Structure Dynamics and Function at Micelle and

20

connection. Additionally, cooperative work with the University of Bergen is underway to

use NMR to measure which specific amino acid residues are affected most.

Page 34: Protein Structure Dynamics and Function at Micelle and

21

4 Methods and Materials

The methods described in this section are divided into the sub-objective for which they

were designed to achieve. The first objective is to find suitable model enzyme and

surfactant systems which induce faster enzyme activities and compare them to those that

do not. The primary technical challenge of this objective will be to devise or adapt a

suitable method for quantifying enzyme activity. The second objective is to elucidate the

interactions between surfactant and enzymes, or the physical characteristics of the

surfactant-enzyme aggregates that form upon mixing the two components. The third

objective aims to investigate the role of the enzyme’s structural mobility in determining

activity and how surfactants maybe affecting this structural dynamic.

4.14.14.14.1 Objective 1 Objective 1 Objective 1 Objective 1 ---- Find suitable model system to test enzyme hyperactivity via Find suitable model system to test enzyme hyperactivity via Find suitable model system to test enzyme hyperactivity via Find suitable model system to test enzyme hyperactivity via

interaction with surfactants interaction with surfactants interaction with surfactants interaction with surfactants

The first objective of this study was to select enzyme and surfactant pairings that will

result in measureable changes in catalytic activity. For this purpose two enzymes,

subtilisin protease and horse radish peroxidase (HRP) were studied in the presence of 6

surfactants: two commonly used anionic surfactants, sodium dodecyl sulfate and linear

alkylbenze sulfonate, and four non-ionic surfactants, alkyl polyglucoside, dodecyl

maltoside, Polyoxyethylene (4) lauryl ether (Brij-30) and its larger headgroup

Page 35: Protein Structure Dynamics and Function at Micelle and

22

counterpart, Polyoxyethylene (23) lauryl ether. The rationale for selecting these enzymes

and surfactants is explained in the following sub-sections.

4.1.14.1.14.1.14.1.1 Enzyme selectionEnzyme selectionEnzyme selectionEnzyme selection

The enzymes subtilisin and HRP were both chosen because their structural and functional

characteristics have been well studied and documented. Subtilisin proteases are

commonly used in commercial and industrial applications. Likewise, horseradish

peroxidase is commonly used for bio-chemical research applications. Furthermore, both

of these enzymes are readily available in large, relatively pure quantities. The subtilisin

families of enzymes are a non-specific protease initially obtained from the Bacillus

Subtilis bacteria strains and are commonly used for a large variety of application [65],

including as a components in detergent formulation to aid in protein based stain removal.

Over the last decades, several variations of the subtilisin protein have been isolated.

Subitilisin Carlsberg refers to enzyme derived from Bacillus Subtilis bacteria. Subtilisin

BPN’ originates from the B. Subtilis strain N’ and subtilisin Novo refers to the enzyme

from the Novo pharmaceuticals company. For this study, subtilisin Carlsberg is

predominately used and will be referred to from here on in as subtilisin unless specified

otherwise.

Page 36: Protein Structure Dynamics and Function at Micelle and

23

Subtilisin Carlsberg has a total molecular weight of 27,287 in the form of a single peptide

chain devoid of cysteine or cysteine, phosphorous and metals and an isoelectric point of

9.4 kpH. The tertiary structure is approximately 30% alpha-helical resulting in a protein

that is close to spherical and approximately 42Å in diameter[66]. The catalytic center is a

Asp-His-Ser catalytic triad with the nucleophilic serine-OH group located in a shallow

groove in the enzyme surface[40]. The subtilisin used for this study was all products

ordered from Sigma Aldrich in different forms. Initially P4860 was used, this is a

subtilisin product produced by Novozymes under the trademarked name of Alcalase.

However, since the exact composition of the proteins and suspension additives were not

known, P4860 was later replaced with P5459, another subtilisin product suspended in a

polyethylene glycol solution. For the time resolved fluorescence experiments P5459 was

initially used, but the polyethylene was thought to interfere with the probe attachment

process, so P5380 (Sigma-Aldrich), a lypholized subtilisin, was used predominately for

those experiments.

Horseradish peroxidase is a glycoprotein isolated from plants and is commonly used in

biochemistry applications. The majority of reactions catalyzed by HRP can be

summarized as such:

H2O2 + 2AH2 � 2H2O + 2AH●

Page 37: Protein Structure Dynamics and Function at Micelle and

24

Where AH2 is a substrate, typical aromatic phenols, phenolic acid, amines or sulfonates

and AH● is the radical product.

There are several isoforms of HRP[60], but the most commonly studied is the ‘C’ variant

derived from the horseradish plant. The primary structure of HRP consists of 308 amino

acid residues forming a single polypeptide chain. Unlike subtilisin, HRP incorporates

carbohydrates and metals into its structure, as well as multiple cysteine disulfide bridges

to maintain structure. Four disulfide bridges are found in HRP between eight cysteine

residues. Three metal centers are also present in the HRP structure, an iron (III)

protoporphyrin (heme) group located near the center between two clearly defined

domains of the structure. On either side of the heme group are two calcium ion binding

sites, one distal and the other proximal to the heme group. Both of these metal groups are

essential to the peroxidase structural integrity while the heme group is also central to the

catalytic mechanism. The weight of the polypeptide chain is 33,890 Da, the heme group

is 572 Da, and the calcium ions 80 Da. The total weight of HRP C is approximately 42

kDa assuming 20% contribution from the carbohydrate moieties[60].

4.1.24.1.24.1.24.1.2 SurfactantsSurfactantsSurfactantsSurfactants

Several surfactants have been reported in literature to improve enzyme activity. The large

majority of these are non-ionic structures including: Tween-20, Tween-80, and Brij-30.

In our own preliminary experiments, we have identified improved subtilisin activity in

solutions of alkyl polyglucosides (APG). For this study the synergistic properties of

APG’s as well as Brij-30 on enzymes were further studied. Dodecyl maltosides and Brij-

Page 38: Protein Structure Dynamics and Function at Micelle and

25

35 were also added to compare the effects of long chain ethoxylate chains of varying

lengths vs. the sugar groups of APG and DM on the enzyme activity.

Figure 4-1 - Dodecyl maltoside

The alkyl polyglucosides are a class of non-ionic surfactants derived from sugar sources

and usually in the form of a mixture consisting of a range of linear tail chain lengths and

sugar head group oligomer size. The APG used in this study was a commercial product

supplied by the Cognis specialty chemicals company with a tail chain length between 12

and 14 carbons and head group sizes between one and three sugar groups. However the

exact composition of tail and head group size distributions are not known, for this reason,

beta-D-maltoside a pure, single sugar, is also used in this study to better control the

surfactant composition in solution.

The polyethylene glycol dodecyl ethers known by their trade name, Brij, has been used in

several of the previous papers that have reported increased enzyme activity in non-ionic

surfactants. Brij-30 (polyoxyethylene (4) lauryl ether was most commonly cited of the

Brij surfactants and is used in this study as well. The effects of larger head group sizes

Page 39: Protein Structure Dynamics and Function at Micelle and

26

were also studied with Brij-35, which also features a linear polyoxyethylene headgroup,

but several times longer with 23 oxyethylene groups. In parallel to these four non-ionic

surfactants, two ionic surfactants were tested to compare the effects of charged

surfactants on protein structure. Sodium dodecyl sulfate (SDS) and linear alkylbenzne

sulfonate (LAS) were selected for this reason.

4.1.34.1.34.1.34.1.3 Subtilisin assay:Subtilisin assay:Subtilisin assay:Subtilisin assay:

Subtilisin activity was measured primarily in two ways, product formation assays and

initial rate activities. Product formation assays were measured by the amount of tyrosine

liberated from casein substrate by subtilisin in buffer and various surfactants. The

liberated tyrosine reacts with Folin-Ciocalteu reagent, turning it from from yellow to

blue. The color change is measured with UV-Vis and presented below normalized versus

the enzyme in buffer only.

Page 40: Protein Structure Dynamics and Function at Micelle and

27

Figure 4-2 - Example of colormetric activity assay

4.1.44.1.44.1.44.1.4 Subtilisin Subtilisin Subtilisin Subtilisin initial rate activities and kineticinitial rate activities and kineticinitial rate activities and kineticinitial rate activities and kineticssss

The second method of quantifying enzyme catalyzed proteolysis rates was done via a

method modified from Broos etal [67], which utilizes fluorescein labeled casein as a

substrate of proteolysis. In its native state the fluorophores are in very close proximity

and their fluorescence is quenched by adjacent fluorescein. Changes in FITC

concentration, which is released from the casein substrate during the course of the

reaction, is measured as a function of time during the relation. Fluorescence intensity is

monitored with Horriba Yvoin fluorolog -3 fluorometer. The reaction takes place in a 3ml

quartz cuvette (1cm path length) where an aliquot of the enzyme (1.0ml) is added to the

FTIC casein suspension (2.0ml) under constant stirring. Alternately, the reaction is

Page 41: Protein Structure Dynamics and Function at Micelle and

28

started in 1.5 ml plastic vials by adding 250µl of enzyme to 500µl of substrate, shaking

and quickly transferring to a 0.56 ml reduced volume quartz cuvette with a 4mm path

length (Starna cell). A series of emission intensity measurements at 519nm is taken once

every 2 seconds for 10 minutes, starting shortly after the enzyme addition. This reaction

is measured at 22 degrees C.

Figure 4-3 - Schematic illustrating liberation of free fluoresein probe from labled casein

4.1.54.1.54.1.54.1.5 HRP Kinetics: HRP Kinetics: HRP Kinetics: HRP Kinetics:

Kinetics rates were obtained from the aminoantipyrine method [68], spectroscopically

measuring the increase of absorbance at 510nm resulting from the formation of the

colored complex AMPH. Two solutions are prepared: one consisting of 2.5 mM

aminoantipyrine and 170 mM phenol in DI water and 1.7 mM Hydrogen peroxide also

diluted in DI water. 0.01 mg/ml. A 1:1 mixture of the two reagents were prepared shortly

before a small aliquot of HRP in buffer is added to the substrate mixture. The final

concentration of HRP was 0.01 mg/ml. The resulting colorimetric evolution was

monitored by UV-Vis spectrometer at 510nm.

Fluorescein isothiocyanate

Page 42: Protein Structure Dynamics and Function at Micelle and

29

4.1.64.1.64.1.64.1.6 Subtilisin assay side note: Behavior of Fluorescein Probe (optimizing subtilisin Behavior of Fluorescein Probe (optimizing subtilisin Behavior of Fluorescein Probe (optimizing subtilisin Behavior of Fluorescein Probe (optimizing subtilisin

assay method)assay method)assay method)assay method)

Fluorescein is a common fluorophore dye used in a wide variety of applications, ranging

from biology and geology to medicine and oil recovery. It has also been critical to our

enzyme kinetics experiments done up until this point, as activity is actually the measure

the release of fluorescein as an indicator of enzyme activity. Previously, we have reported

the results as changes in intensity over time, but this is only relative information. Using a

concentration standard these relative intensity results can be correlated to quantitative

“freed” flourescein concentrations.

Furthermore, it is important to find the concentration parameters in which fluorescence

intensity and probe concentration deviates from Beer-Lamberts law. An additional

concern when using the fluorescien probe is that they can interact with one another,

quenching emission and giving rise to another limitation to the experimentally observable

concentration range. These properties need to be carefully considered when selecting

experimental concentration ranges and interpreting the intensity results.

MethodsMethodsMethodsMethods:

0.33 g of fluorescein was added to 50mM potassium phosphate buffer with pH adjusted

to 7.5. Mixture stirred for 60 minutes allowing probe to dissolve and saturate solvent, any

after an hour any undissolved fluorescein was centrifuged, filtered out, dried and weighed

Page 43: Protein Structure Dynamics and Function at Micelle and

30

to determine actual probe solution concentration. 0.18 grams of fluorescein probe

dissolved into the buffer, making a max solution concentration of 25mM.

This concentrated solution was measured with steady state fluorescence. Excited by 490

nm light and emission spectrum between 500 and 600 nm was measured. Lamp and

detector slit widths where 0.5 and 1 mm respectively. Each successive measurement was

diluted to half the concentration of the previous and measured again. The resulting

spectra are shown below.

Figure 4-4 - Fluoresence spectroscopy of fluorescein fluorophore dilution series

0

100000

200000

300000

400000

500000

600000

500 520 540 560 580 600

Inte

nsi

ty

wavelength

25

12.5

6.25

3.125

1.5625

0.78125

0.390625

0.1953125

mMMax intensity at

approximately 100x dilution

Page 44: Protein Structure Dynamics and Function at Micelle and

31

These over lapping plots are hard to visually deconstruct two crucial elements are

immediately observable: 1) The highest concentration does NOT result in greatest

emission intensity and the 2) fluorescence peak is shifting with concentration.

First we plot changes in plot intensity with concentration. Below 0.1 mM, the peak

emission wavelength is approximately at 512 nm, the value reported in literature. At

higher concentrations, however, we see an exponential shift in the peak positions which

strongly indicate probe excimer formation, resulting in energy transfer between adjacent

probes producing emitted light of lower energy.

Figure 4-5 - Fluorescein peak wavelength vs. concentration

Next, we look at the peak intensities with concentration: As mentioned before, the

greatest emission intensity occurs at 0.2 mM and quickly declines afterwards. This

Page 45: Protein Structure Dynamics and Function at Micelle and

32

concentration threshold also corresponds to the point where the peak intensities begin to

drop off. These results suggest that, not only are excimers formed after 0.2 mM, but they

also self-quench.

Figure 4-6 - Fluorescein peak intensity vs. Concentration

The figure below indicates that the linear concentration/intensity relationship breaks

down above 0.075mM, hence any measurements above 300k cps should be reevaluated

as it may provide a misleading measurement of actual probe concentration.

Page 46: Protein Structure Dynamics and Function at Micelle and

33

Figure 4-7 - Divergence from intensity / concentration linarity starting at 0.1 mM of flourescein

In conclusion: These properties are important to know for using fluorescein probe.

Concentrations should be kept below the 0.075 mM range to avoid quenching effects and

peak position shifts.

Obtaining reliable enzyme kinetics parameters such as binding constants and max

reaction rates requires repeated acquisition of data points over a range of substrate

concentrations. In addition, we are studying the effect of surfactants on these kinetics

parameters, which adds an additional dimension to our work on enzyme kinetics.

Previously, a Novozymes product provided from Sigma Aldrich (P5860) was used. This

is a serine based subtilisin listed under the trade name Esperase. However, because it is a

commercial product the suspension medium and total enzyme concentration not

disclosed.

Page 47: Protein Structure Dynamics and Function at Micelle and

34

To eliminate possible interactions with unknown stabilizing agents, enzyme kinetic

experiments are repeated with lyophilized Subtilisin Carslberg protease (p5380) of

known wt% and activity. Since the product is received as a dry powder, they must be

freshly dissolved in buffer solution before each experiment. Steady state fluorometery is

used to monitor enzyme catalyzed proteolysis of fluorescein labeled casein insitu. While

bound to the casein substrate, the probe’s fluorescence is quenched due to the close

proximity of adjacently bound probe molecues. As the protease acts on the substrate,

probe is released into solution and can freely fluoresce. The resulting increase in emission

intensity can be repeatedly measured over the course of the reaction to measure the

progress of protein hydrolysis.

Initially, the reaction profiles of subtilisin catalyzed casesin hydrolysis are measured for

casein concentrations from 0.001 to 0.1 wt%

Figure 4-8 - Reaction profiles of subtilisin catalyzed proteolysis (0.001 to 0.1 mM substrate concentrations)

0

100000

200000

300000

400000

500000

600000

0 50 100 150 200 250 300 350

Inte

nsi

ty (

CP

S)

Time (s)

.001

.0025

.005

.01

.025

.05

.1.001

.0025

.005

.01

.025

.05

.1

Page 48: Protein Structure Dynamics and Function at Micelle and

35

The reaction profiles indicate more products being formed up until 0.025 wt%. Above

this concentration the amount of product rapidly drops off. A comparison of these final

intensities can be seen in the figure 4.11. This apparent loss of enzyme function might be

mistaken for an inhibition mechanism that would require additional study. However, our

previous experiments with the fluorescein probe indicate that we have far exceeded the

ideal concentration parameters (not to exceed 300k CPS intensity), and that this loss in

intensity is trivial, due to an over saturation of liberated fluorescein probe and not

indicative of enzyme reaction kinetics.

Figure 4-9 - End of reaction fluoresence intensity vs. substrate concentration (.0001 to 0.1 mM)

0

100000

200000

300000

400000

500000

600000

0 0.02 0.04 0.06 0.08 0.1 0.12

Inte

nsi

ty @

10

min

ute

s

Substrate wt %

End reaction intensity vs Substrate conc.

Page 49: Protein Structure Dynamics and Function at Micelle and

36

To correct this, the substrate concentration window was adjusted to less than 0.025 wt%

or approximately 320nM of casein substrate. This would ensure that the amount of

liberated fluorescein will not exceed a concentration where probe aggregates can form.

The amount of enzyme used is also reduced by 10 times, to ensure that substrate

saturation, or Vmax, can be achieved.

Figure 4-10 - Casein hydrolysis reaction profile (0.0005 to 0.025 mM substrate concentration)

0

50000

100000

150000

200000

250000

300000

350000

400000

0 100 200 300 400

Inte

nsi

ty (

CP

S)

Time (s)

.0005

0.001

0.0025

0.005

0.01

.025

A

Page 50: Protein Structure Dynamics and Function at Micelle and

37

Figure 4-11 - End of reaction fluoresence intensity (low substrate concentration 0.0005 to 0.025 mM)

Above are reaction rate profiles of the reduced substrate concentration range. To the left

is a plot of the final intensities after the 10 minute reaction.

In this optimized concentration window, we see a linear correlation between substrate

concentration and total amount of probe freed after the 10minute reaction for

concentrations below 0.01 wt%. The highest concentration deviates from this linearity;

however, since the intensity remains below 300k cps, this deviation can be attributed to

enzyme saturation phenomenon and not an aberration of probe interactions. This

kinetics experiment was repeated three times to establish an error range. Enzyme kinetics

0

50000

100000

150000

200000

250000

300000

350000

400000

0 0.005 0.01 0.015 0.02 0.025 0.03

Inte

nsi

ty (

CP

S)

Substrate wt %

B

Page 51: Protein Structure Dynamics and Function at Micelle and

38

parameters, Vmax and Km are needed to comparatively study enzyme efficiency. Kcat,

or the turn over number, is derived from Vmax, while Kcat/Km is a quantification of the

enzyme’s overall efficiency. These parameters are determined by plotting the initial

steady state portion of the reaction profile (change in intensity over the first 20 seconds)

as a function of substrate concentration. Below is such a plot derived from the reaction

profiles of subtilisin carlsburg illustrated by figure 4.13 on page 44.

Figure 4-12 - Set of initial rates plotted against substrate concentration drawn with error distribution.

Page 52: Protein Structure Dynamics and Function at Micelle and

39

Fitting the Michaelis-Menten equation to the data gives us the kinetics parameters Vmax

and Km, (Vmax = .00015 mM/sec and Km= 45 um.) for subtilisin Carlsberg in buffer

solution. Moving forward we repeated the same experiment in different surfactant

solution and inorganic particle suspensions of various concentrations. In the meantime

the reaction rates of subtilisin in surfactant have been measured at just one substrate

concentration in the presence of varying concentrations of Brij-30. The substrate

concentration tested here is relatively high, (160 um) so the reaction rates can be

indicative of changes in Vmax but a range of substrate concentrations need to be tested to

ascertain Km.

Page 53: Protein Structure Dynamics and Function at Micelle and

40

4.24.24.24.2 OOOObjective 2: bjective 2: bjective 2: bjective 2: Characterizing surfactant / particle properties in protein Characterizing surfactant / particle properties in protein Characterizing surfactant / particle properties in protein Characterizing surfactant / particle properties in protein

solutionssolutionssolutionssolutions

In addition to studying how surfactants affect protein conformation, conformation

dynamics and function, we are also investigating how additions of proteins ultimately

affect surfactant solutions properties. Because proteins are essentially bio-polymers, with

domains of varying hydrophobicity and polarity, they are expected to interact with

surfactants much in the way any polymer would. Some examples may include, shifting

CMC, affecting micelle shape and size or affecting foaming properties.

To achieve this particular objective this study will incorporate dynamic light scattering

(DLS) for particle size measurements and surface tension to measure surface activity of

surfactant/protein mixtures.

4.2.14.2.14.2.14.2.1 DLSDLSDLSDLS

Dynamic light scattering measurements were taken of surfactant solutions in 200mM

phosphate buffer with a Malvern Zetasizer Nano ZS90 and Brookhaven instrument

consisting of a 324 nm laser coupled with a BI-9000AT autocorrelator. DLS

measurements are often done to characterize the size and distribution of sub-micrometer

sized particles. The technique involves shining a monochromatic and coherent source of

light onto a sample suspension. The resulting Reyleigh scattering of the photons are

dependent on the size and potion of the particles off which the light is scattered. The

scattered photons then undergo either constructive or destructive interference. The

Page 54: Protein Structure Dynamics and Function at Micelle and

41

fluctuation of this pattern contains information of the Brownian motions of the particles.

Extrapolation of the particle size and distribution is commonly done with either the

Malvern or Brooklyn haven software utilizing the non-negative least squares (NNLS) or

CONTIN methods. Dynamic light scattering techniques have several limitations,

however; high particle concentrations will result in indirect scattering that cannot be

distinguished from the initial source, requiring DLS be done on relatively dilute

suspensions. Also our DLS instruments are they are can only assume particles of

spherical shapes.

Rayleigh scattering techniques are useful for particles smaller than the wavelength of

visible light, approximately 50 nm and smaller. The resulting scattered light is isotropic,

i.e. uniform in intensity regardless of the angle from the incident beam. This is feasible

for this system of micelles and proteins if there is no interaction between each other and

the colloids remain dispersed. However, if there is any further aggregation between the

surfactants and/or protein, then the final particle size may easily cause the scattering

behavior to transition out of the Rayleigh scattering size range into particles sizes that

may inelastically scatter radiation. In this case, Mie light scattering theory is considered.

Mie theory is a more general treatment of Rayleigh theory and covers larger particles

which absorb light and would maybe of more practical interests. For this application, the

detector position can be rotated about the sample to gather anisotropic scattering

intensities.

Page 55: Protein Structure Dynamics and Function at Micelle and

42

4.2.24.2.24.2.24.2.2 Surface tensionSurface tensionSurface tensionSurface tension

Surface tension is a phenomenon that occurs at the interphase between two continuous

bulk phases. The anisotropic forces that occur as a result of this transition is commonly

expressed in units of area (dynes / cm2) or as a linear form (mN/m). Surface tension

applies to all phases of matter, but most commonly studied at liquid-gas interface or the

interfaces of non-miscible liquids. Surface tension measurements are especially relevant

in studying detergent applications because of an amphiphillic nature of detergents. These

molecules will find their way to the water-air or water-oil interface of the solution they

are in, lowering the surface tension. This trend of decreasing surface tension vs.

increasing concentration continues until the interface is saturated with surfactant

molecules. At this point, the surface tension ceases to decrease, as the excess detergent

molecules can no longer fit at the interface. Instead, they begin to form aggregates. These

aggregates are referred to as “micelles”, and form in such a way to minimize exposure of

the hydrophobic tails with the aqueous solvent. The most commonly represented form of

such aggregates is in monolayer spheres, consisting of a hydrophobic tail core, and

hydrophilic headgroups in the solvent. This is not always the case, and structures such as

elongated rods, bilayer sheets, or hollow vesicles are often formed.

The Whilhelmy plate technique is used to measure the surface tension of surfactant

solutions as a function of concentration to determine the point at which surface tension no

longer decreases with concentration, indicating the critical micelle concentration (CMC).

This is particularly useful for this project in determining if the non-ionic surfactants are

affecting enzyme activity through free monomers or in their aggregated micelle form.

Page 56: Protein Structure Dynamics and Function at Micelle and

43

4.2.34.2.34.2.34.2.3 Electron paramagnetic resonance Electron paramagnetic resonance Electron paramagnetic resonance Electron paramagnetic resonance

Electron paramagnetic resonance (EPR), also known as Electron spin resonance (ESR) is

a probe technique that takes advantage of the Zeeman Effect wherein an unpaired

electron is oriented by a magnetic field. In measuring surfactant aggregation behavior,

ESR probes can provide information on the degree of molecular crowding about its

immediate environment.

Figure 4-13 - Schematic of Zeeman Effect energy splitting

The ESR absorption characteristic is heavily influenced by the unpaired electrons

environment and is employed as a probe to measure complex molecular structures and

aggregates.

The interactions of the probe influence the shape and intensities of an ESR spectral line

which can yield information about local viscosity adjacent to the probe and quantified in

Page 57: Protein Structure Dynamics and Function at Micelle and

44

the form of a rotational correlation time. Common classes of probes used in surfactant

and protein research are stabilized nitroxyl radical compounds functionalized to form

specific covalent bonds, or attached along the carbon chains of stearic acid. Common

examples of the latter are 16 and 5 Doxyl stearic acids (DSA), where a 4,4-dimethyl-3-

oxazolidinyloxy group is attached to either to 16th or 5th carbon of the hydrophobic chain.

A typical ESR spectrum of 16-DSA is shown in figure 3a

The probe mobility can be determined from this spectrum by comparing the ratios of

hyperfine splitting peaks. This can be calculated manually from the equations:

Figure 4-14 - Typical nitroxyl group ESR spectrum and peak parameters used to calculate rotational correlation

times

Page 58: Protein Structure Dynamics and Function at Micelle and

45

These functionalized surfactant probes are extensively used in many areas of colloid

research as in the study of self-assembled co-polymers [69] and SDS aggregates on

alumina surfaces [70]. Probe mobility is a common indicator of surfactant packing or

protein flexibility in these studies. This is quantified by comparing the hyperfine splitting

peaks of the nitrogen radical. Intensities of these peaks vary as the head group is hindered

by adjacent compounds, the degree of interaction can quantified as a rotation correlation

time by the relationship given in figure 3b and further defined by equation 3b. This

technique is also applied to measuring protein structure flexibility[35].

EPR spectrum of 0.1, 2.5, 5.0, and 20 mM surfactant (APG or DM) was measured in 200

mM sodium phosphate buffer as well as with the addition of 0.01 mg/ml HRP in also in

buffer solution. 16-DSA probe was introduced to the system at a final concentration of

1.0x10-5 M. For this investigation, a Bruker EMX-EPR spectrometer that operated at a

microwave field of 9.5 GHz (X-band). Dynamic light scattering measurements were

taken of surfactant solutions in 200mM phosphate buffer with a Malvern Zetasizer Nano

ZS90. Spectrum were analyzed with “Topspin” software installed onto Matlab. Initial

parameters: a, g, tcorr. Line fitting was done using Nelder/Mead simplex until at least .05

RMSD was achieved.

Page 59: Protein Structure Dynamics and Function at Micelle and

46

4.34.34.34.3 Objective 3Objective 3Objective 3Objective 3: : : : Measuring structural dynamics of model proteinsMeasuring structural dynamics of model proteinsMeasuring structural dynamics of model proteinsMeasuring structural dynamics of model proteins

An important aspect of this project focuses on measuring the “flexibility”, or rate and

range of internal motion, if it is affected by the selected surfactants, and whether these

changes in dynamics correlate with changes in enzyme kinetics explored in the first

objective. However, these must be measured in various surfactant solutions, so an

analytical method capable of measuring these dynamic structure properties in an aqueous

solution.

4.3.14.3.14.3.14.3.1 Life time Life time Life time Life time fluorescencefluorescencefluorescencefluorescence anisotropyanisotropyanisotropyanisotropy

Previous works have used electron spin resonance or time-resolved fluorescence to

measure subtilisin flexibility in the presence of different organic solvents [17, 43, 45, 67]

This study focuses on time resolved florescence decay, measuring protein flexibility

through probe anisotropy and intrinsic fluorescence decay.

Figure 4-15 - Dansyl fluoride structure

Anisotropic fluorescence decay is a probe mediated technique wherein a fluorophore

probe is attached to a polymer structure; in this case the enzyme. Irradiated by vertically

polarized light the directional (anisotropy) aspect of the fluorescent emission will decay

over time as the probe randomly spins, reorienting in space. This is typically done by

Page 60: Protein Structure Dynamics and Function at Micelle and

47

comparing the fluorescence decay of vertically polarized component of this emission

versus the horizontal[71].

Equation 4-1 anisotropy as a function of vertical and horizontally polarized fluorescence decay

Where VV(t) and VH(t) denote fluorescence decay kinetics of the sample measured using

vertically polarized light from the excitation source and emission component is vertically

(V) or horizontally (H) polarized. VV(t) and VH(t) are often referred to as parallel and

perpendicular polarized decays. The measure of anisotropy is the ratio between the

difference between parallel and perpendicular decays over the sum of all emission.

Figure 4-16 - example of parallel (VV) and perpendicular(VH) fluorescence decays with instrument response

function (IRF)

Page 61: Protein Structure Dynamics and Function at Micelle and

48

Figure 4-17 - anisotropy (r(t)) derived from VV(t) and VH(t). [62]

The anisotropic decay kinetics can be attributed to rotation diffusion of a spherical

particle and described as :

Equation 4-2 - anisotropy as a function of initial anisotropy, time and decay rate

Ro is the initial anisotropy value, t is elapsed time and τr is the decay correlation time. In

the case of labeled proteins, this decay function is primarily the sum of two rotation

mechanisms: A fast decay, originating from the spin of the probe about its covalent bond

to the protein structure. A much slower rotation occurs as the whole protein tumbles in

solution. The fast rotation is indicative of the flexibility about the probes immediate area

and therefore of most interest in this study.

Page 62: Protein Structure Dynamics and Function at Micelle and

49

This study emulates a procedure described by Broos and coworkers[72] where the

flexibility of subtilisin carlsburg is measured in different solvent environments via

fluorescence anisotropy decay. A major advantage of this procedure is in the dansyl

fluoride probe used; it specifically binds to the serine residue at the enzyme active site

allowing observation of changes to the active site’s flexibility. This specific interaction is

useful as active-site conformational changes can play a major role in enzyme activity.

Some additional factors that need to be considered include the treatment of anisotropy

decay time model [71, 73] and geometric considerations, such as the probe rotation in a

confined angle [74, 75].

From this portion of the work we explored the possibility that non-ionic surfactants also

affects enzyme structure dynamics, which opens the possibility of a new mechanism by

which enzymes are regulated by surrounding structures. Secondly, we investigated a

correlation between enzyme activity and flexibility as a function of colloid concentration.

4.3.24.3.24.3.24.3.2 Life time fluorescence decayLife time fluorescence decayLife time fluorescence decayLife time fluorescence decay

Fluorescence decay time of the single tryptophan residue in the HRP structure can also be

used as a qualitative measure of this protein’s overall flexibility. This decay is

characterized by the summation of three distinct decay parameters, dominated by a

picosecond-scale lifetime component, with small contributions from two other lifetime

components in the nanosecond range. The fastest decay time is the result of energy

transfer via Förster theory (FRET) from the excited Trp to the heme moiety located at the

Page 63: Protein Structure Dynamics and Function at Micelle and

50

enzyme active center[76, 77]. FRET transfer efficiencies are highly dependent on the

distances from acceptor and donor with transfers typically occurring between 10-70Å.

Figure 4-18 - Structure of horseradish peroxidase. Trp residue and heme group are highlighted.

� = 11 + ��|� 6

Equation 4-3 - relationship between energy transfer efficency and acceptor donor distance

Within this spatial range, energy transfer efficiency, E, is dictated by proportional to r,

the distance between the excited donor and acceptor, and by the ‘Forster distance’, Ro,

the point at which the transfer efficiency is 50%. Ro itself is defined by

Page 64: Protein Structure Dynamics and Function at Micelle and

51

Equation 4-4 Determination of Forster distance

Where Qo is the fluorescence quantum yield of the donor in the absence of an acceptor.

k2 is the dipole orientation factor, n is the refractive index of the medium and Na is

Avogadros’ number. The energy transfer efficiency can be related to fluorescence

lifetimes via the following relationship:

Equation 4-5 - Forster resonance energy transfer efficiency (E) as a function of fluorescence decay times

Where τ’D and τD are the donor decay life times with and without acceptors, respectively.

Many studies have taken advantage of the relationship between FRET efficiency and

decay times, by using time resolved fluorescence spectroscopy as a method of monitoring

protein unfolding. However, in this work, we are interested in changes in steady state

protein dynamics such as flexibility and fluctuations, on which life time fluorescence can

also provide information.

General thermodynamic expressions for energy and volume fluctuations about the mean

have been proposed over the past 50 years since revealing the dynamic nature of proteins,

Page 65: Protein Structure Dynamics and Function at Micelle and

52

such as the following models for the distribution functions of internal energy and

volumes of a protein reported by A. Cooper in 1976:

� ������� = ������

Equation 4-6 Discrete changes in protein's free energy

�������� = �����

Equation 4-7 Thermodynamically driven changes in protein volume

Where m and V are the mass and volume of the system, respectively, Cv is the heat

capacity at constant volume, �� the isothermal bulk compressibility, T the absolute

temperature and k is the Boltzmann’s constant. The change in protein radius resulting

from global protein structure fluctuations may be large enough to fall into the range that

effects FRET efficiency, and therefore decay time resulting from FRET mechanism, as is

the case of intrinsic fluorescence decay of HRP, which is predominately driven by the

energy transfer from Trp to the heme group of the enzyme active site. The range and rate

of atomic fluctuations will drive the distribution function of distances between Trp and

Heme at any given time. A more flexible structure with greater degrees of atomic

mobility will result in a spread in this distribution function. While the mean distance

between groups may not change from more a more “flexible” protein structure, the

Page 66: Protein Structure Dynamics and Function at Micelle and

53

overall time two residue will spend in closer proximity will increase, decreasing the

overall Trp fluorescence decay time with increasing protein local structure mobility.

While donor/acceptor distance, r, is a commonly studied parameter using time resolved

fluorescence. Ro is also an important, and at times controversial, parameter in measuring

FRET efficiency between donors and acceptors within proteins. A degree of uncertainty

arises when considering the dipole orientation factor, κ2. If the donor and acceptor are

both freely and isotropically rotating, then κ2 is averaged to 2/3. However, this is unlikely

to be the case in horseradish peroxidase, where both the donor (tryptophan) and acceptor

(heme group) are intrinsically bound within the protein structure. Molecular dynamic

simulation studies of intrinsic Trp in tetracycline repressor indicated the possible rotation

modes of the fluorescence donor and calculated a κ2 value of 0.36. A lower κ2 value of

hindered acceptors and donor would ultimately lower the transfer efficiency of the

Forster radiative energy transfer. Alternatively, should the flexibility of the protein

increase so will the mobility of the fluorophore and quencher, thus the transfer efficiency

will increase as a function of protein flexibility. Through measuring changes in the

intrinsic fluorescence decay rates of horseradish peroxides, we can determine changes in

the proteins relative structural flexibility.

Page 67: Protein Structure Dynamics and Function at Micelle and

54

4.44.44.44.4 OOOObjective 3bjective 3bjective 3bjective 3aaaa: Dynamics and function changes at inorganic : Dynamics and function changes at inorganic : Dynamics and function changes at inorganic : Dynamics and function changes at inorganic particle surfaces particle surfaces particle surfaces particle surfaces

A parallel objective is to measure protein flexibility and structure changes at TiO2 and

AlPO4 surfaces; TiO2 because it is a common electron scavenger employed in dye

sensitized solar cell (DSSC) and proposed bio sensitized solar cell designs; AlPO4

because it is prominently used as an adjuvant compound in vaccine formulations [78, 79].

A common problem in vaccine formulations featuring these particles is their aggregation

during long term storage; currently we hypothesize that gradual changes in protein and

polysaccharide structure and structure dynamics on particle surface create shifts in

electrostatics, hydrodynamics and steric interactions, these gradual shifts can be

influenced by the particle surfaces and lead to protein-protein attraction [80, 81].

Similarly, in an effort to utilize bacteriorhodopsin as a light sensitizer in greener

photovoltaic cells [82, 83], the protein needs to be attached to an electrode. In DSSC, this

is usually in the form of a sintered network of TiO2 nanoparticles. While it has been

established through circular dichromism that the secondary structures remain intact, the

internal dynamics that drive the proton pumping have not yet been studied at this particle

interface. For this work we will also employ time-resolved fluorescence and ESR to

measure flexibility changes in protein structures.

Page 68: Protein Structure Dynamics and Function at Micelle and

55

Figure 4-19 - Structure of CRM 197 carrier protein used in vaccine formulations

Page 69: Protein Structure Dynamics and Function at Micelle and

56

4.4.14.4.14.4.14.4.1 UVUVUVUV----Vis: Vis: Vis: Vis:

UV-Vis detection of proteins via tyrosine and tryptophan is a well-known method [84,

85]. For this project we will use it to quantify the concentration of proteins in solutions

and how much is adsorbed onto the AlPO4 particle surfaces. Using BSA as a model

protein, standard curve was prepared to correlate concentration with intensity. With that

curve we were able to determine the concentration of BSA adsorbed into AlPO4 by

measuring the concentration left in solution after mixing protein and particles and

separating the particles out. By subtracting from the known concentration of the original

solution, we determined that approximately 48 mg of BSA is absorbed onto ever 1 gram

of AlPO4

Figure 4-20 - Concentration standards of BSA measured by UV-Vis at 280nm

Page 70: Protein Structure Dynamics and Function at Micelle and

57

4.4.24.4.24.4.24.4.2 ZetaZetaZetaZeta potential measurementpotential measurementpotential measurementpotential measurement

A Zeta-Meter electrophoresis instrument and a Malvern Zetasizer have been used to

characterize the zetapotential of AlPO4 suspension as a function of pH and ionic strength

to determine the IEP of adjuvants alone. The change in surface charge of AlPO4 by

adsorbed protein have been studied as a function of concentration of protein, pH and

aging. The adsorption of protein and modification of surface charge can have an effect on

the agglomeration behavior of adjuvant-protein complex.

In electrophoresis measurement the electrodes are positioned at the opposite ends of a

glass capillary tube and electric current is applied to the dilute solution. The mobility of

particles under the influence of the electric field is measured manually using a

microscope. This technique is limited to approximately 1-micron size particles due to the

limitations of simple microscopy. A light scattering based Zetasizer instrument has been

used to measure the average zeta potential of micron to submicron AlPO4 particles. It can

yield an electrophoretic mobility distribution.

4.4.34.4.34.4.34.4.3 FTIR FTIR FTIR FTIR

A second method used to examine secondary structures in proteins (with and without

adjuvants) is infrared spectroscopy. FTIR analyzes the normal modes of vibrations of

covalent bonds, changes in dipole moment are necessary for IR absorption intensities to

be observed and most commonly involve stretching and bending modes of

nonsymmetrically bonded atoms. In the case of proteins, the amide I, II, and III bands

Page 71: Protein Structure Dynamics and Function at Micelle and

58

near 1700–1600 (C=O str), 1575–1475 (N-H str), and 1300–1220 cm−1 (C-N and N-H)

are the most commonly employed signals to monitor secondary structure in solutions. By

employing different sample geometry, Attenuated Total Reflectance (ATR), the changes

in protein structure in conjugation with adjuvant particles has been monitored as a

function of pH and degree of adsorption onto alumina phosphate particles.

Page 72: Protein Structure Dynamics and Function at Micelle and

59

5 Results and discussion

The results and discussions described in this section follow the four main objectives

outlined in previous sections. 1) Establishing a model system and methods for

quantifying enzyme activity in surfactant solutions. 2) Elucidating the nature of

interaction between surfactant, and surfactant aggregates with enzyme structures. 3)

Investigating changes in enzyme structure flexibility as a function of surfactant structure

and concentration. 4) Comparing how these interactions differ in situations where

proteins interact with a solid surface as opposed to surfactant micelle solutions.

5.15.15.15.1 Objective 1: Establishing a model systemObjective 1: Establishing a model systemObjective 1: Establishing a model systemObjective 1: Establishing a model system

In this first objective we present the data from enzyme “assays” and wherein the enzyme

catalyzed reaction is allowed to continue for a set number of minutes and the amount of

product formed during this time period is quantified and compared. This is described and

discussed in section 5.1.1. In contrast, the bulk of our enzyme activity measurements

referred to in this project involve real time measurement of product formation as the

reaction progresses, and the initial rates of these reaction profiles are what are compared

to each other. The term enzyme “activity” refers to this latter form of enzyme function

quantification.

Page 73: Protein Structure Dynamics and Function at Micelle and

60

5.1.15.1.15.1.15.1.1 5.1.1 Enzyme 5.1.1 Enzyme 5.1.1 Enzyme 5.1.1 Enzyme product yieldproduct yieldproduct yieldproduct yield assays. assays. assays. assays.

Activity is measured by the extent of polyphenolic residues such as tryptophan and

tyrosine liberated from the casein substrate through subtilisin catalyzed proteolysis in

buffer and various surfactants. The liberated tyrosine changes the color of Folin-

Ciocalteu reagent from yellow to blue. The color change is measured with UV-Vis and

presented below normalized versus the enzyme in buffer only.

When the color treated solutions were analyzied with UV-Vis, significant drops in

tyronise concentration were observed as protease solutions were exposed to 20mM

solutions of SDS. In the figure 5.1, the amount of tyrosine liberated by enzyme in SDS solution

is compared to that in pure water. It can been seen that the activity rate decreases significantly

faster in the presence of the anionic surfactant than in buffer alone.

Figure 5-1 - Subtilisin activity with and without SDS anionic surfactant as a function of exposure time

Page 74: Protein Structure Dynamics and Function at Micelle and

61

The effect of concentration was also studied. Enzyme was introduced to a 40 mM

solution of SDS, the activity tested before another 40 mM of SDS was added to the

solution. This was repeated until the enzyme was in a 120 mM solution of SDS.

Surprisingly, the enzyme activity *increased* with surfactant concentration. The error of

these experiments are determined by two additional repeats of the experiment.

Since it was additions of surfactant to a solution as opposed to three separate solutions, it

is implied that the deactivation seen in the 40 mM is actually *reversed* upon addition of

SDS. This will need to be confirmed and further studied.

Page 75: Protein Structure Dynamics and Function at Micelle and

62

20 40 60 80 100 1200

10

20

30

40

50

60

70

80

90

100

Tyro

sin

e L

ibera

ted

(u

M)

SDS Conc (mM)

Protease activity vs. SDS concentration

Figure 5-2 - Subtilisin activity with and without SDS anionic surfactant as a function of detergent concentration

This subtilisin assay experiment were carried out in solutions of linear alkyl benzene

(LAS), Brij-30 (Poly(oxyethylene)(4) lauryl ether), and APG (alkyl polyglucosides).

Page 76: Protein Structure Dynamics and Function at Micelle and

63

SDS LDS Brij-30 APG

-10

0

10

20

30

40

50

60

70

Activity r

ela

tive

to

su

btilis

in in

bu

ffe

r

Sufactant

Figure 5-3 - Enzyme activity assays of subtilisin in various surfactants (relative to enzyme in buffer)

Both the anionic surfactants, sodium dodecyl sulfonate (SDS) and linear alkylbenze

sulfonate (LAS) reduced the amount of product formed while the two non-ionic

surfactants Brij-30 and alkyl polyglucoside (APG) improved the yield. However, these

activity assays provide only a measure of product formed after a given time. Moving

forward, this dissertation will focus on activity quantifications that can provide rates of

product formation that is more applicable to extrapolating kinetics data.

5.1.25.1.25.1.25.1.2 Initial rate assaysInitial rate assaysInitial rate assaysInitial rate assays –––– Subtilisin proteaseSubtilisin proteaseSubtilisin proteaseSubtilisin protease

As previously described by section 4.1.3 in methods and materials section, the majority

of enzyme assays are carried out with casein substrate labeled with fluorescein

fluorphores. It is hydrolyzed by the subtilisin enzyme. The concentration of liberated

Page 77: Protein Structure Dynamics and Function at Micelle and

64

fluorophores is measured with a steady state fluorescence spectrometer at regular time

intervals over the course of the reaction. This procedure is repeated over a range of

substrate concentrations. The initial reaction rates are plotted as a function of substrate

concentration. Initial experiments are conducted with P4860, a commercial product of

subtilisin A from Novozymes in 50 mM potassium phosphate buffer solution. A <1 ml>

aliquot of <150ng/ml> enzyme solution is introduced to 2 ml of the substrate solution of

<0.005 wt%> FITC casein. The resulting changes in fluorescence are recorded by a

Horiba-Yvon Fluorolog 3, fluoro-spectrometer at 490nm excitation. Initial fluorescence

spectrum were measured between 500 and 580 nm to determine the correct emission

range as well as changing intensity with time. The experiment was then repeated and the

intensity at 520nm was measured every 30 seconds for the duration of the experiment,

1000 seconds in all. This experimental method will now be the standard measurement of

subtilisin activity for the remainder of this project, either as an absolute value of

▲intensity/second, Δ moles of fluorophore released/second or a normalized value

relative to the activity rate of the enzyme in buffer without surfactant.

Page 78: Protein Structure Dynamics and Function at Micelle and

65

Figure 5-4 - Fluorescence spectrum of casein substrate proteolysis as a function of time. Peak intensity is found at

519 nm.

Figure 5-5 - Fluorescence spectrum of casein substrate proteolysis as a function of time. Peak intensity is found at

519 nm.

0.00E+00

5.00E+05

1.00E+06

1.50E+06

2.00E+06

2.50E+06

3.00E+06

3.50E+06

4.00E+06

480 500 520 540 560 580 600

Inte

nsi

ty (

cou

nts

)

Emission wavelenth (nm)

Spectra of FTC-Casein w/ subtilisin protease

20 minutes

5 mintues

1 minute

1.30E+07

1.50E+07

1.70E+07

1.90E+07

2.10E+07

2.30E+07

2.50E+07

0 200 400 600 800 1000

Inte

nsi

ty (

cou

nts

)

Time (s)

Absorption @ 519 nm

Initial rate

measured

Page 79: Protein Structure Dynamics and Function at Micelle and

66

With an experimental method of measuring enzyme activity established, the next portion

of the first objective will be to confirm activity changes in the chosen surfactant systems.

Of our initial assay tests, APG and Brij-30 exhibited increased enzyme activity over

subtilisin in water alone. Thus, these surfactants were added to the system and the

changes in initial rate were measured. This initial experiment was done with P5459

subtilisin enzyme in 5mM of surfactant solution versus in water suspension. The reaction

profiles can be seen in figure 5.6, which demonstrates the difference between initial rates

of subtilisin as influenced by APG. To further investigate the dependence of subtilisin

mediated proteolysis rates on surfactant concentration, changes in initial activity are

measured over a concentration range of APG from 0.2 to 20mM. The reaction profiles

and corresponding initial rate measurements, normalized to the activity in buffer, is

plotted in figures fig 5.7 and 5.8 below. The differences of intensity at time zero can be

attributed to the delay between addition of enzyme to substrate and the start of

measurement. This interval is maintained to 10 seconds.

Page 80: Protein Structure Dynamics and Function at Micelle and

67

Figure 5-6 - Initial subtilisin rate profile experiments confirm faster initial proteolysis rates in alkyl polyglucoside

solutions.

Figure 5-7 - Subtilisin catalyzed proteolysis reaction profile in APG solutions

0

10000

20000

30000

40000

50000

60000

70000

80000

0 100 200 300 400 500 600 700

Inte

nsi

ty (

CP

S)

Time (seconds)

In APG

0.02

0.025

0.03

0.035

0.04

0.045

0.05

0.055

0.06

0 100 200 300 400 500 600 700

Pro

du

ct f

orm

ed

(m

M)

Time (seconds)

Subtilisin in APG

buffer

0.2mM

2.0 mM

5.0 mM

10 mM

20 mM

In Buffer only

Page 81: Protein Structure Dynamics and Function at Micelle and

68

Figure 5-8 - Normalized values of the initial rates derived from reaction profiles shown in Fig. 5.7

This experiment was also run in concentrations of Brij-30 at 0.007, 0.067, 0.67 and 6.7

mM of surfactant. At the lowest concentrations of this surfactant, there appears to be a

slight loss in relative activity at 0.95 relative units. The activity increase with surfactant

concentration, although not as much as APG, only reaching 1.24 relative activity units at

6.7mM. The reaction profiles and corresponding initial rate measurements, normalized to

the activity in buffer, are plotted in figures 5.9 and 5.10 below. These activity

experiments in APG were repeated to obtain experimental error which is reported in table

5.1

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

0.2 2 5 10 20

rela

tiv

e a

ctiv

ity

un

it

APG concentration (mM)

APG

Page 82: Protein Structure Dynamics and Function at Micelle and

69

Figure 5-9 - Subtilisin catalyzed proteolysis reaction profile in Brij-30 surfactant

Figure 5-10 - Initial reaction rates of subtilisin mediated proteolysis in Brij-30 non-ionic surfactants

0.00E+00

5.00E+06

1.00E+07

1.50E+07

2.00E+07

2.50E+07

0 100 200 300 400 500 600 700

Flu

ore

sce

in m

ark

er

inte

nsi

ty (

CP

S)

Reaction time (s)

Water

0.007 mM

0.067 mM

0.66 mM

6.7mM

0.951.03 1.07

1.24

0

0.2

0.4

0.6

0.8

1

1.2

1.4

0.007 0.067 0.67 6.7

No

rma

lize

d i

nit

ial

rea

ctio

n r

ate

Initial rate relative to enzyme in water

Page 83: Protein Structure Dynamics and Function at Micelle and

70

To compare the activity in a surfactant with a similar structure that is extensively used for

protein extraction and preservation This experiment was also run in concentrations of

Brij-35 at 0.026, 0.26, 2.62 and 26.0 mM of surfactant. At the lowest concentrations of

this surfactant, there appears to be a slight loss in relative activity at 0.95 relative units.

The activity increase with surfactant is even lower, only reaching 1.15 relative activity

units at 0.26mM, not significantly greater than the activity in buffer alone. At 2.6 and

26mM of Brij-35 surfactant the activity dramatically drops to 0.58 and 0.53 relative

activity units respectively. The reaction profiles and corresponding initial rate

measurements, normalized to the activity in buffer, is plotted in figures 5.8 and 5.9

below.

Figure 5-11 - Proteolysis reaction profile in Brij-35 surfactant

0

10000

20000

30000

40000

50000

60000

70000

0 100 200 300 400 500 600 700

Reaction time (seconds)

Digestion of 0.01 wt% casein via subtilisin in conc. of

Brij 35

Water

0.026 mM

0.26 mM

2.60 mM

26.0 mM

Page 84: Protein Structure Dynamics and Function at Micelle and

71

Figure 5-12 - Initial reaction rates of subtilisin catalyzed proteolysis in Brij-35 surfactant

To investigate the effects of free surfactant monomers on subtilisin function, the activities

were recorded in APG concentrations below CMC, at 0.015, 0.03, 0.06, 0.12, 0.24, 0.5

and 1.0 mM (figures 5.13). A rapid loss in activity was observed, reaching a minimum of

0.092 relative activity units at 0.1 mM of surfactant. This trend reverses upon increasing

surfactant concentration. It is important to note that the CMC of APG is measured to be

between 0.2 and 0.3 mM. This preliminary was the first indication that the activity is

negatively influenced by interactions with surfactant monomers, but also suggest that the

hyperactivity observed is linked to the formation of micelles. The reaction rate is

compromised at low Brij-30 (fig. 5.14) concentrations below CMC (0.3 mM) but can be

seen increasing at higher surfactant concentrations suggesting that enzymatic

hyperactivity is linked to micelle concentration and not surfactant monomer. However,

0

0.2

0.4

0.6

0.8

1

1.2

1.4

0 0.026 0.26 2.6 26

Init

ial

rate

re

lati

ve

to

en

zym

e i

n b

uff

er

Brij - 35 (mM)

Enzyme activity in Brij-35

Page 85: Protein Structure Dynamics and Function at Micelle and

72

the presence of micelles does not exclude the presence of surfactant monomers still in

solution and interactions between monomers and enzyme should not be ignored.

Figure 5-13 - Normalized reaction rates of Subtilisin in low concentrations. APG CMC is approximately 0.2 mM.

Figure 5-14 - Subtilisin activity in Brij-30 solutions

0

0.1

0.2

0.3

0.4

0.5

0.6

0.01 0.1 1

Act

ivit

y (

rela

tiv

e t

o b

uff

er)

APG concentration (mM)

0

0.00005

0.0001

0.00015

0.0002

0.00025

0.0003

0.00035

201020.020

Re

act

ion

ra

te (

mM

/ S

ec)

Brij-30 concentration (mM)

Page 86: Protein Structure Dynamics and Function at Micelle and

73

5.1.35.1.35.1.35.1.3 HRP activity in BrijHRP activity in BrijHRP activity in BrijHRP activity in Brij----30, APG and DM 30, APG and DM 30, APG and DM 30, APG and DM

To compare the effects of surfactant on another model protein of different size, catalytic

mechanism and substrate, horseradish peroxidase was selected and activities tested in the

presence of surfactants APG, dodecyl maltoside and Brij-30. Figures 5.12, and 5.13 show

the product formation profile and extrapolated initial reaction rates for HRP in 2.0mM

and 0.5mM of APG. Further trials were measured for APG concentrations 0.17, 0.33,

0.66, 3.3, 6.7 and 20mM and their values are listed in Table 5.1. The relative activity

units of those reactions are plotted in figure 5.14.

Figure 5.15. Reaction profile of horseradish peroxidase mediated phenol oxidation.

0

0.2

0.4

0.6

0.8

1

1.2

1.4

0 1 2 3 4 5 6

Pro

du

ct f

orm

ati

on

Time (minutes)

Buffer

2.0 mM APG

.5 mM APG

Page 87: Protein Structure Dynamics and Function at Micelle and

74

Figure 5-15 - HRP initial reaction rates in APG surfactant

APG

conc

trial 1 2 3 Avg StD

Measured Buffer 0.21 0.23854 0.244 0.23 0.02

04/16,

04/17

.17mM 0.23 0.23988 0.22 0.23 0.01

.33 mM - 0.21165 0.19 0.20 0.01

.67 mM 0.24 0.25051 0.24 0.24 0.01

APG

conc.

trial 1 Trial 2 Trial 3 Avg StD

Measured Buff 2 0.18 0.17 - 0.18 0.006

18-Apr 3.3mM 0.26 0.25 - 0.25 0.003

6.7 mM 0.19 0.20 - 0.19 0.008

20 mM 0.19 0.20 - 0.19 0.001

Table 5.1 - Results of multiple HRP activity tests.

0.2

0.205

0.21

0.215

0.22

0.225

0.23

0.235

0.24

0.245

Buffer 2.0 mM APG .5 mM APG

Init

ial

rea

ctio

n r

ate

Page 88: Protein Structure Dynamics and Function at Micelle and

75

Figure 5-16 - - Initial reaction rates of HRP in APG (columns) and surface tension measurements of APG (diamonds)

As seen with the subtilisin protease model enzyme, initial introduction of APG surfactant

decreased HRP enzyme activity up until the critical micelle concentration. Upon

formation of detergent aggregates, this shift in activity so closely linked to the

surfactant’s CMC strongly implicates surfactant aggregates, or micelles, as the primary

driver in the observed enzyme hyper-activity, whereas interactions with surfactant

monomers appear to suppress it. Surface tension measurements were also conducted to

investigate the effect of added enzyme has on micelle formation. Preliminary tests did not

indicate an appreciable shift in CMC, however additional experiments will have to be

carried out to evaluate the validity of those initial findings.

0.6

0.7

0.8

0.9

1

1.1

1.2

1.3

1.4

1.5

27

29

31

33

35

37

39

41

43

0.01 0.1 1 10 100

HR

P a

ctiv

ity

(n

orm

ali

zed

)

Su

rfa

ce t

en

sio

n n

N/m

APG concentration mM

.3 mM

Page 89: Protein Structure Dynamics and Function at Micelle and

76

0.1 1

0.85

0.90

0.95

1.00

No

rma

lize

d H

RP

in

itia

l re

actio

n r

ate

Dodecyl Maltoside Conc. (mM)

D

Figure 5-17 - Normalized initial rate values of HRP activity in dodecyl maltoside concentrations

Page 90: Protein Structure Dynamics and Function at Micelle and

77

0.1 1 10

0.6

0.8

1.0

1.2

1.4

1.6

Enyzm

e A

ctivity (N

orm

aliz

ed r/r

o)

APG conc. (mM)

Figure 5-18 - Normalized initial reaction rates of subtilisin (red) compared to those of HRP enzyme (black)

0.1 1 10

0.7

0.8

0.9

1.0

1.1

1.2

1.3

1.4

1.5

En

zym

e a

ctivity (

no

rma

lize

d r

ate

)

Surfactant Concentration (mM)

Figure 5-19 - Figure 5.17. Comparison of normalized initial rates of HRP enzyme in APG (black squares) vs DM (red

triangles)

Page 91: Protein Structure Dynamics and Function at Micelle and

78

5.25.25.25.2 Discussion of subtilisin and HRP assay and activity resultsDiscussion of subtilisin and HRP assay and activity resultsDiscussion of subtilisin and HRP assay and activity resultsDiscussion of subtilisin and HRP assay and activity results

The primary objective of this portion of work was to devise methods for measuring

enzyme activity and changes of enzymatic catalytic rates in the presence of the selected

surfactants: APG, DM, Brij-30 and Brij-35. Initially, activity of subtilisin was measured

by a colorimetric method of reduction of Foilin-Ciocalteu reacgent via liberated phenolic

residues. However, this method required that the reaction be manually halted at discrete

intervals and could not provide reasonably accurate reaction rates required for the fast

subtilisin mediated proteolysis. To overcome these limitations, a real time assay method

was adapted, which utilized the fluorescein-labled casein as a proteolysis substrate. This

allowed us to observe the reaction profile as it was occurring in real-time. Furthermore, it

was a much more streamlined process that was easier to measure reaction rates over

substrate concentrations to extract kinetics parameters. These kinetics experiments are

discussed in the next section. However, the parameters for using fluorescein needed to be

optimized. This process is described in the methods and materials section in the methods

and materials section

Using this assay method, we were able to reliably measure the initial proteolysis rates

mediated by subtilisin and perform a systematic study of enzymatic activity in the

presence of APG, Brij-30 and Brij-35 at various concentrations. As was demonstrated by

the initial assay tests, all three surfactants were shown to increase enzyme activity to

some degree, 48% in APG, 24% in Brij-30 and 15% in Brij-35. This already illustrates a

fundamental difference between how the various surfactant structures interact with either

the substrate, enzyme or combination of the two. What is also important to note is that the

Page 92: Protein Structure Dynamics and Function at Micelle and

79

overall increase in reaction rate occurs at specific surfactant concentrations. APG is the

most intense example, where the 48% increase in relative reaction rate occurs at 2mM

and quickly drops after this concentration range, returning to normal or sub-normal

activities at 5, 10 and 20 mM. Subtilisin activity in Brij-30 also appears to reach a

possible maximum of an additional 24% at 6.7 mM surfactant concentration. However

the Brij-30 cloudpoint, a temperature above which non-ionic surfactant micelles self-

aggregate, is below room temperature. As a result, higher concentrations of Brij-30 result

in increased turbidity, causing complications in the spectroscopic methods that we have

adapted. Alternatively, Brij-35, with its larger polyexthoxylate head group has a lower

cloud point. Like APG it appears to increase enzyme activity the most at a particular

concentration, only by 15% at 0.26 mM. However, further repeats of this experiment will

be needed to confirm if this value is significantly greater than the normal value. It is also

worthy to note that there is some activity loss at the very low concentration of 0.026 mM

of Brij-35 surfactant. While this also needs to be retested to determine if the difference is

significant, it prompted us to return to the APG system and measure enzyme activity

below the critical micelle concentration.

Since APG is a mixed system of surfactants, ranging in headgroup and tail group sizes,

the CMC can vary depending on the APG mixture. Using surface tension measurements

and pyrene probe fluorescence techniques to be discussed in the next section, it was

determined that micelle formation begins at 0.2 mM of the APG sample we were using.

This is confirmed in literature for APG mixtures ranging from 12-14 carbon units of

length. Measuring the activity below this concentration at 0.015, 0.03, 0.06, 0.13, 0.25,

Page 93: Protein Structure Dynamics and Function at Micelle and

80

0.5 and 1.0 mM indicates rapidly diminishing activity with surfactant concentration, up to

0.13 mM. This is near the CMC range and indeed, the trend in activity loss reverses at

0.25 mM, continuing to rise with each increase of surfactant concentration. This loss of

activity in pre-micelle concentrations, as well as the apparent dependence of subtilisin

activity on surfactant concentration strongly suggests that it is the micelles that play a

major role in the observed enzyme hyper activity phenomena. Moreover, free surfactant

monomers may be responsible for negatively affecting enzyme activity. The concept of

micelles being responsible for enzyme hyperactivity has been proposed by theoretical

models reported by Viparelli, et al. However, direct observation through experimentation

has been lacking, and, as previously mentioned, a precise physical reason explaining why

or how micelle surfaces should influence the enzyme’s catalytic properties is also

missing. Finally, the activity experiments presented in this thesis demonstrate an activity

increase maxima, most notably in the APG mixed surfactant system. Theoretical models

do not take into account possible complex interactions in surfactant aggregate structures,

such as to rod-like or liquid crystal formations. This will be explored in subsequent

sections of this dissertation.

Horseradish peroxidase (HRP) was chosen as an alternative model protein to be studied

in parallel with subtilisin. Aside from being well characterized, both structurally and

functionally, HRP differs from subtilisin in several key factors, such as having disulfide

bonds, metal complexation sites, a heme group, and most importantly, requires hydrogen

peroxide as a oxidizing reagent. For an HRP mediated oxidation, both substrate and

oxidant are water soluble molecules, unlike the large casein protein used in our subtilisin

Page 94: Protein Structure Dynamics and Function at Micelle and

81

activity tests, where the surfactant may have a significant impact on the reaction rate

through interaction with the substrate, by either improved dispersion and/or partially

denaturing the substrate before proteolysis. The differences of activity between these two

enzyme/substrate systems in various surfactants have provides valuable comparison

points for final elucidation of the surfactant / protein interactions.

A comparison of subtilisin and HRP activities in the presence of a range of APG

concentrations yield two very similar activity vs. concentration profiles. As was shown

previously, subtilisin activity suffers a loss at lower concentration ranges before the

CMC. Between 0.7mM and 1.5 mM of APG, this decrease is reversed and reaches a

maximum in activity centered around 2.0 mM. This activity drops back to less than

normal rates at 5 and 10 mM. The relative activities of HRP in APG solutions follow a

very similar pattern, with activity rates that decrease with additions of surfactant up until

the critical micelle concentration. Upon the formation of micelles, as was in the case of

subitilisin, the activity begins to increase reaching a maximum of 1.43 relative activity

units at 2mM and, also like the protease system, decreases after this point to 1.1 and 1.1

relative activity unites for 5.0 and 10.0 mM of APG respectively. The results can be seen

in figures 5.16 and 5.17.

The implications of these similar activity profiles again strongly indicate that micelle

structure plays an essential role in this mechanism of enzyme activity enhancement.

However, APG is a mixed surfactant system, known to undergo micelle structure

Page 95: Protein Structure Dynamics and Function at Micelle and

82

transitions and form worm-like structures at higher concentrations. Furthermore, the APG

product we started experiments with was sourced from an industrial product with an

unknown mixture composition, only an estimate provided by the manufacturer. For this

reason, dodecyl maltoside was also added to HRP reaction solution to observe induced

changes in the enzyme activity and how it compares to the APG mixed surfactant system.

Contrary to our initial assumption that the enzyme in dodecyl maltoside would react

similarly as in APG, the activities of HRP steadily decreased with additions of the

surfactant, from 0.015 to 5.0mM. This is a stark contrast from enzyme activity in APG,

despite having similar headgroup structures. In subsequent sections 5.3 and 5.4, this

thesis will explore the differences in micelle structure between the two sugar based

surfactants and how these differences correlate to these diverging activities. This next

section 5.2.1, will describe our investigation into how enzyme activity kinetics are

affected by the introduction of surfactants.

5.2.15.2.15.2.15.2.1 Sub sectionSub sectionSub sectionSub section:::: kinetics parameters Km and Vmaxkinetics parameters Km and Vmaxkinetics parameters Km and Vmaxkinetics parameters Km and Vmax

Vmax and Km are important kinetic parameters which reflect the binding efficiency and

forward rate of reaction in catalytic reactions. These two variables fall under the

Michaelis-Menten model of enzyme kinetics :

Equation 5-1 – Michaelis-Menten relationship between

Where υ is the rate of the reaction, P is the concentration of product, S is the

concentration of substrate, Vmax is the rate of reaction at substrate saturation and Km is

Page 96: Protein Structure Dynamics and Function at Micelle and

83

the substrate concentration at half Vmax. The overall scheme of an enzyme mediated

reaction is as follows:

K cat is the rate of forward reaction, also known as the turnover number, and is associated

with Vmax by:

Where [E]0 is the enzyme concentration. The Michaelis constant is the substrate

concentration at half Vmax and is a measure of the affinity of the enzyme for the substrate

molecule. The only controlled variable in this model is the substrate concentration. In the

subsequent experiments we used the fluorescence activity measurement techniques to

monitor changes in initial reaction rates of subtilisin on a casein substrate in substrate

concentrations ranging from 20, 40, 80 and 160 and sometimes 320 ng/ml. The data from

these experiments were then fitted with the Micealis-Menten equation in Origins 8.5

software, giving us the values of Km and Vmax for subtilisin acting on casein in buffer.

Page 97: Protein Structure Dynamics and Function at Micelle and

84

Figure 5-20 - Subtilisin mediated proteolysis reaction profiles as a function of substrate concentration

0 20 40 60 80 100 120 140 160 180

0.00004

0.00005

0.00006

0.00007

0.00008

0.00009

0.00010

0.00011

0.00012

Data: Data1_B

Model: Dhyperbl

Equation:

y = P1*x/(P2 + x) + P3*x/(P4 + x) + P5*x

Weighting:

y No weighting

Chi^2/DoF = 1.9857E-11

R^2 = 0.98737

P1 0.00015 ±0.00001

P2 45.4249 ±7.88149

P3 0 ±0

P4 0 ±0

P5 0 ±0

Initia

l re

action r

ate

(m

M/s

ec)

Casesin Substrate (ng/ml)

B

Figure 5-21 - Non-linear regression fitting of reaction rates of subtilisin in buffer vs. casein substrate concentration

done in Origins 8.0 software

0

0.005

0.01

0.015

0.02

0.025

0.03

0.035

0.04

0.045

0 100 200 300 400 500 600 700

Est

pro

du

ct f

orm

ati

on

(m

M)

Time (s)

20 ng/ml

40

80

160

Page 98: Protein Structure Dynamics and Function at Micelle and

85

To measure the effects of various surfactants on Vmax and Km, the same series of

experiments were conducted in solutions of APG (2, 10 and 20 mM), DM (10 and 20

mM), Brij-30 (2,5, 10, and 20mM) and Brij-35 (5, 10 and 20mM). The initial rates and

the Michaelis-Menten model fittings can be seen for the APG 10 and 20 experiments in

figures 5.20 and 5.21. In figures 5.22 a-d, the Vmax values are plotted as a function of

APG, DM, Brij-30 and Brij-35 surfactant concentration, respectively. The Km values are

plotted in Figures 5.23 a-d.

0 20 40 60 80 100 120 140 160 180

0.00002

0.00004

0.00006

0.00008

0.00010

0.00012

0.00014

0.00016

Data: Apg10mM_B

Model: Hyperbl

Equation:

y = P1*x/(P2 + x)

Weighting:

y No weighting

Chi^2/DoF = 1.0838E-10

R^2 = 0.96282

P1 0.0002 ±0.00003

P2 58.71577 ±20.21745

Substrate conc. nM

B

Figure 5-22 - Initial rates of subtilisin catalyzed proteolysis vs substrate concentration in 10mM of APG

Page 99: Protein Structure Dynamics and Function at Micelle and

86

Figure 5-23 - Initial rates of subtilisin catalyzed proteolysis vs substrate concentration in 20mM of APG

Page 100: Protein Structure Dynamics and Function at Micelle and

87

5.2.1.15.2.1.15.2.1.15.2.1.1 VVVVmaxmaxmaxmax

The changes in Vmax reflect the overall activity changes that have been demonstrated in

the previous sections. Activity tests of subtilisin in APG solutions results in higher Vmax

values, and therefore faster overall Kcat, for intermediate APG concentrations, 5.0 and

10.0 mM, and returns to nominal levels at higher concentrations. As seen in the activity

tests with HRP, dodecyl maltosides do not do much to increase the Vmax of subtilisins,

however, there is not strong evidence that it significantly hinders it either. Brij-30 also

increases subtilisin Vmax but only after an initial loss in before 5mM. Brij-35, as with the

activity tests, demonstrated little effects on the Vmax.

Figure 5-24 - Subtilisin Vmax as a function of APG concentration

1E-08

6E-08

1.1E-07

1.6E-07

2.1E-07

2.6E-07

0 5 10 15 20 25

Max r

eaction r

ate

(M

/sec)

Surf. Conc (mM)

APG

Page 101: Protein Structure Dynamics and Function at Micelle and

88

Figure 5-25 - Subtilisin Vmax as a function of DM concentration

Figure 5-26 - Subtilisin Vmax as a function of Brij-30 concentration

1E-08

6E-08

1.1E-07

1.6E-07

2.1E-07

2.6E-07

0 5 10 15 20 25

Ma

x R

ea

ctio

n r

ate

(M

/se

c)

Surf Conc. (mM)

Dodecyl Maltoside

1E-08

6E-08

1.1E-07

1.6E-07

2.1E-07

2.6E-07

0 5 10 15 20 25

Re

ac

tio

n R

ate

(M

/se

c)

Surf. Conc. (mM)

Brij-30

Page 102: Protein Structure Dynamics and Function at Micelle and

89

Figure 5-27 - Subtilisin Vmax as a function of Brij-35 concentration

1E-08

6E-08

1.1E-07

1.6E-07

2.1E-07

2.6E-07

0 5 10 15 20 25

Vm

ax R

ea

ctio

n r

ea

te (M

/ s

ec)

Surf. Conc (mM)

Brij-35

Page 103: Protein Structure Dynamics and Function at Micelle and

90

5.2.1.25.2.1.25.2.1.25.2.1.2 KmKmKmKm

For the most part, Km, which is an inverse measure of the binding efficiency between

substrate and enzyme, appears to be proportional with Vmax. That is, there is an inverse

relationship between the binding efficiency of subtilisin and Vmax in the presence of the

surfactants. In the case of dodecyl maltosides and Brij-35, which did not significantly

improve Vmax at any concentration, their Km values are increased significantly,

particularly at 5.0 and 10 mM concentrations. This is also true for APG, although this

decrease in binding efficiency is matched with an increase in Vmax. Brij-30 is the only

surfactant that appears to improve binding efficiency, lowering Km between 2.5 and 10

mM. This trend reverses and by 20mM Km is again higher than subtilisin in buffer alone.

Figure 5-28 - Subtilisin Km as a function of APG concentration

0.00E+00

2.00E-08

4.00E-08

6.00E-08

8.00E-08

1.00E-07

0 5 10 15 20 25

Km

(M

)

Surf. Conc

Alkyl Polyglucoside

Page 104: Protein Structure Dynamics and Function at Micelle and

91

Figure 5-29 - Subtilisin Km as a function of DM concentration

Figure 5-30 - Subtilisin Km as a function of Brij-35 concentration

0.0E+00

2.0E-08

4.0E-08

6.0E-08

8.0E-08

1.0E-07

0 5 10 15 20 25

Km

(M

)

Surf. Conc (mM)

Dodecyl Maltoside

0.0E+00

1.0E-08

2.0E-08

3.0E-08

4.0E-08

5.0E-08

6.0E-08

7.0E-08

8.0E-08

9.0E-08

1.0E-07

0 5 10 15 20 25

Km

(M

)

Surf Conc. (mM)

Brij-35

Page 105: Protein Structure Dynamics and Function at Micelle and

92

Figure 5-31 - Subtilisin Km as a function of Brij-30 concentration

0.0E+00

2.0E-08

4.0E-08

6.0E-08

8.0E-08

1.0E-07

0 5 10 15 20 25

Km

(M

)

Surf. Conc. (mM)

Brij-30

Page 106: Protein Structure Dynamics and Function at Micelle and

93

5.2.1.35.2.1.35.2.1.35.2.1.3 Discussion of kinetics in surfactants: Discussion of kinetics in surfactants: Discussion of kinetics in surfactants: Discussion of kinetics in surfactants:

The changes in Vmax in the four surfactants tested reflect what was demonstrated in the

activity tests done at a single substrate concentration (section 5.2): APG induces higher

maximum reaction rates at 2 and 5 mM and this increased rate drops back to normal

values at higher concentrations. Brij-30 increases the forward catalytic rate the most at

higher surfactant concentrations (20mM). These increases suggest that the surfactant

maybe interacting with the enzyme and/or substrate. Since Vmax is the product of forward

catalysis rate, Kcat, and enzyme concentration, one likely explanation is that the

introduction of surfactants helps disperse the enzyme, increasing the overall availability

per unit of enzymes introduced to the system. This hypothesis is explored in later sections

with particle sizing experiments. However, this is possibly an incorrect explaination, as

one might expect to see binding efficiency improve if more active site were exposed

through surfactant interactions. Alternately, the surfactant may be dispersing the

substrate, as opposed to the enzyme. As with dispersion of enzymes, dispersion of

substrates will increase the number of binding sites per unit of substrate added, which

would theoretically lower Km. Furthermore, at higher substrate concentrations, the

relationship between initial rate and substrate reaches a zero-order relationship and thus

Vmax should not be affected by the availability of more substrate attachment sites. Both of

these are not reflected the changes of kinetics in APG, which demonstrates greater Vmax

AND Km at 5 and 10mM. Both of these phenomena, especially the increase in Km suggest

that the surfactant is directly affecting the enzyme to increase Kcat while also lowering

binding efficiency. One physical explanation for these two kinetics parameters is that the

surfactant is in some way changing the flexibility of the subtilisin structure. Specifically,

Page 107: Protein Structure Dynamics and Function at Micelle and

94

a stiffer enzyme structure may account for further reduction of activation energy,

inducing a faster Kcat, while making binding with the substrate more difficult, thus

increasing the Km parameter.

Brij-30 is the only surfactant seen to lower Km. As mentioned before, this can be the

result of dispersion, either of the enzyme or substrate. Alternatively, the binding

efficiency is directly affected at those particular surfactant concentrations. This is another

distinct possibility, as Vmax is seen to decrease from a buffer only solution at those

surfactant concentrations. This maybe the opposite of the hypothetical “structure-

stiffening” scenario proposed earlier to explain greater Vmax and greater Km. A “softer”,

more “flexible” enzyme structure can conceivably have a greater substrate binding

affinity while slowing the forward rate of catalysis.

Finally, in these kinetics experiments, it is apparent that Vmax and Km are affected only at

particular concentrations of APG and Brij-30, while DM and Brij-35 with their similar

head group chemical compositions did not. This, combined with the drop in activities at

pre-micelle concentrations demonstrated in activity assays, strongly indicates that the

changes in enzyme activities are driven by micelle formations and presumably, the size,

shape and structure of these surfactant aggregates. The next section will describe our

efforts in elucidating the differences in the micelle properties of the four surfactants and

if and how they might correlate to the changes in activities that we have observed.

Page 108: Protein Structure Dynamics and Function at Micelle and

95

5.35.35.35.3 Objective 2: AggregatObjective 2: AggregatObjective 2: AggregatObjective 2: Aggregation /ion /ion /ion / micelle properties micelle properties micelle properties micelle properties

The second objective of this project is to elucidate the nature of surfactant – enzyme

interactions. Enzyme activity vs. surfactant concentration measurements discussed in

section 5.2 indicate a strong correlation between changes in enzyme activity and the

formation of APG micelles. The differences between APG and DM aggregates will be

measured by DLC, ESR, NMR and surface tension, and the results discussed in this

following section.

5.3.15.3.15.3.15.3.1 Surface tension: Surface tension: Surface tension: Surface tension:

The results of APG surface tension vs concentration can be seen in figure 5.32. The

break-point which indicates the concentration of micelle formation can be identified at

0.3 mM. However, surface tension continues to decrease up to 3mM of surfactant

concentration. This phenomena has been observed in mixed surfactant APG systems and

is indicative of micelle structure transitions which result from the varying solubility of

the heterogeneous surfactants[86].

Page 109: Protein Structure Dynamics and Function at Micelle and

96

In comparison, dodecyl maltoside do not exhibit this transition range. The surface tension

break point occurs at 0.1 mM and remains constant at all higher surfactant

concentrations. This is consistent with our previous assertion that the transition phase

seen in APG is the result of APG being a mixture. Furthermore, the surface activity of

DM is lower when compared to APG figure 5.25.

CMC I CMC II

Figure 5-32 - Surface tension vs APG concentration

0.1 1 10

30

40

Surf

ace T

ensio

n (

mN

/m)

APG Conc. (mM)

Page 110: Protein Structure Dynamics and Function at Micelle and

97

Figure 5-33 - Surface tension comparisons of dodecyl maltoside (single surfactant) to alkyl polyglucosides, a

surfactant mixture of similar composition.

5.3.25.3.25.3.25.3.2 Micelle sizeMicelle sizeMicelle sizeMicelle size::::

In order to better understand the interactions between the surfactants and enzyme

structure DLS was used to measure the differences between APG and DM micelles. 1.5

and 3.0 mM APG micelles appear to agglomerate into large particles, 1.4 and 1.7 in

diameter respectively. This concentration range is within the ‘transition’ phase between

0.3 and 3.0 mM as identified by surface tension. At 5 mM, these large aggregates

disperse and the micelle sizes rapidly reduce to 50-60nm in size. This dimension is still

large for a monolayer, spherical micelle which should only range between 5-10nm for

APG. The formation of rod-like micelles is a common for APGs and 60nm is a likely size

for the length of such conformations. The rod width was not picked up by this method,

however.

25

30

35

40

45

50

0.01 0.1 1 10 100

Su

rfa

ce t

en

sio

n (

mN

/m)

Surfactant Conc. (mM)

APG DM

Page 111: Protein Structure Dynamics and Function at Micelle and

98

Figure 5-34 - APG micellesize vs concentration

In comparison, the micelle size of DM at all concentrations above CMC were between 6

and 8 nm, which would be consistent with monolayer, spherical micelles.

Figure 5-35 - DM micelle size vs concentration

1443

1770

56 52 56

0

200

400

600

800

1000

1200

1400

1600

1800

2000

1.5 3 5 10 20

Pa

rtic

le d

iam

ete

r (n

m)

APG conc. (mM)

8.59

8.29

0

2

4

6

8

10

12

2 5 10 20

Mic

ell

e S

ize

(n

m)

DM conc. (mM)

Page 112: Protein Structure Dynamics and Function at Micelle and

99

5.3.35.3.35.3.35.3.3 ESR investigation of micelle structure: APG vs DM ESR investigation of micelle structure: APG vs DM ESR investigation of micelle structure: APG vs DM ESR investigation of micelle structure: APG vs DM

To further investigate the nature of these micelles and how the differences in these

structures ultimately influence the enzyme activity we performed electron spin resonance

experiments to measure the rigidity and packing density of the two surfactants. In figure

5.28 the correlation time parameter of the nixtroxile stearic acid probe, which is a

measure of the probe’s mobility, is plotted as a function of both APG and DM surfactant

concentration.

Figure 5-36 - Rotational correlation times (measure of local viscosity) as a function of surfactant (APG and DM)

concentration in buffer only.

0

5E-10

1E-09

1.5E-09

2E-09

0.01 0.1 1 10

Ro

t. C

orr

ela

tio

n t

ime

(s)

Surfactant Conc. (mM)

APG only

DM only

Page 113: Protein Structure Dynamics and Function at Micelle and

100

Figure 5-37 - Rotational correlation times (measure of local viscosity) as a function of surfactant (APG and DM)

concentration in buffer with 0.01 mg/ml HRP.

Figure 5.28 further highlights a significant difference in the evolution of APG and DM

micelle aggregates. In both surfactants, the transition from free monomers to aggregates

is apparent starting at 1mM. This is higher than the expected CMC values of APG and

DM (0.2 and 0.3 mM respectively) and this may be due to the significant concentration of

16-DSA probe at these lower concentrations.

Above 1.0 mM, the local viscosity around the probe in dodecyl maltoside solution

increases dramatically plateauing after a quick transition between 1.0 and 2.5 mM. In

contrast, the packing density of APG makes a gradual transition starting at 1.0 mM and

appears to begin to plateau by 5mM. Currently, it is hypothesized that Dodecyl maltoside

makes an immediate transition to spherical, monolayer micelles at CMC, while, APG

undergoes an extended transition phase before the bulk concentration of the mixed

0.00E+00

5.00E-10

1.00E-09

1.50E-09

2.00E-09

0.01 0.1 1 10

Ro

t. c

orr

ela

tio

n t

ime

(s)

Surfactant concentration (mM) + 0.01mg/ml HRP

APG +

HRP

Page 114: Protein Structure Dynamics and Function at Micelle and

101

surfactant is sufficient to form rod like micelles. The reasoning and implications of this

structure shift are further discussed in Section 5.6, result discussions.

5.3.45.3.45.3.45.3.4 NMR:NMR:NMR:NMR: (Cooperative work with C. Totland of University of Bergen)(Cooperative work with C. Totland of University of Bergen)(Cooperative work with C. Totland of University of Bergen)(Cooperative work with C. Totland of University of Bergen)

NMR was used as a tool to understand the precise mechanisms of interaction between

surfactant and micelles. Its major advantage over probe techniques such as electron spin

resonance, is that it can give detailed data on all molecular interactions in the system, not

only about the probe molecule. However, the resulting spectrum of complex systems,

such as in surfactant-enzyme solutions, can be very dense and difficult to analyze. With

the help of Dr. C. Totland[87, 88] from the university of Bergen, NMR was successfully

used to investigate APG and HRP interactions. Furthermore, NMR revealed unique head-

tail group interactions which implies a unique anti-parallel packing that has not yet been

observed in APG systems.

5.3.4.15.3.4.15.3.4.15.3.4.1 Line WidthsLine WidthsLine WidthsLine Widths

The line width data strongly suggest the formation of rod-shaped aggregates. Line widths

are inversely proportional to the T2 (spin-spin) relaxation time of those protons. In

contrast to T1 (spin-lattice) relaxation, which requires fluctuations near the Larmor

frequency (rotational speed of the nucleus) to occur, T2 is affected by slower motions as

well. Hence, while changes in T1 upon micellization reflect changes in the local

environment of that proton, changes in T2 (or line width) primarily reflect changes in

Page 115: Protein Structure Dynamics and Function at Micelle and

102

mobility of the aggregate as a whole. An increase in line-width therefore means that the

micelle rotation is slowing down due to increased size.

Furthermore, a recent study [89] showed using both gemini and conventional surfactants

that large increases in line width at concentrations above CMC only occur when a sphere-

to-rod transition takes place. Surfactants without the ability to form rods does not display

such line width increases. A comparison with more sophisticated techniques such as

NMR diffusometry showed that line width data qualitatively gives the same information.

Figure 5-38 - Comparison of the line widths of the main APG CH2 peak (2-12) for the 0.15 and 12 mM APG sample in

0.01 mg/ml HRP in 100mM potassium phosphate buffer

10080604020020406080100 Hz

0.15 mM

12 mM

Page 116: Protein Structure Dynamics and Function at Micelle and

103

Figure 5.30 shows the line widths of APG just below CMC (0.15 mM) and well above

the hypothesized sphere-to-rod transition and illustrates the differences in line width

between a mobile surfactant tail and those slowly spinning in a confined micelle. Figures

5.31 and 5.32 plot the T2 relaxation times derived from these peak widths as a function of

surfactant concentration.

Figure 5-39 - The line width here is the line width of the main CH2 peak (2-12) in the NMR spectrum. The line

widths are measured by using the deconvolution tool in the Topspin software (version 1.3).

10

15

20

25

30

35

40

45

50

0.01 0.1 1 10 100

APG + HRP

APG

[APG] mM

Lin

e w

idth

(H

z)

Page 117: Protein Structure Dynamics and Function at Micelle and

104

Figure 5-40 - Overlay of the aggregate mobility data presented in figure 5.31 with the activity of HRP in APG

(triangles)

Figure 5.31 shows that the sphere-to-rod transition of APG (white diamonds). Starting at

0.4 mM the average aggregate mobility begins to increase steadily until ~4.0 mM

wherein the rate of aggregate size increase vs. surfactant concentration decreases. This

can be attributed to the shift from one micelle transition to a final rod like conformation.

The presence of HRP may delay this transition with APG concentration (0.6 mM without

HRP, about 1 mM with HRP). This may be due to adsorption of APG monomers to HRP

prior to the transition, thus reducing the APG concentration in the bulk. Furthermore, the

overall line width is higher when HRP is present below the sphere-to-rod transition. This

may be caused both by an increase in viscosity and by interaction between APG and

HRP.

10

15

20

25

30

35

40

45

50

0.01 0.1 1 10 100

APG + HRP

APG

[APG] mM

Lin

e w

idth

(H

z)

Page 118: Protein Structure Dynamics and Function at Micelle and

105

5.3.4.25.3.4.25.3.4.25.3.4.2 NOESY (Nuclear Overhouser Effect Spectroscopy)NOESY (Nuclear Overhouser Effect Spectroscopy)NOESY (Nuclear Overhouser Effect Spectroscopy)NOESY (Nuclear Overhouser Effect Spectroscopy)

Figure 5-41 - NOESY spectra with water signal suppression using excitation sculpting with gradients, at different

concentrations of APG at constant HRP concentration. A possible interpretation in terms of APG aggregate

structure is presented to the left.

Head and tail closer in space

ppm

0.51.01.52.02.53.03.54.04.55.0 ppm

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

NOESY - 3 mM APG + HRP

NOESY - 1 mM APG + HRP

ppm

0.51.01.52.02.53.03.54.04.55.0 ppm

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

ppm

0.51.01.52.02.53.03.54.04.55.0 ppm

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

NOESY - 9 mM APG + HRP

Page 119: Protein Structure Dynamics and Function at Micelle and

106

As shown in Figure 5.41 the NOESY spectra show that the sphere-to-rod transition

results in a cross-peak in the 3 and 9 mM APG samples which suggests that the tail

methyl group is fairly close to the head group. This further supports the formation of a

lamellar bilayer structure. Put together with the rod formation indicated by the line width

data, the above combined results point towards the formation of tubes. These line

broadening effects observed in phospholipid bilayer systems by NOSEY NMR have also

been reported by Chen and Stark [90]. They have also reported a degree of interaction

between the CH2N of the head group and CH3 from the alkyl chain that is unexpected for

the traditional bilayer model. They have also proposed an intercalated, anti-parallel,

spacing of the lipids so that the tail end of one layer extends through the bilayer to

interact with the head group of the other surface, much in the same way as our proposed

APG tube like structure. In addition to this formation, Chen and Stark also suggest the

possibility of tail chains bending backwards to interact with its own head group.

5.3.55.3.55.3.55.3.5 IR investigation of BioIR investigation of BioIR investigation of BioIR investigation of Bio----Surfactant Surfactant Surfactant Surfactant degradationdegradationdegradationdegradation via enzyme hydrolysisvia enzyme hydrolysisvia enzyme hydrolysisvia enzyme hydrolysis: : : :

In a parallel study, the possibility of enzymes digesting bio-surfactants that have been

proposed for use in detergents was also investigated. IR spectroscopy of surfactant-

enzyme combinations were carried out to determine if there is any bond breaking or

formation when enzyme and surfactants are introduced to each other. Two different

surfactants were studied, n-Dodecyl β,D Maltoside and SDS. IR spectrum were taken of

Page 120: Protein Structure Dynamics and Function at Micelle and

107

both individually and then each with approximately 5.0 g/L concentration of cellulase

enzyme. Enzyme alone in water was also studied.

Figure 5-42 - Sodium Dodecyl Sulfate

Figure 5-43 - n-Dodecyl β,D Maltoside

Page 121: Protein Structure Dynamics and Function at Micelle and

108

Figure 5-44 - IR spectrum of SDS and Dodecyl Maltoside between 1350 and 850 cm-1 wavenumber range.

SDS peaks can be identified by S=O, and S-O stretching in the 1250-1212 and 972 – 955

Cm-1 range respectively. There are a series of C-O-C stretching modes that can attribute

to the series of peaks between 1200 and 950 cm-1.

Comparing the IR spectrum of dodecyl maltoside before and after being introduced to the

enzyme:

1370.0 1340 1320 1300 1280 1260 1240 1220 1200 1180 1160 1140 1120 1100 1080 1060 1040 1020 1000 980 960 940 920 900 870.0cm-1

A

SDSS=O Stretch

S-O Stretch

Dodecyl MaltosidePossible C-O-C stretching

Page 122: Protein Structure Dynamics and Function at Micelle and

109

Figure 5-45 - IR spectrum of SDS and Dodecyl Maltoside between 1160 and 960 cm-1 wavenumber range.

In this plot most of the major peaks associated with the DM head group are present when

in the presence of enzyme. Some notable differences worth further investigation are shifts

in peaks at 1038 and 1027, as well as the total disappearance of a peak at 1054. cm-1

Thorough investigation into the cellulase cleavage mechanisms will be necessary to give

any relevance to these peaks shifts. Initial searches indicate that the enzyme cleaves the

o-glycosidic bonds via a donation of a proton by a carboxyl group of glutamic acid to the

one to be cleaved. There is a stabilization of the carbonium ion intermediate by aspartic

acid. The reaction sequence is finished by an atack of nucleophilic water.

1161.7 1150 1140 1130 1120 1110 1100 1090 1080 1070 1060 1050 1040 1030 1020 1010 1000 990 980 970 959.1cm-1

A

DM Before Enzyme

DM W/ enzyme 1075.701075.70

1040.75

992.85

1024.18

1107.78

1148.30

1075.70

1038.041027.05

1055.21

1107.78

1148.30

995.26

Page 123: Protein Structure Dynamics and Function at Micelle and

110

From this understanding, it might be possible to speculate that the peak at 1054 that

disappears may belong to the intermediate o-glycosidic bond shown in figure 5.39 and

disappears after introduction to the enzyme and it is cleaved. This is speculative at this

point and will require more study to confirm.

Figure 5-46 - Schematic of cellulase hydrolysis mechanism

HATR Spectroscopy of Dodecyl Maltoside and Enzyme mixtures further suggest

enzymatic activity and cleavage of the head group. Comparing the IR spectrum of

dodecyl maltoside before and after being introduced to the enzyme:

Page 124: Protein Structure Dynamics and Function at Micelle and

111

1200 1100 1000 900

0.000

0.005

0.010

0.015

0.020

0.025

-0.010

-0.005

0.000

0.005

0.010

0.015

0.020

0.025

DM

DM

cm-1

DM + Enzyme

DM

+ E

nzym

e

Dissapearing Peak After Enzyme Exposure

Figure 5-47 - FTIR of Dodecyl maltoside with a-Cellulase indicating possible loss of C-O-C bond

This disappearing peak may be attributed to the C-O-C stretching connecting the sugar

groups in the DM head structure.

Figure 5-48 - Proposed reaction of surfactant catalysis

Page 125: Protein Structure Dynamics and Function at Micelle and

112

Evidence of enzymatic activity can also be seen in the IR spectrum of the cellulase

enzyme before and after introduction to dodecyl maltoside surfactant. C-O Stretching

peaks are indicative of the carboxyl group responsible for glycosidic bond catalysis. In

the plot below (Figure 5.37), we can see a shift in this peak after introduction of DM

surfactant strongly suggesting there is a deprotonation of the carboxyl group, which is

step in the catalysis process.

1800 1750 1700 1650 1600 1550 1500

-0.004

-0.003

-0.002

-0.001

0.000

0.001

0.000

Care

nzym

e C

ellu

lase

Wavenumber (cm-1)

Shift after exposure

to DM surfactant

Figure 5-49 - FTIR from 1800 to 1500 cm-1 possibly showing the carboxyl group located at the cellulase active site.

Page 126: Protein Structure Dynamics and Function at Micelle and

113

Evidence of enzymatic activity can also be seen in the IR spectrum of the cellulase

enzyme before and after introduction to dodecyl maltoside surfactant. C-O Stretching

peaks are indicative of the carboxyl group responsible for glycosidic bond catalysis. In

the above plot, we can see a shift in this peak after introduction of DM surfactant strongly

suggesting there is a deprotonation of the carboxyl group, which is step in the catalysis

process.

Page 127: Protein Structure Dynamics and Function at Micelle and

114

5.45.45.45.4 Objective 3: Objective 3: Objective 3: Objective 3: Protein structure dynamics:Protein structure dynamics:Protein structure dynamics:Protein structure dynamics:

The third objective of this study is to investigate the “flexibility” of an enzyme, as a

physical characteristic that is being affected by interactions with surfactants. This next

section describes two time resolved fluorescence techniques used to study the structural

mobility of subtilisin and horseradish peroxidase enzymes used for this work.

5.4.1 Life-time fluorescence: Anisotropy

The anisotropic decay of dansyl probe attached to subtilisin active site has been measured

in buffer solutions and compared to a solution of dansyl probe in ethanol. Fitting a decay

profile to the anisotropy curve is done through Datamax software. The number of decay

correlation times between 1 and 3 were tested and the parameters with the chi-squared

values closest to 1 are displayed in the chart below.

Column1 Probe protein Protein plus 8mM brij 30

chi sq 1.005 1.020 1.090

θ 1(s) 4.90E-10 6.80E-10 5.19E-10

θ 2(s) 3.60E-08 2.30E-06

θ 3(s) 6.99E-06

Table 5.2- Significant anisotropy decays from labled subsitlin protease

Page 128: Protein Structure Dynamics and Function at Micelle and

115

Table 5.2 Shows the decay correlation times for the models which best fit the anisotropy

loss of dansyl labeled subtilisin. The first column denotes decay times for probe

suspended in ethanol indicating a singular decay rate. Center: probe labeled subtilisin

yield two decay rates. Presumably one for the probe spinning about a covalent bond and

the second a result of the whole protein rotation. Right: Labeled protein in 8mM

surfactant solution. Here, the fast motion of the probe is less hindered relative to the

probe without surfactant.

A singular decay time is the best fit for the probe in solution, shown above as θ1. This is

expected as the unbound probe should spin randomly at a uniform rate. Moreover, the

decay time is rapid, about half a nanosecond; this indicates that the probe in solution is

relatively unhindered by its local environment. Alternatively, a function that is the sum of

two decay times’ best fit the dansyl labeled enzyme. A decay function with multiple

decay correlation times is typically an indication of non-spherical particles. However,

considering that subtilisin is a globular protein, and that the differences between the two

decay times are vastly different (two orders of magnitiude), it is more likely that the

faster correlation time corresponds to the probe rotating about the covalent bond while

the slow rotation can be attributed to the whole protein tumbling.

It is important to note the decay rate increase in θ1 from the free probe, indicating a

slower rotation of the probe. This is consistent with the probe being attached to the

enzyme active site, hindered by adjacent peptide residues. Repeating this experiment at a

relatively high concentration of Brij-30 surfactant resulted in a reduction in rotational

Page 129: Protein Structure Dynamics and Function at Micelle and

116

correlation time, from .68ns to .52ns, indicating that the probe is more freely spinning.

This is change is regarded as the active site becoming more flexible, possible influenced

by the non-ionic surfactant. It might be argued that the surfactant is unfolding the protein

structure, hence the increased probe mobility. But, enzyme activity if found to be

improved in this surfactant range, strongly suggesting that the structure is preserved.

Further experiments include expanding the selection and concentration range of

surfactants as discussed in objective 1. It should also be noted that, in the presence of the

Brij-30 surfactant, a three exponential decay was the best fitted solution with a chi-sq

value of 1.09. The global rotation of protein is slowed by two orders of magnitude, to the

order of microseconds, presumably from interaction with surfactant micelles.

This initial experiment demonstrates subtilisin interaction with the surfactant resulting in

both slower global rotation of the protein and faster probe mobility about the active site.

The next set of these experiments were to monitor changes in the active site mobility as a

function of surfactant concentration.

Page 130: Protein Structure Dynamics and Function at Micelle and

117

Figure 5-50- Rotational correlation time (blue diamonds) and subtilisin activity (orange squares) as a function of

Brij-30 concentration.

Figure 5.50 plots the fast rotational correlation time as a function of surfactant

concentration (blue) for reference the initial activities of subtilisin are also plotted here

(orange). The initial correlation time as previously measured was .68 ns and .52 at the

highest Brij-30 concentration measured. The correlation times measured at the

concentrations in between 0.0 and 8.0 mM indicate an exponential increase in probe

mobility with increasing surfactant concentration. In parallel, the reaction rates of

subtilisin increase at a similar rate with additional surfactant. This correlation strongly

indicates a connection between active site flexibility and enzyme activity as a function of

surfactant.

0.9

0.95

1

1.05

1.1

1.15

1.2

1.25

1.3

0.5

0.54

0.58

0.62

0.66

0.7

0.001 0.01 0.1 1 10 100

Reacti

on

Rate

(r/

r 0)

Co

rrela

tio

n t

ime (

ns)

Brij - 30 Conc. (mM)

Page 131: Protein Structure Dynamics and Function at Micelle and

118

The same preparation and measurements were carried out for APG, the sugar based

surfactant also shown to increase enzyme activity. Like Brij-30, additions of APG

reduced the local viscosity of the protease active site with the addition of surfactant.

Figure 5-51 - Decay correlation times of dansyl probe attached to subtilisin as a function of APG detergent.

Unlike Brij-30, APG activity increases in a specific concentration at 1.0 – 2.0 mM and

drops back to normal activity values above 3 mM (see figure 5.16). We attribute this drop

in activity to the formation of rod-like micelles that increase the solution viscosity,

slowing down the mass transfer kinetics, limiting enzyme – substrate interactions.

5.70E-10

5.80E-10

5.90E-10

6.00E-10

6.10E-10

6.20E-10

6.30E-10

0.1 1 10

Anis

otr

opy D

ecay C

orr

ela

tion (

s)

APG Conc (mM)

Page 132: Protein Structure Dynamics and Function at Micelle and

119

5.4.25.4.25.4.25.4.2 Time resolved fluorescence to measure flexibility of horseradish peroxidase: Time resolved fluorescence to measure flexibility of horseradish peroxidase: Time resolved fluorescence to measure flexibility of horseradish peroxidase: Time resolved fluorescence to measure flexibility of horseradish peroxidase:

Horse radish peroxidase is another model enzyme that was used in this study. It has been

shown in the enzyme activity portion of this thesis that HRP is also positively influenced

by APG and Brij-30 surfactant, (figure 5.16). However, there is no serine residue located

at the active site for convenient dansyl probe attachment, limiting the application of

fluorescence anisotropy decay analysis. However, as discussed in the methods section,

the intrinsic decay of HRP’s naturally occurring fluorescence can be measured to

compare overall mobility of the structure. Figure 5.40a is the raw data of this decay of

HRP in buffer. Also in this graph is the fitted model of three exponential decays used to

extrapolate decay correlation times τ.

Figure 5-52 -Top. Raw data (Blue) and fitting of sum of 3 exponential decay functions (orange) of hrp in buffer.

Bottom: residual of fitted line vs raw decay data.

0.1

1

10

100

1000

10000

0 1000 2000 3000 4000 5000

Co

un

ts

channels

Decay

Fit

-10

0

10

0 1000 2000 3000 4000 5000

Residuals

Page 133: Protein Structure Dynamics and Function at Micelle and

120

Exponential decay of fluorescence intensity as a function of time, I(t), can be expressed

as a sum of exponential decays

Equation 5-2

Where k is the number of distinct decay rates, t is the time in seconds, τ is the decay

parameter in seconds, and ak is the contribution factor for that particular decay function.

From the data above the decay parameters and percent contributions are listed below.

T1 (s) a1 T2 (s) a2 T3 (s) a3

2.60E-09 12% 9.10E-09 8% 5.50E-11 80%

Table 5.3 - Three parameters of intrinsic HRP fluorescence decay. The majority of the decay is attributed to the τ3

parameter.

The majority of the decay is attributed to the τ3 parameter occurring at 5.5e-11 seconds,

which is two orders of magnitude faster than the second fastest, happening at 2.6

nanoseconds. This fast decay is attributed to resonant energy transfer of from the excited

tryptophan to the heme group located at the enzyme’s active site.

Page 134: Protein Structure Dynamics and Function at Micelle and

121

Using this technique we will measure changes in the decay parameter of HRP in dodecyl

maltoside and alkyl polyglucosides. Referring back to <section> it was shown that APG

will induce greater enzyme activity while DM, with a very similar head group structure,

does not.

The fast decay parameter was measured and plotted over the range of these surfactant

concentrations to further investigate the correlation between activity and flexibility in

HRP. The results are plotted below in figures 5.47 and 5.48. To reiterate, I hypothesize

that a faster decay rate can be attributed to a more flexible enzyme structure. The

opposite is also postulated to be true: that a slower decay rate is indicative of a more rigid

structure. At the lowest concentration measured (0.05) mM the decay correlation times of

HRP in both DM and APG have decreased (i.e. faster decay) to ~4.2e-11s. The decay

data and the multiple exponential fitting, as well as the residual data can be seen below

for APG concentrations: 0.05, 2.0 and 10.0 mM.

Page 135: Protein Structure Dynamics and Function at Micelle and

122

Figure 5-53 - Decay data and multiple exponential fitting (including residuals below) of HRP enzyme in 0.05 mM

APG.

-2000

0

2000

4000

6000

8000

10000

12000

2E-08 2.5E-08 3E-08 3.5E-08 4E-08

Fit Decay

-6

-4

-2

0

2

4

2E-08 2.5E-08 3E-08 3.5E-08 4E-08

Page 136: Protein Structure Dynamics and Function at Micelle and

123

Figure 5-54 - Decay data and multiple exponential fitting (including residuals below) of HRP enzyme in 2.0 mM APG.

-2.00E+03

0.00E+00

2.00E+03

4.00E+03

6.00E+03

8.00E+03

1.00E+04

1.20E+04

2E-08 2.5E-08 3E-08 3.5E-08 4E-08

Fit Decay

-8

-6

-4

-2

0

2

4

2E-08 2.5E-08 3E-08 3.5E-08 4E-08

Page 137: Protein Structure Dynamics and Function at Micelle and

124

Figure 5-55 - Decay data and multiple exponential fitting (including residuals below) of HRP

enzyme in 10.0 mM APG.

-2.00E+03

0.00E+00

2.00E+03

4.00E+03

6.00E+03

8.00E+03

1.00E+04

1.20E+04

2E-08 2.5E-08 3E-08 3.5E-08 4E-08

Fit Decay

-10

-5

0

5

2E-08 2.5E-08 3E-08 3.5E-08 4E-08

Page 138: Protein Structure Dynamics and Function at Micelle and

125

0.1 1

0.6

0.8

1.0

1.2

1.4

Avg (

AP

G)

Concentration (mM)

Figure 5-56 - Fluorescence decay times as a function of surfactant (APG) concentration

Page 139: Protein Structure Dynamics and Function at Micelle and

126

0.1 1

0.6

0.8

1.0

1.2

1.4avg (

DM

)

Concentration (mM)

Figure 5-57 - Fluorescence decay times as a function of surfactant (dodecyl maltoside) concentration

Page 140: Protein Structure Dynamics and Function at Micelle and

127

5.55.55.55.5 Objective 3aObjective 3aObjective 3aObjective 3a: : : : Protein flexibility and orientation at particle surfacesProtein flexibility and orientation at particle surfacesProtein flexibility and orientation at particle surfacesProtein flexibility and orientation at particle surfaces: : : :

Model protein CRM 197 was introduced to AlPO4 in pH controlled buffer solution. UV-

Vis was used to quantify the amount of protein adsorbed onto the particle surfaces.

Figure 5-58 - Amount of CRM-197 adsorbed onto AlPO4 particles at pH 4 and 5 as a function of initial protein

concentration.

These results indicate that pH 5 is the optimal condition for binding of the two. CRM

adsorbs onto the particles at a linear rate up to 150 ppm of initial protein concentration.

Knowing the optimal conditions for adsorption, time resolved fluorescence and FTIR

techniques were employed to probe the orientation of the adsorbed protein, and perhaps

change in protein structure dynamics on the particle surface.

0

0.05

0.1

0.15

0.2

0.25

0

20

40

60

80

100

120

140

0 50 100 150 200 250 300 350

Ad

sorb

ed

CR

M (

g/g

AlP

O)

Ad

sorb

ed

CR

M (

pp

m)

Initial CRM 197 (ppm)

pH 5

Series2pH 4

Page 141: Protein Structure Dynamics and Function at Micelle and

128

Intrinsic time resolved fluorescence is a suitable technique to measure conformation

changes of CRM-197 or the protein’s interactions with particle surfaces because of the

few Trp residues located throughout the structure. Trp 206 and 281 are located within an

alpha-helix dense region, Trp 398 amongst a series of beta sheet conformations, while

Trp 150 and 53 are attached to a random coil at different region of the protein. Changes

in the fluorescence of these residues can give information such as the polarity of its

immediate environment. Lifetime-fluorescence will provide the ability to differentiate

between the decays of the 5 Trp throughout the structure, allowing us to better define

which regions are interacting with solvent, or particle surfaces.

Figure 5-59 - CRM-197 structure with Trp residues highlighted

Page 142: Protein Structure Dynamics and Function at Micelle and

129

Time resolved fluoresence data for CRM in buffer at pH 7.0 and 5.8 as well as CRM on

AlPO4 particles at pH 5.8.

Figure 5-60 - Fluorescence decay times resolved from CRM in pH 7 and 5.8 buffer and adsorbed onto AlPO4 at pH

5.8

1.6E-091.9E-09

6.1E-095.8E-09

4.5E-09

6.1E-11 9.9E-110

1E-09

2E-09

3E-09

4E-09

5E-09

6E-09

7E-09

CRM pH 7 CRM pH 5.8 CRM pH 5.8 ALPO

De

cay

lif

e t

ime

(s)

t1

t2

t3

Page 143: Protein Structure Dynamics and Function at Micelle and

130

Figure 5-61 - The decay contribution of each decay time.

Figure 5.44a illustrates the time decays that are resolved from the time resolved

fluoresence data of CRM-197 at pH 7, 5.8 and on a particle surface. In buffer, at a neutral

pH, there are two distinct decay times, one at 1.6 ns and the other 6.1 ns contributing

30% and 70% of the decay respectively. Both of these decay rates are similar to each

other within the nanosecond range, which is typical for free tryptophan in solution. The

slower decay time can be attributed to tryptophans buried further in the protein’s

structure, away from the polar solvent.

Adjusting the pH to 5.8 did not drastically affect these decay times. However a new,

much faster decay time (6.1e-11s), arises. This new decay pathway contributes

30.0% 30.0%

70.0%

55.0%

20.0%15.0%

80.0%

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

CRM pH 7 CRM pH 5.8 CRM pH 5.8 ALPO

De

cay

Pa

ram

ete

r C

on

trib

uti

on

(%

)

t1

t2

t3

Page 144: Protein Structure Dynamics and Function at Micelle and

131

significantly to the over decay profile (15%), taking away from the slower of the two

original decay contributors (T2) which is now reduced from 70% to 55% from lowering

the pH.

The appearance of such a short decay time can be attributed to a tryptophan species

becoming more mobile as the enzyme structure as a whole gains greater flexibility.

Page 145: Protein Structure Dynamics and Function at Micelle and

132

5.65.65.65.6 Objective 3b:Objective 3b:Objective 3b:Objective 3b: Modeling enzyme activity as a function of surfactant Modeling enzyme activity as a function of surfactant Modeling enzyme activity as a function of surfactant Modeling enzyme activity as a function of surfactant

interaction and concentration.interaction and concentration.interaction and concentration.interaction and concentration.

Viparelli and coworkers created mathematical models to describe changes in enzyme

activity as a function of surfactant concentration. These models were heavily based on the

kinetics of protein/micelle interactions as well as the partition of enzymes between bound

water / free water / and micelle core. Ultimately, changes in enzyme activity would be a

function of the degree of interaction between the various phases of water found about a

micelle or bulk solution. In this work, the volume of bound water about the micelle is

given as:

�� = ����� ��� !"#$

Equation 5-3 - Determination of bound water volume in a surfactant system

Where Dn is the concentration of surfactant, ψ is the surfactant number per micelle, ϴ,

the number of bound water molecules, and �� !"#$ the molecular volume of the bound

water. Furthermore the partition between enzyme at a micelle bound water layer and

those in bulk water can be expressed as:

%�&' + �()� *+⇔ ����

Equation 5-4 - Equilibrium reaction between micelles and free enzyme.

Where [Ef] is the concentration of free micelles, [Mn] is the concentration of micelles,

[Eb] is the concentration of bound enzyme, and KE is the reaction constant of binding.

Viparelli derived equations for the concentrations of free and bound enzymes as follows:

Page 146: Protein Structure Dynamics and Function at Micelle and

133

%�&' = ��"�1 + -.�()�

Equation 5-5 - Concentration of free enzyme for a given, Micelle concentration

���� = �-.�()���"�1 + -.�()�

Equation 5-6 - Concentration of enzyme in bound water for a given concentration of micelles

What is missing from this theoretical work is the mechanism that would explain why

enzyme activity should change in these different water environments is not explored, nor

is there an explanation as to the driving force that would compel the enzyme to seek out

the micelle core, surface layer or solution bulk.

This dissertation investigated and identified those mechanisms, allowing for the

expression of a more comprehensive model of interaction that would take into account

the physical factors that drive enzyme partition throughout the water layers as well as

identifying changes in flexibility as the mechanism that drives the actual change in

enzyme activity. Furthermore, we now know that micelles that induce more flexible

protein structures are those that have lower cloud point temperatures, that is, their bound

water layers are less stable than those micelles that do not increase enzyme flexibility and

activity. These findings further strengthen past assertions that the bound water layers

about the micelles are responsible for changes in enzyme activity in surfactant/enzyme

systems. This work has demonstrated that changes in protein structural flexibility is a key

factor that links bound water partitions to changes in enzyme activity. However, this

leads to an additional question not yet explored by this work: Why does the hydration

sphere about a micelle affect the flexibility?

Page 147: Protein Structure Dynamics and Function at Micelle and

134

The relationship between solvent dynamics and protein function is still a contentious one.

With many proponents arguing that small and fast motions of the protein structure add up

to quantifiable effects on enzyme catalysis [54], while others argue that they are driven

primarily through changes in electro-static interactions and bond-formation energies [91,

92]. Since our system does not involve a change in ionic strength, rather a change in

water structure induced by non-ionic surfactants, I have further investigated changes in

dynamics as the link between micelle structure and enzyme activity.

A recently proposed theory solvent slaving links solvent mobility with protein processes

and is explored by Fenimore, Fraudenfelder, and coworkers in 2002 [93, 94]. That study

proposed that there are two classes of protein motions, those that follow the classic

Arrhenius relationship between heat and molecular mobility and those that are linked to

the rotational frequency of solvent motions. Myoglobin was cited as an example of this

phenomenon, where the release of CO is proportional to the dielectric relaxation times of

the surrounding solvent. The rate of the release and of the respective protein motions

where related to solvent motions by the following equation:

��� = �01#2 3��⁄

Equation 5-7 - Frudenfelder relationship between solvent slaved processes and

solvent relaxation rates

Where ���is the slaved protein function, kdiel is the relaxation rate of the solvent’s

dielectric constant, and n(T) is a correlation constant translating the faster water rotation

rates to the slower motions of protein dynamics.

Page 148: Protein Structure Dynamics and Function at Micelle and

135

Solvent-slaving theory

Putting together elements from Viparelli’s and Fraudenfelder’s work, a model that

describes surfactant modified enzyme catalytic rate (����) as a function of micelle

number and how it relates to the volume of pseudo phase water around the micelle

surface. By combining equations 5.3 and 5.7, we now have an expression for the enzyme

reaction rate we a function of enzyme concentration in micelle bound water phase and in

the bulk:

Equation 5-8 - Enzyme reaction rate as a function of enzyme in micelle bound and water regions

Where kdb and kdf are the dielectric relaxation rates of the solvent in bound water and bulk

regions, respectively. NT is the relaxation correlation factor that translates the rate of

solvent motion to observed protein function rate. By plugging in equations 5-5 and 5-6

into bound and free enzyme concentrations, giving a new relationship between enzyme

activity as a function of micelle number:

Equation 5-9 - enzyme reaction rate as a function of micelle number

What has not yet been considered is the actual very small percentage of bound water

volume in the APG surfactant concentrations in which enhanced activity is observed.

���� = 5�0�3� 6 ���� + 7�0&3� 8 %�&'

���� = 5�0�3� 6 9�-��(:���;�1 + -��(:� < + 7�0&3� 8 9 ��;�

1 + -��(:�<

Page 149: Protein Structure Dynamics and Function at Micelle and

136

Also, what factors that arise between densely packed and loosely packed DM and APG

surfactants have not yet been differentiated. Some mechanism of interaction between

micelles that drives enzyme and micelles together so that their water structures would

affect one another.

To reconcile this inconsistency we further investigate water structure and its role in

determining aggregation. K. Collins proposed a theory on how water structure is a

stronger driving force that influences ion attraction than previously assumed. Collins

proproses ‘law of matching water affinity’ where the similarity of water affinity of

oppositely charged ions play just a critical role in their ability to form pairs as do their

electrostatic interactions[95, 96]. Ultimately, ‘like sees like’, small ions with greater

charge densities result in highly ordered water about them, but their electro-static

attractions overcome this boundary and push out the intermediate water. Alternatively,

large ions of low charge density poorly adhere to water, thus big ions of opposite charges

also tend to form pairs. It is small, well hydrated ions, and large loosely hydrated ions

that do not form pairs.

It is not unreasonable that the same phenomenon is happening in this system of micelles

and enzymes. Non-ionic surfactants do not have strong electro static forces that induce

attraction or repulsion, so sphere of hydration is the only factor that determines

dispersion. In this, it is possible that micelles carrying strongly bound water will remain

dispersed. Those with weakly bound hydration shells will succumb to Van Der Waals

forces, and aggregate to minimize the surface area of weakly bound water. This would

not be unlike the forces that result in the surfactant “clouding” phenomenon commonly

observed in non-ionic surfactant systems. In the case of APG’s, the greater degree of

Page 150: Protein Structure Dynamics and Function at Micelle and

137

head group exthoxylation resulting in higher cloud point temperatures, i.e. greater

temperature needed to disrupt hydration layer and induce clouding. Also longer chain

lengths lower cloud point. This is consistent with what has been observed in

heterogeneous APG where micelles initially form from the longest chain lengths with the

shortest head groups. This aggregation between micelle and enzyme would be reflected

in the ‘interaction rate’ parameter Ke

Figure 5-62 - a) (left) hydration sphere of densely packed micelles such as that of dodecyl maltoside and high conc.

APG consists of tightly bound water. b) (right) loosely packed micelles result in more mobile hydration sphere

Figure 5-63 - Minimization of loosely bound hydration layers leading to enzyme/micelle interaction

Page 151: Protein Structure Dynamics and Function at Micelle and

138

Expanding on this concept, the water structuring not only around the micelle but relative

to that in the bulk needs also be considered. Because of the high ionic strength of

potassium phosphate buffer, the bulk water is rather ordered. This results in further

increasing the driving force that draws hydration water away from the softly formed APG

micelles, resulting in diminished hydration spheres.

At this point it is speculated that the hydration about the enzyme is also relatively

disordered and weakly bound, and thus seeks out the spaces near the micelle. The waters

near the micelles maybe fewer and very likely more mobile. It’s likely the one or both of

these factors interact with the enzyme structure to globally effect fluctuation modes of the

overall structure.

5.75.75.75.7 Results and discussion conclusionsResults and discussion conclusionsResults and discussion conclusionsResults and discussion conclusions

In this dissertation the effects of various surfactants on enzymes have been measured via:

• Enzyme activity & kinetics

• Surface tension

• Time resolved fluorescence

• Electron spin resonance

• FTIR

• Dynamic light scattering

Page 152: Protein Structure Dynamics and Function at Micelle and

139

The main goals of this project was to 1) Identify which surfactants induced increased

enzyme activity 2) Elucidate the structures or surfactant properties that are associated

with this enhancement and 3) Explore protein flexibility as a mechanism by which

surfactants are affecting enzyme activities.

From the activity experiments, several non-ionic surfactants were shown to improve the

product yield within the experiment time, long chain exthoxylate surfactants Brij-30 and

Brij-35, as well as an alkyl polyglucoside, sugar-based surfactant mixture, provided to us

by an industrial partner. Of the three surfactants, interaction with APG resulted in the

greatest degree of enzyme activity increases in the two model enzymes tested, subtilisin

protease and horseradish peroxidase. In addition, APG induced enzyme hyper-activity in

both enzymes in a narrow surfactant concentration, peaking at 2-3mM. In comparison,

the activities horseradish peroxidase was also tested in the lab grade, single surfactant,

dodecyl maltoside, also a sugarbased surfactant with a sized head and tail group similar

to the average of what would be found in the APG mixture. This control surfactant did

not induce significant changes to the activity of the HRP enzyme.

To better understand the differences between these two surfactants that would cause one

to effect greater enzyme activity while the other not, surface tension, dynamic light

scattering, NMR and ESR were employed to characterize aggregation behavior of the two

surfactants. APG surface tension results reveal two breaks in the surface tension vs

concentration plot, suggesting that there are two significant aggregation events that occur

Page 153: Protein Structure Dynamics and Function at Micelle and

140

as bulk concentration increases. We will refer to these transition points as critical micelle

concentrations (CMC) I and II.

The first CMC at 0.3 mM indicates an initial formation of surfactant aggregates. After

this point, the surface tension is expected to remain constant with surfactant

concentration. In the case of this APG sample, the surface tension continues to decrease

with concentration from 0.3 mM to 5.0 mM. This phenomenon is common in mixed

surfactant systems (ref mixed systems APG) where the Gibbs free energy of micelle

formation is not uniform for each component of the mixture. This results in the less

soluble moieties in the mixture to precipitate into micelles first, while detergents with

higher CMC continue to stay in monomer form, collecting at the surface and lowering

surface tension. This trend continues until the solubility limit for the most soluble

surfactant moiety is reached, signified by the “second” CMC depicted in this APG

system. It is also worthy to note that it is during this micelle transition phase between 0.3

and 5.0 mM where HRP activity is shown to increase, peaking at 3mM of surfactant

concentration.

Alternatively, DM demonstrates a singular inflection point suggesting that the

homogeneous distribution of surfactants immediately forms one structure of micelle

above the CMC and there is no further change in that structure with increasing surfactant

concentration, only an increase in the number of aggregates.

Page 154: Protein Structure Dynamics and Function at Micelle and

141

Particle size measurements were done of these surfactant systems to further elucidate the

differences between APG and DM. APG at higher concentrations, 5, 10 and 20 mM form

micelles that range from 60 to 70 nm. This is considered a large micelle, approximately

10 times larger than expected if the surfactant were to aggregate into a spherical,

monolayer configuration. Since rod-like morphologies of APG has been well observed,

(ref on APG) these large particle sizes observed can be attributed to the formation of such

elongated structures.

However, these well-defined, rod-like structures, can only be seen at concentrations

greater than 5 mM, that is, past the second “CMC” identified with surface tension. At

3mM and below, DLS can only pick up extremely large particles averaging in 1.5

microns in size, approximately 2 orders of magnitude larger than the already large rod-

like configuration formed at higher concentrations. A continuous structure with a core of

hydrophobic tails and surface of hydrophilic head is unlikely at this scale. This

phenomenon suggests a “clouding” phenomenon [97] wherein micelles of non-ionic

surfactants further aggregate with each other because they lack the surface charge nor a

stable hydration cage to overcome the forces attracting them to neighboring aggregates.

The result is “super-aggregates” large enough to render the solution cloudy. This is would

explain the hazy appearance of the solution at these middle concentration ranges.

Dodecyl maltoside did not exhibit any of the multi-phase size shifts as seen in the APG

mixtures. DLS indicated that the single surfactant formed small (~8-10 nm) micelles

upon increasing the concentrations past CMC. Further additions of surfactant did not

change micelle size, but increased DLS signal counts per second, indicating and increase

in micelle count.

Page 155: Protein Structure Dynamics and Function at Micelle and

142

Thus far, surface tension and particle sizing experiments have demonstrated significant

physical differences between the aggregation characteristics of APG and DM. Moreover,

it appears that the transition phase between the two CMC’s observed in APG has greatest

effect on the enzyme activities. Alternately, while dodecyl maltoside has similar head-

group structure and tail lengths, the micelle structures that this homogeneous surfactant

formed are significantly different than the mixed APG system, which correlates well to

the differences in enzyme activity that the two surfactants induce. ESR and NMR were

used to further investigate the differences in physical properties between the aggregates

of these two surfactants, and how they might ultimately affect enzyme activity.

ESR was used to probe the formation of micelles and their packing density with and

without the presence of HRP enzyme. Surface tension measurements indicate micelle

formation starts around 0.2 – 0.3mM for both APG and DM. However, ESR indicated the

formation of micelles occurring at a higher concentration, 1mM, for both surfactants.

This discrepancy may be attributed to the stearic acid probe, which maybe suppressing

micelle formation. Above 1mM the differences in micelle evolution can be observed. In

DM, the probe mobility rapidly decreases (inverse of rotation correlation time) and then

plateaus, supporting surface tension and DLS observations of formation of spherical,

monolayer micelles as the final conformation of the micelle aggregates. For APG, it was

inferred from surface tension and particle sizing that an intermediate micelle structure

phase is formed before the surfactant concentration is sufficiently high to support the

final, rod-like, micelle structures. ESR further supports this hypothesis, demonstrating

that the micelle packing increases gradually after 1mM appearing to stabilize as the

Page 156: Protein Structure Dynamics and Function at Micelle and

143

concentration approaches 10mM. This slow increase in packing also supports the

proposed hypothesis of gradual incorporation of surfactants from the APG mixture, with

less soluble portions aggregating in the bulk first. As surfactant concentration increases

the more soluble portions start to integrate into the micelles, increasing packing density

of the aggregate as reflected in ESR. Moreover, the mobility of the probe in the DM

micelles is more restricted that those in APG. This agrees with the initial assertion that

DM forms mono-layered, spherical micelles, a configuration that most densely packs the

hydrophobic tail portions. Local viscosity of the probe in APG does not reach the same

degree of immobility and the 2D NMR spectrum indicates significant proximity between

the C13-C14 and hydrophobic head groups. Combined with the ~70 micelle size diameter

measured by DLS, we can infer an anti-parallel, bilayer morphology associated with the

final aggregate structure. This can come in the form of a hollow spherical vesicle or tube

like structure. It is important to note that it is not this bilayer formation that induces

greater enzyme activity. When the APG concentration is sufficiently high for these large

structures to form, the activity is seen to return to initial values. Putting the information

from surface tension, DLS and ESR together we propose the following schematic to

illustrate the differences in micelle structure between APG and DM.

Page 157: Protein Structure Dynamics and Function at Micelle and

144

Figure 5-64 - Schematic differentiating micelle formation in DM (top) vs APG (bottom)

NMR also provides greater insights into the formation of APG micelles and how this is

affected by the presence of proteins such as the horseradish peroxidase enzyme. Peak

line-widths indicate changes in slow rotation of whole aggregates and were used to

characterize the evolutions of APG structure as a function of concentration. Mirroring the

ESR results discussed above, NMR revealed a dual transition phase. The first occurs

shortly after the first CMC at approximately 0.4 mM resulting in aggregates that

gradually, but steadily, decrease in mobility with surfactant concentration. This agrees

with the current hypothesis that this concentration range results in super aggregates,

loosely packed micelles that further aggregate with adjacent micelles. Further increasing

the concentration at this range only adds more mass and volume to the super-aggregate,

resulting in an increasingly slower rotation of the particle. The rate of decreasing particle

Page 158: Protein Structure Dynamics and Function at Micelle and

145

motion with respect to concentration slows at 5mM, approximately where the surfactant

is expect to transition to the rod-like conformation.

While surface tension, DLS, ESR and NMR has yielded valuable information on the

distinct differences between the surfactant structures that do initiate enhanced enzyme

activity (APG) versus those that do not (DM), they have not been able to explore the

precise molecular mechanism by which these particular micelle conformations should

enact on the system to increase the rate of enzyme catalyzed reaction. As to date, several

theories have been provided to explain the mechanism that governs the interactions

between these non-ionic surfactants and the enzyme activity, ranging from influencing

local substrate concentrations about the micelles to preventing unproductive adsorption of

enzymes onto a substrate surface. These factors discussed in previous literature will all

play a role in determining enzyme activity, but what has yet to be suggested or

investigated in significant detail is how surfactant interactions with the enzyme may

affect the structural dynamics, or “flexibility” of the protein structure, thus affecting the

activity. This study closely examines the relationship between enzyme internal structure

mobility and function as a possible explanation for the changes in activity as they interact

with the surfactants and seeks to find a generic correlation between surfactant interaction

and enzyme activity that will be applicable to all systems involving a functional protein

and surface active agents.

Page 159: Protein Structure Dynamics and Function at Micelle and

146

Time resolved fluorescence was used to measure the changes in the internal structural

mobility of the two model enzymes, subtilisin protease and horseradish peroxidase. For

the subtilisin, a dansyl fluoride probe was selectively attached to the serine residue of the

active site, and its rotational mobility was measured with time resolved anisotropy

fluorescence. Exposing the labeled proteinase to two surfactants known to initiate an

increased activity response, APG and Brij-30, resulted in faster rotational correlation

times (i.e. increased mobility) of the dansyl probe.

HRP does not have a convenient attachment site for a fluorescent probe to take advantage

of. However, it does have a heme-group that quenches the intrinsic fluorescence of local

tryptophan residues. The rate of fluorescence decay is attributed to the physical distance

between between the trp and heme group and has been used by other groups as a means

of measuring conformation changes. This work theorizes that changes to this decay time

can also be indicative to changes in the rate and range of internal fluctuations. The greater

the degree of these motions, indicate a more flexible structure and would result in faster

energy transfers reflected by time resolved fluorescence measurements. Additions of

APG to the system increased the fluorescence decay rate, indicating a more flexible

structure, much in the same way that APG affected the subtilisin enzyme. As a

comparison, dodecyl maltoside, which did not induce faster enzyme activities, was

introduced to the fluorescence experiment and did not increase the HRP flexibility. This

is the first indication that these surfactants control the internal mobility of protein

structures.

Page 160: Protein Structure Dynamics and Function at Micelle and

147

Currently, we have rationalized that the interactions which induce greater enzyme activity

are primarily between the enzyme and micelles, as activity is closely linked to surfactant

concentration in APG, brij-30 and Brij-35 post CMC, this is especially evident in APG

systems, where the activities changes significantly according to the conformation of the

surfactant micelle. This opens a significant question as to how does a non-charged

micelle surface, or any surface for that matter, interact with a protein structure in such a

way as to manipulate the dynamics of its internal structures. Enzyme activity and

flexibility measurements indicate that this enhancement phenomena occurs at a very

specific concentration range for APG which is also associated with an increase in the

enzyme flexible mobility. While the factors that drive the inter-micellar aggregation

between non-ionic surfactant micelles, or “clouding”, are still not entirely clear, water

structure is commonly cited as the primary motivator in determining the stability of non-

ionic surfactant suspensions. More specifically, hydration spheres of ordered water about

a non-ionic surfactant provided the necessary repulsion to prevent non-ionic micelles

from aggregating. If this hydration sphere were disrupted by the addition of salts or an

increase in temperature above that surfactants “cloud-point”, this hydration sphere is

disrupted and micelle aggregation occurs, typically resulting in cloudy solutions.

This work as demonstrated that the APG surfactant mixture transitions through a micelle

morphology that results in a low cloud point, and it is during this surfactant concentration

range that increases enzyme flexibility, increasing activity. A wealth of studies highlight

Page 161: Protein Structure Dynamics and Function at Micelle and

148

the importance that hydration and solvent dynamics have on the structure and function of

proteins[39, 43, 46, 49]. In this work I hypothesize that the hydration spheres determined

by micelle interfaces dictate the solvent structure and dynamics which in turn interact

with the protein. The APG surfactants between 0.2 and 2.0 mM result in a depleted

hydration sphere, resulting in a localized volume of solvent which is more mobile than in

the bulk. The model that this work proposes describes protein function rates as a function

of two aspects of micelle – bound water – protein interactions: volume of bound water as

a function of surfactant concentration, and interaction kinetics between micelle and

protein. This interaction kinetics is an indication of the attractive forces between micelle

and protein which has been proposed by this work to arise from the ‘likenes’ of hydration

layers about protein and micelles. This concept of attraction of like hydration was

discussed by Collins and coworkers and creates a physical frame work which predicts ion

pairs form based not only on charge, but the similarities of their hydration

Page 162: Protein Structure Dynamics and Function at Micelle and

149

6 Conclusions and Next Steps

This project was initially focused on the possible degradation of bio-based surfactants by

enzymes often included in detergent formulations. Preliminary observations with FTIR

indicate the loss of the β-acetyl linkage of n-dodecyl-β-d-maltoside upon introduction of

a cellulase enzyme. In parallel, we also measured the activity of subtilisin protease in the

presence of various detergents. We anticipated that anionic detergents such as SDS would

eventually bind to, and denature the enzyme, reducing catalytic activity, while non-ionic

surfactants such as Brij-30 (Polyethylene glycol dodecyl ether), and APG (alkyl

polyglucoside) will preserve the enzymes native structure and thus, activity, as these

findings are often described in literature. Indeed, prolonged exposure (>24 hrs) of

subtilisin to the charged surfactants resulted in diminished activity relative to the enzyme

in buffer alone. While the precise mechanisms for this activity drop in anionic surfactant

has not been investigated experimentally in this work, literature consistently cites

denaturation of proteins in ionic surfactants. Unexpectedly, enzyme in solutions of the

non-ionic surfactants did not just maintain their activity, but exhibited increased activity

relative to that of the buffer only normal. This result could not be as easily explained as

with the loss of activity induced by the charged surfactants. Thus it is this “enhancement”

phenomenon and mechanism behind it that has become the central point of investigation

of this work.

Page 163: Protein Structure Dynamics and Function at Micelle and

150

Our group is not the first to observe this enzymatic “hyperactivity” induced by non-ionic

surfactants. Within the last decade several groups have reported similar results with

different enzymes including amylases and celluases, in the presence of a range of non-

ionic surfactants, such as: several from the Brij series of poly ethylene ethers, and Triton

X-100. While the phenomenon of enhanced enzyme activity in non-ionic surfactant

solutions were consistently observed by all of these studies, the explanations provided by

each one rarely agreed with one another. The goal of this work was then to propose a

mechanism that would universally explain this enzyme hyperactivity, and to test this

novel hypothesis. One mechanism that the previous studies have not considered was the

effect of the enzymes’ conformational mobility in determining overall activity. As such,

this was the concept central to this thesis: the observed increase in enzyme activity is the

result of changes to the enzyme’s structural dynamics induced by interactions with the

surfactants.

The pathway to testing this hypothesis was divided into three distinct objectives. 1) To

determine appropriate test systems of enzyme, substrate, surfactant and buffer, as well as

to develop method to measure enzyme activity of each system. 2) Elucidate the nature of

surfactants and enzyme aggregates if there are any. 3) Measure enzyme structure

flexibility and determine if there is a correlation between surfactant structure, flexibility

and enzyme activity.

Page 164: Protein Structure Dynamics and Function at Micelle and

151

By working towards these three objectives, we were able to: 1) define two model

enzymes that exhibited varied enzyme activity in certain surfactants, and a variety of

surfactants that generated different responses in this enzyme activity. 2) Identify the

differences between these surfactants, specifically those of their micelle structures, and

how these differences may be responsible for their interactions with enzymes affecting

enzyme activity. 3) Identify changes in enzyme structure flexibility as a prominent

mechanism by which certain surfactants are manipulating enzyme activity. In parallel to

these main objectives, we have also explored the effects that a hard- particle surface has

on the structural flexibility of adsorbed proteins and proposed a model on how Forster

resonant energy transfer can be affected by changes in protein flexibility.

Ultimately, completion of these objectives revealed a positive connection between

surfactant interaction with a protein structure flexibility and the impact that these changes

have on enzyme activity. They also illustrated the differences between the interaction of

enzymes with non-ionic surfactant micelles and protein’s adsorbed onto hard particle

surfaces via electro-static interactions. While the former increased the protein’s structural

flexibility, the latter is seen to decrease it. It is hypothesized in this work that bound

water interactions are a primary path way of interaction that allows non-ionic micelles to

increase flexibility of the enzyme structure and we have built model linking changes in

rates of enzyme function to the volume of water pseudo-phases generated by micelle

surfaces. The claim that solvent dynamics is the primary conduit between micelle

structure and enzyme flexibility is further substantiated by our initial enzyme activity

tests in Hoffmeister series ion solutions. Experiments to conclusively explore the role of

Page 165: Protein Structure Dynamics and Function at Micelle and

152

bound water structure between surfactant micelles and enzyme structures are planned for

the next steps for this study, but with the work described in this study, we have

determined that surfactants can affect the activity enzyme activity by manipulating

structural flexibility, a mechanism that should be considered in any system where a

surfactant is seen to have an impact on the activity of an enzyme or a functional protein.

Other studies have also observed differences in activity and flexibility as a function of the

Hoffmeister series [98].

6.16.16.16.1 NNNNext steps: ext steps: ext steps: ext steps:

In this body of work we have investigated the mechanisms that drive changes enzyme

activity in non-ionic surfactants. This behavior might easily be attributed to unfolding

and conformation changes of the protein structure in the presence of charged surfactants,

except that non-ionic surfactants typically do not disrupt the steady state structure of

proteins. On the contrary, many are used to preserve proteins Ex-Vivo. And only

recently, has increased enzyme activity in the presence of non-ionic surfactants have been

observed. This phenomenon hints that a proteins immediate environment has more

influence on its functions than previously considered. The majority of the studies that

have reported this enzyme enhancement via surfactant interaction phenomena have

attributed this to factors external to the protein itself, i.e. prevention of non-specific

interactions, concentrating substrate through volume exclusion, or dispersion of substrate.

This study is the first to consider the effects that the surfactants might have on the

protein’s structure directly, that is through manipulation of the structures overall

Page 166: Protein Structure Dynamics and Function at Micelle and

153

flexibility without disrupting secondary or tertiary native structures. In the future, a

simpler, simple-cleavage peptide substrate should be considered for future activity

measurements, as casein substrate used in the experiments described in this dissertation

are large complex globular proteins, which may result in large deviations in activity

results. While it was demonstrated in this work that flexibility is being affected by

interaction with particles and surfactant micelles, what remains to be conclusively proven

is that these changes in the protein’s structural dynamics are motivated by water

structural dynamics and a disruption of hydration boundaries extending outwards from

non-ionic surfaces. To conclusively test for such a mechanism, a superfast spectroscopy

technique will be needed to measure the dynamics of water in different surfactant

environments. Fortunately such techniques exist, such as the super-fast, time resolved IR

techniques pioneered by Professor Fayer [51, 99, 100] In parallel, molecular dynamics

simulations [37, 101] may provide precise data on how the motions of solvent molecules

impact the motions of proteins.

Page 167: Protein Structure Dynamics and Function at Micelle and

154

7 References

1. Antrim, R.L., et al., Industrial Enzymes: Enzymes in Nonfood Applications, in Enzymes in

Industry. 2004, Wiley-VCH Verlag GmbH & Co. KGaA. p. 155-244.

2. Sivakumar, G., et al., Bioethanol and biodiesel: Alternative liquid fuels for future

generations. Engineering in Life Sciences, 2010. 10(1): p. 8-18.

3. Wilson, D.B., Cellulases and biofuels. Current Opinion in Biotechnology, 2009. 20(3): p.

295-299.

4. Leech, D., P. Kavanagh, and W. Schuhmann, Enzymatic fuel cells: Recent progress.

Electrochimica Acta, 2012. 84(0): p. 223-234.

5. Somasundaran, P., Chin, M., Latosiewicz, U. T., Tuller, H. L., Barbiellini, B. and

Renugopalakrishnan, V, Nanoscience and engineering for Robust Biosolar Cells. Bio-

nanotechnology: Global Perspectives II, ed. R. D.E. Vol. 2. 2012, Boca Raton Fl: CRC

Press.

6. Moore, P.N., S. Puvvada, and D. Blankschtein, Role of the Surfactant Polar Head

Structure in Protein−Surfactant Complexa&on:  Zein Protein Solubilization by SDS and by

SDS/C12En Surfactant Solutions. Langmuir, 2003. 19(4): p. 1009-1016.

7. Hoshino, E. and A. Tanaka, Enhancement of enzymatic catalysis of &lt;i&gt;Bacillus

amyloliquefaciens α-amylase by nonionic surfactant micelles. Journal of Surfactants and

Detergents, 2003. 6(4): p. 299-303.

8. Martinek, K., et al., The principles of enzymes stabilization: VI. Catalysis by water-soluble

enzymes entrapped into reversed micelles of surfactants in organic solvents. Biochimica

et Biophysica Acta (BBA) - Enzymology, 1981. 657(1): p. 277-294.

9. Rodríguez, V.B., et al., Modification of the activity of an α-amylase from Bacillus

licheniformis by several surfactants. 2006. 2006.

10. Kabanov, A.V., et al., Enzymes entrapped in reversed micelles of surfactants in organic

solvents: A theoretical treatment of the catalytic activity regulation. Journal of

Theoretical Biology, 1988. 133(3): p. 327-343.

11. Eriksson, T., J. Börjesson, and F. Tjerneld, Mechanism of surfactant effect in enzymatic

hydrolysis of lignocellulose. Enzyme and Microbial Technology, 2002. 31(3): p. 353-364.

12. Mizutani, C., et al., Effect of a nonionic surfactant on Trichoderma cellulase treatments

of regenerated cellulose and cotton yarns. Cellulose, 2002. 9(1): p. 83-89.

13. Kaar, W.E. and M.T. Holtzapple, Benefits from Tween during enzymic hydrolysis of corn

stover. Biotechnology and Bioengineering, 1998. 59(4): p. 419-427.

14. Shome, A., S. Roy, and P.K. Das, Nonionic Surfactants:  A Key to Enhance the Enzyme

Activity at Cationic Reverse Micellar Interface. Langmuir, 2007. 23(8): p. 4130-4136.

15. Shi, Y.-F., X.-J. Yu, and Q. Yuan, Enhancement Effect of Reverse Micelles on Enzymatic

Synthesis of Cephalexinaa. Annals of the New York Academy of Sciences, 1998. 864(1):

p. 497-501.

16. Deo, N., et al., Surfactant Interactions with Zein Protein. Langmuir, 2003. 19(12): p.

5083-5088.

17. De, S., A. Girigoswami, and S. Das, Fluorescence probing of albumin–surfactant

interaction. Journal of Colloid and Interface Science, 2005. 285(2): p. 562-573.

Page 168: Protein Structure Dynamics and Function at Micelle and

155

18. Ruso, J.M., N. Deo, and P. Somasundaran, Complexation between Dodecyl Sulfate

Surfactant and Zein Protein in Solution. Langmuir, 2004. 20(21): p. 8988-8991.

19. Bru R., S.-F.A., Garcia-Carmona F., A theoretical study on the expression of enzymic

activity in reverse micelles. Biochem. J., 1989. 259: p. 355-61.

20. Viparelli P., A.F., Cantarella M., , Models for Enzyme Superactivity in Aqueous Solutions

of Surfactants. Biochem. J., 1999. 344: p. 765-773.

21. Lanyi, J.K., Bacteriorhodopsin. Annual Review of Physiology, 2004. 66(1): p. 665-688.

22. Hayashi, S., E. Tajkhorshid, and K. Schulten, Molecular Dynamics Simulation of

Bacteriorhodopsin's Photoisomerization Using Ab Initio Forces for the Excited

Chromophore. Biophysical Journal, 2003. 85(3): p. 1440-1449.

23. Edholm, O., O. Berger, and F. Jähnig, Structure and Fluctuations of Bacteriorhodopsin in

the Purple Membrane: A Molecular Dynamics Study. Journal of Molecular Biology, 1995.

250(1): p. 94-111.

24. Hampp, N., Bacteriorhodopsin as a Photochromic Retinal Protein for Optical Memories.

Chem Rev. , 2000. 100(5).

25. Audette, G., Lombardo S., Dudzik, J., Arruda, T.M., Kolinski M., Filipek, S., Mukerjee, S.,

Kannan, A.M., Thavasi, V., Ramakrishna, S., Chin, M., Somasundaran, P., Viswanathan,

S., Keles R.S., and Renugopalakrishnan V., Protein Hot Spots at Bio-nano Interfaces.

Materials Today, 2011. 14.

26. Koshland, D.E., Conformational changes: How small is big enough? Nature Medicine

1998. 4: p. 1112-1114.

27. Koshland, D.E., Application of a Theory of Enzyme Specificity to Protein Synthesis.

Proceedings of the National Academy of Sciences, 1958. 44(2): p. 98-104.

28. Daniel, R.M., et al., THE ROLE OF DYNAMICS IN ENZYME ACTIVITY. Annual Review of

Biophysics and Biomolecular Structure, 2003. 32(1): p. 69-92.

29. Agarwal, P., Enzymes: An integrated view of structure, dynamics and function. Microbial

Cell Factories, 2006. 5(1): p. 2.

30. Tang, K.E. and K.A. Dill, Native protein fluctuations: the conformational-motion

temperature and the inverse correlation of protein flexibility with protein stability.

Journal of biomolecular structure & dynamics, 1998. 16(2): p. 397-411.

31. Allen, T.W., Andersen O.S., Roux B., On the Importance of Atomic Fluctations, Protein

Flexibility, and Solvent in Ion Permeation. Journal of General Physiology, 2004. 124(6): p.

679-690.

32. Cooper, A., Conformational change, fluctuation and drift in biological macromolecules:

An empirical langevin approach. Journal of Molecular Liquids, 1988. 39(0): p. 195-206.

33. Lee, A.L. and A.J. Wand, Microscopic origins of entropy, heat capacity and the glass

transition in proteins. Nature, 2001. 411(6836): p. 501-504.

34. Koshland, D.E., The Key–Lock Theory and the Induced Fit Theory. Angewandte Chemie

International Edition in English, 1995. 33(23-24): p. 2375-2378.

35. Watanabe, K., T. Yoshida, and S.-i. Ueji, The role of conformational flexibility of enzymes

in the discrimination between amino acid and ester substrates for the subtilisin-

catalyzed reaction in organic solvents. Bioorganic Chemistry, 2004. 32(6): p. 504-515.

36. Fischer, S. and C.S. Verma, Binding of buried structural water increases the flexibility of

proteins. Proceedings of the National Academy of Sciences, 1999. 96(17): p. 9613-9615.

37. Tuckerman, M.E. and G.J. Martyna, Understanding Modern Molecular Dynamics: 

Techniques and Applications. The Journal of Physical Chemistry B, 1999. 104(2): p. 159-

178.

38. Vickers, T. Diagram illustrating the induced fit model. 2006.

Page 169: Protein Structure Dynamics and Function at Micelle and

156

39. Yang, L., J.S. Dordick, and S. Garde, Hydration of Enzyme in Nonaqueous Media Is

Consistent with Solvent Dependence of Its Activity. Biophysical Journal, 2004. 87(2): p.

812-821.

40. Hudson, E.P., et al., Active-Site Motions and Polarity Enhance Catalytic Turnover of

Hydrated Subtilisin Dissolved in Organic Solvents. Journal of the American Chemical

Society, 2009. 131(12): p. 4294-4300.

41. Griebenow, K. and A.M. Klibanov, Can conformational changes be responsible for solvent

and excipient effects on the catalytic behavior of subtilisin Carlsberg in organic solvents?

Biotechnology and Bioengineering, 1997. 53(4): p. 351-362.

42. Marzola, P. and E. Gratton, Hydration and protein dynamics: frequency domain

fluorescence spectroscopy of proteins in reverse micelles. The Journal of Physical

Chemistry, 1991. 95(23): p. 9488-9495.

43. Affleck, R., et al., Enzymatic catalysis and dynamics in low-water environments.

Proceedings of the National Academy of Sciences, 1992. 89(3): p. 1100-1104.

44. Fitzpatrick, P.A., et al., Enzyme crystal structure in a neat organic solvent. Proceedings of

the National Academy of Sciences, 1993. 90(18): p. 8653-8657.

45. Broos, J., et al., Flexibility of Enzymes Suspended in Organic Solvents Probed by Time-

Resolved Fluorescence Anisotropy. Evidence That Enzyme Activity and Enantioselectivity

Are Directly Related to Enzyme Flexibility. Journal of the American Chemical Society,

1995. 117(51): p. 12657-12663.

46. Bone, S., Electrostatic interactions in protein structures: The influence of water on

enzyme fluctuations. Journal of Electrostatics, 1988. 21(2–3): p. 245-256.

47. Fernández, A. and H.A. Scheraga, Insufficiently dehydrated hydrogen bonds as

determinants of protein interactions. Proceedings of the National Academy of Sciences,

2003. 100(1): p. 113-118.

48. Baldwin, R.L., Temperature dependence of the hydrophobic interaction in protein folding.

Proceedings of the National Academy of Sciences, 1986. 83(21): p. 8069-8072.

49. Kurkal, V., et al., Low frequency enzyme dynamics as a function of temperature and

hydration: A neutron scattering study. Chemical Physics, 2005. 317(2–3): p. 267-273.

50. Fenn, E.E., et al., Dynamics of Water at the Interface in Reverse Micelles: Measurements

of Spectral Diffusion with Two-Dimensional Infrared Vibrational Echoes. The Journal of

Physical Chemistry B, 2011. 115(40): p. 11658-11670.

51. Thielges, M.C. and M.D. Fayer, Protein Dynamics Studied with Ultrafast Two-

Dimensional Infrared Vibrational Echo Spectroscopy. Accounts of Chemical Research,

2012.

52. Bey Temsamani, M., et al., Fourier Transform Infrared Investigation of Water States in

Aerosol-OT Reverse Micelles as a Function of Counterionic Nature. The Journal of

Physical Chemistry B, 1998. 102(18): p. 3335-3340.

53. Orecchini, A., et al., Low-frequency Vibrational Anomalies in β-Lactoglobulin: 

Contribution of Different Hydrogen Classes Revealed by Inelastic Neutron Scattering. The

Journal of Physical Chemistry B, 2001. 105(48): p. 12150-12156.

54. Henzler-Wildman, K.A., et al., A hierarchy of timescales in protein dynamics is linked to

enzyme catalysis. Nature, 2007. 450(7171): p. 913-916.

55. Havsteen, B.H., A new principle of enzyme catalysis: Coupled vibrations facilitate

conformational changes. Journal of Theoretical Biology, 1989. 140(1): p. 101-127.

56. McCammon, J.A., B.R. Gelin, and M. Karplus, Dynamics of folded proteins. Nature, 1977.

267(5612): p. 585-590.

Page 170: Protein Structure Dynamics and Function at Micelle and

157

57. Doster, W., Viscosity scaling and protein dynamics. Biophysical Chemistry, 1983. 17(2):

p. 97-103.

58. McCammon, J.A. and M. Karplus, The dynamic picture of protein structure. Accounts of

Chemical Research, 1983. 16(6): p. 187-193.

59. Karplus, M., Role of conformation transitions in adenylate kinase. Proceedings of the

National Academy of Sciences, 2010. 107(17): p. E71.

60. Textor, L.C., et al., Purification, crystallization and preliminary crystallographic analysis

of peroxidase from the palm tree Chamaerops excelsa. Acta Crystallographica Section F,

2011. 67(12): p. 1641-1644.

61. Becker, T., et al., General Production Methods, in Enzymes in Industry. 2004, Wiley-VCH

Verlag GmbH & Co. KGaA. p. 37-82.

62. Alexov, E.G. and M.R. Gunner, Incorporating protein conformational flexibility into the

calculation of pH-dependent protein properties. Biophysical Journal, 1997. 72(5): p.

2075-2093.

63. Bickley, R.I., et al., A structural investigation of titanium dioxide photocatalysts. Journal

of Solid State Chemistry, 1991. 92(1): p. 178-190.

64. Kipp, J.E., The role of solid nanoparticle technology in the parenteral delivery of poorly

water-soluble drugs. International Journal of Pharmaceutics, 2004. 284(1–2): p. 109-122.

65. Smith, E.L., et al., Subtilisin Carlsberg. Journal of Biological Chemistry, 1968. 243(9): p.

2184-2191.

66. Bott, R., et al., The three-dimensional structure of Bacillus amyloliquefaciens subtilisin at

1.8 A and an analysis of the structural consequences of peroxide inactivation. Journal of

Biological Chemistry, 1988. 263(16): p. 7895-7906.

67. Twining, S.S., Fluorescein isothiocyanate-labeled casein assay for proteolytic enzymes.

Analytical Biochemistry, 1984. 143(1): p. 30-34.

68. Saito Y., M.M., Nakashima S., Odo J., Tanaka Y., Chikuma M., Tanaka H., Determination

of Hydrojen Peroxide with Phenol and 4-Aminoantipyrine by the Use of a Resin Modified

with Manganese-Tetrakis(sulfophenyl)porphine. Analytical Sciences 1987. 3.

69. Beghein, N., et al., Characterization of self-assembling copolymers in aqueous solutions

using Electron Paramagnetic Resonance and Fluorescence spectroscopy. Journal of

Controlled Release, 2007. 117(2): p. 196-203.

70. Waterman, K.C., et al., Use of a nitroxide spin probe to study the structure of the

adsorbed layer of dodecyl sulfate at the alumina-water interface. The Journal of Physical

Chemistry, 1986. 90(26): p. 6828-6830.

71. Cross, A.J. and G.R. Fleming, Analysis of time-resolved fluorescence anisotropy decays.

Biophysical Journal, 1984. 46(1): p. 45-56.

72. Broos, J., et al., Flexibility of enzymes suspended in organic solvents probed by time-

resolved fluorescence anisotropy. Evidence that enzyme activity and enantioselectivity

are directly related to enzyme flexibility. Journal of the American Chemical Society, 1995.

117(51): p. 12657-12663.

73. Wahl, P., Analysis of fluorescence anisotropy decays by a least square method.

Biophysical Chemistry, 1979. 10(1): p. 91-104.

74. Kinosita Jr, K., A. Ikegami, and S. Kawato, On the wobbling-in-cone analysis of

fluorescence anisotropy decay. Biophysical Journal, 1982. 37(2): p. 461-464.

75. Beechem, J.M., The Global Analysis of Fluorescence Intensity and Anisotropy Decay

Data: Second-Generation Theory and Programs. 2002, Springer. p. 241-305.

76. Das, T.K. and S. Mazumdar, pH-Induced Conformational Perturbation in Horseradish

Peroxidase. European Journal of Biochemistry, 1995. 227(3): p. 823-828.

Page 171: Protein Structure Dynamics and Function at Micelle and

158

77. Lasagna, M., et al., Apohorseradish peroxidase unfolding and refolding: intrinsic

tryptophan fluorescence studies. Biophysical Journal, 1999. 76(1 Pt 1): p. 443-450.

78. Bai, S. and A. Dong, Effects of immobilization onto aluminum hydroxide particles on the

thermally induced conformational behavior of three model proteins. International

Journal of Biological Macromolecules, 2009. 45(1): p. 80-85.

79. Braun, L.J., et al., Development of a freeze-stable formulation for vaccines containing

aluminum salt adjuvants. Vaccine, 2009. 27(1): p. 72-79.

80. Wang, J., Sonocatalytic damage of bovine serum albumin (BSA) in the presence of

nanometer titanium dioxide (TiO~ 2) catalyst. 中國化學快報, 2005. 16(8): p. 1105.

81. Nel, A.E., et al., Understanding biophysicochemical interactions at the nano-bio

interface. Nat Mater, 2009. 8(7): p. 543-557.

82. Griep, M.H., Walczak, K., Winder E., Lueking D. R., Friedrich C.R., Quantum dot

enhancement of bacteriorhodopsin-based electrodes. Biosensors and Bioelectronics,

2010. 25.

83. Alfinito, E.R.L., Charge Transport in bacteriorhodopsin monolayers: The contribution of

conformational change to current-voltage characteristics. European Physica Lett., 2009.

85.

84. Layne, E., Spectrophotometric and Turbidimetric Methods of Measuring Proteins.

Methods in Ezymology, 1957. 3.

85. Stoscheck, C., Qantitation of Protein Methods in Ezymology, 1990. 182.

86. Jiang, L.C., et al., Self-assembly behaviour of alkylpolyglucosides (APG) in mixed

surfactant-stabilized emulsions system. Journal of Molecular Liquids, 2011. 158(3): p.

175-181.

87. Totland, C., R.T. Lewis, and W. Nerdal, 1H NMR relaxation of water: A probe for

surfactant adsorption on kaolin. Journal of Colloid and Interface Science, 2011. 363(1):

p. 362-370.

88. Totland, C., et al., Water Structure and Dynamics at a Silica Surface: Pake Doublets in 1H

NMR Spectra. Langmuir, 2011. 27(8): p. 4690-4699.

89. Tehrani Bagha, A.R., et al., Interactions of gemini cationic surfactants with anionic azo

dyes and their inhibited effects on dyeability of cotton fabric. Dyes and Pigments, 2007.

72(3): p. 331-338.

90. Chen, Z.-J. and R.E. Stark, Evaluating spin diffusion in MAS-NOESY spectra of

phospholipid multibilayers. Solid State Nuclear Magnetic Resonance, 1996. 7(3): p. 239-

246.

91. Pisliakov, A.V., et al., Enzyme millisecond conformational dynamics do not catalyze the

chemical step. Proceedings of the National Academy of Sciences, 2009. 106(41): p.

17359-17364.

92. Kamerlin, S.C.L. and A. Warshel, At the dawn of the 21st century: Is dynamics the missing

link for understanding enzyme catalysis? Proteins: Structure, Function, and

Bioinformatics, 2010. 78(6): p. 1339-1375.

93. Fenimore, P.W., et al., Slaving: Solvent fluctuations dominate protein dynamics and

functions. Proceedings of the National Academy of Sciences, 2002. 99(25): p. 16047-

16051.

94. Frauenfelder, H., et al., Protein folding is slaved to solvent motions. Proceedings of the

National Academy of Sciences, 2006. 103(42): p. 15469-15472.

95. Collins, K.D., Ions from the Hofmeister series and osmolytes: effects on proteins in

solution and in the crystallization process. Methods, 2004. 34(3): p. 300-311.

Page 172: Protein Structure Dynamics and Function at Micelle and

159

96. Collins, K.D., Why continuum electrostatics theories cannot explain biological structure,

polyelectrolytes or ionic strength effects in ion–protein interactions. Biophysical

Chemistry, 2012. 167(0): p. 43-59.

97. Balzer, D., Cloud point phenomena in the phase behavior of alkyl polyglucosides in

water. Langmuir, 1993. 9(12): p. 3375-3384.

98. Tóth, K., et al., Flexibility and enzyme activity of NADH oxidase from Thermus

thermophilus in the presence of monovalent cations of Hofmeister series. Biochimica et

Biophysica Acta (BBA) - Proteins and Proteomics, 2008. 1784(5): p. 789-795.

99. Pal, S.K., J. Peon, and A.H. Zewail, Biological water at the protein surface: Dynamical

solvation probed directly with femtosecond resolution. Proceedings of the National

Academy of Sciences, 2002. 99(4): p. 1763-1768.

100. Hamm, P., M. Lim, and R.M. Hochstrasser, Structure of the Amide I Band of Peptides

Measured by Femtosecond Nonlinear-Infrared Spectroscopy. The Journal of Physical

Chemistry B, 1998. 102(31): p. 6123-6138.

101. Paci, E. and M. Marchi, Intrinsic compressibility and volume compression in solvated

proteins by molecular dynamics simulation at high pressure. Proceedings of the National

Academy of Sciences, 1996. 93(21): p. 11609-11614.