copyright by jeffrey philip potratz 2012

180
Copyright by Jeffrey Philip Potratz 2012

Upload: others

Post on 28-Oct-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Copyright by Jeffrey Philip Potratz 2012

Copyright

by

Jeffrey Philip Potratz

2012

Page 2: Copyright by Jeffrey Philip Potratz 2012

The Dissertation Committee for Jeffrey Philip Potratz certifies that this is the

approved version of the following dissertation:

Local and Global Investigations into DEAD-box Protein Function

Committee:

Rick Russell, Supervisor

Kenneth Johnson

Alan Lambowitz

Scott Stevens

Jessie Zhang

Page 3: Copyright by Jeffrey Philip Potratz 2012

Local and Global Investigations into DEAD-box Protein Function

by

Jeffrey Philip Potratz, B.S.

Dissertation

Presented to the Faculty of the Graduate School of

The University of Texas at Austin

in Partial Fulfillment

of the Requirements

for the Degree of

Doctor of Philosophy

The University of Texas at Austin

May 2012

Page 4: Copyright by Jeffrey Philip Potratz 2012

Dedication

To my wife and parents

Page 5: Copyright by Jeffrey Philip Potratz 2012

Acknowledgements

My time at UT has been positively influenced by many individuals that have

helped make this dissertation possible. I would like to thank my advisor, Rick Russell, for

his exceptional patience and clear, effective teaching and communication style. His

logical thinking and attention to detail set an excellent example for his mentees to follow.

Dr. Alan Lambowitz deserves my gratitude for the collaborative scientific efforts

in which I was included and for the generous offerings of his laboratory supplies. I thank

Dr. Kenneth Johnson for teaching me the foundation of enzyme kinetics. Dr. Scott

Stevens and Dr. Jessie Zhang, along with the aforementioned faculty, are my committee

members for whom I would like to show appreciation for their willingness to serve on my

committee and their scientific input.

My fellow lab mates in the Russell lab are also to be acknowledged. The former

members of Cindy Chen, Hari Bhaskaran, Amanda Chadee, and Yaqi Wan made me feel

welcome in the lab and kindly taught me lab protocols. In addition, Pilar Tijerina was a

great source of knowledge for me when I joined the lab. Current lab members Brian

Cannon, Woongsoon Choi, Inga Jarmoskaite, David Mitchell, and Cynthia Pan are

appreciated for fostering a relaxed and comfortable environment in which to work and for

their willingness to help me mature as a scientist through scientific discussions and

critiques.

Special thanks to my mother and father who supported my decision to venture

down to Texas from Wisconsin and to my wife who encourages me, supports me daily,

and is the reason I have called Texas home for the past five years.

Page 6: Copyright by Jeffrey Philip Potratz 2012

vi

Local and Global Investigations into DEAD-box Protein Function

Jeffrey Philip Potratz, Ph.D.

The University of Texas at Austin, 2012

Supervisor: Rick Russell

Numerous essential cellular processes, such as gene regulation and tRNA

processing, are carried out by structured RNAs. While in vitro most RNAs become

kinetically trapped in non-functional misfolded states that render them inactive on a

biologically-relevant time scale, RNAs folding in vivo do not share this same outcome.

RNAs do indeed misfold in the cell; however, chaperone proteins promote escape from

these non-native states and foster folding to functional conformations. DEAD-box

proteins are ATP-dependent RNA chaperone proteins that function by disrupting

structure, which can facilitate structural conversions. Here, studies with both local and

global focuses are used to uncover mechanistic features of DEAD-box proteins CYT-19

and Mss116p. Both of these proteins are general RNA chaperones as they each have the

ability to facilitate proper folding of multiple structured RNAs.

The first study probes how DEAD-box proteins interact with a simple duplex

substrate. Separating the strands of a duplex is an ATP-dependent process and is central

to structural disruption by DEAD-box proteins. Here, how ATP is utilized during duplex

separation is monitored by comparing ATP hydrolysis rates with strand separation rates.

Results indicate that one ATP molecule is sufficient for complete separation of a 6-11

base pair RNA duplex. Under some conditions, ATP binding in the absence of hydrolysis

is sufficient for duplex separation.

Page 7: Copyright by Jeffrey Philip Potratz 2012

vii

Next, focus is shifted to a more global perspective as the function of Mss116p is

probed in the folding of a cognate group II intron substrate, aI5γ, under near-

physiological conditions. Three catalytically-active constructs of aI5γ are used and

catalysis serves as a proxy for folding. Folding of all constructs is promoted by the

presence of Mss116p and ATP. In vitro and in vivo results indicate that a local unfolding

event is promoted by Mss116p, stimulating formation of the native state. Lastly,

orthogonal methods that probe physical features of RNA are used to provide insight into

the structural intermediates with which Mss116p acts.

Page 8: Copyright by Jeffrey Philip Potratz 2012

viii

Table of Contents

List of Tables ........................................................................................................ xii

List of Figures ...................................................................................................... xiii

List of Abbreviations ........................................................................................... xvi

Chapter 1: RNA folding: A Problem and a Solution .............................................1

1.1 Importance of non-coding RNA ...............................................................1

1.2 The folding problem .................................................................................1

1.3 Does misfolding happen in the cell? .........................................................3

1.4 Chaperones assist proper folding ..............................................................3

1.5 Dead-box proteins are RNA chaperones ...................................................5

1.5.1 Structure ........................................................................................5

1.5.2 Not traditional helicases ................................................................5

1.5.3 Mechanism and functions .............................................................6

1.5.4 Specific and general chaperones ...................................................7

1.5.4.1 Specific chaperones ..........................................................8

1.5.4.2 General chaperones ...........................................................8

1.6 Two model general RNA chaperones .......................................................9

1.6.1 CYT-19 .........................................................................................9

1.6.2 Mss116p ......................................................................................11

1.7 Research plan: local and global methods ...............................................12

Chapter 2: DEAD-box proteins can completely separate an RNA duplex using a

single ATP ....................................................................................................19

2.1 Introduction .............................................................................................19

2.2 Materials and methods ............................................................................21

2.2.1 Materials .....................................................................................21

2.2.2 Determination of RNA and nucleotide concentrations ...............22

2.2.3 RNA strand separation ................................................................22

2.2.4 ATP hydrolysis ...........................................................................23

Page 9: Copyright by Jeffrey Philip Potratz 2012

ix

2.2.5 Analysis of strand separation ......................................................23

2.2.6 Stimulation of strand separation by bound ATP .........................25

2.2.7 Derivation of an equation relating the ATP utilization value to the

ATP concentration ......................................................................26

2.2.8 Analysis of data...........................................................................29

2.2.9 Simulations .................................................................................30

2.2.10 Full equation, including intrinsic dissociation of the duplex ....31

2.3 Results .....................................................................................................32

2.3.1 Enhancement of strand separation by bound ATP without hydrolysis

.....................................................................................................34

2.3.2 Increased ATP requirement for longer or more stable duplexes 36

2.3.3 Similar ATP utilization by other DEAD-box proteins ...............37

2.4 Discussion ...............................................................................................38

2.4.1 Strand separation depends on ATP binding, not hydrolysis .......38

2.4.2 What is the role of ATP hydrolysis? ...........................................40

2.4.3 Implications for physiological activities .....................................41

2.5 Footnotes .................................................................................................42

Chapter 3: RNA catalysis as a probe for chaperone activity of DEAD-box helicases

.......................................................................................................................58

3.1 Catalytic activity as a probe of RNA folding..........................................58

3.1.1 Catalytic activity distinguishes the native state from all other

conformations .............................................................................58

3.1.2 Catalytic activity can be used to study chaperone-assisted folding59

3.2 Self-splicing as a readout for native state formation ..............................60

3.2.1 Interpreting chaperone-promoted changes in observed splicing rate

.....................................................................................................60

3.2.2 Potential complications ...............................................................61

3.3 Substrate cleavage as a readout for native state formation .....................63

3.3.1 Setting up a discontinuous assay: folding and catalysis stages...63

3.3.2 Interpreting results from the catalysis stage ................................65

3.3.3 Using the discontinuous assay to probe chaperone-assisted folding

.....................................................................................................67

Page 10: Copyright by Jeffrey Philip Potratz 2012

x

3.4 Other applications of the discontinuous assay ........................................68

3.4.1 Unfolding native structure ..........................................................69

3.4.2 Integrating results with other methods ........................................69

Chapter 4: ATP-dependent roles of the DEAD-box protein Mss116p in group II

intron splicing in vitro and in vivo ................................................................76

4.1 Introduction .............................................................................................76

4.2 Materials and methods ............................................................................81

4.2.1 Recombinant Plasmids ................................................................81

4.2.2 RNA preparation .........................................................................82

4.2.3 Preparation of Mss116p ..............................................................82

4.2.4 Splicing reactions ........................................................................82

4.2.5 Discontinuous catalytic activity assay for D135 RNA folding ...84

4.2.6 S. cerevisiae Northern hybridizations and immunoblotting .......85

4.3 Results .....................................................................................................86

4.3.1 Splicing of LE and SE constructs and Mss116p acceleration .....86

4.3.2 Mutants that are deficient in RNA-unwinding activity ...............88

4.3.3 Two-stage, discontinuous catalytic activity assay for folding of D135

RNA ............................................................................................89

4.3.4 Acceleration of D135 folding by Mss116p .................................92

4.3.5 The role of ATP in acceleration of D135 folding by Mss116p ..93

4.3.6 Disruption of native D135 ribozyme by Mss116p ......................95

4.3.7 Mss116p-promoted splicing of aI5 in vivo ................................96

4.4 Discussion ...............................................................................................98

4.4.1 Requirement for ATP binding and hydrolysis and effects of exon

length on Mss116p-mediated splicing in vitro............................98

4.4.2 The SAT/AAA mutant is compromised for splicing SE and LE

constructs ..................................................................................100

4.4.3 The roles of ATP in Mss116p-promoted intron folding ...........101

4.4.4 Disruption of the native D135 ribozyme ..................................102

4.4.5 Requirement for ATP binding and hydrolysis by Mss116p in vivo103

4.5 Conclusions and implications ...............................................................104

Page 11: Copyright by Jeffrey Philip Potratz 2012

xi

4.6 Footnote ................................................................................................105

Chapter 5: Rapid structure formation within the aI5γ group II intron ................128

5.1 Introduction ...........................................................................................128

5.2 Materials and Methods ..........................................................................130

5.2.1 RNA preparation .......................................................................130

5.2.2 Footprinting data acquisition and analysis ................................130

5.2.3 SAXS data acquisition and analysis .........................................132

5.3 Results ...................................................................................................133

5.3.1 The unfolded state of aI5γ is readily accessible to DMS modification

...................................................................................................133

5.3.2 Addition of Mg2+

results in rapid formation of structural elements

...................................................................................................134

5.3.3 Native state is significantly more protected than early folding time

point state ..................................................................................134

5.3.4 Addition of Mg2+

gives rapid formation of structure in D135

ribozyme ...................................................................................135

5.3.5 SAXS data reveal compaction of D135 ribozyme ....................136

5.3.6 Time-resolved data indicate early compaction .........................137

5.4 Discussion .............................................................................................137

5.5 Footnote ................................................................................................138

Appendix ..............................................................................................................149

A.1 Preliminary DMS footprinting of D135 ribozyme using capillary

electrophoresis ...................................................................................149

Bibliography ........................................................................................................152

Page 12: Copyright by Jeffrey Philip Potratz 2012

xii

List of Tables

Table 2.1: ATP utilization for CYT-19-mediated separation of the 6-bp P1 duplex

...........................................................................................................51

Table 2.2: Dependence of ATP utilization by CYT-19 on Mg2+

concentration and

duplex length .....................................................................................52

Table 2.3: Dependence of strand separation and ATPase rates on duplex length53

Table 2.4: RNA strand separation and ATP hydrolysis rate for duplexes composed

solely of canonical Watson-Crick base pairs ....................................54

Table 2.5: Temperature dependence of ATP utilization by CYT-19 during strand

separation ..........................................................................................55

Table 2.6: ATP utilization by the DEAD-box proteins Mss116p and Ded1p ...56

Table 2.7: Range of conditions for measurements of ATP utilization ...............57

Table 5.1: Nucleotides modified by DMS. ......................................................148

Page 13: Copyright by Jeffrey Philip Potratz 2012

xiii

List of Figures

Figure 1.1: RNA folding pathway .......................................................................14

Figure 1.2: DEAD-box proteins are involved in all aspects of RNA metabolism15

Figure 1.3: Helicase core of DEAD-box proteins and ancillary domains ...........16

Figure 1.4: Structure of Mss116p ........................................................................17

Figure 1.5: Duplex unwinding by a traditional helicase and a DEAD-box protein18

Figure 2.1: ATP hydrolysis and RNA strand separation by CYT-19..................44

Figure 2.2: Effects of a double-stranded extension on RNA strand separation and

ATPase activity .................................................................................45

Figure 2.3: Strand separation by CYT-19 is independent of ‘chase’ CCCUCUA5

concentration (1–10 µM) ..................................................................46

Figure 2.4: ATP-independent strand separation by CYT-19 ...............................47

Figure 2.5: ATP hydrolyzed by CYT-19 per separation event of the 6-base-pair P1

duplex with low Mg2+

concentration (2 mM) ...................................48

Figure 2.6: Model for duplex separation by DEAD-box proteins .......................49

Figure 2.7: Duplex constructs used for measurements of ATP utilization during

unwinding (top, constructs 1-6) and for control experiments (bottom,

constructs 7-9) ...................................................................................50

Figure 3.1: Group I and group II introns .............................................................71

Figure 3.2: Self-splicing constructs .....................................................................72

Figure 3.3: The discontinuous assay ...................................................................73

Figure 3.4: Examples of catalytic reactions with the D135 and Azoarcus ribozymes

...........................................................................................................74

Page 14: Copyright by Jeffrey Philip Potratz 2012

xiv

Figure 3.5: The discontinuous assay with the D135 ribozyme and the DEAD-box

helicase Mss116p ..............................................................................75

Figure 4.1: RNA constructs ...............................................................................106

Figure 4.2: Self-splicing of the LE and SE constructs ......................................107

Figure 4.3: Self-splicing of the LE construct (panel A) and SE construct (panel B) at

different temperatures .....................................................................108

Figure 4.4: Mss116p-stimulated splicing of the LE and SE constructs ............109

Figure 4.5: Mss116p-stimulated splicing of LE and SE constructs ..................111

Figure 4.6: Stimulation of splicing by the SAT/AAA mutant of Mss116p.......112

Figure 4.7: Mss116p motif I mutants K158A and K158R in splicing reactions of the

LE and SE constructs (panels A and B, respectively) ....................113

Figure 4.8: Two-stage catalytic activity assay to monitor folding of D135 ribozyme

.........................................................................................................114

Figure 4.9: Testing of 30 °C, 100 mM Mg2+

for conditions of stage 2 in the

discontinuous catalytic activity assay .............................................116

Figure 4.10: Prefolded and nonfolded reactions for monitoring D135 RNA folding at

42 °C ...............................................................................................117

Figure 4.11: Mss116p accelerates native folding of the D135 ribozyme ............118

Figure 4.12: Progress curves of D135 ribozyme folding ....................................120

Figure 4.13: Proteolysis of Mss116p after incubation with D135 RNA in the absence

of ATP .............................................................................................121

Figure 4.14: Proteinase K digestion of Mss116p in experiment shown in Figure 4.13

.........................................................................................................123

Figure 4.15: Proteolysis of Mss116p after incubation with D135 RNA .............124

Figure 4.16: Unfolding of native D135 RNA by Mss116p .................................126

Page 15: Copyright by Jeffrey Philip Potratz 2012

xv

Figure 4.17: Northern hybridization and correlated immunoblot comparing the ability

of wild-type and mutant Mss116p to promote splicing of aI5 in vivo127

Figure 5.1: Representative DMS footprinting gel .............................................139

Figure 5.2: Many DMS accessible nucleotides in the unfolded state ................140

Figure 5.3: Nucleotides protected and exposed in the native and early folding time

point states ......................................................................................141

Figure 5.4: DMS footprinting results for domain IV ........................................142

Figure 5.5: DMS footprinting profiles of full-length aI5γ ................................144

Figure 5.6: Kratky plots of static samples reveal different compaction peak heights

.........................................................................................................145

Figure 5.7: Time-resolved SAXS data indicate an early compaction event......146

Figure A.1: DMS footprinting profile of the D135 ribozyme ............................150

Page 16: Copyright by Jeffrey Philip Potratz 2012

xvi

List of Abbreviations

AMP-PNP- adenosine 5´-(,-imido) triphosphate

bp(s)- base pair(s)

EDTA- ethylenedinitrilotetraacetic acid

MOPS- 3-(N-morpholino)propanesulfonic acid

RNP- RNA-protein complex

SF1- helicase superfamily 1

SF2- helicase superfamily 2

DTT- dithiothreitol

LE construct- self-splicing construct of the aI5γ intron with long 5 and 3 exons of 293

and 321 nucleotides, respectively

Mss116p SAT/AAA- mutant of Mss116p with S305A and T307A substitutions in motif

III

mt- mitochondrial

nt(s)- nucleotide(s)

PVDF- polyvinylidene fluoride

SDS- sodium dodecyl sulfate

SE construct- self-splicing construct of the aI5γ intron with short 5 and 3 exons of 28

and 15 nucleotides, respectively

Tris- tris(hydroxymethyl)aminomethane

WT- wild-type

ssRNA- single-stranded RNA

tRNA- transfer RNA

rRNA- ribosomal RNA

Page 17: Copyright by Jeffrey Philip Potratz 2012

1

Chapter 1: RNA folding: A Problem and a Solution

1.1 IMPORTANCE OF NON-CODING RNA

RNA has long been known as the middle man in the process of transforming a

DNA sequence into a functional protein. However, through the years the importance of

non-coding RNA has been recognized and, in addition to serving as a template for protein

production, RNA is involved in many other vital cellular processes (1-4). For example,

RNA composes a large percentage of the ribosome which carries out the cellular

production of protein. Additionally, riboswitches are RNA molecules that change

structure upon binding small ligands and are able to regulate gene expression (5, 6).

The cellular roles attributable to RNA greatly increased with the discovery of

catalytic RNA in the early 1980’s by Thomas Cech and Sidney Altman (7, 8). This

finding revealed that RNA can fulfill enzymatic roles in the cell. These catalytic RNA

molecules include group I and group II introns which self-splice out of precursor

messenger RNA. The processing of tRNA molecules is also carried out by catalytic RNA

found in RNaseP, a complex composed of RNA and protein whose RNA component is a

ribozyme and is responsible for catalytic and enzymatic activity (8).

1.2 THE FOLDING PROBLEM

In order for any non-coding RNA to perform its function, it must first fold to a

correct three-dimensional structure. RNA folding follows a progressive, hierarchical

model in which the correct conformation is achieved through a pathway of intermediate

states that have increasing stability. Because secondary structural elements form

extremely rapidly and can persist in the absence of tertiary interactions, molecules with

secondary structure are considered the first intermediates in the folding pathway. Next,

these pre-formed secondary structures interact to form tertiary contacts that result in a

Page 18: Copyright by Jeffrey Philip Potratz 2012

2

further increase in molecular stability and a correctly-folded molecule (9-12) (Figure

1.1). This correct final structure is commonly referred to as the native state. Interestingly,

one RNA sequence can fold into two structurally-distinct conformations that are both

native, as in the case of riboswitches (1). In addition, it has also been suggested that the

native state is not necessarily one particular conformation with minimal global free

energy but is rather composed of an ensemble of native conformations that have deep

local minima that interconvert rather slowly (13). While a pathway of forming the native

state through interactions by pre-formed secondary structural elements is logical, it is not

without obstacles. RNA contains intrinsic properties which complicate folding to the

native state and result in most RNAs studied in vitro becoming trapped in a misfolded

state(s) (14).

The hierarchical folding of RNA, combined with the limited variety of its primary

sequence, leads to trapped, misfolded RNA conformations. RNA consists of only four

standard bases, which endow it with minimal information content. When an RNA

molecule hundreds of nucleotides in length forms secondary structure, the limited

information content of its bases permits a multitude of different combinations of duplex

elements to form. Some of these duplex elements will form quickly even if they are not

present in the native state of the RNA. These duplexes form under kinetic control and are

not found in the minimal free energy conformation of the native state (15). Clearly, these

non-native duplex elements must be disrupted before the RNA can attain its native

conformation. However, RNA duplexes can persist on the time scale of hours or days

even in isolation, making escape from non-native structures and subsequent folding to the

native state quite slow. In addition, tertiary interactions can form that reinforce the

stability of the non-native duplexes (16). Lastly, incorrect tertiary contacts can form or

incorrect strand orientations can occur that would also necessitate correction and lead to

Page 19: Copyright by Jeffrey Philip Potratz 2012

3

slow folding to the native state (17). Therefore, on a biological time scale, the folding

process is often not completed successfully and leaves the RNA trapped in one or more

misfolded states, which prevents the RNA from functioning (15).

1.3 DOES MISFOLDING HAPPEN IN THE CELL?

The RNA folding problem has been well documented in vitro and has been shown

to impact nearly every RNA studied (14). However, in vitro misfolding does not prove

that RNA is prone to misfolding in its cellular environment. There are several rather

apparent differences between the environments of a test tube and the cell in which RNA

folds physiologically. In the test tube, RNA starts as a generally linear molecule in the

absence of cations and forms secondary and tertiary structure upon the addition of

monovalent cations and divalent cations, respectively. These cations reduce the

electrostatic repulsion of the nucleotide phosphate groups, allowing the RNA strand to

interact with itself. Conversely, within the cellular environment the folding process does

not need to wait for the addition of cations and occurs as the RNA is still being

transcribed. It is influenced by events such as pausing of the RNA polymerase and single

stranded binding proteins interacting with the nascent transcript (18). While these factors

undoubtedly influence the folding pathway of RNA, there is compelling evidence that

RNA misfolding occurs within the cell, as described further below.

1.4 CHAPERONES ASSIST PROPER FOLDING

Certainly, if most cellular RNA was trapped in misfolded, inactive conformations,

the cell could not function. Subsequently, most cellular RNA must have a way to escape

misfolded states if they become trapped. Chaperone proteins provide an escape pathway.

They allow RNA to escape misfolded conformations by promoting the disruption of

structural elements which in turn promotes structural conversions that lead to proper

Page 20: Copyright by Jeffrey Philip Potratz 2012

4

folding (19, 20). However, one could make the argument that within the cellular

environment RNA simply avoids misfolding and chaperone proteins are not strictly

necessary. Refuting this argument is the fact that at least one chaperone protein is

required for practically every process performed by structured RNA, (21-24) indicating

that RNA is prone to misfolding inside the cell just as it is outside the cellular

environment and needs assistance to fold properly.

RNA chaperone proteins function distinctly from specific RNA-binding proteins,

even though both proteins ultimately assist the RNA in reaching its native state (19, 25).

Specific RNA-binding proteins, such as CYT-18 and CBP2, function by promoting

formation of correct structural elements or by stabilizing the native structure once

independently formed (25). After either function, these proteins remain bound to the

native state. Without the continued presence of the protein, the RNA would revert back to

a non-functional structure. Conversely, chaperones interact with the RNA transiently,

exert their influence on the folding pathway, and are then dispensable (25). They

accelerate structural transitions between alternative conformations thereby disrupting

misfolded states and giving the RNA another opportunity to fold correctly (15, 26, 27).

Importantly, chaperone proteins have been shown to facilitate RNA folding both in vitro

and in vivo (18, 25, 28-31).

In addition to escaping misfolded conformations, some RNAs have multiple

conformationally-distinct native states which must be populated temporally. The small

nuclear RNAs found in the spliceosome are examples of RNAs that populate multiple

native, functional states during their existence (32, 33). As in the case with misfolded

conformations, RNA chaperone proteins can promote structural disruptions to facilitate

transitions that lead to interconversion of one native structure into another (20).

Page 21: Copyright by Jeffrey Philip Potratz 2012

5

1.5 DEAD-BOX PROTEINS ARE RNA CHAPERONES

One group of proteins that act as RNA chaperones are DEAD-box proteins. They

belong to helicase superfamily 2 (SF2) and make up its largest class (21, 34). There are

38 DEAD-box proteins in humans while Saccharomyces cerevisiae has 25 (21).

Additionally, these proteins are prevalent in all forms of life and are ubiquitous in all

stages of RNA metabolism (22) (Figure 1.2). Among their diverse biochemical activities

is the ability to elicit conformational changes in RNA through ATP-dependent separation

of RNA duplexes.

1.5.1 Structure

The minimal helicase core of all DEAD-box proteins consists of two domains that

are similar in appearance to RecA and are connected by a flexible linker (21, 34-36). The

core may be flanked on either or both sides with non-conserved ancillary domains

(Figure 1.3). Within the helicase core are eleven conserved motifs, including domain II

(D-E-A-D) for which the group is named. These motifs align themselves on the interface

of the two domains when ligands are present (12, 37-42) (Figure 1.4). DEAD-box

proteins cooperatively bind an ATP molecule or analog and a single stranded RNA

segment (43-47). Motifs I, II, III, V, VI, and Q are positioned near the ATP moiety and

help bind and regulate it while motifs Ia, Ib, GG, IV, QxxR, and V bind the single

stranded RNA.

1.5.2 Not traditional helicases

DEAD-box proteins are not traditional helicases, which bind their substrates via a

single-stranded overhang and use ATP hydrolysis to translocate along one strand of

nucleic acid with a specific polarity, either 3’-5’ or 5’-3’, displacing the other strand

along the way (48, 49) (Figure 1.5). This helicase mechanism is quite useful in separating

Page 22: Copyright by Jeffrey Philip Potratz 2012

6

long, continuous duplex segments such as those found in both host and viral genomes.

However, applying this mechanism to chaperone RNA folding would be

counterproductive, as all the helices formed in a structured RNA would be disrupted,

regardless of whether they were native or non-native. This action would require the RNA

to essentially start folding again from a linear molecule.

Rather, DEAD-box proteins are unique ‘helicases’ as they appear to lack

translocation in their strand separation mechanism. The lack of translocation and

concurrent lack of processivity is made apparent by their inability to unwind duplexes

that are longer than 20-25 base pairs (50, 51). In addition, while single-stranded

overhangs can increase the efficiency of duplex separation by DEAD-box proteins, (52,

53) there is no polarity preference for these overhang regions, again indicating no

processivity (51, 54-56). Furthermore, the inclusion of single-stranded overhangs is not

strictly required for duplex separation (57). The unique duplex separation strategy

utilized by DEAD-box proteins is called local strand separation (58) and allows them to

bind directly to one strand of an RNA duplex, absent of any single-strand overhangs, and

efficiently disrupt short duplex elements (59). The duplex lengths disrupted by DEAD-

box proteins are generally less than ten base pairs long, lengths typically found in

structured non-coding RNAs (20, 60). This mechanism of structural disruption allows the

chaperone to function on local areas of the RNA without necessarily disrupting the entire

structure.

1.5.3 Mechanism and functions

The mechanism of a DEAD-box protein is strongly tied to the use of ATP, as

DEAD-box proteins are ATP-dependent strand separators (21). The protein undergoes

changes in ssRNA binding affinity throughout the ATPase cycle. Initially, binding of

Page 23: Copyright by Jeffrey Philip Potratz 2012

7

ATP and ssRNA is cooperative and induces a large conformational change as the two

domains of the helicase core form a ‘closed’ conformation around the nucleotide and

RNA (43-47). In the ATP and/or ADP-Pi state, the protein has a high affinity for the

ssRNA, while after hydrolysis and release of Pi, the protein has a low affinity for the

ssRNA (44, 61, 62). This cycle of ssRNA affinity changes allows DEAD-box proteins to

tightly bind one strand of a duplex in the presence of ATP and/or ADP-Pi. The interaction

with one of the duplex strands induces a conformation in which a duplex cannot exist and

thus promotes duplex dissociation (38). Once the weak binding state of ssRNA is

promoted through Pi release, the protein releases the single strand and is available to

interact with another duplex element.

DEAD-box proteins use this general property of tight binding to ssRNA to

perform multiple functions. In addition to duplex separation activity, DEAD-box proteins

disrupt RNP complexes (63-69). Disruption of RNP complexes is thought to proceed

similarly to duplex disruption as the protein binds tightly to an RNA strand, which leads

to dissociation of the RNA from the protein in the RNP complex. Yet another function

DEAD-box proteins can perform via tight binding to ssRNA is highlighted in the exon

junction complex. In the EJC, the DEAD-box protein eIF4AIII is bound to ssRNA and

acts as a clamp to stabilize the complex. The protein is stably bound to the RNA because

the EJC holds eIF4AIII in the ‘closed’ conformation, which prevents the release of Pi. In

the absence of Pi release, the protein remains in the state which retains high affinity for

ssRNA and cannot let go (70).

1.5.4 Specific and general chaperones

DEAD-box proteins associate with all the cellular processes of RNA from

transcription to decay, and often the same protein is involved in multiple processes (22)

Page 24: Copyright by Jeffrey Philip Potratz 2012

8

(Figure 1.2). Some DEAD-box proteins find their particular substrates with the help of

non-conserved ancillary domains, which flank the helicase core on the N and/or C-

terminus (71, 72) (Figure 1.3). Others use interactions mediated through the helicase core

to find their substrates (40, 73-76). DEAD-box proteins fall into two categories when

grouped according to their substrates, either specific or general. They can either interact

specifically with a target substrate or are capable of interacting with and chaperoning

multiple substrates.

1.5.4.1 Specific chaperones

The DEAD-box protein DbpA from Escherichia coli and its ortholog YxiN from

Bacillus subtilis are examples of chaperones that are directed to a specific substrate, the

23S rRNA. Both DbpA and YxiN interact with it through ancillary C-terminal domains

(77, 78). More descriptively, the C-terminal domain forms an RNA recognition motif and

binds to a secondary structure feature in the 23S rRNA termed hairpin 92 (79, 80). This

additional RNA binding site outside of the helicase core tethers the chaperone in place

and allows the helicase core to bind nearby duplexes and separate them (79) (Figure 1.5).

While the tethering interaction is specific in this case, a similar tethering mechanism

appears to function with general chaperones.

1.5.4.2 General chaperones

There are significantly more distinct RNA sequences in the cell than chaperone

proteins. Further, each RNA can presumably adopt more than one misfolded structure or

contact, so that the number of misfolded structures that may require assistance from

chaperones is very large. Therefore, it logically follows that some DEAD-box proteins

function on multiple substrates non-specifically (20, 26, 30, 81, 82). These proteins are

referred to as general RNA chaperones. Similar to how specific chaperones have been

Page 25: Copyright by Jeffrey Philip Potratz 2012

9

shown to use ancillary domains to anchor near target substrates, general chaperones also

appear to tether near substrates via ancillary domains (53, 83, 84) (Figure 1.5). Ded1p is

an example of a general RNA chaperone found in S. cerevisiae and thought to function in

translation initiation (85-87). A possible role suggested for Ded1p is removing secondary

structures present in mRNA that prohibit the ribosome from reaching the start codon (88,

89). Disrupting structural elements of mRNAs to allow ribosome binding is a process that

inherently would require the protein to act non-specifically because each RNA sequence

would naturally be unique as it encodes for a different protein.

1.6 TWO MODEL GENERAL RNA CHAPERONES

Two DEAD-box proteins that exhibit general chaperone activity have become

model general RNA chaperones. The first is CYT-19 from Neurospora crassa and the

second is Mss116p from S. cerevisiae. Both function physiologically to assist folding of

introns and appear to use an ancillary domain containing many basic residues, referred to

as the ‘basic tail’ or the basic ‘C-tail’, to bind non-specifically to their substrates (84)

(Figure 1.3).

1.6.1 CYT-19

A significant event in the history of DEAD-box protein research occurred in 2002

when CYT-19 was shown genetically and biochemically to function in the folding of

three cognate mitochondrial group I introns (28). While in vivo RNA chaperone activity

had been demonstrated earlier using the nucleocapsid protein and StpA, (90) this was the

first study in which a DEAD-box protein functioned as a chaperone in vivo to disrupt

misfolded structure and promote proper folding of physiologically relevant RNA. In

addition, the finding that one protein could function on three different group I introns

highlighted the general nature of its activity. The generalness of action was demonstrated

Page 26: Copyright by Jeffrey Philip Potratz 2012

10

further when it was shown that CYT-19 can be expressed in S. cerevisiae and at least

partially suppress all the defects in splicing that result from loss of the cognate DEAD-

box protein Mss116p (91). Therefore, not only is CYT-19 able to promote group I intron

folding, but it can promote non-cognate folding of both group I and group II introns

which have different overall architectures.

CYT-19 contains an approximately 50 amino acid ‘C-tail’ that is highly basic and

is a common feature in various sub-families of DEAD-box proteins (92) (Figure 1.3).

This basic C-tail appears to be responsible for non-specific tethering of CYT-19 to its

numerous substrates. The positively charged tail binds to the negative charge on the

phosphate backbone of nucleic acid and tethers the helicase core near adjacent duplexes

(84). Comparison of wild type CYT-19 and a mutant missing the C-tail reveals that the

strand separation activity of the mutant is not enhanced by the presence of a nucleic acid

extension adjacent to the substrate duplex, whereas the same extension enhances strand

separation activity of the wild type protein. This and additional results suggested a model

in which the C-tail can tether the protein to the extension and position the protein in close

proximity to the substrate duplex (93).

CYT-19 has been shown to resolve a misfolded state found in the group I intron

ribozyme derived from Tetrahymena thermophila, (31) further demonstrating its non-

specific activity as a chaperone. Importantly, the misfolded species of the ribozyme is

well characterized, is almost as compact as the native state, and contains the same

secondary and tertiary interactions characteristic of the native state (17, 94). Observing

how the chaperone interacts with the similar misfolded and native states allows features

of its mechanism to be dissected. For instance, does CYT-19 recognize and act on the

misfolded state only or does it also act on the native state? Because the native state and

misfolded state of the ribozyme are so comparable, it is hard to imagine how CYT-19

Page 27: Copyright by Jeffrey Philip Potratz 2012

11

might interact specifically with the misfolded state. In fact, CYT-19 is able to disrupt the

native structure under destabilizing conditions (26). This disruption of native structure

indicates that CYT-19 is so non-specific that it cannot differentiate between the native

state and a misfolded state. Rather, it ‘blindly’ acts on the RNA, with the misfolded

structure being disrupted with greater efficiency because of its lower stability compared

to the native structure, and does not influence the refolding of the disrupted structure.

Therefore, CYT-19 promotes net formation of the native state from a population of

misfolded molecules by kinetically redistributing the RNA and allowing it further

chances to fold to the more stable native state, where it is not productively acted on (26).

While the study referenced above indicates that general chaperones cannot

discriminate between native and misfolded structural features, the native state may indeed

be protected from disruption. CYT-19 appears unable to separate a duplex element of the

Tetrahymena ribozyme when it is involved in a tertiary contact within the core of the

ribozyme (95). This protection offered by a tertiary contact may represent a general

strategy to protect native structure. As is the case for the Tetrahymena ribozyme, native

states may be more tightly packed than misfolded states which may bias chaperone

proteins to productively function predominantly on misfolded states because their

structural elements are more exposed. The misfolded states may contain structural

features that are exposed enough and unstable enough to be disrupted while the native

states may contain structural elements that are protected by tertiary contacts and/or are

too stable to be disrupted.

1.6.2 Mss116p

Mss116p is a general chaperone and physiologically functions in splicing of all

introns found in the mitochondrial genome of S. cerevisiae (91). There are nine group I

Page 28: Copyright by Jeffrey Philip Potratz 2012

12

introns and four group II introns in the mitochondrial genome that have varying overall

architectures. The ability of one protein to help fold RNAs which have varying

architectures displays the non-specific nature of Mss116p activity. Like CYT-19,

Mss116p contains a basic C-tail which may be involved in tethering the chaperone to its

substrates (84).

The crystal structure of Mss116p was solved in 2009 and is very reminiscent of

other solved structures of DEAD-box proteins complexed with an ATP analogue and

ssRNA, with one notable exception (37) (Figure 1.4). The ssRNA is bent twice in this

structure compared to only one bend seen in the previous structures (12, 37-42). This

‘extra’ bend, which results in the RNA being ‘crimped’, is caused by the positioning of

the C-terminal extension.

1.7 RESEARCH PLAN: LOCAL AND GLOBAL METHODS

The overarching goal of my research is to uncover key facets to how DEAD-box

proteins function. The studies used to probe the mechanism start with a narrow focus and

then broaden out to gain a more comprehensive perspective. In the first study, the action

of chaperones is probed through the use of the most basic substrate, a duplex element (see

Chapter 2). The study quantitatively dissects how DEAD-box proteins use ATP to disrupt

this simple secondary structural element. Several general chaperones, including CYT-19

and Mss116p, from different organisms are used on a target substrate derived from the

Tetrahymena ribozyme. In this way, the results are unbiased by any physiological context

and provide a simple look at how many ATPs are required to separate duplex elements.

The second project uses catalytically active RNA (see Chapter 3) and takes a

more global look at how proper RNA folding is promoted through chaperone activity.

The study dissects the role of Mss116p in the folding pathway of the cognate group II

Page 29: Copyright by Jeffrey Philip Potratz 2012

13

intron aI5γ (see Chapter 4). In this way, physiologically relevant chaperone activity is

observed. Lastly, methods used to probe for structural features of RNA, DMS

footprinting and small angle x-ray scattering (SAXS), are used to identify physical

characteristics of intermediate conformations found along the aI5γ folding pathway (see

Chapter 5).

Page 30: Copyright by Jeffrey Philip Potratz 2012

14

Figure 1.1: RNA folding pathway

RNA folding follows a progressive pathway filled with intermediates of increasing

stability. Secondary structure forms first kinetically, followed by modular assembly into

tertiary contacts and a final folded conformation. This figure originally published in (12)

by Landes Bioscience.

Page 31: Copyright by Jeffrey Philip Potratz 2012

15

Figure 1.2: DEAD-box proteins are involved in all aspects of RNA metabolism

Yeast and human ortholog DEAD-box proteins (green circles) are shown connected to

the RNA processes (white boxes) in which they participate. Other SF2 family members

are also depicted (red, light blue, dark blue, and yellow colored circles). This figure

originally published in (22) by Elsevier.

Page 32: Copyright by Jeffrey Philip Potratz 2012

16

Figure 1.3: Helicase core of DEAD-box proteins and ancillary domains

The minimal helicase core shared by all DEAD-box proteins is labeled as ‘DEAD-box’

and shown in light blue. The ‘C-tail’ for CYT-19, Mss116p, and Ded1p is labeled BT for

basic tail and is dark blue. Additional ancillary domains are depicted with various colors

and labeled as follows: RRM, RNA-recognition motif; Di, dimerization domain; RBD,

non-conserved RNA-binding domain; αHR, α-helical region; SMN, survival of motor

neurons protein; SF-1, steroidogenic factor 1 protein; RS arginine-serine-rich domain;

RGG, arginine-glycine-glycine repeats; IQ, calmodulin-binding domain. This figure

originally published in (82). Reproduced with permission by The Royal Society of

Chemistry.

Page 33: Copyright by Jeffrey Philip Potratz 2012

17

Figure 1.4: Structure of Mss116p

(A) Cylindrical representation of the Mss116p sequence. Regions found in the crystal

structure are shown as gray cylinder while regions not found in the crystal structure are

black lines. The abbreviations below the cylinder are as follows: NTE, N-terminal

extension; Domain 1 and 2, the helicase core; CTE, C-terminal extension; BT, basic tail

or ‘C-tail.’ The conserved motifs are highlighted with different colors and labeled above

the cylinder. (B) The crystal structure (37) with domain 1 in blue, domain 2 in green, C-

terminal extension in gray, ssRNA in magenta, and AMP-PNP in orange. C) The crystal

structure (37) in the same orientation as panel B but with the individual motifs now

colored as in panel A. Note the ‘crimp’ in the ssRNA as it is bent twice. This figure

originally published in (96) by Wiley.

Page 34: Copyright by Jeffrey Philip Potratz 2012

18

Figure 1.5: Duplex unwinding by a traditional helicase and a DEAD-box protein

(A) A traditional helicase binds to a single-stranded extension (red) and uses ATP

hydrolysis to translocate directionally through a duplex region (black) and separate the

strands. This model helicase translocates 3’-5’. (B) A DEAD-box ‘helicase’ binds to the

RNA duplex (black) directly with its helicase core (large orange circle) and uses ATP to

separate it with minimal if any translocation. This illustration shows a DEAD-box protein

which possesses an ancillary domain (smaller orange circle) that can bind a region

adjacent to the substrate duplex, depicted here as an adjacent duplex (red), and tether the

helicase core near substrate duplex regions The ancillary domain could bind specifically

or non-specifically to the structure adjacent to the substrate duplex and could be a ‘C-tail’

domain, an RRM domain, or other motifs. This figure originally published in (12) by

Landes Bioscience.

Page 35: Copyright by Jeffrey Philip Potratz 2012

19

The text below and following figures of chapter 2 were originally published by National

Academy of Sciences. Copyright © by the National Academy of Sciences 2008:

Chen, Y*., Potratz, J.P*., Tijerina, P. Del Campo, M., Lambowitz, A.M., Russell, R .

(2008) DEAD-box proteins can completely separate an RNA duplex using a single ATP.

Proc Natl Acad Sci U S A 105(51):20203-20208.

*Chen, Y. and Potratz, J.P. are co-first authors.

Mark Del Campo from the Alan M. Lambowitz lab purified proteins.

Yingfeng Chen and Pilar Tijerina from the Rick Russell lab helped collect data.

Rick Russell, Yingfeng Chen, and Alan Lambowitz helped analyze the data and wrote the

paper.

Chapter 2: DEAD-box proteins can completely separate an RNA

duplex using a single ATP

2.1 INTRODUCTION

To gain insight into the mechanism of CYT-19 activity, we took advantage of the

observation that it can use its non-specific chaperone activity to efficiently separate the

six-base-pair helix termed P1, formed between group I introns and their 5´-exon-intron

junction, and that this unwinding efficiency (kcat/KM) is enhanced by two orders of

magnitude when the duplex is covalently linked to the ribozyme compared to the same

duplex in solution (31). Then, using simple constructs based on group I intron structure,

we found that the activity was also enhanced by simple extensions to the helix.

Interestingly, a single-stranded extension gave a smaller enhancement than a double-

stranded flanking region, and both gave smaller enhancements than the highly structured

intact group I intron.

Whereas conventional DNA and RNA helicases commonly require a 5- or 3-

single-stranded region, which serves as a starting point for translocation into and through

the duplex, results above suggested that the increased activity arose instead from an

Page 36: Copyright by Jeffrey Philip Potratz 2012

20

additional and distinct interaction of CYT-19 with the RNA. Further, the enhancement

under subsaturating conditions (kcat/KM) suggested that this interaction is maintained in

the transition state for strand separation. Additional work showed that the enhancement is

nearly eliminated by deletion of 49 amino acids from the highly basic C-terminus of

CYT-19, whereas strand separation of a duplex that lacks an extension is essentially

unaffected, most simply suggesting that this additional interaction is mediated by the C-

terminal region (93).

Together, these findings led to a model in which interactions with adjacent RNA

structure tether DEAD-box proteins in proximity to exposed helices or perhaps other

elements of RNA structure, where binding of the core domain and ATP-dependent

conformational changes give strand separation (31, 35, 53). Although early studies

demonstrating that DEAD-box proteins can readily separate duplexes of approximately

one helical turn or less but are essentially inactive for duplexes of two or more turns (56,

97) indicated a lack of processivity, the tethering model suggests a more radical

difference in mechanism from processive helicases. This is because continuous formation

of a tethering interaction during duplex unwinding would most simply suggest the

absence of any translocation during the unwinding process.

Strong independent support for essential features of this model has come from

studies in the Jankowsky lab using an elegant set of model duplex substrates. First, they

demonstrated conclusively that a flanking sequence can enhance activity for DEAD-box

proteins without serving as a starting point for translocation by showing that a single-

stranded segment can still give activation if it is not linked to the target duplex but is

instead bound through biotin-mediated interactions with the protein streptavidin (53).

Second, they showed that even model duplexes with an RNA segment flanked on both

sides by DNA can be efficiently separated by DEAD-box proteins, whereas the same

Page 37: Copyright by Jeffrey Philip Potratz 2012

21

proteins are not active on fully DNA substrates, indicating that strand separation can be

initiated internally, without translocation (98). The main features of the model are also

indicated for the E. coli DEAD-box protein DbpA by studies from Uhlenbeck and

colleagues, with the important difference that this protein uses an ancillary domain to

recognize a particular structure within the large subunit ribosomal RNA rather than

interacting with structured RNA more generally (77-80).

In the current work, we have tested and extended this model by measuring the

number of ATP molecules utilized by CYT-19 and other DEAD-box proteins as they

separate RNA duplexes. In the most extreme form of the model, with strand separation

accomplished in the presence of a continuously-formed tethering interaction, it would be

possible that the complete reaction would be accomplished in a single cycle of ATP-

dependent conformational changes and would therefore give hydrolysis of only one ATP.

Indeed, we obtain this result for duplexes of 6–11 bp, characteristic of helices present in

structured RNAs. Further, under some conditions, a significant fraction of strand-

separation events occur in the absence of any ATP hydrolysis. Nevertheless, these events

are dependent on ATP, indicating that bound ATP favors a protein conformation that

promotes local strand separation even prior to ATP hydrolysis.

2.2 MATERIALS AND METHODS

2.2.1 Materials

Oligonucleotides were purchased from Dharmacon. CCCUCUA5 was 5´-end-

labeled with [-32P]ATP by using T4 polynucleotide kinase and gel-purified (99). CYT-

19, Mss116p, and Ded1p were expressed and purified as described (55, 83). AMP-PNP

was treated to remove any contaminating ATP as described (60).

Page 38: Copyright by Jeffrey Philip Potratz 2012

22

2.2.2 Determination of RNA and nucleotide concentrations

The ATP utilization values are linearly dependent on both the duplex and ATP

concentrations, so it is important that the concentrations be measured carefully. We

determined these concentrations spectrophotometrically using extinction coefficients

(260 nm) calculated from base composition: ATP and AMP-PNP, 1.54 104 M

–1 cm

–1;

CCCUCUA5, 1.3 105 M

–1 cm

–1; CCCUCCA5, 1.2 10

5 M

–1 cm

–1; RNA/DNA hybrid

strand, 3.64 105 M

–1 cm

–1 (for hybrid strands with uridine extensions, values were 3.64

105 M

–1 cm

–1 + 0.99 10

4 M

–1 cm

–1 per uridine). Measurements in the presence of 7 M

urea, which disrupts the secondary structure of the P2 extension (31), gave at most small

increases in absorbance (<5%), so concentrations were calculated directly from

measurements made in water. Each concentration was measured at least twice, and

measured values varied by less than 5%.

2.2.3 RNA strand separation

Unless otherwise indicated, reaction conditions were 25 °C, 50 mM Na-MOPS,

pH 7.0, 10 mM MgCl2, 50 µM ATP-Mg2+

, 50 mM KCl, and 5% glycerol. Reactions were

initiated by adding pre-formed duplex (final concentrations of 0.5 µM RNA/DNA hybrid

oligonucleotide, 0.2 µM CCCUCUA5, and trace 32P-labeled CCCUCUA5) to CYT-19 (2

µM), followed by addition of 1–5 µM unlabeled CCCUCUA5 to give a final duplex

concentration of 0.5 µM. Control reactions showed that varying the concentration of

unlabeled CCCUCUA5 across and beyond this range did not affect the rate constant for

strand separation (Figure 2.3), and thus 5 µM CCCUCUA5 was used typically to increase

the signal for strand separation. At various times, aliquots were quenched by adding 70

mM MgCl2 and 1 mg/ml Proteinase K, and then loaded on a 20% nondenaturing

polyacrylamide gel run at 5 ºC. We confirmed that the quench solution was effective, as

CYT-19 did not promote strand separation under the quench conditions (data not shown).

Page 39: Copyright by Jeffrey Philip Potratz 2012

23

Gels were dried, visualized with a phosphorimager (GE Healthcare), and quantitated with

ImageQuant 5.2 (GE Healthcare). Time courses were fit by a single exponential equation

(Kaleidagraph, Synergy Software). Rate constants were converted to steady-state rates by

multiplying by the duplex concentration, 0.5 µM. A control experiment in which

unlabeled chase CCUCUA5 was added before labeled CCCUCUA5, such that CYT-19-

mediated strand separation of the unlabeled P1 duplex was monitored by detecting the

formation of labeled duplex, gave the same rate constant within error (data not shown).

Thus, the presence of the 5’-phosphoryl group on the radiolabeled CCCUCUA5 does not

affect the rate of CYT-19-mediated strand separation.

2.2.4 ATP hydrolysis

Conditions were as above except that reactions included trace [-32P]ATP instead

of 32P-labeled CCCUCUA5. Unlabeled CCCUCUA5 was 1 µM (see above). Aliquots

were quenched with 100 mM EDTA, applied to a polyethyleneimine (PEI) cellulose thin-

layer chromatography plate, developed in 1 M formic acid, 0.5 M LiCl, and quantitated

as above.

2.2.5 Analysis of strand separation

To determine the number of ATP molecules hydrolyzed per stand separation

event, we compared steady-state rates of ATP hydrolysis and strand separation. We

measured the steady-state ATPase rate directly, but for strand separation we measured the

displacement of radiolabeled CCCUCUA5 in a pulse-chase experiment. Thus, it was

necessary to convert the rate constant obtained from a fit by an exponential equation to a

steady-state rate. This conversion was achieved by multiplying the rate constant by the

duplex concentration, typically 0.5 µM (the concentration of the limiting species, the

RNA/DNA hybrid strand). This simple conversion is possible because, although we

Page 40: Copyright by Jeffrey Philip Potratz 2012

24

monitored dissociation only of the labeled CCCUCUA5, the unlabeled CCCUCUA5

present in the reaction is expected to rapidly re-form the duplex, producing a steady state

in which the duplex is largely formed (100-102). Indeed, at the concentrations of

oligonucleotides used, we confirmed that the duplex forms much faster than it is

separated by CYT-19 (data not shown), indicating that the total duplex concentration is

maintained at a steady-state level close to the concentration of the hybrid strand (0.5

µM). This unlabeled duplex is then subject to multiple rounds of unwinding by CYT-19

(31), even though only the first round – separation of the labeled duplex – is monitored

experimentally. The ‘dilution’ of the labeled duplex by continuous formation of

unlabeled duplex is not expected to affect the measured rate constant for strand

separation because CYT-19 is present at subsaturating concentration for binding to the

duplex. The observation that the ATPase rate remains constant at times long after the first

round of strand separation is completed (Figure 2.1B and C) provides strong support for

the interpretation that the duplex concentration remains relatively constant, as control

experiments indicate that the separated strands would give much lower ATPase rates

(Figure 2.1B).

The fraction of labeled CCCUCUA5 that migrated as the duplex at the start of

reactions was typically 0.6–0.7 both in the presence and absence of CYT-19, as shown in

Figure 2.1C. This incomplete retention appears to be caused principally by dissociation of

a modest fraction of the duplex in the gel rather than incomplete duplex formation in the

reaction, as increases in the concentration of the complementary strand up to 4-fold (2

µM) produced little or no increase in the fraction of labeled CCCUCUA5 retained in the

duplex (data not shown). This small loss of duplex is not expected to affect the measured

values of the rate constants because it is not expected to affect the shape of the strand

separation curves, but only to decrease the amplitudes.

Page 41: Copyright by Jeffrey Philip Potratz 2012

25

2.2.6 Stimulation of strand separation by bound ATP

To explore whether bound ATP stimulates unwinding even when it is not

hydrolyzed, we fit the measured values of ATP utilization during unwinding of the 6-bp

duplex (25 °C, 2 mM Mg2+

) by models that included or excluded a pathway that does not

result in ATP hydrolysis, yet gives more rapid strand separation than the pathway

mediated by nucleotide-free CYT-19. The equations relating ATP utilization value to

ATP concentration were derived from Scheme 1, in which S1 and S2 represent the

separated strands, and all separation events are shown as irreversible because of the

presence in the experiment of an excess of one of the unlabeled strands. The rate constant

kATPADP represents the second-order process of CYT-19•ATP binding to the duplex and

separating it to single strands in a process that results in hydrolysis of the ATP. Similarly,

kATP represents the second order rate constant for CYT-19•ATP binding to the duplex and

separating the strands, but in a process that does not result in ATP hydrolysis. The rate

constant kno nuc reflects the second order process of CYT-19 binding to the duplex and

Page 42: Copyright by Jeffrey Philip Potratz 2012

26

separating the strands without bound ATP, and kintrinsic is the first-order rate constant for

strand separation in the absence of CYT-19.

Scheme 1 also includes the following assumptions. 1) ATP binding by CYT-19 is

assumed to be in rapid equilibrium, as suggested by the relatively high KM value

determined previously in steady-state ATPase assays (28). 2) CYT-19 is assumed to be

subsaturating for duplex binding, as indicated by an increasing rate with increasing CYT-

19 concentration across the range used in this and previous studies (10 nM – 2 µM) (31,

83). 3) Release of the products, both the single-stranded products and ADP and Pi are not

shown explicitly; the second-order rate constants are defined as the rate constant up to

and including the rate-limiting step(s). This treatment does not assume that product

release steps are faster than hydrolysis. If one or both of the product release steps are

rate-limiting and hydrolysis is readily reversible, the overall rate constant shown reflects

the equilibrium for ATP hydrolysis and the rate constant for the rate-limiting product

release step.

2.2.7 Derivation of an equation relating the ATP utilization value to the ATP

concentration

The ATP utilization value relates the rate of duplex-dependent ATP hydrolysis to

the rate of CYT-19-dependent strand separation. To derive an equation that relates the

ATP utilization value to the ATP concentration, we assumed for simplicity that all

duplex-dependent ATP hydrolysis events are coupled to successful strand separation

under the conditions of Figure 2.5 (2 mM Mg2+

, 6-bp duplex), which give less than one

ATP hydrolyzed per duplex separated. This assumption allows the ATP utilization value

to be expressed as the rate of ATP-hydrolysis-dependent strand separation events relative

to the total strand separation rate produced by all pathways, or in other words as the

fraction of strand separation events that are coupled to ATP hydrolysis (eq. 1). It is

Page 43: Copyright by Jeffrey Philip Potratz 2012

27

possible that some ATP hydrolysis is ‘wasted’, as is suggested to occur with longer

duplexes and at higher Mg2+

concentration (see Table 2.2). If so, the fraction of

unwinding events that proceed through the ATP-hydrolysis-dependent pathway would be

overestimated, while the fraction that rely on bound ATP without its hydrolysis would be

underestimated. (The total fraction of ATP-dependent events is measured directly from

the increase in unwinding rate in the presence of ATP). Thus, this analysis provides a

lower limit on the acceleration of strand separation provided by ATP when it is not

hydrolyzed.

Equation (1)

Under the conditions of our experiments, the non-enzymatic dissociation of the

duplex can be omitted from the equation (kintrinsic), as this background rate was subtracted

from the raw data (see eq. 5 below for the form of the final equation that includes non-

enzymatic duplex dissociation). Further, this process is substantially slower than the

CYT-19-promoted processes and does not substantially affect the fitted parameters. Eq. 1

can therefore be simplified by omitting this term and by dividing the numerator and

denominator by [Duplex] to give eq. 2.

Equation (2)

The concentration of ATP-bound CYT-19 can be related to the ATP

concentration by dividing the numerator and denominator by [CYT-19] and substituting

ATP utilization=[Duplex][CYT -19 • ATP]kATPADP

[Duplex][CYT -19 • ATP]kATPADP [Duplex][CYT -19]kno nuc [Duplex][CYT -19 • ATP]kATP [Duplex]kintrinsic

[CYT -19 • ATP]kATPADP

[CYT -19 • ATP]kATPADP [CYT -19]kno nuc [CYT -19 • ATP]kATP

Page 44: Copyright by Jeffrey Philip Potratz 2012

28

terms using the expression for equilibrium ATP binding to give the following (KD is the

equilibrium constant for ATP binding):

Equation (3)

Eq. 3 can be rearranged by combining terms and multiplying the numerator and

denominator by KD/(kATPADP + kATP) to give a form that is analogous to the Michaelis-

Menten equation (eq. 4), describing a hyperbolic dependence of the ATP utilization value

on ATP concentration.

Equation (4)

The prominent features revealed by eq. 4 are: 1) the plateau value of the ATP

utilization value with increasing ATP concentration reflects the ratio of rate constants for

ATP hydrolysis-dependent and hydrolysis-independent pathways involving ATP-bound

CYT-19, and 2) the ATP concentration that gives the half-maximal ATP utilization value

(the KM-like term) is composed of the KD for ATP binding and a factor that reflects the

increase in rate afforded by bound ATP. This results in the half-maximal value being

lower than the KD because the increased activity of CYT-19 upon ATP binding causes

most of the separation events to be mediated by ATP-bound CYT-19 even at ATP

concentrations that give ATP binding to only a relatively small fraction of CYT-19.

[ATP]kATPADP

KD[ATP]kATPADP

KD kno nuc

[ATP]kATP

KD

[ATP]kATPADP

kATPADP kATP

KDkno nuc

kATPADP kATP

[ATP]

Page 45: Copyright by Jeffrey Philip Potratz 2012

29

2.2.8 Analysis of data

The ATP utilization data in Figure 2.5 are shown with fits by the model of eq. 4

and two related models. The solid curve shows a fit with eq. 4, which includes the

possibility of a rate acceleration by bound ATP without its hydrolysis (kATP). The fit

gives a maximal value of 0.5 that is well defined by the data, indicating that ATP is

hydrolyzed in only half of the unwinding events mediated by CYT-19•ATP; i.e. kATP is

approximately equal to kATPADP. Further, if the value of KD for ATP binding is assumed

to be equal to the KM value measured previously in steady-state ATPase assays under

similar conditions (200 µM, ref. (28), it is possible to calculate values of kATP and

kATPADP using eq. 4. The best-fit of the data in Figure 2.5 gives a half-maximal value of

5 µM (the KM-like term), which from eq. 4 is equal to KD (kno nuc/(kATPADP + kATP)).

Using the measured value of kno nuc of 2 105 M

–1 min

–1 (converted from the steady-state

rate of 0.2 µM/min (see section 2.3.1) with 0.5 µM duplex and 2 µM CYT-19) and the

relationship that kATP = kATPADP, the values of kATP and kATPADP are determined to be 4

106 M

–1 min

–1 (see Scheme 2 below). Thus, bound ATP is calculated to give a 40-fold

enhancement overall, 20-fold via each of two pathways that include or exclude its

hydrolysis. It should be noted that this overall enhancement represents a lower limit; we

directly observe the strand separation rate to increase linearly to at least 150 µM ATP,

indicating that KM is at least ~200 µM, but we are unable to measure strand separation

above 150 µM ATP because it becomes too fast to measure with hand pipetting.

Nevertheless, the previous determination of KM for ATP of 200 µM in steady-state

ATPase assays under similar conditions (28) and our measurement of 200 µM for KD of

AMP-PNP by inhibition (Figure 2.4) suggests that the KD for ATP under our conditions

is likely to be close to 200 µM.

Page 46: Copyright by Jeffrey Philip Potratz 2012

30

In contrast to the good fit obtained with the model above, the data are not well-

described by a model that excludes the ATP-dependent, hydrolysis-independent pathway

by omitting kATP from the equation, giving clear systematic deviation from the long-

dashed curve in Figure 2.5. Excluding this pathway has the effect of forcing the ATP

utilization value to plateau at unity with increasing ATP concentration, with the physical

implication that any strand-separation process mediated by ATP-bound CYT-19 gives

hydrolysis of the bound ATP. The data are also not well-described by a model in which

strand separation by ATP-bound CYT-19 is permitted without ATP hydrolysis, but bound

ATP does not accelerate strand separation unless it is hydrolyzed (short dashed curve in

Figure 2.5). In this model, the value of kATP is required to be equal to the value of kno nuc.

To be compatible with our observed data, the ATP utilization value is required to plateau

at a value of 0.9 or larger (the plateau value is kATPADP/(kATPADP + kno nuc)), as we

observed directly that the ATP-dependent unwinding rate constant kATPADP is at least 10-

to 20-fold larger than the rate constant for ATP-independent unwinding (kno nuc).

2.2.9 Simulations

To confirm the utility of eq. 4, we performed simulations using Scheme 2, which

Page 47: Copyright by Jeffrey Philip Potratz 2012

31

is modified from Scheme 1 by including values of the rate constants obtained from fitting

eq. 4 to the data on ATP utilization as described above (Figure 2.5). Our prediction was

that, if eq. 4 accurately described the conversion of the fitting parameters (KM- and kcat-

like parameters) to the values of the relevant rate constants (kno nuc, kATPADP, and kATP), a

simulation using these values for the rate constants should return simulated ATP

utilization values that are consistent with the data and the fitted curve.

Simulations were performed using Kinetic Explorer (Kintek) with the rate

constants shown above and the following initial concentrations: CYT-19, 2 µM; S1, 0.5

µM; S2, 1 µM. The duplex was allowed to form with a rate constant of 109 M

–1 min

–1

(100). At several ATP concentrations, steady-state rates of ATPase activity and strand

separation were recorded. From these rates, ATP utilization values were calculated and

plotted on the data and fits from Figure 2.5. The ATP utilization values calculated from

the simulations were in good agreement with the curve fit to the data (data not shown),

indicating that eq. 4 is valid. Its use returned rate constants that give the observed

behavior.

2.2.10 Full equation, including intrinsic dissociation of the duplex

Eq. 5, shown below, is analogous to eq. 4 but includes non-enzymatic duplex

separation (kintrinsic). Terms including the first-order rate constant kintrinsic are divided by

the total CYT-19 concentration, [CYT-19]t, instead of the equivalent form of multiplying

each of the second-order rate constants by [CYT-19]t, to allow the form of eq. 5 to more

closely resemble the simpler eq. 4. (Eq. 5 was not used in any analysis herein but is

included for completeness. It could be necessary to include this term under conditions

that give little or no stimulation of strand separation by CYT-19 in the absence of

nucleotide, such as Mg2+

concentrations of 10 mM or higher.)

Page 48: Copyright by Jeffrey Philip Potratz 2012

32

Equation (5)

2.3 RESULTS

In designing a duplex substrate for these studies, we took advantage of our

previous results that CYT-19 can efficiently separate the P1 duplex of the Tetrahymena

group I ribozyme, leading to dissociation of the oligonucleotide substrate, and the

separation is much more efficient when P1 is covalently linked to the intron or to another

RNA helix (31). This increased activity allows robust experimental signals for strand

separation and duplex-dependent ATPase activity (ref. (31) and results herein). However,

attachment of the intron or even a second RNA duplex would inextricably complicate the

analysis because the additional RNA would be expected to interact with CYT-19 and

stimulate its ATPase activity, and it would not be possible to determine how much of the

total ATPase activity arose from separation of the P1 duplex.

We therefore generated a construct in which the P1 duplex is formed from an

RNA oligonucleotide (CCCUCUA5) and a hybrid RNA/DNA oligonucleotide, resulting

in P1 being flanked by a DNA duplex (Figure 2.1A). This dsDNA extension gave the

same activation of strand separation activity as an equivalent RNA extension and, as

expected from earlier work (106, 107), an oligonucleotide containing only the DNA

[ATP]kATPADP

kATPADP kATP kintrinsic

[CYT -19]t

KD

kno nuckintrinsic

[CYT -19]t

kATPADP kATP kintrinsic

[CYT -19]t

[ATP]

Page 49: Copyright by Jeffrey Philip Potratz 2012

33

portion did not stimulate ATPase activity, implying that it is not actively unwound by the

helicase core (Figure 2.2).

Using this substrate, we measured ATPase activity by CYT-19 under defined

conditions (10 mM Mg2+

; Figure 2.1B). Nearly all of the ATPase activity arose from

interactions with the duplex under these conditions, as the rate was ~10-fold lower in the

presence of either strand alone (Figure 2.1B). Because the duplex separation reactions

included a small excess of CCCUCUA5 and were performed under subsaturating

conditions, we subtracted from the total the rate in the presence of 1 µM CCCUCUA5,

which approximates its free concentration in reactions including the duplex. After

subtracting this background, we obtained a duplex-dependent rate of 0.74 ± 0.05 µM/min

(Figure 2.1B and Table 2.1).

We then measured P1 duplex separation under the same conditions using a pulse-

chase gel mobility shift assay (Figure 2.1C and Figure 2.3). After subtracting the CYT-

19-independent separation, we obtained a rate constant of 1.4 ± 0.1 min–1

from a fit by a

first-order rate equation (Figure 2.1C and additional replicates not shown). To compare

the rate of strand separation with the steady-state rate of ATPase activity measured

above, we converted the rate constant to a steady-state rate by multiplying it by the

duplex concentration (see Section 2.2.5). This conversion gave a steady-state rate of 0.69

± 0.05 µM/min (Table 2.1). Strikingly, the rates of ATPase activity and strand separation

are the same within error, giving a ratio of 1.1 ± 0.1 ATP molecules hydrolyzed per

duplex separated. This ratio, or ATP utilization value, was the same within error across

the range of experimentally accessible ATP concentrations (5–150 µM, data not shown;

KM = 200 µM, ref. (28) and across the more limited range of accessible CYT-19

concentrations (0.5–2 µM, data not shown, see Section 2.5)*. Thus, under these

conditions (10 mM Mg2+

, 25 °C), a single cycle of ATP-dependent conformational

Page 50: Copyright by Jeffrey Philip Potratz 2012

34

changes apparently gives complete strand separation of this 6-bp duplex. If any of the

base pairs are not disrupted directly by the enzyme during this cycle, they must dissociate

spontaneously.

2.3.1 Enhancement of strand separation by bound ATP without hydrolysis

To explore whether hydrolysis of ATP is uniformly required for strand separation

by DEAD-box proteins, we decreased the Mg2+

concentration, shown previously to

increase the strand separation activity of CYT-19 (31, 83). With 5 mM Mg2+

, the ATP

utilization value remained approximately one (Table 2.1). With lower Mg2+

concentration

(2 mM), however, it decreased to 0.45, implying that about half of the strand separation

events proceeded without ATP hydrolysis. ATP-independent separation of this duplex

has been shown for the related DEAD-box protein Mss116p (60), and we confirmed that,

with 2 mM Mg2+

but not 10 mM Mg2+

, CYT-19 gives detectable strand separation in the

absence of nucleotide (with a rate of 0.2 µM/min; Figure 2.4). This activity may be

achieved by ‘strand capture’, analogous to the RNA chaperone activity of proteins that

are not ATPases (60, 108, 109).

We therefore considered a model in which the ATP utilization value of 0.45

reflected a balance between strand separation mediated by ATP-bound CYT-19, which

would result in ATP hydrolysis, and nucleotide-free CYT-19, which would of course not

give ATP hydrolysis. However, it was not clear that the unwinding rate in the absence of

ATP (0.2 µM/min; above) was large enough to support this model. In order for half of the

separation events in the presence of ATP to be mediated by nucleotide-free CYT-19, the

ATP-independent rate would have to be at least half of the total CYT-19-dependent rate

of 0.97 ± 0.05 µM/min (Table 2.1). Although the difference between the expected and

Page 51: Copyright by Jeffrey Philip Potratz 2012

35

observed values is not large, it raised the possibility that bound ATP stimulates the rate of

strand separation by CYT-19 even when it is not hydrolyzed.

We tested this possibility by systematically varying the ATP concentration under

the same 2 mM Mg2+

conditions (Figure 2.5). If the low ATP utilization value resulted

from activity of nucleotide-free CYT-19, it would increase with increasing ATP

concentration, reflecting the increased fraction of CYT-19 bound to ATP, and would

approach unity with saturating ATP (see Section 2.2.6). Instead, the ATP utilization

reached a plateau value of 0.5-0.6, giving no further increase with ATP concentration,

even as the rate of strand separation increased to a value ≥ 40-fold larger than without

ATP. This behavior indicates that ATP is hydrolyzed only in approximately half of the

strand separation events, even when it is bound, and that bound ATP strongly accelerates

strand separation even when it is not hydrolyzed (≥ 20-fold, see Section 2.5)‡. These

results strongly suggest that ATP elicits or stabilizes a protein conformation that induces

or captures local strand separation events (see below and Figure 2.6).

Interestingly, we found that the enhancement from bound ATP is not mimicked

by the non-hydrolyzable analog AMP-PNP, which gave no acceleration of strand

separation beyond the basal level in the absence of nucleotide (Figure 2.4A). The lack of

activity is not due to a failure to bind CYT-19, because AMP-PNP inhibited ATP

stimulation with a KI of 200 µM (data not shown), 5-fold lower than the concentration

used to test for stimulation. Thus, these results suggest that AMP-PNP binding does not

elicit the same conformation as ATP, a conclusion that is strongly supported by recent

work using a series of nucleotide analogs (110).

Page 52: Copyright by Jeffrey Philip Potratz 2012

36

2.3.2 Increased ATP requirement for longer or more stable duplexes

Regardless of the mechanism of strand separation, it would be expected that more

ATP hydrolysis would be necessary for longer duplexes. An increased ATP requirement

is further suggested for DEAD-box proteins by previous observations that longer

duplexes are separated with greatly reduced rates. However, to our knowledge, the

dependence of duplex length on the ATPase activity of DEAD-box proteins has not been

systematically investigated.

Therefore, we extended the 5´-end of the hybrid strand with uridines to form P1

duplexes of seven, nine, or eleven base pairs (see Figure 2.1A and Figure 2.7). As

expected, the ATP requirement increased with duplex length (Table 2.2). This increase

arose from a decreased strand separation rate, while the ATPase rate remained essentially

constant (Table 2.3). The insensitivity of the ATPase rate suggests a simple model in

which CYT-19 manipulates both the longer and shorter duplexes by using a single cycle

of ATP-dependent conformational changes to induce local strand separation, but for the

longer duplexes a fraction of these events do not lead to complete strand separation,

allowing re-zipping of the duplex after the core domain of CYT-19 dissociates (see

Figure 2.6 and refs (52, 61, 97). These experiments also established 11 base pairs as a

lower limit on the unwinding that can be accomplished using a single ATP, as this duplex

gave an ATP utilization value of 1.2 ± 0.1 (2 mM Mg2+

).

We next increased the duplex stability, without increasing the length, by changing

the natural G•U wobble pair within the P1 duplex to a G-C pair (Figure 2.7). The G•U

pair has been shown previously to destabilize the P1 helix, relative to the G-C, in part by

disrupting base stacking (111, 112). We tested the effects of this base pair in the context

of the 6-bp and 11-bp duplexes. Under standard conditions (10 mM Mg2+

), the ATP

utilization value for the 6 bp duplex increased 8-fold (from 1.1 ± 0.1 to 9 ± 3, Table 2.4),

Page 53: Copyright by Jeffrey Philip Potratz 2012

37

and in the context of the 11-bp duplex it increased 5-fold (from 11 ± 2 to 56 ± 11, see

section 2.5)§ Increases were also observed at lower Mg2+

concentrations, although the

changes were smaller (2–4-fold, Table 2.4). These results underscore the established links

between duplex stability and the efficiency of separation by DEAD-box proteins (59),

and they indicate that the less efficient unwinding for a more stable duplex does not

simply result from slower action by DEAD-box proteins, but is accompanied by an

increase in ATP consumption. This increase would not be expected for a conventional

helicase and, as above, may reflect ATP hydrolysis events that are non-productive

because the core domain of CYT-19 dissociates before strand separation is complete (see

Discussion Section 2.4).

A model involving non-productive ATPase cycles would suggest that conditions

that weaken CYT-19 binding or stabilize the duplex would give increased ATP

requirements. Although changes in experimental conditions invariably give complex

effects with multiple physical origins, changing Mg2+

concentration and temperature

generally conformed to these expectations. With increased Mg2+

concentration, strand

separation rates decreased substantially (Tables 2.1, 2.3, and 2.4) and the ATP utilization

values increased. These changes most likely include contributions from increased duplex

stability (31, 57) and weaker binding by CYT-19 (S. Mohr and A.M.L, unpublished

data). Lower temperatures also gave increased ATP requirements (Table 2.5),

presumably in part from increased duplex stability.

2.3.3 Similar ATP utilization by other DEAD-box proteins

To explore whether the insights obtained for CYT-19 extend to other DEAD-box

proteins, we tested the Saccharomyces cerevisiae proteins Mss116p and Ded1p. Like

CYT-19, these proteins separated a 9-base-pair version of the P1 duplex at low Mg2+

Page 54: Copyright by Jeffrey Philip Potratz 2012

38

concentrations with hydrolysis of at most a single ATP (Table 2.6). Further, both proteins

gave an ATP utilization of less than one at 2 mM Mg2+

, indicating that they, like CYT-

19, are capable of strand separation without ATP hydrolysis. With increasing Mg2+

concentration, the ATP utilization by each protein increased, analogous to the behavior of

CYT-19. Mss116p consistently hydrolyzed less ATP per separation event than did Ded1p

or CYT-19, perhaps reflecting tighter binding of Mss116p to single-stranded RNAs and

intermediates formed during strand separation (55).

2.4 DISCUSSION

By quantitatively comparing rates of ATPase activity and strand separation by the

DEAD-box protein CYT-19 and related proteins, we have measured the number of ATP

molecules hydrolyzed during separation of short helices under a range of solution

conditions and temperatures (summarized in Table 2.7). Although analogous

measurements have been made for several processive SF1 and SF2 helicases (reviewed in

ref. (113), few have been reported for DEAD-box proteins (52), and none for DEAD-box

proteins using short helices that are characteristic of structured RNAs. Our central

conclusions are, first, that DEAD-box proteins can separate short duplexes in a single

cycle of ATP-dependent conformational changes. Second, the process of strand

separation is initiated, and sometimes even completed, while the proteins remain in the

ATP-bound form. As described below, these insights provide critical new mechanistic

information on strand separation by DEAD-box proteins, supporting and extending

models for their action on physiological substrates (Figure 2.6).

2.4.1 Strand separation depends on ATP binding, not hydrolysis

Upon binding an RNA duplex, a conformational change is suggested to result in

tight binding of the protein to one strand of the RNA in a conformation that is

Page 55: Copyright by Jeffrey Philip Potratz 2012

39

incompatible with a duplex (Figure 2.6, top left). This conformational change depends on

ATP binding but not hydrolysis, consistent with prior findings of cooperativity between

binding of ATP and ssRNA (44, 45). It may induce local strand separation or trap a

single-stranded segment that emerges due to transient ‘breathing’ of the duplex. After this

initial separation, which is suggested from structural analysis to be limited to five or six

base pairs (38, 39, 41), additional base pairs can apparently dissociate spontaneously to

allow complete separation of duplexes up to at least nine base pairs in the absence of any

ATP hydrolysis (pathway shaded blue in Figure 2.6). The existence of this pathway is

indicated by the ATP utilization values of less than one.

The conclusion that nucleotide-dependent unwinding does not require hydrolysis

is strongly supported by the finding from Jankowsky and colleagues that the non-

hydrolyzable analog ADP-BeFx also promotes unwinding (110), and our demonstration

that ATP gives this enhancement strongly supports the conclusion that ADP-BeFx

provides a good approximate model for the ATP-bound state of DEAD-box proteins

(110). Both studies also show that the enhancement by ATP is not mimicked by AMP-

PNP. The Jankowsky group reports that unwinding of the longer duplexes in their study

is undetectable with AMP-PNP, consistent with earlier work for several DEAD-box

proteins and indicating that any stimulation by AMP-PNP is much less than that of ATP

(79, 114, 115). Our work extends these results; because CYT-19-mediated separation of

the 6-bp duplex can be monitored in the absence of any nucleotide, we are able to show

that AMP-PNP provides no stimulation relative to this basal level. Whereas the lack of

detectable activity with AMP-PNP has been suggested to indicate a requirement for ATP

hydrolysis, the present results indicate that the defect arises at least in part from

differences between the ATP-bound and AMP-PNP-bound states. A similar suggestion

was made previously for the eIF4A protein from differences between ATP and AMP-

Page 56: Copyright by Jeffrey Philip Potratz 2012

40

PNP in RNA binding and crosslinking experiments (44). These results cast doubt on the

general applicability of AMP-PNP as an analog of ATP for interactions with DEAD-box

proteins. Although AMP-PNP can give tight binding of ssRNA to DEAD-box proteins

(38, 45, 110, 116), this tight complex may not reflect an on-pathway intermediate for

ATP-dependent strand separation. Alternatively, it may be on-pathway but attained with

poor efficiency when starting from a double-stranded RNA and bound AMP-PNP.

2.4.2 What is the role of ATP hydrolysis?

Although some strand separation occurs in the absence of ATP hydrolysis, the

ATPase activities of all three DEAD-box proteins are substantially higher in the presence

of a duplex than with either strand alone (Figure 2.1B and data not shown). Thus,

interactions with the duplex stimulate ATP hydrolysis, suggesting that ATP is sometimes

hydrolyzed during strand separation. This ATP hydrolysis could occur from an

intermediate complex in which both strands remain present, in which case it could give

disruption of additional base pairs, or it could follow complete separation and function

solely to facilitate dissociation of the helicase core from the tightly bound strand to allow

additional cycles of unwinding (Figure 2.6 and ref. (110).

Although the precise role remains an open question, it is tempting to suggest that

ATP can be hydrolyzed prior to complete strand separation, as such premature hydrolysis

could account for the observations by us and others that separation of longer duplexes is

accompanied by hydrolysis of more than one ATP (52). If ATP were hydrolyzed prior to

complete strand separation, premature dissociation of the helicase core could then allow

re-zipping of the duplex and ‘wasted’ ATP hydrolysis (boxed pathway in Figure 2.6).

Supporting this interpretation, the ATPase rate remains constant as the length of the

duplex increases, whereas the strand separation rate decreases (52), suggesting that the

Page 57: Copyright by Jeffrey Philip Potratz 2012

41

steps up to and including ATP hydrolysis are not affected by length, but that the longer

duplexes are then less likely to become completely separated. It should be noted that

alternative models for increased ATP utilization are possible, at least for longer duplexes

where the participation of multiple functional units of protein could be imagined.

However, even a duplex as short as 6-bp, which is almost certainly bound by only one

monomer (38), can have an ATP requirement exceeding unity (Table 2.4). These results

lead us to favor the non-productive cycles shown in Figure 2.6 as a central origin of the

increased ATP requirements.

Notably, both the ATP-hydrolysis-dependent and -independent pathways can give

complete strand separation in a single cycle, yielding ATP utilization values of one or

lower and ruling out a general requirement for multiple cycles of ATP-hydrolysis-

dependent translocation along the duplex (see Section 2.2.11). Importantly, the same

general model could apply to tertiary contact disruption or protein displacement, as the

critical feature is that a single strand of RNA is bound tightly and prevented from

interacting with alternative partners.

2.4.3 Implications for physiological activities

These mechanistic features of DEAD-box proteins are likely to be critical for their

roles in manipulating structured RNAs and RNPs. A tethering interaction, which for

CYT-19 and Mss116p appears to be mediated by the C-terminal domain (83, 92),

positions the helicase core to disrupt nearby RNA structure. For CYT-19 and Mss116p,

this interaction is relatively non-specific, whereas for DEAD-box proteins that function

with a defined RNA or RNP, this interaction is likely to be specific (77, 79). For both

classes, it has been proposed that the tethering interaction may be maintained during local

strand separation (31, 79). Here we have tested this model and provide data in strong

Page 58: Copyright by Jeffrey Philip Potratz 2012

42

support of it. A single cycle of ATP-dependent conformational changes is sufficient to

give complete disruption of short helices, and such a cycle can easily be envisioned to

occur while the tethering interaction is maintained. This continued interaction may allow

DEAD-box proteins to disrupt the same local structure repeatedly, which may be

necessary to resolve misfolded RNAs such as group I introns (26), or to remain poised to

facilitate rearrangements of newly-formed intermediates that would otherwise revert

rapidly to non-productive structures. The same mechanism may also permit DEAD-box

and related proteins that function in such processes as pre-mRNA splicing to mediate

rapid and repeated interconversion of alternative sets of contacts, improving fidelity by

allowing sampling of alternative conformations and intermolecular contacts (33, 64, 117).

2.5 FOOTNOTES

*The CYT-19 concentration can be varied only over a limited range with our

current methods because at low concentrations the ATPase rate becomes too small to

measure the fraction of ATP hydrolyzed, and at high CYT-19 concentrations the strand

separation becomes too fast for hand pipetting. Across the accessible range (0.5–2 µM,

with CYT-19 in 4-fold excess of the duplex), the rate constant for strand separation

increased approximately linearly, indicating that CYT-19 is subsaturating with respect to

the duplex, and the ATP utilization value was unchanged within the expected limits of

uncertainty (data not shown).

‡This value arises from the strand separation rate in the presence of ATP under

conditions that do not favor its hydrolysis. The rate is 2–4 µM/min with 150 µM ATP,

with no indication of saturation (data not shown), and therefore at least 8 µM/min with

saturating ATP. ATP is hydrolyzed in only half of the strand-separation events. Thus, the

pathway that involves bound ATP but not its hydrolysis must give half of the total rate (4

Page 59: Copyright by Jeffrey Philip Potratz 2012

43

µM/min), ≥ 20-fold faster than CYT-19-dependent strand separation in the absence of

ATP.

§It is interesting that the G•U to G-C change increased the ATP requirement

approximately equally as a terminal pair (6-bp duplex) or an internal pair (11-bp duplex),

because G•U is more destabilizing as an internal pair in model duplexes (111). The

substitutions also gave similar decreases in the CYT-19-mediated unwinding rate, ~20-

fold. It is possible that the ‘extra’ destabilization from the internal G•U pair is reduced or

absent under our solution conditions (the previous study used 1 M NaCl). Alternatively or

in addition, the terminal G•U may generate local destabilization and a preferred entry

point for CYT-19, compensating for its smaller intrinsic effect on duplex stability.

Page 60: Copyright by Jeffrey Philip Potratz 2012

44

Figure 2.1: ATP hydrolysis and RNA strand separation by CYT-19

(A) Structure of the duplex substrate, derived from the P1 duplex of the Tetrahymena

group I intron. RNA nucleotides are red and DNA nucleotides are black. The DNA

portion has the equivalent sequence of the P2 helix, which is adjacent to P1 in the natural

intron. (B) P1 duplex-dependent ATP hydrolysis by CYT-19. ATP hydrolysis was

measured in the presence of the P1 duplex by including both oligonucleotides (), or

with the same concentrations of the RNA/DNA oligonucleotide (0.5 µM, ) or the

excess CCCUCUA5 (1 µM, ) alone. (C) Strand separation of the P1 duplex construct in

the presence () or absence () of CYT-19. Excess CCCUCUA5 (5 µM) was present to

prevent re-annealing of the 32P-labeled CCCUCUA5. An equivalent reaction with 1 µM

CCCUCUA5 gave the same rate constant within error (Figure 2.3), but the higher

concentration allowed more precision by increasing the extent of displacement of the

labeled CCCUCUA5. Experimental conditions for panels B and C were 25 °C, pH 7.0, 10

mM Mg2+

, 50 µM ATP-Mg2+

, and 2 µM CYT-19.

Page 61: Copyright by Jeffrey Philip Potratz 2012

45

Figure 2.2: Effects of a double-stranded extension on RNA strand separation and

ATPase activity

(A) Separation of the 6-base-pair P1 duplex. Rate constants are plotted against the CYT-

19 concentration for the P1 duplex alone (construct 7 in Figure 2.7, , 2.6 105 M

–1

min–1

) or with an adjacent helix composed of RNA (◊, 3.2 106 M

–1 min

–1) or DNA (,

4.4 106 M

–1 min

–1) with 2 mM ATP-Mg

2+ (constructs 8 and 1, respectively, in Figure

2.7). The enhancement in strand separation activity conferred by the additional helix, 10–

20-fold relative to the P1 duplex alone, is the same within error whether the additional

helix is composed of DNA or RNA. (Data for constructs 7 and 8 are reprinted from ref.

(31) for comparison. All reactions were performed at 25 °C, pH 7.0, 10 mM Mg2+

). (B)

ATPase activity under identical solution conditions (25 °C, pH 7.0, 10 mM Mg2+

) in the

presence of 50 µM ATP-Mg2+

. Symbols are the same as above. After subtracting the

basal ATP hydrolysis rate, 0.05 µM/min, ATPase rates are 0.02 µM/min for P1 alone

(), 1.6 µM/min for P1 with an activating RNA helix (◊), and 0.6 µM/min for P1 with an

activating DNA helix (). The activating DNA helix gives less enhancement of ATP

hydrolysis by CYT-19 than the activating RNA helix. Further, the increase in ATPase

activity from the activating DNA helix (30-fold) is in the same range as the increase in

strand separation, suggesting that the increased ATPase activity arises from the increased

rate of strand separation of the attached P1 RNA helix, not from direct ATPase

stimulation by the DNA portion. (C) The DNA extension does not itself stimulate ATP

hydrolysis. Progress of ATP hydrolysis is shown in the presence of 0.5 µM of the

standard RNA/DNA chimeric oligonucleotide (present in constructs 1 and 5 of Figure

2.7, ), the same concentration of an oligonucleotide that includes only the DNA portion

(construct 9 in Figure 2.7, ), or CYT-19 alone ().

(Panels A and B are the work of Cindy Chen.)

Page 62: Copyright by Jeffrey Philip Potratz 2012

46

Figure 2.3: Strand separation by CYT-19 is independent of ‘chase’ CCCUCUA5

concentration (1–10 µM)

Progress curves for separation of the 6-bp P1 duplex construct (Figure 2.1A and construct

1 in Figure 2.7) are shown in the presence of 1 µM (, 1.61 min–1

), 2 µM (∆, 1.60 min–1

),

5 µM (, 1.60 min–1

), and 10 µM (, 1.41 min–1

) unlabeled CCCUCUA5 chase. An

equivalent reaction in the absence of CYT-19 is also shown (, 0.08 min–1

). Conditions

for all reactions were 25 °C, pH 7.0, 10 mM Mg2+

, 50 µM ATP-Mg2+

, and 0.5 µM of the

P1 duplex construct. The independence of the rate constant on the concentration of

unlabeled CCCUCUA5 indicates that CCCUCUA5 does not inhibit strand separation at

concentrations ≤10 µM. Thus, 1 µM CCCUCUA5 was used in ATPase measurements to

minimize the ATP hydrolysis arising from interactions of free CCCUCUA5 with CYT-19,

and 5 µM CCCUCUA5 was used in strand separation experiments to maximize the extent

of visible strand separation. Lower concentrations of chase are expected to give smaller

extents of visible strand separation, as shown here, because a substantial fraction of

labeled material remains in the duplex when equilibrium is reached. These fractions were

somewhat lower than calculated from the relative concentrations of chase CCCUCUA5

and the duplex, apparently because a fraction of the duplex dissociates during gel

electrophoresis (see Section 2.2.5).

(This figure is the work of Cindy Chen.)

Page 63: Copyright by Jeffrey Philip Potratz 2012

47

Figure 2.4: ATP-independent strand separation by CYT-19

(A) Separation of the 6-base-pair P1 duplex with the DNA extension (Figure 2.1A) in the

presence of 2 mM Mg2+

. The rate constants are 0.12 min–1

in the absence of CYT-19 (∆),

0.46 min–1

in the presence of CYT-19 alone (), 0.47 min–1

in the presence of CYT-19

and 1 mM AMP-PNP (), and 2.30 min–1

in the presence of CYT-19 and 50 µM ATP-

Mg2+

(). The CYT-19-dependent strand separation rate in the absence of ATP is

calculated by subtracting the CYT-19-independent rate constant and multiplying by the

duplex concentration: (0.46 min–1

– 0.12 min–1

) 0.5 µM = 0.2 µM/min. These results

indicate that CYT-19 possesses ATP-independent strand separation activity under these

conditions and that AMP-PNP does not enhance this activity. CYT-19 was bound to

AMP-PNP in our experiments, because AMP-PNP inhibits ATP-dependent strand

separation with a Ki value of 200 µM (data not shown), 5-fold lower than the

concentration used here. (B) Separation of the 6-base-pair P1 duplex under standard

conditions with 10 mM Mg2+

. The rate constants were 0.08 min–1

in the absence of CYT-

19 (∆), 0.09 min–1

for CYT-19 alone (), and 1.8 min–1

for CYT-19 with 50 µM ATP-

Mg2+

(). Under these conditions, no significant ATP-independent strand separation by

CYT-19 was detected.

(These figures are the work of Cindy Chen.)

Page 64: Copyright by Jeffrey Philip Potratz 2012

48

Figure 2.5: ATP hydrolyzed by CYT-19 per separation event of the 6-base-pair P1

duplex with low Mg2+

concentration (2 mM)

Open and filled circles show results from two identical experiments. The solid curve

shows the best fit by a model that includes stimulation of unwinding by bound ATP

without hydrolysis (see Section 2.2.6). Dashed curves show best fits to discarded models

in which only free CYT-19 gives ATP hydrolysis-independent strand separation (long

dashes) or in which ATP-bound CYT-19 can give strand separation without hydrolysis,

but with the same efficiency as nucleotide-free CYT-19 (short dashes).

(This figure is the work of Cindy Chen.)

Page 65: Copyright by Jeffrey Philip Potratz 2012

49

Figure 2.6: Model for duplex separation by DEAD-box proteins

Interactions are formed between the ATP-bound helicase core and the RNA (radiolabeled

RNA indicated by an asterisk), and a tethering interaction is formed adjacent to the core

by an ancillary site, as shown, or by an additional protomer (53). Concomitant with or

subsequent to initial binding by the helicase core, a conformation that is dependent on

ATP binding but not hydrolysis induces or captures local strand separation. Complete

strand separation can be achieved without ATP hydrolysis (left, shaded blue) or with

ATP hydrolysis (pathways down and to the right), which accelerates dissociation of the

helicase core and may induce additional strand separation. Premature dissociation of the

helicase core after ATP hydrolysis leads to a futile cycle (counterclockwise within box).

Throughout the entire strand separation process, the tethering interaction may remain

intact, as shown, allowing the protein to perform multiple cycles of structure disruption

on the same RNA without being lost to solution.

(Figure prepared by Cindy Chen.)

Page 66: Copyright by Jeffrey Philip Potratz 2012

50

Figure 2.7: Duplex constructs used for measurements of ATP utilization during

unwinding (top, constructs 1-6) and for control experiments (bottom,

constructs 7-9)

All constructs for ATP utilization measurements included the same DNA sequence

(black). The P1 duplex was increased in length by using different DNA/RNA hybrid

strands that included various numbers of uridine nucleotides at the 5-end as shown

(constructs 2-4 and 6). The all-Watson-Crick duplexes (5-6) were constructed by using

CCCUCCA5 instead of CCCUCUA5.

(Figure prepared by Cindy Chen.)

Page 67: Copyright by Jeffrey Philip Potratz 2012

51

[Mg2+

]

, mM

Total

ATPase rate

Duplex-

dependent

ATPase rate†

Total strand

separation rate

CYT-19-

dependent

strand

separation rate‡

ATP

hydrolyzed per

strand

separation

2 1.10 ± 0.03 0.50 ± 0.03 1.05 ± 0.05 0.97 ± 0.05 0.45 ± 0.04

5 1.25 ± 0.04 1.10 ± 0.05 1.25 ± 0.10 1.20 ± 0.10 0.90 ± 0.09

10 0.80 ± 0.05 0.74 ± 0.05 0.73 ± 0.05 0.69 ± 0.05 1.1 ± 0.1

20 0.13 ± 0.01 0.11 ± 0.01 0.10 ± 0.01 0.05 ± 0.01 2.0 ± 0.4

Table 2.1: ATP utilization for CYT-19-mediated separation of the 6-bp P1 duplex

Reactions were performed at 25 °C, 50 mM Na-MOPS, pH 7.0, 50 µM ATP-Mg2+

, 0.5

µM duplex, 1.2 µM total CCCUCUA5, 2 µM CYT-19. All rates are µM/min. Values are

averages and standard deviations from two to four independent determinations.

†Values are the total ATPase rate minus the ATPase rate in the presence of the

approximate concentration of excess CCCUCUA5 expected to be single-stranded (0.7 – 1

µM). This difference represents the rate of ATPase activity arising from CYT-19

interacting with the duplex. The background ATPase activity was measured in the

presence of 1 µM CCCUCUA5 to allow for the possibility of incomplete duplex

formation, and control reactions established that the background rate depends only

weakly on CCCUCUA5 concentration between 0.7 and 1 µM (<20% at 2 mM Mg2+

and

less than 10% at 10 mM Mg2+

; data not shown).

‡Values are the observed rate constant for strand separation minus the basal strand

separation rate constant in the absence of CYT-19, with the difference multiplied by the

duplex concentration (0.5 µM) to give a steady-state rate.

(This table is the work of Cindy Chen.)

Page 68: Copyright by Jeffrey Philip Potratz 2012

52

[Mg2+

],

mM 6 bp 7 bp 9 bp 11 bp

2 0.45 ± 0.04 0.4 ± 0.1 1.1 ± 0.1 1.2 ± 0.1

5 0.90 ± 0.09 1.3 ± 0.4 1.4 ± 0.4 4.4 ± 0.7

10 1.1 ± 0.1 1.3 ± 0.2 5 ± 2 11 ± 2

20 2.0 ± 0.4 2.0 ± 0.4 14 ± 2 18 ± 2

Table 2.2: Dependence of ATP utilization by CYT-19 on Mg2+

concentration and

duplex length

Values are ATP hydrolyzed per duplex separated (ATP utilization value). Longer

duplexes were derived from that shown in Figure 2.1A by extending the 5´-end of the

RNA/DNA strand with uridine nucleotides. All reactions were performed at 25 °C, 50

mM Na-MOPS, pH 7.0, 50 µM ATP-Mg2+

, 0.5 µM duplex, and 2 µM CYT-19.

(This table is the work of Cindy Chen.)

Page 69: Copyright by Jeffrey Philip Potratz 2012

53

duplex

length

[Mg2+

],

mM

Total ATPase

rate

Duplex-dependent

ATPase rate†

Total strand

separation rate

CYT-19-

dependent strand

separation rate‡

ATP

hydrolyzed per

strand

separation

7 bp¶ 2 1.10 ± 0.05 0.65 ± 0.05 1.8 ± 0.3 1.8 ± 0.4 0.4 ± 0.1

5 1.6 ± 0.3 1.4 ± 0.5 1.2 ± 0.3 1.1 ± 0.3 1.3 ± 0.4

10 0.9 ± 0.2 0.9 ± 0.3 0.58 ± 0.02 0.55 ± 0.04 1.3 ± 0.2

20 0.12 ± 0.03 0.10 ± 0.03 0.07 ± 0.02 0.038 ± 0.003 2.0 ± 0.4

9 bp 2 2.0 ± 0.4 1.7 ± 0.4 1.4 ± 0.2 1.4 ± 0.2 1.1 ± 0.1

5 2.1 ± 1.0 2.1 ± 1.3 1.3 ± 0.4 1.3 ± 0.4 1.4 ± 0.4

10 1.3 ± 0.4 1.2 ± 0.4 0.28 ± 0.07 0.27 ± 0.08 5 ± 2

20 0.30 ± 0.03 0.25 ± 0.10 0.020 ± 0.002 0.014 ± 0.002 14 ± 2

11 bp 2 2.0 ± 0.2 1.7 ± 0.3 1.8 ± 0.5 1.8 ± 0.6 1.2 ± 0.1

5 3.2 ± 0.2 3.00 ± 0.05 0.9 ± 0.4 0.9 ± 0.4 4.4 ± 0.7

10 1.7 ± 0.2 1.6 ± 0.3 0.15 ± 0.05 0.14 ± 0.05 11 ± 2

20 0.15 ± 0.05 0.10 ± 0.03 0.006 ± 0.001 0.005 ± 0.001 18 ± 2

Table 2.3: Dependence of strand separation and ATPase rates on duplex length

Reaction conditions were 25 °C, 50 mM Na-MOPS, pH 7.0, 0.5 µM duplex, 2 µM CYT-

19. Rates are µM/min. Values are averages and standard deviations from at least three

independent determinations. Duplexes are as shown in Figure 2.1A, with longer versions

achieved by extending the 5´-end of the chimeric RNA/DNA oligonucleotide with

uridine nucleotides (constructs 2–4 in Figure 2.7).

†Duplex-dependent ATPase rate is the total ATPase rate minus the ATPase rate in the

presence of CYT-19 and the approximate concentration of CCCUCUA5 expected to be

present as ssRNA (1 µM).

‡CYT-19-dependent strand separation rate is the strand separation rate constant minus

the basal dissociation rate constant in the absence of CYT-19, both multiplied by the

duplex concentration.

¶Although this construct has sequence potential to form a 7-bp duplex (compare

constructs 1 and 2 in Figure 2.7), recent work using the same duplex extension in the

context of the intact Tetrahymena ribozyme suggests that the single-base extension does

not form an additional base pair (111). Consistent with this interpretation, the unwinding

rates are the same within error for this construct and the 6-bp duplex construct that lacks

this single-nucleotide extension (Table 2.1).

(This table is the work of Cindy Chen.)

Page 70: Copyright by Jeffrey Philip Potratz 2012

54

Table 2.4: RNA strand separation and ATP hydrolysis rate for duplexes composed

solely of canonical Watson-Crick base pairs

Reaction conditions were 25 °C, 50 mM Na-MOPS, pH 7.0, 50 µM ATP-Mg2+

, 0.5 µM

duplex, and 2 µM CYT-19. Rates are µM/min. Values represent averages and standard

deviations from two independent determinations.

†The duplex constructs are the same as shown in Figure 2.1A, except that the P1 portion

of the constructs used here includes a G-C base pair instead of the natural G•U pair

shown in Figure 2.1A. The 6-bp and 11-bp duplex constructs are numbered 5 and 6,

respectively, in Figure 2.7.

‡The values shown are the ATP utilization values (ATP hydrolyzed per strand separation,

second-to-right column) divided by the corresponding values for the duplex of the same

length that includes the natural G•U pair instead of a G-C pair.

(This table is the work of Cindy Chen.)

Duplex

length†

[Mg2+

]

, mM

Total

ATPase

rate

Duplex-

dependent

ATPase rate

Total strand

separation rate

CYT-19-

dependent

strand

separation rate

ATP

hydrolyzed per

strand

separation

ATP

utilization

rel. to G•U

duplex‡

6 bp 2

1.6 ±

0.5 1.4 ± 0.4 1.3 ± 0.5 1.3 ± 0.5 1.2 ± 0.3 2.7 ± 0.7

5

2.2 ±

0.8 2.1 ± 0.8 0.9 ± 0.5 0.9 ± 0.5 2.9 ± 0.7 3.2 ± 0.8

10

0.31 ±

0.11 0.28 ± 0.10 0.035 ± 0.01 0.032 ± 0.01 9 ± 3 8.2 ± 2.8

11 bp 2

1.4 ±

0.6 1.3 ± 0.5 0.29 ± 0.11 0.29 ± 0.11 4.3 ± 1.5 3.6 ± 1.3

5

1.7 ±

0.7 1.6 ± 0.7 0.10 ± 0.05 0.10 ± 0.06 17 ± 7 3.9 ± 1.8

10

0.35

±0.06 0.31 ± 0.06 0.006 ± 0.0007 0.006 ± 0.0007 56 ± 11 5.1 ± 1.4

Page 71: Copyright by Jeffrey Philip Potratz 2012

55

T (ºC)

Total

ATPase rate

Duplex-

dependent

ATPase rate

Total strand

separation rate

CYT-19-

dependent

strand

separation rate

ATP hydrolyzed

per strand

separation

15 0.22 0.19 0.006 0.005 38

25 0.49 0.45 0.036 0.033 14

37 0.18 0.13 0.21 0.14 0.93

Table 2.5: Temperature dependence of ATP utilization by CYT-19 during strand

separation

Reaction conditions were 50 mM Na-MOPS, pH 7.0 (at 25 °C), 0.5 µM duplex, 2 µM

CYT-19, 10 mM Mg2+

, and 50 µM ATP-Mg2+

. Rates are µM/min. Experiments

monitored separation of the 6 bp duplex that included a G-C base pair instead of the

natural G•U pair (construct 5 in Figure 2.7). A reaction at 25 °C, whose results are

included here, was performed in parallel with reactions at 15 °C and 37 °C, and gave

similar, but not identical, results to those shown in Table 2.4 from independent,

equivalent experiments.

(This table is the work of Cindy Chen.)

Page 72: Copyright by Jeffrey Philip Potratz 2012

56

[Mg2+

], mM Mss116p Ded1p CYT-19

2 0.5 ± 0.1 0.52 ± 0.09 1.1 ± 0.1

5 0.8 ± 0.3 1.5 ± 0.9 1.4 ± 0.4

10 0.8 ± 0.2 1.9 ± 0.5 5 ± 2

20 2.6 ± 0.5 6 ± 3 14 ± 2

Table 2.6: ATP utilization by the DEAD-box proteins Mss116p and Ded1p

Values are ATP hydrolyzed per duplex separated. Separation of a 9-bp duplex was

monitored (construct 3 in Figure 2.7). Conditions were identical to experiments for CYT-

19 except that lower concentrations of proteins, substrates, and ATP were used to

compensate for increased activity of these proteins. For experiments with Mss116p, the

protein concentration was 100 nM, the duplex concentration was 25 nM, and the ATP

concentration was 0.5–100 µM, depending on the Mg2+

concentration. For experiments

with Ded1p, the protein concentration was 100–500 nM, the duplex concentration was

25–125 nM (maintaining a constant ratio of Ded1p and duplex concentrations), and the

ATP concentration was 1–200 µM. In all cases, the ATP utilization values reported are

saturating with respect to ATP concentration (data not shown). Corresponding values for

CYT-19 are reproduced from Table 2.2 for comparison.

(Pilar Tijerina contributed data used in this table.)

Page 73: Copyright by Jeffrey Philip Potratz 2012

57

Variable Range Other conditions*

CYT-19 concentration 0.5 – 2 µM 10 mM Mg2+

, duplex 1 (6 bp) (not shown)

ATP concentration 1 – 150 µM

5 – 150 µM

2 mM Mg2+

, duplex 1 (6 bp), Figure 2.5

10 mM Mg2+

, duplex 1 (6 bp) (not shown)

All ATP concentrations tested are subsaturating for

CYT-19 binding, based on a previous determination

of KM of 200 µM (28) and our results that strand

separation rate increases with ATP concentration

across the range tested (data not shown).

Mg2+

concentration 2 – 20 mM

2 – 10 mM

Duplexes 1–4 (6–11 bp), Tables 2.1 and 2.3

Duplexes 5 and 6 (6 and 11 bp), Table 2.4

Duplex length 6 – 11 bp G•U background (duplexes 1-4), 2–20 mM Mg2+

,

Tables 2.1 and 2.3

G-C background (duplexes 5 and 6), 2–10 mM Mg2+

,

Table 2.4

Duplex stability G•U or G-C 2, 5, and 10 mM Mg2+

, 6 and 11 bp (duplexes 5 & 6),

Tables 2.1, 2.3, and 2.4

Temperature 15 – 37 °C Duplex 5 (6 bp), 10 mM Mg2+

, Table 2.5

Protein identity CYT-19,

Mss116p

Ded1p

See above

Duplex 3 (9 bp), 2 – 20 mM Mg2+, 100 nM

Mss116p,

0.5 – 100 µM Mg2+

, Table 2.6

Duplex 3 (9 bp), 2 – 20 mM Mg2+, 100 – 500 nM

Ded1p,

1 – 200 µM Mg2+

, Table 2.6

Table 2.7: Range of conditions for measurements of ATP utilization

* Duplex numbering refers to Figure 2.7.

(This table is the work of Cindy Chen.)

Page 74: Copyright by Jeffrey Philip Potratz 2012

58

The text below and following figures of chapter 3 were published originally by Elsevier.

Copyright © by Elsevier Inc. 2012. All rights reserved.

Potratz, J.P. and Russell, R. (2012) RNA catalysis as a probe for chaperone activity of

DEAD-box helicases. Methods in Enzymology. 511, 111-130.

Rick Russell helped organize the paper content.

Chapter 3: RNA catalysis as a probe for chaperone activity of DEAD-

box helicases

3.1 CATALYTIC ACTIVITY AS A PROBE OF RNA FOLDING

To track productive folding of an RNA, a signal must exist that allows the native

state of the RNA to be distinguished from all other conformations. Functional assays

provide a powerful probe, because even folding intermediates with extensive native

structure can be readily distinguished if they are unable to function. Catalytic RNAs are

well suited for this purpose, as their native states can be readily detected by monitoring

chemical conversion of a substrate to its corresponding product.

3.1.1 Catalytic activity distinguishes the native state from all other conformations

The central goal of using catalytic activity to monitor RNA folding is to measure

the fraction of the RNA that is present in the native state unambiguously and

quantitatively. This information can then be combined with other approaches to gain

insights into the structural properties of folding intermediates. In the case of the ribozyme

derived from a group I intron in Tetrahymena thermophila, a misfolded conformation

exists that contains the full set of native secondary and tertiary contacts and is difficult to

distinguish from the native state by physical approaches (17, 118). However, it is

straightforward to distinguish between the two structures in a catalytic activity assay

because the natively folded ribozyme, but not the misfolded ribozyme, can cleave an

Page 75: Copyright by Jeffrey Philip Potratz 2012

59

oligonucleotide substrate in trans. Using this assay, the transition from the misfolded state

to the native state has been measured (17, 119-121).

3.1.2 Catalytic activity can be used to study chaperone-assisted folding

The ability to track native RNA folding allows the influence of chaperone

proteins to be monitored. While this chapter highlights group I and group II introns and

the DEAD-box helicase chaperone proteins, the general techniques can be applied to a

broad range of catalytically active RNAs and chaperone proteins. Indeed, RNA catalytic

activity of the hammerhead ribozyme was used nearly two decades ago to study

chaperone activities of the HIV NC protein (122, 123), and group I introns have been

used to monitor in vitro and in vivo chaperone activities of a variety of RNA binding

proteins (90, 124-126).

Group I and group II introns are mobile genetic elements that catalyze their own

excision from precursor RNA via two transesterification splicing reactions (Figure 3.1A).

In order for these splicing events to occur, the RNA must fold to an active three-

dimensional conformation. Introns can be converted to ribozymes with constructs that

lack the exons (Figure 3.1B,C). Ribozymes of group I and group II introns cleave

oligonucleotide substrates that include a 5´ splice site in reactions that mimic the first step

of splicing (99, 127-131).

Group I and group II introns and their corresponding ribozyme constructs are well

suited for studying chaperone-assisted RNA folding because they have extensive

networks of secondary and tertiary structures (Figure 3.1B,C), and RNAs from both

groups have been shown to fold slowly and to accumulate intermediates (132-140). These

RNAs are sufficiently complex to include diverse sets of intermediates and corresponding

kinetic barriers during folding, allowing detailed probing of the abilities of chaperones to

Page 76: Copyright by Jeffrey Philip Potratz 2012

60

assist in overcoming these barriers, while remaining simple enough to deeply probe the

folding processes and pathways.

3.2 SELF-SPLICING AS A READOUT FOR NATIVE STATE FORMATION

A straightforward way of measuring native folding is to follow self-processing of

intron-containing constructs after folding is initiated by addition of Mg2+

. The reaction

can be followed in the format of a continuous assay, in which folding and the catalytic

steps take place concurrently and in the same reaction. If the rate of native RNA structure

formation is slower than the subsequent catalytic steps, the observed splicing rate

provides a good measure of the rate of folding to the native state.

It is common practice to body label the RNA to visualize all the products of intron

splicing. This can be accomplished by including a radiolabeled nucleotide, commonly [α-

32P] UTP, to label nucleotides throughout the RNA (30, 91, 141). The reaction is then

monitored by denaturing polyacrylamide gel electrophoresis, which allows the unspliced

precursor to be separated from the splicing products (Figure 3.2A). The loss of precursor

RNA over time is quantitated to indicate the observed rate of splicing, and in turn folding

to the native state (30, 81, 142) (Figure 3.2B).

3.2.1 Interpreting chaperone-promoted changes in observed splicing rate

A simple reaction scheme for intron folding and splicing is shown below and is

instructive for understanding how a chaperone protein affects the splicing process.

Scheme 1

Page 77: Copyright by Jeffrey Philip Potratz 2012

61

To identify whether an RNA helicase influences folding, the observed rate

constant (kobserved) is compared in the absence and presence of multiple protein

concentrations and in the presence and absence of ATP. If the helicase increases the rate

of folding to the native state (kfold), most likely by accelerating resolution of one or more

intermediates (I), this increase will lead to an increase in kobserved if the splicing rate

constant (ksplice) is larger than the folding rate constant. For introns with robust self-

splicing activity this condition is generally met because large RNAs usually fold slowly

in vitro (143, 144).

The observed rate constant is typically plotted as a function of protein

concentration to determine the efficiency with which the chaperone promotes folding

(Figure 3.2C). The observed rate constant may increase with protein concentration at low

concentrations and then level or begin to decrease at higher protein concentrations

(Figure 3.2C). Possible physical sources for this inhibition are trapping of intermediate

conformations in a protein-bound state (shown in Scheme 1 as a single intermediate (I)

for simplicity) and chaperone-induced unfolding of the native state. Although unfolding

of native RNAs may be physiologically important for RNAs that must cycle between

functional structures, this inhibition complicates analysis of the effects of the chaperone

on the folding process toward the native state. Thus it is most straightforward to interpret

the data in terms of protein-accelerated folding at relatively low protein concentrations

where these additional effects are minimized (Figure 3.2C).

3.2.2 Potential complications

Since the chaperone is present during the splicing reaction, it could affect the

catalytic steps of splicing (ksplice) in addition to the folding steps. If the catalytic steps

(ksplice) are rate-limiting, influencing them will alter the overall rate constant (kobserved)

Page 78: Copyright by Jeffrey Philip Potratz 2012

62

even if the protein does not affect the rate of native structure formation (kfold). However,

this will not be an issue provided that folding remains slower than splicing. Group I

introns allow the rate-limiting step to be determined. A folding incubation is performed in

the absence of the exogenous guanosine cofactor and then guanosine is added to permit

splicing. If this reaction gives a larger splicing rate than the reaction in which guanosine

was present continuously, it indicates that folding is rate limiting.

A second issue concerns the multi-step nature of the splicing process. Although it

is shown as a single step for simplicity in scheme 1, some group I and group II introns

display rapid, reversible first steps of splicing, giving accumulation of intermediates in a

fast phase that is followed by slower completion of the second step and formation of

products (145-148). If native folding is slower than the reversible first step but faster than

the second step, the precursor will be lost with an initial rate constant that reflects

folding, as in the simple case with a more rapid second step. A slow step will result in

additional loss of precursor. This loss could mistakenly be assigned to an additional

folding pathway that suggests slower folding of the precursor, whereas it actually reflects

completion of the second splicing step. This behavior can be identified by the appearance

and subsequent disappearance of the intermediates.

A final issue concerns the choice of RNA constructs. Although emphasis is

typically on the intron for folding, exon length and composition can cause substantial

effects (142, 149, 150). The exons may misfold themselves and/or stabilize structure

within the intron, and either of these effects may be important biologically. Care must be

taken when comparing results from constructs that differ in the properties of the exons.

Page 79: Copyright by Jeffrey Philip Potratz 2012

63

3.3 SUBSTRATE CLEAVAGE AS A READOUT FOR NATIVE STATE FORMATION

Many of the potential complications associated with using self-splicing constructs

in continuous assays can be avoided by using ribozyme versions. Lacking exons,

ribozyme constructs allow interpretation of all folding processes to reflect the intron

domains. Further, although ribozyme versions can be used in continuous assays, the fact

that they cleave oligonucleotide substrates in trans makes them well suited for

discontinuous assays, in which folding and catalytic activity are separated into two

discrete stages (Figure 3.3A). The main advantage of this separation is the ability to

directly assess the fraction of native ribozyme during folding. This allows productive

folding to be dissected and unfolding of the native structure to be probed, (see Section

3.4.1, below). In addition, the discontinuous assay permits the use of folding conditions

that do not support robust catalysis, an option not available using a continuous assay.

3.3.1 Setting up a discontinuous assay: folding and catalysis stages

The discontinuous assay is composed of two stages, the folding stage (stage 1)

and the catalysis (or cleavage) stage (stage 2) (Figure 3.3A). The folding stage contains

the ribozyme alone or with the chaperone under conditions desired for chaperone-assisted

RNA folding. To prevent catalytic activity at this stage, the oligonucleotide substrate is

omitted. In the catalysis stage, conditions are changed such that further folding to the

native state is blocked, and radiolabeled oligonucleotide substrate is added, typically in

small excess of the ribozyme (2-3-fold over the ribozyme). The cleavage reaction is

allowed to proceed for various times (t1), and the substrate and product from each time

point are separated on a denaturing polyacrylamide gel (Figure 3.3B). The fraction of

product, normalized by the substrate and ribozyme concentrations, is plotted against

cleavage time (t2) (Figure 3.3B,C). Most commonly, the cleavage stage produces a burst

of product formation, with the amplitude reflecting the fraction of ribozyme in the native

Page 80: Copyright by Jeffrey Philip Potratz 2012

64

state (see Section 3.3.2, below). This fraction increases as a function of time spent in the

folding stage (t1), giving a rate constant for native folding of the ribozyme (Figure 3.3D).

After initial experiments have been performed and the cleavage rate constant is known, a

single time point in the cleavage reaction (stage 2) may be sufficient to determine the

burst amplitude (see Figure 3.4A). This timepoint should be chosen after completion of

the burst phase, but before significant contribution from subsequent turnovers is seen.

A key advantage of the discontinuous assay is that any set of conditions can be

used for the folding stage. However, setting up the catalysis stage requires care, as

conditions must support enzymatic activity while blocking further native folding. This is

because ribozyme that reached the native state during stage 2 would produce cleavage

products and cause the calculated fraction of native ribozyme to be artificially high.

The ability of the catalysis stage to prohibit folding is probed by comparing

cleavage reactions from a ribozyme that has been prefolded to the native state and a

ribozyme that is transferred directly to the catalysis stage, omitting the folding step. (In

practice, testing for suitable conditions for the catalysis stage can be undertaken

simultaneously with learning how to prefold ribozyme to the native state, as described

below.) The burst amplitude of product from ribozyme that skipped the folding stage

should be much less than that from the prefolded ribozyme, indicating little formation of

native ribozyme during the cleavage reaction. It may not be possible to completely block

native folding in stage 2, but a relatively small fraction of native ribozyme produced in

this stage can be accounted for by normalization (142).

Although optimal conditions for stage 2 are likely to be different for different

catalytic RNAs, some general guidelines are applicable. A study using a ribozyme

engineered from the group II intron aI5γ from S. cerevisiae, D135 ribozyme (Figure

3.1C), used a high pH in the catalysis stage to accelerate cleavage, and high Mg2+

Page 81: Copyright by Jeffrey Philip Potratz 2012

65

concentration (100 mM) and low temperature (15 C ) to enhance the arrest of folding

(Figure 3.4A) (see Section 4.3.3) (142). Cleavage is not particularly fast under these

conditions (10-2

min-1

). However, slow cleavage is acceptable for the catalysis stage in a

discontinuous assay as long as cleavage is significantly faster than folding. In addition,

proteinase K is included at the catalysis stage to ensure that protein transferred from the

folding stage has no effect on substrate cleavage.

To compare a ribozyme prefolded to the native state with an unfolded ribozyme, it

must be known how to fold the ribozyme to the native state. Establishing how to fold the

ribozyme to the native state can be undertaken concurrently with exploring conditions to

use for the catalysis stage. Folding of a ribozyme (typically by adding Mg2+

) is initiated

and aliquots are removed at different times from the folding stage. These aliquots are

transferred to stage 2 conditions in the presence of a small excess of substrate. The

product burst amplitudes are plotted against folding time. When the burst amplitude no

longer increases as a function of folding time, the ribozyme has been folded to the native

state as fully as it can be under that set of conditions. Care must be taken to avoid

misinterpreting a slower phase of folding as a plateau indicating complete folding (Figure

3.4B).

3.3.2 Interpreting results from the catalysis stage

To make optimal use of the discontinuous assay, it is critical to interpret the burst

amplitude quantitatively in order to determine the fraction of native ribozyme. Selected

examples from work involving chaperones will be covered below, and a more thorough

guide to interpreting results from catalytic rate measurements was published last year

(151).

Page 82: Copyright by Jeffrey Philip Potratz 2012

66

The interpretation of the burst amplitudes depends on the relative rate constants of

different steps in the catalytic cycle, which is shown in Scheme 2.

Scheme 2

In general, the cleavage stage is performed under multiple turnover conditions,

and the fraction of product is normalized by the substrate and ribozyme concentrations,

giving the ratio or product to ribozyme. This ratio is plotted against cleavage time (t2)

(Figure 3.3B,C). The reaction products are bound by base-pairing, with or without

additional tertiary contacts, and for many ribozymes their dissociation rate constants

(kP5´release and kP3´release) are small and limit the rates of subsequent turnovers. In all, there

are three possible regimes: (1) both products are released quickly relative to the catalytic

step (kcleavage); (2) both are released slowly; (3) one is released quickly and the other

slowly (151). Regimes 2 and 3 result in bursts of product formation, and the amplitude of

this burst can be used to calculate the fraction of native ribozyme, as described in two

examples.

The D135 ribozyme was used in a catalytic reaction with three-fold excess

substrate. When performed with a ribozyme prefolded to the native state, the cleavage

stage resulted in a burst of product with an amplitude approximately equal to one

turnover of the ribozyme (Figure 3.4A) (see Section 4.3.3) (142). This is consistent with

regime 3, in which there is a burst of product formation with a rate constant equal to the

cleavage rate (kcleavage) and a subsequent linear phase dictated by the slow release of one

of the products. Because the 5´ product of the oligonucleotide substrate forms twelve

Page 83: Copyright by Jeffrey Philip Potratz 2012

67

Watson-Crick base pairs with the ribozyme, product release might be slow. Under this

reaction regime, the amplitude of the burst phase is equal to the fraction of D135

ribozyme folded to the native state.

The ribozyme engineered from a group I intron from Azoarcus evansii (Figure

3.1B) was used in a catalytic reaction with two-fold excess substrate (136). When the

ribozyme was prefolded to the native state, the cleavage stage resulted in a burst of

product with an amplitude approximately equal to half of the ribozyme concentration

(Figure 3.4B). This is consistent with regime 2, in which the slow release of both

products allows equilibrium between the cleavage (kcleavage) and ligation (kligation) to be

reached. The burst amplitude is smaller than the fraction of native ribozyme and must be

corrected by the value of the internal equilibrium to reveal the fraction of native

ribozyme.

3.3.3 Using the discontinuous assay to probe chaperone-assisted folding

When optimal conditions for stage 2 are established and the relationship of the

burst amplitude to the fraction of native ribozyme is understood, the discontinuous assay

can provide important insights into chaperone-assisted RNA folding. The most

straightforward experiment is to compare the folding reaction in the presence of various

concentrations of chaperone, plotting the fraction of native ribozyme against time (t1)

(Figure 3.3D). A chaperone may increase the rate of native state formation (31, 136), an

effect that most likely arises from accelerated resolution of one or more kinetically-

trapped intermediates. This interpretation is particularly clear for RNAs that are known to

misfold (31, 134, 136, 152). Further confirmation can be obtained by allowing the RNA

to misfold first and then adding the chaperone (31, 136).

Page 84: Copyright by Jeffrey Philip Potratz 2012

68

It is possible for a chaperone to increase the fraction of ribozyme that reaches the

native state rapidly without having a significant impact on the observed rate constant

(Figure 3.5A,B) (142). This effect may arise from an influence of the chaperone early in

folding, which decreases the probability of misfolding at later folding steps.

Alternatively, resolution of an intermediate can become sufficiently fast so that this

pathway is then indistinguishable from pathways that avoid the intermediate.

Analogous to effects on self-splicing constructs, higher protein concentrations can

inhibit folding of ribozymes to the native state (Figure 3.5C). The physical processes

responsible for inhibition are presumably the same for ribozymes and self-splicing

constructs, but the discontinuous assay allows the origins of the inhibition to be

distinguished. For self-splicing constructs, inhibition by trapping protein-bound non-

functional intermediates and by unfolding native RNA both lead to a decrease in the

observed splicing rate (Figure 3.2C). In a discontinuous assay unfolding of native

ribozyme decreases the endpoint of the folding curve (Figure 3.5C). Trapping of folding

intermediates, without unfolding of natively-folded RNA, would result in a decrease in

the rate of native ribozyme formation but not a decrease in the endpoint.

3.4 OTHER APPLICATIONS OF THE DISCONTINUOUS ASSAY

The discontinuous assay is amenable to a diverse set of experiments. It can be

used to monitor a decrease in the native ribozyme, and an additional stage can be

included to probe the role of ATP in chaperone-mediated folding (see Section 4.3.5) (55,

142). Further, the progress of native ribozyme formation obtained from the assay can also

complement insight from other powerful physical approaches, and thus provide a more

complete understanding of the action of RNA chaperones.

Page 85: Copyright by Jeffrey Philip Potratz 2012

69

3.4.1 Unfolding native structure

Many chaperones function non-specifically and are capable of disrupting the

native states of RNAs as well as misfolded states (26). For probing the mechanisms of

chaperone activity in structure disruptions, it can be very useful to monitor the native

state because it is relatively homogeneous and the structure may be known. In contrast,

folding intermediates may be heterogeneous and their structure poorly defined.

To monitor loss of the native ribozyme, the ribozyme is first pre-folded to the

native state, and then the chaperone protein is added. Even for general chaperones, the

level of activity may be reduced for the native structure because it is typically highly

stable, so it may be necessary to lower the Mg2+

concentration and/or to use relatively

high protein concentrations to detect net unfolding. A decrease in the fraction of native

ribozyme over time indicates that the chaperone protein has mediated at least partial

unfolding of the native ribozyme, giving intermediates that do not readily refold to the

native state upon transfer to the stage 2 conditions. To ensure that the loss of native

ribozyme reflects reversible unfolding, the protein should be proteolyzed and the

ribozyme again folded to the native state (Figure 3.5D and see Section 4.3.6) (142).

3.4.2 Integrating results with other methods

The materials and methods required to employ the use of RNA catalytic activity

are standard for many laboratories, and the ease of implementing this method makes it a

convenient tool for obtaining a kinetic view of the fraction of native ribozyme. Results

can be highly complementary to those from methods that provide physical information on

folding intermediates (see Chapter 5).

Three physical probes that have been used extensively for RNA folding studies

are chemical footprinting, small angle X-ray scattering (SAXS), and single-molecule

FRET. Time-resolved chemical footprinting can be performed with several different

Page 86: Copyright by Jeffrey Philip Potratz 2012

70

probes and provides a highly specific view of nucleotides that are engaged in secondary

or tertiary contacts during a folding process (153-158). Hydroxyl radical footprinting has

been particularly valuable for probing structured intermediates and elucidating folding

pathways (159-161). The orthogonal information provided by catalytic activity

measurements – how much of the ribozyme is in the native state – is tremendously

valuable because it can be used to place constraints on the folding pathways modeled

from footprinting data (161, 162). SAXS provides rich information on the overall size

and shape of RNA as it folds, which is highly complementary to footprinting (163, 164),

and again coordinated activity measurements under the same conditions can assist greatly

in constraining physical descriptions of intermediates (136, 161-163, 165). Last, single-

molecule FRET experiments are uniquely powerful for detecting and characterizing

intermediates that do not accumulate in bulk experiments (140, 166-170), and the

concurrent detection of catalytic activity can be used to tremendous advantage for

distinguishing the native state from folding intermediates (167-169, 171).

Page 87: Copyright by Jeffrey Philip Potratz 2012

71

Figure 3.1: Group I and group II introns

(A) Self-splicing reactions. Left panel, Group I intron splicing reaction. An exogenous

guanosine attacks the 5´ splice site in the first step, and the 5´ exon attacks the 3´ splice

site in the second step. Right panel, Group II intron splicing reaction. A bulged adenosine

near the 3´ end of the intron attacks the 5´ splice site, generating a lariat intermediate. The

5´ exon then attacks the 3´ splice site to ligate the exons together. (B) Secondary structure

of the Azoarcus group I intron ribozyme. The nine-nucleotide substrate is lowercase. The

cleavage site is indicated by a thin arrow, and two tertiary contacts are indicated by thick

arrows. (C) Secondary structure of the aI5γ group II intron from S. cerevisiae (131). The

domains shown in black are present in the D135 ribozyme. Tertiary interactions are

indicated by Greek letters. Interaction sites between exon and intron sequences are

indicated by the abbreviations IBS and EBS (Intron Binding Site and Exon Binding Site,

respectively). The 24-nt substrate is shown at the right and the cleavage site is indicated

by an arrow.

Page 88: Copyright by Jeffrey Philip Potratz 2012

72

Figure 3.2: Self-splicing constructs

(A) Denaturing gel showing splicing products for a group II intron: (from top to bottom)

lariat intron, unspliced precursor, linear intron, and spliced exons. (B) Simulated plot of a

splicing reaction showing the fraction of precursor as a function of time. The simulated

data are fit by a single exponential curve to obtain rate constants for splicing in the

presence (circles) and absence (diamonds) of protein. (C) Simulated plot showing how

the observed splicing rate varies with protein concentration. The rising linear portion of

the data (circles) is fit with a line to obtain a second order rate constant for chaperone-

accelerated folding. The plateau and decrease in rate (diamonds in gray area) reflect

inhibition by the chaperone at higher concentrations (see Section 3.2.1).

Page 89: Copyright by Jeffrey Philip Potratz 2012

73

Figure 3.3: The discontinuous assay

(A) Reaction schematic. The ribozyme folds in the first stage, and then it cleaves the

oligonucleotide substrate (S) in the second stage under reaction conditions that block

further native folding. (B) Denaturing gel showing the results of cleavage of a 5´ labeled

oligonucleotide substrate by native ribozyme. (C) Simulated plot of multiple cleavage

reactions representing different folding times prior to the cleavage reaction. The curves

with larger bursts represent longer folding times, giving greater accumulation of native

ribozyme. The amplitude values are shown with filled symbols. (D) Simulated plot

showing the fraction of native ribozyme (fN) plotted as a function of folding time (t1).

The burst amplitudes from simulated cleavage reactions in panel C are plotted as a

function of folding time. In this simulated scenario, the folding progress can be fit by a

single exponential function, giving a single rate constant for native state formation.

Page 90: Copyright by Jeffrey Philip Potratz 2012

74

Figure 3.4: Examples of catalytic reactions with the D135 and Azoarcus ribozymes

(A) Identifying conditions that block native folding in the catalysis stage of the

discontinuous assay. Comparing burst amplitudes resulting from both D135 ribozyme

placed into stage 2 conditions with (circles) or without (triangles) a prior incubation in

Mg2+

-containing buffer to allow prefolding to the native state. The burst is much smaller

without the preincubation, indicating that these conditions for stage 2 (pH 8.0, 100 mM

Mg2+

, 500 mM KCl, 15 °C) effectively block folding. The solid circle indicates how a

single time point can be sufficient to determine the fraction of native ribozyme if the

kinetics of cleavage are known. (B) Ribozyme prefolding to the native state. The burst

amplitudes from cleavage reactions of the Azoarcus ribozyme are smaller than they

would be for stoichiometric product formation, and further work showed that this results

from an equilibrium between substrate cleavage and ligation (136). Note also that the

bursts from ribozyme prefolded at 37 °C and 10 mM Mg2+

for 15-45 min are identical,

suggesting that 15 minutes is sufficient for complete folding (see Section 4.1). Panel A

reprinted from (142) with permission from Elsevier. Panel B adapted from (136). This

research was originally published in Journal of Biological Chemistry. Selma Sinan,

Xiaoyan Yuan, and Rick Russell. The Azoarcus Group I Intron Ribozyme Misfolds and

Is Accelerated for Refolding by ATP-dependent RNA Chaperone Proteins. JBC. 2011;

286:37304-37312. © the American Society for Biochemistry and Molecular Biology

Page 91: Copyright by Jeffrey Philip Potratz 2012

75

Figure 3.5: The discontinuous assay with the D135 ribozyme and the DEAD-box

helicase Mss116p

(A) Cleavage time courses initiated with ribozyme folded in the presence of Mss116p for

the times indicated. The increase of the burst amplitude with folding time (t1) indicates

productive folding to the native state. (B) Comparison of ribozyme folding in the

presence (diamonds) and absence (circles) of Mss116p. The protein changes the folding

profile from multi-phasic to a single exponential phase. (C) High Mss116p concentrations

inhibit the native conformation. At 500 nM (diamonds) and 1000 nM (inverted triangles)

Mss116p, the fraction of native ribozyme at steady state is lower than in the presence of

lower protein concentrations (gray symbols). (D) The native conformation of D135 can

be unfolded by Mss116p. Ribozyme prefolded to the native state (black/gray solid circle)

is divided into reaction tubes containing a high concentration of Mss116p (black circles)

or no Mss116p (gray circles). After approximately 10 min, the protein is proteolyzed and

the ribozyme is treated to allow it to refold to the native state (black and gray diamonds),

demonstrating that the loss of native ribozyme arises from unfolding and not an

irreversible process. Panels A-D adapted from (142) with permission from Elsevier.

Page 92: Copyright by Jeffrey Philip Potratz 2012

76

The text below and following figures of chapter 4 were originally published by Elsevier.

Copyright © by Elsevier 2011.

Potratz JP, Del Campo M, Wolf RZ, Lambowitz AM, & Russell R (2011) ATP-

Dependent Roles of the DEAD-Box Protein Mss116p in Group II Intron Splicing In Vitro

and In Vivo. J Mol Biol 411(3):661-679.

Mark Del Campo performed experiments with proteins and self-splicing constructs.

Rachel Wolf performed the in vivo experiments.

Mark and Rachel are both in the Alan. M. Lambowitz lab.

Rick Russell and Alan M. Lambowitz helped analyze the data and write the paper.

Chapter 4: ATP-dependent roles of the DEAD-box protein Mss116p in

group II intron splicing in vitro and in vivo

4.1 INTRODUCTION

Autocatalytic group I and group II introns have proven to be valuable model

systems for understanding RNA folding, structure, and function (14, 172-175). The

mitochondrial (mt) genome of S. cerevisiae encodes nine group I introns and four group

II introns, all of which require the DEAD-box protein Mss116p for efficient splicing in

vivo (91). These introns differ substantially in their structural features and global

architectures, suggesting that the roles played by Mss116p in their splicing reflect non-

specific interactions with RNA. Further, other DEAD-box proteins are able to mitigate

the defects from loss of functional Mss116p (30, 55, 81, 176). These proteins include the

cytoplasmic S. cerevisiae protein Ded1p, which functions in cellular compartments that

lack group I and II introns and therefore does not function naturally in folding of these

introns.

The mechanism by which DEAD-box proteins promote group I intron splicing

has been analyzed biochemically. Mss116p and its Neurospora crassa homolog CYT-19

Page 93: Copyright by Jeffrey Philip Potratz 2012

77

interact with cognate and non-cognate group I introns and accelerate conformational

transitions, including those from kinetically-trapped, misfolded intermediates to the

native states (26, 28, 31, 55, 177). These observations and the established propensity of

group I introns to misfold (14, 20, 178) have led to models in which the principal

function of DEAD-box proteins in group I intron folding is to disrupt structure non-

specifically, allowing misfolded intermediates additional opportunities to fold to the

native state (20, 25, 28, 81, 143).

Analogous models have been proposed for Mss116p and CYT-19 in group II

intron folding (30, 81, 82, 91, 175). Group II introns consist of six domains with a

complex set of local and long-range tertiary contacts that generate a functional structure

with an active site for splicing, which is in many cases stabilized by the binding of

specific proteins (175, 179-181). To understand how Mss116p functions, attention has

focused on the yeast mtDNA group II introns that are its natural substrates. Two of these

introns, aI1 and aI2, are closely related group IIA introns and encode maturase proteins,

which are required for structural stabilization during RNA splicing. In the absence of

Mss116p in vivo, unspliced precursor RNA accumulates in a complex with the maturase,

suggesting that Mss116p can function at a step after stable maturase binding,

hypothesized to be the resolution of kinetic traps (91). The other two yeast mt group II

introns, aI5 and bI1, are small subgroup IIB introns that do not encode maturases, and

their splicing is accelerated by Mss116p and CYT-19 in vivo and under near-

physiological conditions in vitro (30, 55, 91, 176). Although the specific folding steps

accelerated by Mss116p remain to be established, biochemical studies showed that

Mss116p functions on bI1 as an RNA chaperone, as it promotes ATP-dependent

formation of an active intron structure and is then dispensable for activity (30).

Page 94: Copyright by Jeffrey Philip Potratz 2012

78

The mechanisms of acceleration for the remaining yeast group II intron, aI5,

have been the subject of debate. This intron was the first for which self-splicing via lariat

formation was demonstrated (182) and thereafter has been used as a model for RNA

folding and catalysis (174). Early studies used a shortened version of the intron termed

D135, which lacks exons and some intron domains and functions as a ribozyme by

cleaving an RNA oligonucleotide substrate, and were done at elevated temperature and

high ion concentrations (42 °C, 100 mM Mg2+

, 500 mM KCl). Under these conditions,

hydroxyl radical footprinting and catalytic activity measurements suggested concerted

folding in 1-2 min (0.6 min–1

) (183, 184). The denaturant urea did not accelerate folding,

consistent with the absence of a rate-limiting kinetic trap.(183)

In contrast, aI5 splicing is accelerated by Mss116p and other DEAD-box

proteins in vitro under near-physiological conditions (30 °C, 8 mM Mg2+

, 100 mM KCl)

(30, 55, 60, 176). Under these conditions, a native gel shift assay showed that folding of a

D1356 ribozyme derivative to compact species is slower and more complex, giving as

many as four kinetic phases with rate constants spanning at least three orders of

magnitude (>1 min–1

to <10–3

min–1

) (137, 138). The multiple phases suggest multiple

pathways and rate-limiting steps, and the time scale of hours for major populations would

be surprising for intramolecular diffusive processes. Nevertheless, the compaction rates

of the slow, dominant phases are unaffected by urea, suggesting that if there are rate-

limiting kinetic traps, they do not require substantial regions of buried RNA to become

exposed to solvent in the transition state ensemble (137).

Starting with the conclusion that the folding of aI5γ is not rate-limited by the

resolution of kinetic traps, Solem et al. suggested that Mss116p and other DEAD-box

proteins promote splicing of aI5 without unwinding RNA by binding and stabilizing an

on-pathway intermediate required for intron compaction (176). A central piece of

Page 95: Copyright by Jeffrey Philip Potratz 2012

79

evidence was the behavior of an Mss116p mutant in which the SAT sequence of motif III

is mutated to AAA (abbreviated SAT/AAA) (185). This mutant was reported to promote

in vitro splicing of aI5 -wild-type efficiency but to be inactive in unwinding

of moderately-long RNA duplexes (12–18 bp) (176). A subsequent analysis, however,

revealed that this mutant retains residual ATP-dependent unwinding activity for shorter

RNA duplexes (6–10 bp), of a length more representative of the helices in group II

introns, and that the decrease in unwinding efficiency can be as large as the decrease in

splicing activity under the same solution conditions (8–28-fold) (60). Thus, this mutant

did not provide experimental support for the suggestion that Mss116p promotes aI5γ

splicing without unwinding RNA (60).

The experiments above used a standard construct of aI5, which includes the 887-

nt intron and relatively long exons (~300 nt, termed LE construct). Recently, the same

authors compared splicing of the LE construct and a construct containing short exons of

<30 nt (SE). They concluded that Mss116p functions on the LE construct as an ATP-

dependent RNA chaperone to resolve kinetic traps involving exons but promotes splicing

of the SE construct by binding and stabilizing an on-pathway intermediate (149). A key

basis for this conclusion was the finding that although the SAT/AAA mutant was

compromised in promoting splicing of the LE construct, it appeared to be as efficient as

wild-type Mss116p for the SE construct. Other work showed that Mss116p accelerates

compaction of an RNA that consists only of aI5 domain I (DI), (138) suggesting that this

step is accelerated during folding of the SE construct (138, 176). However, acceleration

of DI compaction is ATP-independent, whereas acceleration of splicing for the SE

construct, like the LE construct, requires ATP. This difference suggests either that one or

more other folding steps for the SE construct require ATP-dependent RNA unwinding or

that ATP hydrolysis is required to accelerate dissociation of Mss116p following an ATP-

Page 96: Copyright by Jeffrey Philip Potratz 2012

80

independent role (138, 149). Finally, recent single molecule studies with the D135

ribozyme revealed that Mss116p-promoted folding involves an initial ATP-independent

step, presumably compaction of DI, and at least one later ATP-dependent step (140). The

authors favored the interpretation that this later step involves dissociation of bound

Mss116p to permit further RNA folding but left open the possibility of ATP-dependent

RNA unwinding to resolve a kinetic trap.

To delineate the roles of Mss116p in aI5 splicing, here we compare Mss116p-

promoted folding of three constructs, all under the same near-physiological conditions. In

addition to the standard LE construct, with 5 and 3 exons of 293 and 321 nt,

respectively, (186) we use a version with greatly shortened exons of 28 and 15 nts (SE

construct), (138, 149, 150) and the D135 ribozyme (Figure 4.1) (131). We show that the

ability of Mss116p to promote folding is ATP-dependent for all of the constructs, and

SAT/AAA and other mutants that are deficient in RNA-unwinding activity are deficient

in promoting splicing of both the SE and LE constructs. Catalytic activity measurements

with the D135 ribozyme and additional splicing assays with the SE and LE constructs

indicate that a major role of ATP is to promote the disruption of RNA structure by

Mss116p. Last, the relative abilities of Mss116p mutants to support ATP-dependent

splicing and RNA unwinding in vitro correlate well with their abilities to support aI5γ

splicing in vivo in a strain that lacks all other mt group I and group II introns. Together

our results indicate that the physiological function of Mss116p in aI5 splicing includes a

critical role for ATP-dependent RNA unwinding to resolve inactive structures.

Page 97: Copyright by Jeffrey Philip Potratz 2012

81

4.2 MATERIALS AND METHODS

4.2.1 Recombinant Plasmids

pMAL-Mss116p, used to express Mss116p from E. coli for biochemical studies,

contains the Mss116p coding sequence (codons 37–664) with an in-frame N-terminal

MalE fusion cloned downstream of a tac promoter in the expression vector pMAL-c2t

(55, 81). pHRH197, used to express Mss116p in S. cerevisiae for genetic analysis,

contains the MSS116 gene with its endogenous promoter cloned in the centromere-

containing (CEN) plasmid vector pRS416 (92). Sequences encoding Mss116p mutations

(K158A, K158R, S305A/T307A) were introduced into these plasmids by Quikchange

mutagenesis (Stratagene). pJD20 encodes the long-exon (LE) construct downstream of a

phage T7 promoter in Bluescribe (186). The LE construct is a 1501-nt precursor RNA

that includes a 293-nt 5 exon, the 887-nt aI5 intron, and a 321-nt 3 exon. The 5 exon

consists of 20 nt derived from the vector (plus three guanosines from the T7 promoter)

and 270 nt of COX1 5-exon sequence, and the 3 exon consists of 291 nt of COX1 exon

sequence followed by 30 nt of vector sequence. Plasmid pUC19::aI5-SE, which encodes

the SE construct, was created by PCR amplification of pJD20 using the primers 5-

TAATACGACTCACTATAGGGACTTACTACGTGGTGGGAC-3 and 5-

TTGATAATACATAGTATCCCGATAGGTAGACC-3, which added a T7 promoter

(underlined). The resulting PCR product was re-amplified with the primers 5-

GCCCcatatgTAATACGACTCACTATAGGG and 5-

GGGCaagcttAATACATAGTATCCCGATAGG to add NdeI and HindIII sites

(lowercase), and cloned between the corresponding sites of pUC19. The SE construct is a

930-nt precursor RNA that includes a 28-nt 5 exon, the 887-nt aI5 intron, and a 15-nt 3

exon. The 5 exon consists of three guanosines from the T7 promoter and 25 nt of COX1

Page 98: Copyright by Jeffrey Philip Potratz 2012

82

5-exon sequence, and the 3 exon consists of 11 nt of COX1 exon sequence followed by

4 nt of vector sequence. pQL71 encodes the D135 ribozyme (see Figure 4.1B) (183).

4.2.2 RNA preparation

The LE and SE RNA constructs were transcribed in vitro from HindIII-digested

plasmids by T7 RNA polymerase in the presence of [α-32P]-UTP (Perkin Elmer) using a

Megascript kit (Ambion). RNA was isolated by phenol-chloroform extraction and size

exclusion chromatography using two consecutive G-50 columns. D135 RNA was

transcribed in vitro from HindIII-digested pQL71 using T7 RNAP and purified via an

RNeasy column (Qiagen). The RNA oligonucleotide substrate for D135

(CGUGGUGGGACAUUUUCGAGCGGU) was 5-end-labeled with [α-32P]-ATP

(Perkin Elmer) by using T4 polynucleotide kinase (New England Biolabs). RNA

concentrations were determined by specific activity of the [α-32P]-UTP precursor or

spectrophotometrically using extinction coefficients at 260 nm of 1.76 107 M

–1 cm

–1 for

the LE construct, 1.17 107 M

–1 cm

–1 for the SE construct, 5.86 10

6 M

–1 cm

–1 for the

D135 construct, and 2.36 105 M

–1 cm

–1 for the D135 substrate oligonucleotide.

4.2.3 Preparation of Mss116p

Wild-type and mutant versions of Mss116p were expressed and purified as

described (55, 60). After purification, protein was dialyzed overnight against storage

buffer solution (20 mM Tris-Cl, pH 7.5, 500 mM KCl, 1 mM EDTA, 1 mM DTT, 50%

glycerol), flash frozen, and stored at –80 °C.

4.2.4 Splicing reactions

Splicing reactions were performed in a thermal cycler in 20 or 50 µl (50 mM Na-

MOPS, pH 7.0, 100 mM KCl, 8 mM MgCl2, and 5% glycerol). The RNA concentration

was 20 nM for reactions shown in Figure 4.2 and 4.3, and it was 1 nM for reactions in the

Page 99: Copyright by Jeffrey Philip Potratz 2012

83

presence of Mss116p (Figure 4.4, 4.6, and 4.7). Control reactions at the lower

concentration in the absence of Mss116p gave rate constants within 2-fold of those at the

higher concentration (Figure 4.4A and data not shown). When indicated, reactions also

included 1 mM ATP, added as a stoichiometric complex with Mg2+

. RNA was heated

briefly in the absence of Mg2+

(92 °C, 1-2 min), cooled rapidly, and then splicing

reactions were initiated by adding RNA to a pre-incubated tube containing splicing buffer

solution or by adding RNA and then MgCl2. Portions of reactions were quenched at

various times by adding 4 µl of 100 mM EDTA to 2 µl of reactions in Figure 4.2 and 4.3

or by adding 5 µl of 50 mM EDTA, 0.1% SDS, 1 mg/ml proteinase K to 3 µl of reactions

in Figure 4.4, 4.5, 4.6, and 4.7. Splicing products were separated on a denaturing 4%

polyacrylamide gel and quantified using a phosphorimager and ImageQuant TL (GE

Healthcare). Rate constants were obtained by fitting the time-dependent decrease in

precursor RNA, relative to reaction products, either by an exponential decay or by a line

for slow reactions of the LE construct in the absence of Mss116p (Kaleidagraph, Synergy

Software). In the latter case the rate constants were inferred from the initial splicing rate.

Unless otherwise indicated, rate constants are reported as the average and standard error

from 2-4 independent determinations. In the presence of higher concentrations of

Mss116p, some splicing reactions included a lag phase. The lag was included in the

analysis by allowing the y-intercept of a single exponential equation to vary. Inclusion or

exclusion of the lag phase did not significantly affect the rate constant for the slower,

major phase of splicing. The origin of the lag is unclear, but its appearance at higher,

inhibitory Mss116p concentrations suggests that it originates from a molecular process

associated with inhibition by Mss116p.

Page 100: Copyright by Jeffrey Philip Potratz 2012

84

4.2.5 Discontinuous catalytic activity assay for D135 RNA folding

The discontinuous activity assay for monitoring D135 ribozyme folding consisted

of two stages. In stage 1, D135 ribozyme (50 nM) was allowed to fold for various times.

The ribozyme was first denatured at 90 °C for 1 min in buffer solution containing 50 mM

Na-MOPS, pH 7.0, and then folding was initiated by addition of D135 to a stage 1

reaction (30 °C, 50 mM Na-MOPS, pH 7.0, 100 mM KCl, and 8 mM MgCl2 unless

otherwise indicated). ATP was also present (1 mM) when indicated, added as a

stoichiometric complex with Mg2+

. In stage 2, substrate cleavage was monitored by

diluting portions of the stage 1 folding reaction 5-fold into a solution with higher Mg2+

concentration and pH and lower temperature (15 °C, 500 mM KCl, 100 mM MgCl2, 80

mM HEPES, pH 8.1, 1 mg/ml proteinase K). This solution also included the radiolabeled

oligonucleotide substrate (30 nM, 3-fold in excess of D135 ribozyme after dilution).

Time points (2 µl) were quenched by adding 4 µl of 100 mM EDTA, and the substrate

and product were separated in a denaturing 20% polyacrylamide gel.

Cleavage time courses were fit by an exponential phase followed by a linear

phase, with the amplitude of the exponential phase reflecting the fraction of native D135

ribozyme (134). After early experiments established the rate constants of the fast and

slow phases, the burst amplitude was determined from a single time point at 300 min

(119). This time was chosen because it allows completion of the fast phase while

minimizing the contribution from the slow phase. Burst amplitudes were normalized and

scaled to reflect the fraction of the population that was native at the time of transfer to

stage 2. First, the raw burst amplitude was corrected by subtracting an amount

representing the fraction of unfolded molecules that fold rapidly in stage 2 (~0.15 under

standard conditions). This value was then divided by the value reflecting full native

folding, determined in reactions with pre-folded ribozyme, after subtracting an equivalent

Page 101: Copyright by Jeffrey Philip Potratz 2012

85

amount (e.g. 1 – 0.15 = 0.85). Folding rate constants are reported as the average and

standard error from at least two independent determinations.

4.2.6 S. cerevisiae Northern hybridizations and immunoblotting

The S. cerevisiae wild-type strain 161/U7-aI5γ is a derivative of 161-U7 MATa

ade1 lys1 ura3, in which the mtDNA contains a single intron aI5 (91). In the isogenic

strain mss116Δ-aI5, the MSS116 gene was replaced by a kanr cassette and the resident

mtDNA was replaced by cytoduction with mtDNA containing only aI5 (91). Wild-type

and mutant versions of Mss116p with a C-terminal myc tag were expressed from the

CEN plasmid pHRH197, which carries a ura3 marker (92). Transformants of the

mss116Δ-aI5 strain containing the CEN-plasmids were selected by plating on 2% agar

plates containing 2% dextrose and Hartwell’s complete (HC) media (1X YNB solution,

1X HC dropout 6 amino acid solution, plus 1% each of lysine, tryptophan, and histidine;

0.1% adenine; and 2% leucine) lacking uracil (187).

For Northern hybridizations, cells were grown at 30 °C to O.D.600 of 1.0 – 1.6 in

Hartwell’s Complete liquid medium containing 2% raffinose and lacking uracil. RNA

was isolated as previously described (92). Samples containing 1.0 µg of RNA were

denatured by incubating with 5.6% glyoxal in 50% DMSO, 0.1 M NaPO4 at 65°C for 15

min and electrophoresed in a 1.5% agarose gel with RNA-grade 1X TAE (40 mM Tris-

acetate, pH 8, 1 mM EDTA) at 25°C. The gels were blotted onto a nylon membrane

(Hybond-XL, GE Healthcare) overnight, hybridized with a 5′-end-labeled DNA

oligonucleotide probe complementary to COX1 exon 6

(GAATAATGATAATAGTGCAAATGAATGAACC), and scanned with a

phosphorimager.

Page 102: Copyright by Jeffrey Philip Potratz 2012

86

For immunoblotting, cells were grown as above, and proteins were precipitated

with trichloroacetic acid as described, (188) except that the resulting protein pellets were

washed once with 0.5 ml of 1 M Tris base before being resuspended in 150 µl of SDS-

PAGE sample buffer. Samples (~60 µg of protein measured by O.D.280) were run in a

pre-cast 4-20% polyacrylamide gradient gel (BioRad) with 0.1% SDS in the running

buffer and transferred to a Sequi-BlotTM PVDF membrane (BioRad) using a BioRad

Criterion blotter apparatus. The blot was probed with anti-Mss116p guinea pig primary

antibody (1:5,000 dilution), (185) developed using an ECL Plus Western Blotting-kit (GE

Healthcare), and imaged using Kodak Biomax XAR film. To confirm equal loading of

protein samples, the PVDF membrane was stripped of antibodies using RestoreTM Plus

Western Blot Stripping Buffer (ThermoScientific) and stained with AuroDye Forte (GE

Healthcare) following the manufacturer’s directions.

4.3 RESULTS

4.3.1 Splicing of LE and SE constructs and Mss116p acceleration

To probe how exons influence the mechanisms of Mss116p-promoted folding of

the aI5γ intron, we first compared self-splicing of the LE and SE constructs. Under near-

physiological conditions in vitro, splicing was considerably faster for the SE construct,

consistent with a recent report using essentially the same constructs and conditions

(Figure 4.2A and Figure 4.3) (149). At 30 °C, the rate constant for the SE construct

determined from the disappearance of precursor and appearance of lariat and linear intron

was 1.4 (± 0.5) 10–3

min–1

, whereas splicing of the LE precursor was 80-fold slower

(1.8 (± 0.5) 10–5

min–1

). Both constructs spliced faster at higher temperatures, with the

LE construct displaying a greater temperature dependence (Figure 4.2B). Thus, one or

both long exons slow splicing and add to the enthalpic barrier, consistent with previous

Page 103: Copyright by Jeffrey Philip Potratz 2012

87

findings that exon sequences can strongly affect aI5γ (150). Figure 4.2B also

shows that the splicing rates of both constructs level and begin to decrease with

increasing temperature, suggesting that the active structures are unstable at higher

temperatures. It is notable that this transition occurs at higher temperature for the LE

construct, suggesting that the longer exons stabilize the native structure in addition to

slowing its formation.

We then measured Mss116p-accelerated splicing of the SE construct. Consistent

with a recent report, (149) the concentrations of Mss116p used previously for the LE

construct (60, 81) gave strong inhibition with the SE construct, and we therefore used

lower concentrations of Mss116p and RNA (1 nM RNA). For comparison, we performed

experiments with the LE construct under these conditions. Splicing of both constructs

was stimulated by low concentrations of Mss116p, just slightly in excess of RNA, in the

presence of ATP (Figure 4.4A and Figure 4.5). In general, the time courses were well

described by single rate constants, with endpoints of >90% for reactions that reached

completion within the observation time (Figure 4.4B). Both constructs required ATP for

maximal stimulation, although SE construct splicing was weakly stimulated in the

absence of ATP (Figure 4.4B). Analogous ATP-independent stimulation was observed

previously for group I intron splicing by Mss116p (55) and at a very low level for the

aI5γ LE construct by the E. coli DEAD-box protein SrmB (81).

As a quantitative benchmark for comparison, we measured the Mss116p

concentration dependences (Figure 4.4C). In earlier work with higher RNA and protein

concentrations, acceleration of the LE construct by Mss116p and other DEAD-box

proteins gave sigmoidal dependences, suggesting roles for multiple protomers, (81)

although the upward curvature was more obvious for the other proteins and had

previously gone undetected for Mss116p (60). Here, across the limited range of

Page 104: Copyright by Jeffrey Philip Potratz 2012

88

stimulatory protein concentrations, the dependence was approximately linear, giving a

second-order rate constant of 4.7 106 M

–1 min

–1. Although the possibility of a higher-

order dependence cannot be excluded, for simplicity we use the linear dependence for

comparison with the SE construct and Mss116p mutants (see Section 4.5).*

We found that Mss116p stimulates splicing of the SE construct at least as

efficiently as it stimulates splicing of the LE construct (>4.2 106 M

–1 min

–1, Figure

4.4C) (60). Only a lower limit could be determined for the SE construct because even at

the lowest attainable protein concentration, the splicing rate was not readily distinguished

from the maximal value. This plateau for the SE construct results from strong inhibition

of splicing by Mss116p at concentrations as low as 4 nM, considerably lower than for the

LE construct (Figure 4.4B,C and Figure 4.5) (60). Together, these results indicate that the

long exons cause or exacerbate slow folding, Mss116p stimulates splicing of both the LE

and SE constructs, and that maximal Mss116p stimulation of both constructs requires

ATP.

4.3.2 Mutants that are deficient in RNA-unwinding activity

To explore further the properties of Mss116p required for acceleration of aI5γ

splicing, we tested three mutants compromised in ATP-dependent RNA unwinding for

stimulation of splicing of the LE and SE constructs. The SAT/AAA mutant gave an

efficiency of 4.2 105 M

–1 min

–1 for the LE construct (Figure 4.6), 10-fold lower than

wild-type Mss116p (WT) and within the range of relative rate constants found previously

(60). For the SE construct, the efficiency for the SAT/AAA mutant was 8.9 105 M

–1

min–1

, at least 4- to 5-fold lower than WT (a minimum estimate because, as noted above,

we could only determine a lower limit for WT). Thus, the decrease in efficiency for the

SAT/AAA mutant with the SE construct is comparable and could be even larger than for

Page 105: Copyright by Jeffrey Philip Potratz 2012

89

the LE construct. Acceleration of the SE construct by Mss116p SAT/AAA is ATP-

dependent, with little or no ATP-independent acceleration up to at least 10 nM protein

(Figure 4.6B). Additionally, two mutants in motif I, K158A and K158R, which are

defective in ATP binding and hydrolysis and in RNA unwinding, (176) failed to stimulate

splicing of either the LE or SE construct at any concentration tested (≤15 nM; Figure

4.7).

These results show that mutations that decrease RNA-unwinding activity

commensurately reduce the ability of Mss116p to stimulate splicing of both the SE and

LE constructs. Our conclusion that the SAT/AAA mutant is similarly deficient in splicing

of both constructs disagrees with the conclusion but not the data from the earlier work

(149) (see Discussion Section 4.4). The similar requirements for the SE and LE

constructs could reflect that Mss116p accelerates a common folding step of these two

constructs or that different folding steps coincidentally give similar dependences on

Mss116p and ATP.

4.3.3 Two-stage, discontinuous catalytic activity assay for folding of D135 RNA

Previously, the folding of the D135 ribozyme was measured by catalytic activity

using a continuous assay, in which folding and substrate cleavage occur simultaneously,

(184) but this assay can give information on the folding rate only under conditions in

which the rate of substrate cleavage exceeds that of folding (e.g. 100 mM Mg2+

at 42 °C)

(183). To follow folding at lower Mg2+

concentrations, where catalytic activity is weaker,

we designed a two-stage, discontinuous activity assay (Figure 4.8A). Folding takes place

in stage 1, and then portions of the folding reaction are transferred at various times into a

second set of conditions (stage 2). This second set of conditions must allow substrate

cleavage by ribozyme that has already folded to the native state at the time of transfer but

Page 106: Copyright by Jeffrey Philip Potratz 2012

90

inhibit the remainder of the population from reaching the native state on the time scale of

the cleavage reaction (189).

Based on previous work on group I intron ribozymes, (17, 119, 121) we tested

whether low temperature and high Mg2+

concentration would provide suitable conditions

for stage 2. When D135 was prefolded to the native state under established conditions (42

°C, 100 mM Mg2+

, 20 min), then transferred to 15 °C with 100 mM Mg2+

, and a small

excess of substrate was added, an initial phase of substrate cleavage gave a rate constant

of 8.7 (± 0.4) 10–3

min–1

and an amplitude approximately equal to one turnover of the

ribozyme (Figure 4.8B). Cleavage of the remainder of the substrate was slower,

presumably because at least one of the cleavage products is released slowly from the

ribozyme and limits subsequent turnovers. The 5 portion of the substrate base pairs with

exon-binding site (EBS) 1 and 2 of the ribozyme and is therefore a strong candidate for

slow release.

Although a cleavage rate constant of ~9 10–3

min–1

indicates a rather slow

reaction at 15 °C, it is within the expected range considering that the reaction proceeds at

1 min–1

under the same solution conditions at 42 °C. We reasoned that 15 °C and 100

mM Mg2+

could be suitable for stage 2 if folding to the native state under these

conditions is even slower than substrate cleavage by the native ribozyme. Indeed, we

found that ribozyme that was transferred directly from buffer solution in the absence of

Mg2+

into stage 2 gave substantially less product formation (Figure 4.8B, termed

‘nonfolded’ control), indicating that most of the ribozyme did not fold to the native state

on the time scale of the cleavage reaction. The low temperature of stage 2 was critical, as

a parallel experiment in which stage 2 was 30 °C gave minimal difference between the

prefolded and nonfolded reactions (Figure 4.9). Nevertheless, even at 15 °C a small

amount of product (15%) appeared in the nonfolded control with roughly the same rate

Page 107: Copyright by Jeffrey Philip Potratz 2012

91

constant as product in the prefolded reaction (Figure 4.8B). This product formation most

likely reflects a subpopulation of the ribozyme that was able to fold rapidly to the native

state in stage 2. In subsequent experiments, we normalized the data to account for this

subpopulation (see Section 4.2.5).

We next measured the kinetics of native folding in stage 1 by incubating the D135

ribozyme for various times and then transferring portions into stage 2 and measuring the

burst amplitude from time courses of substrate cleavage (see Figure 4.8A). As an initial

control, we monitored folding under conditions similar to those used previously in the

continuous activity assay (42 °C, 100 mM Mg2+

). As expected, this experiment gave

essentially the same result as the continuous assay, with complete folding to the native

state occurring with a rate constant of 0.5 ± 0.1 min–1

(Figure 4.8C) (183). As indicated

above, data were normalized by the fraction of ribozyme that reached the native state

rapidly upon transfer from buffer solution in the absence of Mg2+

, which was larger when

the ribozyme was transferred from 42 °C (Figure 4.10).

We next measured folding of D135 at the near-physiological conditions used for

the LE and SE constructs. Under these conditions, folding was slower and more complex,

with at least two phases (Figure 4.8D). A minor fast phase gave a rate constant of 1.0 (±

0.2) min–1

, and a second phase gave a rate constant of 1.4 (± 0.6) 10–3

min–1

with an

apparent endpoint of ~50-60% native ribozyme. These results are crudely consistent with

those from a published gel-shift assay used to measure compaction of the related D1356

ribozyme under similar conditions, which gave two observed phases (0.16 min–1

and

0.006 min–1

), an unresolved fast phase (>1 min–1

), and a slower phase (~10–3

min–1

,

complete in 24 h) (138). It is not clear whether the apparent endpoint in our experiment

reflects equilibration of native and non-native forms or whether an unobserved third

phase would give still more native RNA at longer folding times. Regardless, the results

Page 108: Copyright by Jeffrey Philip Potratz 2012

92

show that under the near-physiological conditions, D135 folds in a complex, multi-phasic

process that differs from the single-phase folding at higher temperature and Mg2+

concentrations (183, 184, 190).

4.3.4 Acceleration of D135 folding by Mss116p

We next used the discontinuous activity assay to probe the effects of Mss116p on

D135 ribozyme folding under near-physiological conditions. Addition of Mss116p and

ATP with Mg2+

gave rapid accumulation of the native ribozyme (Figure 4.11A), reducing

the folding process to a single phase with a rate constant of 0.61 ± 0.07 min–1

and an

apparent endpoint of 0.76 ± 0.04 (Figure 4.11B). While this rate was slightly slower than

that for the minor fast phase without Mss116p, the major effect of Mss116p was to

increase the amplitude of the rapid native state formation. The rate constant did not

change systematically across the accessible range of Mss116p concentrations (Figure

4.12A), preventing determination of a second-order rate constant for acceleration.

However, as observed with the LE and SE constructs, higher Mss116p concentrations

were inhibitory. In contrast to the inhibition of splicing, which gave a decreased rate

constant, here the inhibition gave a decreased endpoint. This result suggested either that

higher concentrations of Mss116p can bind and trap folding intermediates, limiting the

extent of native ribozyme formation, or that Mss116p can unfold the native ribozyme,

generating a steady-state mixture of native and non-native ribozyme.

We then used the endpoint for the single fast phase as a diagnostic tool to

examine the effects of ATP and Mss116p mutations. The enhanced folding by Mss116p

was strongly dependent upon ATP, as 100 nM Mss116p by itself gave only a small

increase in endpoint relative to the fast phase of a reaction without protein (Figure 4.11B,

cf., with Figure 4.8D). As expected from the ATP requirement, the motif I mutants

Page 109: Copyright by Jeffrey Philip Potratz 2012

93

K158A and K158R also gave only small increases in endpoint (Figure 4.11C). The

SAT/AAA mutant was partially active, giving higher endpoints than the other mutants

but lower than WT Mss116p, especially at lower protein concentrations (Figure 4.11C

and 4.12B, cf., with Figure 4.11B). Because the rate constants did not depend

systematically on the concentration of Mss116p SAT/AAA, we could not determine

quantitatively whether the mutant is compromised for D135 folding relative to the WT

protein (Figure 4.12). Nevertheless, the lower endpoints suggest either that the mutant

protein is unable to promote folding of a fraction of the ribozyme population or that it

gives a lower steady-state level of native ribozyme.

4.3.5 The role of ATP in acceleration of D135 folding by Mss116p

Next we used the discontinuous assay to probe the role of ATP in Mss116p-

mediated folding. It has been suggested that for D135 and the splicing constructs,

acceleration of the critical folding step by Mss116p is ATP-independent, but ATP is

required to promote dissociation of Mss116p, allowing rapid folding to the native state

(138, 140, 149). The discontinuous activity assay afforded an incisive test of this model,

which predicts that Mss116p would be active in the absence of ATP if an alternative

means were provided to remove Mss116p prior to the determination of catalytic activity.

Thus, we incubated Mss116p with D135 and Mg2+

for various times in the

absence of ATP and then added proteinase K (1 mg/ml) to degrade Mss116p (Figure 4.13

and Figure 4.14). After additional incubation of up to 60 min to allow further folding of

any ribozyme that was ‘poised’ to form the native state following Mss116p dissociation,

we transferred aliquots to stage 2 and measured the fraction of native ribozyme by

activity. The prediction from the model above was that this reaction would give

accumulation of ribozyme that could quickly reach the native state upon removal of

Page 110: Copyright by Jeffrey Philip Potratz 2012

94

Mss116p. However, with 100 nM Mss116p (2-fold excess over D135 ribozyme), removal

of Mss116p by proteolysis did not promote rapid native ribozyme formation (Figure

4.13A, closed symbols) beyond the level in an equivalent reaction without the proteolysis

step (open symbols). Further, inclusion of ADP or the non-hydrolyzable analog AMP-

PNP with Mss116p did not lead to significant native ribozyme formation after proteolysis

(Figure 4.15A). These results most simply suggested that the role of the ATPase cycle is

not solely to accelerate Mss116p dissociation.

We next considered an extension of the model in which the ATP requirement

stems from Mss116p being sequestered at non-productive sites within the RNA, requiring

multiple cycles of binding and release to bind productively. Thus, we increased the

Mss116p concentration so that sufficient protein would be available even after non-

productive sites were filled. With 400 nM Mss116p, 8-fold excess over D135, we

observed a variable increase in native ribozyme upon proteolysis (10-25% of the

population; Figure 4.13B, closed symbols, different colors indicate replicate

experiments), which reached an intermediate level between reactions with Mss116p in

the presence and absence of ATP (compare with Mss116p+ATP level of ~0.8 shown in

Figure 4.11B). As above, inclusion of ADP or AMP-PNP with Mss116p gave results that

were the same within error as those in the absence of added nucleotide (Figure 4.15B).

These results indicate that for a fraction of the ribozyme, ATP-independent activity of

Mss116p and removal by proteolysis is indeed sufficient to promote folding to the native

state. However, much of the ribozyme population remains non-native, most likely

because it forms additional intermediates that require further activity of Mss116p.56 Still

higher concentrations of Mss116p (up to 1200 nM) did not give more native ribozyme

upon proteolysis than observed with 400 nM Mss116p (data not shown).

Page 111: Copyright by Jeffrey Philip Potratz 2012

95

The variable but significant increase in native D135 ribozyme after Mss116p

binding and proteolysis in the absence of ATP prompted us to investigate whether this

might also occur with the splicing constructs. Incubation of 1 nM RNA with 2 nM or 10

nM Mss116p in the absence of nucleotide, or in the presence of ADP or AMP-PNP,

followed by proteolysis gave no detectable splicing for the LE construct and at most a

small increase in splicing rate for the SE construct (0-10%; data not shown). Together the

results suggest that for a subpopulation of the ribozyme, ATP hydrolysis and product

release are necessary only to accelerate Mss116p dissociation and allow continued

folding as proposed (138). Nevertheless, ATP is required for additional steps in

Mss116p-dependent folding of much of the D135 population and most or all of the

populations of the SE and LE constructs.

4.3.6 Disruption of native D135 ribozyme by Mss116p

A strong prediction of models in which Mss116p functions as a general chaperone

is that it would not specifically recognize non-native structure and would therefore be

able to use ATP to unfold the native intron structure. This inherently non-specific

mechanism can nevertheless lead to native ribozyme accumulation if the native state is

more stable than populated misfolded states and therefore unfolded less efficiently or is

biased to form intermediates after protein-induced unfolding that preferentially refold to

the native state (26).

Thus, we used the discontinuous assay to probe for an Mss116p-dependent

decrease in the fraction of native D135 ribozyme (151). We prefolded D135 to the native

state and then added Mss116p at a concentration that inhibited productive D135 folding

(Figure 4.16). At various times thereafter, we transferred portions of the reaction to stage

2 and determined the fraction of native ribozyme by activity. In the presence of Mss116p,

Page 112: Copyright by Jeffrey Philip Potratz 2012

96

the fraction of native ribozyme decreased rapidly, suggesting that Mss116p can unfold

the native D135 ribozyme to give intermediates that do not readily refold to the native

state in stage 2. Although substantial loss of native ribozyme was also observed for

Mss116p in the absence of ATP, the decrease was faster and larger with ATP. Mss116p

was then proteolyzed and the ribozyme was again incubated under conditions known to

give native folding (42 °C, 100 mM Mg2+

, 20 min). The fraction of native ribozyme

returned to near its original value (Figure 4.16, diamonds), confirming that the Mss116p-

dependent decrease was due primarily to disruption of the native structure rather than a

process giving irreversible inactivation.

4.3.7 Mss116p-promoted splicing of aI5 in vivo

Finally, to determine which activities of Mss116p are physiologically relevant for

aI5 splicing, we tested whether the Mss116p mutants can promote splicing of aI5 in

vivo (Figure 4.17). We used a previously developed in vivo splicing assay in which WT

or mutant Mss116p is expressed from a centromere-containing (CEN) plasmid in an

MSS116 deletion strain (mss116) (91). Because Mss116p is required for the synthesis of

mt respiratory components, null mutants cannot grow on non-fermentable carbon sources,

such as glycerol, but grow well on raffinose, a non-repressing fermentable sugar,

enabling their splicing phenotype to be assessed by Northern hybridization. To eliminate

the possibility of indirect effects from defects in splicing of other introns in the COX1

gene, we used isogenic strains containing mtDNAs with the single intron aI5 (91). The

COX1 pre-mRNAs in these strains contain aI5 with long 5 and 3 exons (~1.5 and 0.5

kb, respectively, to the 5 and 3 ends of COX1 mRNA), and thus most closely resemble

the LE construct used in vitro.

Page 113: Copyright by Jeffrey Philip Potratz 2012

97

Figure 4.17A shows a Northern blot comparing the abilities of WT and mutant

versions of Mss116p to promote splicing of aI5 in vivo. The blot was hybridized with a

COX1 exon probe to detect spliced mRNA and unspliced precursor RNA. Because the

COX1 gene contains only aI5, the blot shows two major COX1 transcripts,

corresponding to spliced mRNA and an unspliced precursor RNA, with the ratio of the

bands providing a measure of the splicing efficiency. As expected, the wild-type strain,

which has a functional chromosomal copy of MSS116, spliced aI5 efficiently, yielding a

predominant band corresponding to COX1 mRNA (lane 1), while the mss116 strain

accumulates unspliced precursor RNA (lane 2). Expression of WT Mss116p from the

CEN-plasmid efficiently complemented the splicing defect in the mss116 strain,

restoring aI5 splicing to nearly the level of the wild-type strain (lane 3). By contrast, the

K158A and K158R mutants were unable to promote splicing of aI5 substantially above

the low residual level in the mss116 strain (null phenotype; lanes 4 and 5), although

K158A gave a very small increase that paralleled a small amount of residual ATPase

activity (176). The SAT/AAA mutant gave an intermediate phenotype, with

approximately equal amounts of unspliced precursor and spliced mRNA (lane 6). We

verified by immunoblotting that the mutant proteins were expressed at or near the level of

WT Mss116p (Figure 4.17B,C). Previous experiments using strains with multiple COB

and COX1 introns similarly showed that K158A and other motif I mutants gave a null

phenotype, and that the SAT/AAA mutant gave an intermediate phenotype for splicing of

aI5 and all other mt group I and II introns examined (91, 185, 191). Collectively, the

findings for the motif I mutants show that ATP binding and hydrolysis are essential for

Mss116p-promoted in vivo splicing of all S. cerevisiae mt group I and II introns. Further,

the splicing efficiencies of the motif I and SAT/AAA mutants agree with their activities

in splicing and folding aI5 RNAs in vitro and, as for all other Mss116p mutants

Page 114: Copyright by Jeffrey Philip Potratz 2012

98

examined, correlate with their residual RNA-unwinding activities for appropriately sized

duplexes (60).

4.4 DISCUSSION

Here we used in vitro and in vivo approaches to probe the mechanisms by which

the yeast DEAD-box protein Mss116p promotes splicing of the group II intron aI5γ. Our

results indicate that a major physiological role of Mss116p in splicing this intron, as with

other group I and group II introns, is to use ATP to promote conformational transitions

that require the transient disruption of RNA structure.

4.4.1 Requirement for ATP binding and hydrolysis and effects of exon length on

Mss116p-mediated splicing in vitro

To probe the functional requirements of Mss116p in aI5γ splicing, we tested

mutants that are compromised in ATP binding, hydrolysis, and RNA unwinding, and we

used the SE and LE constructs to test whether the ATP requirement depends on the

lengths of the flanking exons. The experimental design was essentially the same as in a

recent report, (149) and where overlapping experiments were performed, the data are

largely consistent. Thus, the LE construct splices much slower than the SE construct;

both constructs require Mss116p and ATP for accelerated splicing; and with higher

concentrations of Mss116p, the splicing rates level and then decrease. The plateau

between activation and inhibition occurs at a higher Mss116p concentration and a higher

splicing rate for the LE construct than the SE construct. Also consistent with prior work,

the mutants K158A and K158R, which are deficient in ATP binding and hydrolysis, are

unable to stimulate splicing of either construct (176). Similar results were obtained for

other Mss116p and CYT-19 motif I mutants with the LE construct (30, 55).

Page 115: Copyright by Jeffrey Philip Potratz 2012

99

However, key conclusions differ from those in previous work (149). Zingler et al.

compared the rate constants for the LE and SE constructs at Mss116p concentrations near

the plateau regions and concluded that the LE construct is stimulated to a greater extent

by Mss116p, perhaps reflecting Mss116p recruitment by the exons or facilitation of

protein oligomerization. In our view, it is dangerous to interpret these maximum values

as a measure of activation because they reflect the intersection of activation and

inhibition activities and can be influenced by changes in either or both activities. Instead,

we interpreted the slope of the rising portion of the Mss116p concentration dependence,

which reflects the overall reaction efficiency — i.e. the free energy change in going from

free protein and RNA in solution to the rate-limiting transition states — and is the part of

the concentration dependence that is least affected by inhibition. Making this comparison,

we observe that Mss116p stimulates splicing of the SE construct at least as efficiently as

the LE construct. Thus, rather than stimulating the SE construct less strongly, we

conclude that Mss116p inhibits it more strongly.

The slower, more temperature-dependent splicing of the LE construct suggests

that the long exons cause or contribute to kinetic barriers in RNA folding. This

conclusion is consistent with previous work indicating that sequences in the 5 exon

upstream of the intron-binding sequences (IBS1 and IBS2) can have large effects on aI5γ

splicing (150). The inhibition of splicing from exon sequences could reflect the formation

of misfolded exon structures that must be resolved prior to splicing, as demonstrated for

group I introns (192, 193). Alternatively or in addition, the long exons may stabilize

structure within the aI5γ intron, presumably by interacting with the intron, thereby

increasing barrier heights for disruption of both non-native and native structure. Evidence

in support of this possibility comes from the lower, more temperature-dependent splicing

rate of the LE construct in the absence of Mss116p, indicating a higher enthalpic barrier,

Page 116: Copyright by Jeffrey Philip Potratz 2012

100

and from the greater stability of the LE construct. The latter is suggested by the higher

temperature optimum for LE construct splicing and its greater resistance to inhibition by

Mss116p, which presumably reflects disruption of the native intron structure. Our finding

that stimulation of both the SE and LE constructs by Mss116p depends upon ATP and is

affected similarly by Mss116p mutations raises the possibility of common rate-limiting

steps and barriers for these two constructs, which may be hindered by exon stabilization

in the LE construct. Further work will be necessary to define the structures of folding

intermediates for these constructs and the roles played by exons.

4.4.2 The SAT/AAA mutant is compromised for splicing SE and LE constructs

The motif III mutant of Mss116p (SAT/AAA) follows a design used with eIF4A

and other DEAD-box proteins (43, 47, 194) and was originally described in a Ph.D. thesis

from H.R. Huang in Philip Perlman’s lab, where it was classified as a ‘weak allele’ based

on in vivo analysis of translation and splicing (185). This mutant has since been the

subject of substantial study and contributed to physical models of Mss116p function (60,

176). Recently, the Mss116p SAT/AAA mutant was reported to be impaired in splicing

of the aI5γ LE construct but as efficient as WT Mss116p for splicing of the SE construct

(149). Because it is deficient in RNA-unwinding activity, this result was taken as

evidence that RNA unfolding is required for the exons but not within the intron. In

contrast, our data show that Mss116p SAT/AAA is reduced in activity by at least a

comparable amount for the SE construct as for the LE construct. The difference in

conclusions can be understood from inspection of both sets of data. The previous

conclusion that Mss116p SAT/AAA promotes splicing of the SE construct with WT

efficiency came from a comparison of splicing rates with 15 nM protein, which is well

into the inhibitory regime for WT but much closer to the plateau for the SAT/AAA

Page 117: Copyright by Jeffrey Philip Potratz 2012

101

mutant (Figure 2 of ref. (149)). The reduced efficiency of Mss116p SAT/AAA with the

SE construct is visible at lower protein concentrations, both in our data (Figure 4.4 and

4.6) and in Figure S2B of Zingler et al (149). Thus, the SAT/AAA mutant again fails to

provide evidence that Mss116p can promote splicing of aI5γ without unwinding RNA.

4.4.3 The roles of ATP in Mss116p-promoted intron folding

Using a new two-stage, discontinuous catalytic activity assay, we found that

Mss116p accelerates native folding of the D135 ribozyme under near-physiological

conditions, supporting and extending previous studies of Mss116p on folding of aI5γ

(55, 138, 140). We then used the discontinuous assay to gain new insights

into the roles of ATP in Mss116p-mediated folding of aI5γ. It was suggested previously

that Mss116p promotes folding by stabilizing an on-pathway folding intermediate in a

reaction that is inherently ATP-independent, but that ATP is needed to promote release of

Mss116p after this step (149). A key prediction of this model is that removal of Mss116p

by proteolysis after binding in the absence of ATP would allow the ribozyme to fold to

the native state, which could be detected by activity in stage 2 of our assay. In contrast to

this expectation, we found that with two-fold excess Mss116p over D135, proteolysis of

Mss116p does not give additional native ribozyme. However, with 8-fold excess

Mss116p, proteolysis does allow 10-25% of the ribozyme to reach the native state. The

partial recovery indicates that for a subpopulation of the ribozyme, Mss116p accelerates

native folding by promoting a step, presumably compaction of domain I, (138) without

requiring ATP, and it provides experimental support for models in which Mss116p

dissociation is required for productive folding (149). Acceleration of domain I

compaction may arise from transient binding and stabilization of a folding intermediate,

as suggested by the ATP independence and the ability of several basic proteins to

Page 118: Copyright by Jeffrey Philip Potratz 2012

102

promote this step with low efficiency (140, 149, 176). Alternatively or in addition,

Mss116p and other DEAD-box proteins may accelerate this step by disrupting local

structure in a process that does not strictly require ATP, perhaps giving transient

unstacking of coaxially stacked helices to allow bending of an internal loop within D1

(60, 195, 196).

Nevertheless, even under the most favorable conditions, the yield of native D135

ribozyme upon incubation with Mss116p followed by proteolysis is substantially smaller

than in the presence of Mss116p and ATP, indicating that most of the population requires

Mss116p and ATP for at least one additional step. The fraction that is able to avoid this

additional requirement is smaller for the SE construct and undetectable for the LE

construct. The increased dependence on ATP with increasing exon length supports the

hypothesis that the additional ATP-dependent step(s) include localized RNA unwinding,

because the larger size and complexity of these RNAs are expected to lead to folding via

more complex pathways with more trapped intermediates, and likely with intermediates

of greater stability due to contributions from the exons. It is also reasonable to expect that

binding of Mss116p to at least some sites within the intron would be inhibitory

throughout the folding process, and thus it is likely that acceleration of Mss116p

dissociation is an additional role of the ATPase cycle independent of any structural

stabilization.

4.4.4 Disruption of the native D135 ribozyme

Models for DEAD-box proteins as general RNA chaperones postulate that they

disrupt RNA structure non-specifically, generating a kinetic redistribution of folding

intermediates and additional chances for productive folding (26). A corollary is that there

is no absolute mechanism for distinguishing native from non-native structure. In support

Page 119: Copyright by Jeffrey Philip Potratz 2012

103

of this hypothesis, we used the discontinuous activity assay to show that Mss116p can

disrupt the native D135 ribozyme in an ATP-dependent manner. This result parallels

previous findings for a native group I intron ribozyme (26, 60) and indicates that

Mss116p is capable of promoting native folding of aI5γ by disrupting misfolded

intermediates, which are less stable than the native state. Although it was reported that

Mss116p cannot unfold the native structure of a domain I construct of aI5γ, (138)

unfolding may have gone undetected by non-denaturing gel assays if the isolated domain

I refolds rapidly to the native structure or to alternative compact forms that migrate

similarly to the native structure.

4.4.5 Requirement for ATP binding and hydrolysis by Mss116p in vivo

To investigate which activities of Mss116p are required for its in vivo function,

we used yeast strains that lack all mt group I and group II introns except aI5γ (91). The

major COX1 pre-mRNA in these strains contains aI5 with long 5 and 3 exons (~1.5 and

0.5 kb, respectively) and thus most closely resembles the LE construct used to analyze

the effect of Mss116p in vitro. In other yeast strains that contain multiple COX1 introns,

the COX1 pre-mRNAs are expected to be even longer and more heterogeneous,

depending upon the order in which the introns upstream of aI5 are spliced.

Our Northern blot analysis indicates that the motif I mutants are essentially

inactive for aI5γ splicing, with a very low level of residual activity for K158A, and that

Mss116p SAT/AAA is more active than the motif I mutants but nevertheless significantly

compromised relative to WT Mss116p, as expected from its decreased RNA-unwinding

activity (60). These effects mirror those of the same mutations in vitro, most simply

suggesting similar barriers to group II intron folding in vivo and in vitro.

Page 120: Copyright by Jeffrey Philip Potratz 2012

104

Our results are in agreement with previous in vivo studies of Mss116p mutations

in strains containing multiple introns, which showed that K158A and other motif I

mutations give a null phenotype for splicing of aI5 and all other group I and II introns

examined, and that SAT/AAA and other motif III mutations give an intermediate

splicing-defective phenotype for aI5 and all other group I and II introns examined (91,

185, 191). We note in particular that the initial characterization of the SAT/AAA

mutation in a strain with three COX1 introns showed that “this mss116 allele does not

function as well as the wild-type MSS116 on splicing” (185) (in contrast to statements in

the Discussion of Zingler et al (149)). Although another mutant, Q412A, has been

suggested to promote efficient RNA splicing in vivo despite low RNA-unwinding

activity, (149, 191) its unwinding activity was assayed only with a relatively long duplex

(17 bp) and was greater than that of a motif III mutant (T307A) in the same study. Thus,

in our view, the simplest interpretation is that Q412A supports splicing by virtue of its

residual RNA-unwinding activity, which is expected to be higher for shorter duplexes of

the type found in group I and II intron RNAs.

4.5 CONCLUSIONS AND IMPLICATIONS

Together, our results indicate that a critical activity in Mss116-mediated folding

of the aI5γ intron is ATP-dependent RNA-unwinding activity. ATP is required for

maximal stimulation of folding and splicing of all constructs tested, and Mss116p is

capable of disrupting even the most stable global structure of the intron, the native state.

Further, for all Mss116p mutants analyzed to date, the ability to promote splicing of aI5γ

in vitro and in vivo correlates with RNA-unwinding activity with appropriately sized

duplexes, suggesting that this unwinding activity is necessary for stimulation of aI5γ

splicing. Nevertheless, it remains likely that different RNAs and different folding

Page 121: Copyright by Jeffrey Philip Potratz 2012

105

intermediates require different activities and that DEAD-box proteins use multiple

mechanisms to promote RNA folding. The recent advances in understanding the folding

of aI5γ

transitions accelerated by DEAD-box proteins and the detailed mechanisms through

which the accelerations are achieved.

4.6 FOOTNOTE

* The efficiency reported in the current work is approximately 5-fold larger than

previously reported (60, 81). In principle, this difference could be caused by the

decreased RNA and protein concentrations, changes in buffer from Tris to MOPS, small

accompanying changes in counterion identity and concentration, and/or recent

preparations of protein being somewhat more active than earlier preparations. Further

experiments indicate that the difference is not caused by the change in buffer conditions,

which result in a small decrease in activity (2-fold), whereas there may be contributions

from the changes in RNA and protein concentration (≤ 2-fold) and differences in activity

between different protein preparations (data not shown).

Page 122: Copyright by Jeffrey Philip Potratz 2012

106

Figure 4.1: RNA constructs

(A) Long exon (LE) and short exon (SE) splicing constructs. Each intron domain is

shown in a different color and indicated with a label, and exon lengths are indicated. (B)

D135 ribozyme construct. Truncated and intact domains are shown in the same colors as

in panel A, and the intact domains are labeled. As indicated, the ribozyme includes 37 nts

at the 3 end, which are derived from a multiple cloning site (131). The secondary

structures were generated by modifying a diagram from the Comparative RNA Website

(197).

(Figure prepared by Mark Del Campo.)

Page 123: Copyright by Jeffrey Philip Potratz 2012

107

Figure 4.2: Self-splicing of the LE and SE constructs

(A) Gel images from splicing reactions performed at 33 °C. Splicing was monitored as

the fraction of radiolabeled material present as precursor RNA (see Section 4.2.4). Bands

in this and subsequent gel images are precursor (P), excised intron lariat (I-lar), excised

linear intron (I-lin), and ligated exons (E1-E2). (B) Temperature dependences for self-

splicing of the LE (red) and SE (blue) constructs. Rate constants, as determined from

single exponential fits or from initial rates of slower reactions, are plotted on a log scale

against the inverse of temperature (in Kelvin) multiplied by 1000. Reaction conditions

were 50 mM Na-MOPS, pH 7.0 (determined at 25 °C), 8 mM MgCl2, 100 mM KCl

(including a contribution from Mss116p storage buffer), 1 mM ATP-Mg2+

, and 10%

Mss116p storage buffer (see Section 4.2.4). All values shown reflect the average and

standard error from 2–4 independent determinations except those at 41 °C, which reflect

a single determination. Downward arrows indicate that non-specific products were

observed under those conditions, allowing determination of only an upper limit on the

rate constant for splicing. The regions in which splicing rate increased with temperature

gave ΔH values of 57 kcal/mol and 44 kcal/mol for the LE and SE constructs,

respectively.

Page 124: Copyright by Jeffrey Philip Potratz 2012

108

Figure 4.3: Self-splicing of the LE construct (panel A) and SE construct (panel B)

at different temperatures

Reactions were performed in a thermal cycler with a temperature gradient. Temperatures

for each reaction were estimated by linear interpolation from the temperature settings.

Page 125: Copyright by Jeffrey Philip Potratz 2012

109

Figure 4.4: Mss116p-stimulated splicing of the LE and SE constructs

Page 126: Copyright by Jeffrey Philip Potratz 2012

110

(A) Gel images of LE and SE splicing in the presence and absence of Mss116p (10 nM

and 2 nM Mss116p for the LE and SE constructs, respectively). In this and all other

experiments with Mss116p, the total intensity of radiolabeled RNA in the quantified

region of the gel decreased less than 2-fold during the experiment, indicating minimal

RNA degradation. (B) Progress curves of LE and SE splicing in the presence of various

Mss116p concentrations, as indicated by color. Circles show results from reactions in the

presence of 1 mM ATP and triangles indicate reactions without ATP. Points represent the

fraction of labeled material present as precursor as a function of time, and the curves

reflect the best fit by a first-order rate equation. (C) Mss116p concentration dependences

for splicing of the LE and SE constructs in the presence of ATP. For the SE construct,

Mss116p was saturating or nearly saturating even at low concentrations equivalent to that

of the radiolabeled RNA (~1 nM), preventing determination of a second-order rate

constant for activation.

(This figure is the work of Mark Del Campo.)

Page 127: Copyright by Jeffrey Philip Potratz 2012

111

Figure 4.5: Mss116p-stimulated splicing of LE and SE constructs

Using Tris buffer conditions (10 mM Tris-Cl, pH 7.5, 100 mM KCl, 10% glycerol), with

10 nM RNA. (A and B) Gel images of splicing in the presence and absence of Mss116p

for the LE construct (40 nM Mss116p, panel A) and the SE construct (10 nM Mss116p,

panel B). (C) Concentration dependences of Mss116p stimulation for the LE and SE

constructs. Rate constants from this experiment are 1.8 106 M

–1 min

–1 for the LE

construct and ≥1.8 106 M

–1 min

–1 for the SE construct. (D) Concentration dependences

for Mss116p SAT/AAA for splicing of the LE and SE constructs under Tris buffer

conditions. Rate constants are 1.2 105 M

–1 min

–1 for the LE construct and 4.4 10

5 M

–1

min–1

for the SE construct.

(This figure is the work of Mark Del Campo.)

Page 128: Copyright by Jeffrey Philip Potratz 2012

112

Figure 4.6: Stimulation of splicing by the SAT/AAA mutant of Mss116p

(A) Progress curves of splicing of LE (left) and SE (right) constructs in the presence and

absence of the Mss116p SAT/AAA. Circles indicate reactions in the presence of 1 mM

ATP and triangles indicate reactions without ATP. (B) Concentration dependence of

Mss116p SAT/AAA for stimulation of splicing of the LE and SE constructs in the

presence of ATP.

(This figure is the work of Mark Del Campo.)

Page 129: Copyright by Jeffrey Philip Potratz 2012

113

Figure 4.7: Mss116p motif I mutants K158A and K158R in splicing reactions of the

LE and SE constructs (panels A and B, respectively)

All reactions included 1 mM ATP and were performed under standard near-physiological

conditions at 30 °C (50 mM Na-MOPS pH 7.0, 100 mM KCl, 8 mM MgCl2, and 5%

glycerol). The parallel reactions shown in the absence of Mss116p (black x’s) and in the

presence of WT Mss116p (purple diamonds) were performed side-by-side and gave the

same rates within error as equivalent reactions shown in Figure 4.4.

(This figure is the work of Mark Del Campo.)

Page 130: Copyright by Jeffrey Philip Potratz 2012

114

Figure 4.8: Two-stage catalytic activity assay to monitor folding of D135 ribozyme

(A) Reaction schematic depicting standard, near-physiological conditions for stage 1. In

subsequent panels, results are shown plotted against folding time in stage 1 (t1) or

cleavage time in stage 2 (t2). (B) Prefolded and nonfolded controls. Ribozyme was either

prefolded (circles, 42 °C, 100 mM MgCl2, 500 mM KCl, 50 mM Na-MOPS, pH 7.0, 20

min) or transferred to stage 2 directly from buffer solution (triangles, 30 °C, 50 mM Na-

MOPS, pH 7.0). As expected for a single turnover followed by slow product release, the

reaction with prefolded ribozyme gave a stoichiometric burst of product (see Section

4.2.5) and a second phase reflecting subsequent turnovers of the ribozyme. The

nonfolded ribozyme reaction gave less substrate cleavage, with five determinations

giving an average burst corresponding to 15% of the ribozyme population. (C) Progress

Page 131: Copyright by Jeffrey Philip Potratz 2012

115

of D135 RNA folding in stage 1 under non-physiological conditions similar to those used

previously (42 °C, 100 mM Mg2+

, pH 8.1), (183) with the fraction of native ribozyme

determined from the burst amplitude of substrate cleavage in stage 2. (D) Progress of

D135 folding under near-physiological conditions. Burst amplitudes are normalized by

the maximal burst amplitude from prefolded ribozyme, as shown in panel b. Fifteen

determinations gave a fast phase with an amplitude of 0.22 ± 0.02 and a rate constant of

1.0 ± 0.2 min–1

. A slower phase was also present, and six determinations for this phase

gave a rate constant of 1.4 (± 0.6) 10–3

min–1

and a final amplitude of ~0.6.

Page 132: Copyright by Jeffrey Philip Potratz 2012

116

Figure 4.9: Testing of 30 °C, 100 mM Mg2+

for conditions of stage 2 in the

discontinuous catalytic activity assay

The amount of rapid substrate cleavage from ribozyme transferred directly from buffer

solution lacking Mg2+

at 30 °C (triangles, the nonfolded reaction) is not much less than

that from an equivalent amount of ribozyme that was prefolded (42 °C, 100 mM Mg2+

, 20

min) before being transferred to the stage 2 conditions. A comparison of this plot with

Figure 4.8B demonstrates that the lower temperature in that experiment for stage 2, 15

°C, is more effective for blocking folding of D135 on the time scale of substrate cleavage

by the native ribozyme. Thus, the lower temperature was used for all subsequent

experiments.

Page 133: Copyright by Jeffrey Philip Potratz 2012

117

Figure 4.10: Prefolded and nonfolded reactions for monitoring D135 RNA folding at

42 °C

Black symbols and curves show reactions that were transferred from 42 °C into standard

stage 2 reaction conditions (15 °C, 100 mM Mg2+

, 80 mM Na-HEPES, pH 8.1). The

circles show the progress of substrate cleavage for a reaction that was first prefolded (42

°C, 100 mM Mg2+

, 50 mM Na-MOPS, pH 7.0, 20 min) and then transferred, and the

triangles show results from a reaction that was transferred from 42 °C (80 mM Na-

HEPES, pH 8.1) to the stage 2 conditions without prefolding. HEPES buffer was used

here so that the reaction would be identical to previous work (see Section 4.3.3).(184)

The gray symbols show analogous reactions that were transferred from 30 °C (50 mM

Na-MOPS, pH 7.0, ± 100 mM Mg2+

), either with or without prefolding, reprinted from

Figure 4.8B for comparison. The prefolded reactions are essentially the same as expected,

indicating nearly 100% native ribozyme, but the nonfolded reaction transferred from 42

°C gives substantially more cleavage than the equivalent reaction transferred from 30 °C,

despite the essentially identical conditions in stage 2 for the two reactions.

Page 134: Copyright by Jeffrey Philip Potratz 2012

118

Figure 4.11: Mss116p accelerates native folding of the D135 ribozyme

Page 135: Copyright by Jeffrey Philip Potratz 2012

119

(A) Time courses of substrate cleavage in stage 2 after incubation with Mss116p under

near-physiological conditions in stage 1. D135 ribozyme (50 nM) was incubated with 100

nM Mss116p and 1 mM ATP for the indicated times before substrate cleavage was

measured under stage 2 conditions (see Section 4.2.5). Also shown are a ‘nonfolded’

control reaction, in which D135 was added directly to stage 2 conditions, and a prefolded

control, in which D135 was incubated at 42 °C for 20 min in the presence of 100 mM

Mg2+

to form native ribozyme. (B) Progress of D135 folding in stage 1 in the presence of

100 nM Mss116p with 1 mM ATP (circles, kobs = 0.61 ± 0.07 min–1

, amplitude = 0.76 ±

0.04) or without ATP (diamonds, kobs = 0.64 ± 0.07 min–1

, amplitude = 0.28 ± 0.01). The

fraction of native ribozyme was determined by the burst amplitude in stage 2, as

described in Materials and Methods. (C) Progress of D135 folding in stage 1 with

Mss116p mutants in the presence of 1 mM ATP and 100 nM protein. Proteins are

Mss116p SAT/AAA (blue, kobs = 0.5 ± 0.2 min–1

, amplitude = 0.50 ± 0.09), Mss116p

K158A (orange, kobs = 0.7 ± 0.1 min–1

, amplitude = 0.34 ± 0.06), or Mss116p K158R

(red, kobs = 0.7 ± 0.2 min–1

, amplitude = 0.33 ± 0.01).

Page 136: Copyright by Jeffrey Philip Potratz 2012

120

Figure 4.12: Progress curves of D135 ribozyme folding

50 nM D135 ribozyme in the presence of 1 mM ATP and various concentrations of

Mss116p (panel A) or Mss116p SAT/AAA (panel B).

Page 137: Copyright by Jeffrey Philip Potratz 2012

121

Figure 4.13: Proteolysis of Mss116p after incubation with D135 RNA in the absence

of ATP

Mss116p was added in stage 1 at 100 nM (panel A) or 400 nM (panel B) and incubated

for the indicated times under near-physiological conditions. Proteinase K (1 mg/ml) was

Page 138: Copyright by Jeffrey Philip Potratz 2012

122

then added at the indicated times (closed symbols) and incubated for an additional 8 – 60

min to permit further folding before aliquots were transferred to stage 2 and the fraction

of native ribozyme was determined by measuring the substrate cleavage burst amplitude.

The fraction of native ribozyme did not depend on the incubation time after proteinase K

addition (8–60 min), and the symbols show the average values. Open symbols show

equivalent reactions with Mss116p and without nucleotide, to which proteinase K was

not added. Including or omitting 0.5% SDS with proteinase K to ensure removal of

peptide fragments had no significant effect on the results (data not shown). For the

experiments shown in panel B, 0.5% SDS was added immediately after proteinase K.

Results from independent determinations are shown in different colors. It can be seen that

the increase in native ribozyme upon proteinase K treatment is variable and that the

variation is larger between experiments than within experiments.

Page 139: Copyright by Jeffrey Philip Potratz 2012

123

Figure 4.14: Proteinase K digestion of Mss116p in experiment shown in Figure 4.13

When proteinase K is added to Mss116p in the stage 1 conditions (30 °C, 50 mM MOPS,

pH 7.0, 8 mM Mg2+

), Mss116p is undetectable within 30 s (right lane). The second lane

from the right shows an equivalent reaction to which proteinase K was omitted, and the

third lane from the right shows an equivalent amount of proteinase K alone (at dye front).

Page 140: Copyright by Jeffrey Philip Potratz 2012

124

Figure 4.15: Proteolysis of Mss116p after incubation with D135 RNA

Mss116p was present at 100 nM (Panel A) or 400 nM (panel B). Nucleotides were

present, as indicated, at 1 mM. Closed symbols show the fraction of native ribozyme for

Page 141: Copyright by Jeffrey Philip Potratz 2012

125

reactions in which proteinase K (1mg/ml) was added at the indicated time, followed by

additional incubation of 5 – 30 min before transfer to stage 2. The results did not depend

on this incubation time, and the average values are shown. SDS (0.5%) was added

immediately after proteinase K to ensure that peptide fragments were prevented from

interacting with the ribozyme. Open symbols show equivalent reactions to which

proteinase K was not added. Reactions in the absence of nucleotide (circles) were

performed side-by-side for comparison and gave the same results within error as the

equivalent reactions shown in Figure 4.13. Results from two independent experiments are

shown in red and blue.

Page 142: Copyright by Jeffrey Philip Potratz 2012

126

Figure 4.16: Unfolding of native D135 RNA by Mss116p

The ribozyme (1.8 µM) was prefolded at 42 °C, 100 mM Mg2+

and then diluted to 30 °C

and 8 mM Mg2+

prior to addition of Mss116p (1.2 µM Mss116p; 74 nM D135 after

dilution). Reactions included 1 mM ATP (green) no ATP (red) or the same volume of

storage buffer without Mss116p (blue). Reactions were incubated for the indicated times

and the fraction of native ribozyme was determined by transferring aliquots to stage 2 and

measuring substrate cleavage. After 12 min, proteinase K (1 mg/ml) and additional Mg2+

(100 mM) were added and the RNA was again folded to the native state by incubation at

42 °C for 20 min prior to determining the fraction of native ribozyme as above

(diamonds). This refolding step was included to ensure that the decrease in native

ribozyme upon incubation with Mss116p arose from unfolding rather than an irreversible

process such as RNA degradation.

Page 143: Copyright by Jeffrey Philip Potratz 2012

127

Figure 4.17: Northern hybridization and correlated immunoblot comparing the

ability of wild-type and mutant Mss116p to promote splicing of aI5 in

vivo

(A) Northern hybridization. The blot shows whole-cell RNAs (1.0 µg) from the indicated

strains separated in a 1.5% agarose gel and hybridized with a 32P-labeled oligonucleotide

complementary to COX1 exon 6. Lanes: (1) WT 161-aI5 (WT); (2) mss116-aI5

transformed with CEN plasmid pRS416 (empty vector); (3-6) mss116-aI5 transformed

with CEN plasmids expressing Mss116p mutants K158A, K158R, and SAT/AAA. (B)

Immunoblot. The blot shows TCA-precipitated proteins (~60 µg) from the same strains

as in panel (A) separated in a 4-20% polyacrylamide gradient gel and probed with an

anti-Mss116p antibody. (C) Immunoblot stained with AuroDye Forte to confirm equal

loading. The numbers to the left of the gel in (b) and (c) indicate the positions of size

markers (Precision Plus Protein Dual Color Standards; Bio Rad).

(This figure is the work of Rachel Wolf.)

Page 144: Copyright by Jeffrey Philip Potratz 2012

128

Chapter 5: Rapid structure formation within the aI5γ group II intron

5.1 INTRODUCTION

Using catalytically active RNA to probe folding pathways has been an effective

method for obtaining kinetic information on the fraction of RNA in the native state and

for detecting the presence of intermediate, non-catalytic RNA conformations.

Intermediates may be non-catalytic because they are kinetically-trapped misfolded

conformations (134, 136) or because they are more stable than the native state in the

absence of a cofactor (198, 199). However, without the use of orthogonal techniques that

probe for physical information, studies using catalytic activity are unable to reveal

structural information about the folding intermediates and cannot differentiate between

the two types of non-catalytic conformations mentioned above.

Two powerful techniques that probe physical features of RNA are chemical

footprinting and small angle X-ray scattering (SAXS). Chemical footprinting experiments

use exogenous probes to modify the RNA according to its structural features. For

example, DMS footprinting uses dimethyl sulfate to methylate the base-pairing faces of

accessible adenine and cytosine nucleobases. Formations of base-pairing interactions or

tertiary interactions that involve the base-pairing face of the nucleotides reduce the

accessibility of dimethyl sulfate and provide protection from modification. The

modification pattern of accessible and protected nucleotides can be used to obtain

structural information regarding which features are formed in a particular RNA

conformation. Unlike the mostly local structural information provided by DMS

footprinting, SAXS is able to provide a more global view of folding intermediates. SAXS

is a solution-based method that allows the overall size and shape of RNA to be probed.

While the resolution capabilities of SAXS are not as fine as X-ray crystallography,

compaction events during RNA folding can often be detected (94, 163, 165).

Page 145: Copyright by Jeffrey Philip Potratz 2012

129

In this study, knowledge acquired about the aI5γ group II intron folding pathway

through catalytic activity measurements (see Chapter 4) is combined with DMS

footprinting and SAXS data to probe structural features of any folding intermediates that

might be populated. Folding and splicing of the full-length aI5γ construct in near-

physiological conditions (100 mM KCl, 8 mM Mg2+

, 30 °C) are extremely slow with a

rate constant near 10-5

min-1

(see Section 4.3.1) (142). The DEAD-box protein Mss116p

has been shown to stimulate folding and splicing of aI5γ (see Section 4.3.1 and (60, 81,

142). There is evidence that this stimulation occurs because of and is dependent on the

RNA unwinding ability of the protein. For example, a motif III SAT/AAA mutant of

Mss116p retains ATPase and RNA binding activity but has a decreased efficiency for

separating duplexes compared to wild-type Mss116p (60). Correspondingly, this mutant

has a decreased ability to stimulate folding and splicing of aI5γ constructs, indicating that

Mss116p promotes native folding by promoting a local unfolding event (see Section 4.3.2

and (60)). Importantly, this local unfolding event is also indicated in vivo (see Section

4.3.7). Additionally, a ribozyme version of the intron populates the native state to a

greater degree after interaction with Mss116p and ATP than after interaction with

Mss116p alone, again indicating that an ATP-dependent strand separation event promotes

native state formation (see Sections 4.3.4-5). The requirement for a local unfolding event

implies that there is at least one misfolded conformation populated along the folding

pathway.

To probe the physical characteristics of any intermediates populated along the

folding pathway of aI5γ, DMS footprinting and SAXS strategies were utilized. DMS

footprinting was carried out on full-length aI5γ (~1500 nts) and the derived D135

ribozyme (~600 nts) (see Chapter 4, Figure 4.1). Both constructs have nucleotides that are

in duplex and loop regions in the native state become protected from DMS modification

Page 146: Copyright by Jeffrey Philip Potratz 2012

130

soon after the addition of Mg2+

, indicating rapid formation of structural features.

Additionally, initial SAXS data on the D135 ribozyme reveal a degree of global

compaction upon addition of Mg2+

, again indicating rapid formation of structural features

that underlie this compaction.

5.2 MATERIALS AND METHODS

5.2.1 RNA preparation

Group II intron aI5γ was transcribed in vitro from HindIII-digested plasmid by T7

RNA polymerase using a Megascript kit (Ambion). RNA was isolated by phenol-

chloroform extraction and size exclusion chromatography using two consecutive G-50

columns. D135 RNA was transcribed in vitro from HindIII-digested pQL71 using T7

RNAP and purified via an RNeasy column (Qiagen). RNA concentrations were

determined spectrophotometrically using extinction coefficients at 260 nm of 1.76 107

M–1

cm–1

and 5.86 106 M

–1 cm

–1 for aI5γ and D135, respectively.

5.2.2 Footprinting data acquisition and analysis

DMS footprinting reactions were carried out using 0.5 pmol of full-length aI5γ.

The RNA was denatured at 92°C for 1 minute in the absence of Mg2+

and then transferred

to reaction conditions. The states of the RNA probed were ‘unfolded’ (50 mM Na-MOPS

pH 7.0, 100 mM KCl), ‘native’ (50 mM Na-MOPS pH 7.0, 100 mM KCl, 8 mM MgCl2,

100 nM Mss116p, 2.5 hr+, 30 °C), and an ‘early folding time point’ that was initiated

from the unfolded conditions with the addition of Mg2+

(50 mM Na-MOPS pH 7.0, 100

mM KCl, 8 mM MgCl2, 30 °C, < 5 min). In addition, all reactions contained 1 mM ATP-

MgCl2 and 5% glycerol. After the RNA was folded, it was incubated with 1 µL of 70 mM

DMS for 2.5 minutes at 30 °C. A published protocol was followed using P32 labeled

Page 147: Copyright by Jeffrey Philip Potratz 2012

131

primers to obtain cDNA fragments that were separated on 8% polyacrylamide gels and

visualized with a PhosphorImager (154) (Figure 5.1).

SAFA was used to quantify the individual bands in the footprinting gels (200,

201). These bands were then normalized by cDNA full-length products that were

quantified using ImageQuant (GE Healthcare) and multiplied by 1000. In part because

the cDNA products vary in size depending on if aI5γ has spliced or not, an area that

covered both cDNA product lengths was boxed and counted as full-length product. For

the majority of nucleotides in the intron, data from two or more reactions were averaged

and the standard deviations were determined. However, data from only one reaction was

obtained for nucleotides 632,633, and 725-780 for the native state and nucleotides 862-

886 for the early folding time point state. Additionally, nucleotides 789-886 were not

probed in the native state.

The data for unfolded RNA incubated with DMS were compared to data with

unfolded RNA not incubated with DMS to observe which nucleotides were accessible to

DMS. In order for a nucleotide to be considered accessible, the value +DMS must be ≥ 1

larger than the value –DMS, the –DMS value cannot be larger than 1, and the standard

deviation for the +DMS value cannot overlap with the standard deviation for the –DMS

value. In order for a nucleotide to be considered protected, the unfolded+DMS value

must be ≥ 0.5 larger than the value for the native+DMS or early folding time point+DMS

reactions, the standard deviations cannot overlap, and the –DMS reactions cannot be

larger than 1. Certain nucleotides show greater protection in the native state than in the

early folding time point state. These nucleotides are considered protected in the early

folding time point state and have early folding time point+DMS values ≥ 0.4 larger than

values with the native+DMS reactions. Lastly, some nucleotides are not protected but are

made more accessible to DMS after folding. If a nucleotide had a value in the

Page 148: Copyright by Jeffrey Philip Potratz 2012

132

native+DMS reaction ≥ 0.1 larger than the unfolded+DMS reaction, the standard

deviations did not overlap, and the –DMS reactions were less than 1 then that nucleotide

was labeled as being more accessible after folding.

Experiments probing the native state were carried out both with the addition of

0.1 mg/ml proteinase K after aI5γ incubation with Mss116p and without proteinase K

addition. Because minimal differences in DMS profiles were noted and the amount of

nucleotides protected was similar in the presence and absence of proteinase K, the

experiments were treated as identical during analysis. Additionally, the native state was

confirmed to have significantly spliced by comparing the full-length products

corresponding to unspliced and spliced aI5γ generated in the reverse transcription

reaction (Figure 5.1).

5.2.3 SAXS data acquisition and analysis

SAXS data were collected on 0.5 µM or 1 µM of D135 ribozyme at the Advanced

Photon Source beamline 12-ID-C with twenty 1 second exposures for static

measurements and a variable number of 1 second exposures for the time-resolved data at

a sample-detector distance of 2 meters. The states of the RNA probed statically at room

temperature were ‘unfolded’ (50 mM Na-MOPS pH 7.0), ‘native’ (50 mM Na-MOPS pH

7.0, 500 mM KCl, 50 mM MgCl2, 42 °C for 20’), ‘early time point’ (50 mM Na-MOPS

pH 7.0, 100 mM KCl, 8 mM MgCl2, 30 °C for 20-30’), ‘Mss116p +ATP’ (50 mM Na-

MOPS pH 7.0, 100 mM KCl, 8 mM MgCl2, 30 °C for 20-30’, stoichiometric amount of

Mss116p), and ‘Mss116p –ATP’ (50 mM Na-MOPS pH 7.0, 100 mM KCl, 8 mM

MgCl2, 30 °C for 20-30’, stoichiometric amount of Mss116p). All of the samples also

contained 5% glycerol and 1 mM ATP-MgCl2 (except for the Mss116p –ATP sample,

which lacked ATP-Mg2+

, and the unfolded sample, which only contained 50 mM Na-

Page 149: Copyright by Jeffrey Philip Potratz 2012

133

MOPS pH 7.0). The samples were folded to the aforementioned states at the temperatures

indicated and then were put on ice and exposed to the beam at room temperature.

Conversely, the early time point, Mss116p +ATP, and Mss116p –ATP samples were

observed with time dependence at 30 °C.

Data were collected during two separate visits to APS and were qualitatively in

good agreement when comparing the relative peak heights of Kratky plots of different

states from one visit to the next. Analysis of the SAXS data was performed with IGOR-

Pro (WaveMetrics) and ATSAS (version 2.3) software.

5.3 RESULTS

5.3.1 The unfolded state of aI5γ is readily accessible to DMS modification

First, the unfolded state of aI5γ in the presence of 100 mM KCl was probed

(Figure 5.2 and Figure 5.4A). Predictably, many A and C nucleotides that are thought to

reside in loops or not be base paired in the native state were modified by DMS (nts: 27,

38, 41, etc.). In fact, almost every loop region has at least one nucleotide modified by

DMS, indicating a lack of tertiary structure formation. Somewhat surprisingly, but in

agreement with an earlier study (202), many nucleotides that are base paired in the native

state are also accessible to DMS in the presence of 100 mM KCl (nts: 66,67,129, etc.),

indicating a lack of native secondary structure formation. This could result from Mg2+

being necessary in order for the RNA to form certain secondary structural elements (202)

or from non-native secondary structure forming that prohibits the formation of secondary

structure found in the native state. Because full-length aI5γ is being probed, non-native

secondary structure involving interactions between the exon(s) and intron may form,

which could trap the RNA in a misfolded state (60, 138, 142).

Page 150: Copyright by Jeffrey Philip Potratz 2012

134

5.3.2 Addition of Mg2+

results in rapid formation of structural elements

To determine whether any structural features form early in the folding pathway,

changes in DMS modification patterns were probed soon after folding was initiated. The

conformation being investigated is referred to as an early folding time point state. To

reach the early folding time point conformation, the RNA was folded in near-

physiological conditions (100 mM KCl, and 8 mM Mg2+

at 30 °C for 2’) before DMS

was added. Although this folding incubation leads to much less than 1% of the RNA

reaching the native state and splicing (see Chapter 4, Figure 4.3), significantly more

nucleotides were protected from DMS in the early folding time point state than the

unfolded state (Figure 5.3 and Figure 5.4B, green, blue, and orange nts). While many of

the protected nucleotides appear to simply form secondary structure upon the addition of

Mg2+

(nts: 180, 181, 218 etc.), nucleotides in loop regions involved in tertiary contacts

were also protected (nts: 348 in α’ and 617 in µ). Protection of these nucleotides, and

others in loop regions that do not have identified tertiary contacts (nts: 209, 210, 395,

etc.), indicates tertiary structure formation early in the folding pathway. Therefore, upon

the addition of Mg2+

, some secondary and tertiary structural features form in less than

two minutes.

5.3.3 Native state is significantly more protected than early folding time point state

The last conformation probed was the native state, which was formed by

incubating the RNA with Mss116p for > 2.5 hours. The majority of RNA spliced under

these conditions, as confirmed by the different lengths of the cDNA products generated

from unspliced or spliced RNA (Figure 5.1). Predictably, numerous nucleotides that were

not protected in the early folding time point state are now protected in the native state,

indicating further formation of structure unique to the native state (Figure 5.3 and Figure

5.4B, red nts: 27, 51, 63 etc.). Also, some nucleotides that were protected in the early

Page 151: Copyright by Jeffrey Philip Potratz 2012

135

folding time point state are protected to a greater degree in the native state (Figure 5.3

and Figure 5.4B, orange nts: 187, 191, 195, etc.). Additionally, two nucleotides are made

more accessible to DMS in the native state compared to the unfolded state (Figure 5.3

and Figure 5.4B, magenta nts: 420 and 734). This information leads to a model in which

the native state contains more structural elements than the early folding time point state

and further stabilizes some of the structures that are already formed in the early folding

time point state. Further, the native state has undergone splicing, which may give rise to

some of the changes in DMS modification between the early folding time point and the

native states.

5.3.4 Addition of Mg2+

gives rapid formation of structure in D135 ribozyme

A global comparison of the footprinting profiles for the unfolded, early folding

time point, and native state are shown in Figure 5.5A,B while Table 5.1 indicates all of

the nucleotides that were modified by DMS in full-length aI5γ. One of the most readily

apparent observations from the data with full-length aI5γ is that the early folding time

point state has many more nucleotides protected from DMS modification compared to the

unfolded state, indicating that significant structural features have formed soon after the

addition of Mg2+

(see Figure 5.5A and compare the unfolded profile (black) with the

early folding time point profile (blue)). Correspondingly, preliminary footprinting data

using the derived D135 ribozyme also reveals formation of structural elements soon after

the addition of Mg2+

(see Appendix and Figure A.1). Moreover, many of the structures

that are indicated to have formed in the D135 ribozyme are also formed in the full-length

aI5γ.

Page 152: Copyright by Jeffrey Philip Potratz 2012

136

5.3.5 SAXS data reveal compaction of D135 ribozyme

The full-length aI5γ construct did not provide useful SAXS data at small angles.

In addition, the combined length of the exons is over two thirds the length of the intron

domains and would make data analysis complex. Therefore, the D135 ribozyme was used

to gather SAXS data. The kinetic information on the fraction of natively-folded

molecules of the D135 ribozyme revealed by the discontinuous assay (see Section 4.3 and

(142) can be combined with SAXS data to evaluate global structural features of the native

state and any intermediates that may be populated. First, the SAXS profile for the

unfolded (see Section 5.5)* structure (50 mM Na-MOPS pH 7) was compared to the

profile for D135 ribozyme that was prefolded to the native state‡ (50 mM Na-MOPS pH

7, 500 mM KCl, 50 mM Mg2+

42 °C for 20’). Observing the different profiles on a

Kratky plot plainly reveals the large compaction of the native state relative to the

unfolded state. The native state profile shows a clear, sharp peak indicative of a compact

conformation while the unfolded state profile is relatively flat and has no peak (Figure

5.6, black and gray traces). Analogously, the compaction of D135 ribozymes that have

undergone different prefolding incubations can be probed. States such as an early time

point (50 mM Na-MOPS pH 7.0, 100 mM KCl, 8 mM Mg2+

30 °C for 20’), Mss116p –

ATP (50 mM Na-MOPS pH 7.0, 100 mM KCl, 8 mM Mg2+

, Mss116p, 30 °C for 20’),

and Mss116p +ATP (50 mM Na-MOPS pH 7.0, 100 mM KCl, 8 mM Mg2+

, Mss116p, 1

mM ATP-Mg2+

, 30 °C for 20’) were probed. Previous catalytic activity measurements

showed that these incubations give ~30%, ~40%, and ~80%, respectively, of the

ribozyme folded to the native state. (see Sections 4.3.3-5 and note that here the raw, non-

normalized fraction of active D135 molecules inferred from the discontinuous assay is

used) The height of the peak in the Kratky plot increases as the fraction of native

ribozyme increases (Figure 5.6).

Page 153: Copyright by Jeffrey Philip Potratz 2012

137

5.3.6 Time-resolved data indicate early compaction

In an attempt to follow compaction from the unfolded state to the states

mentioned above, a time-resolved mixing setup was used. RNA and buffer were mixed

with folding conditions to initiate the reaction and data were collected. First, folding to

the early time point state was observed as RNA was folded at 30 °C in 100 mM KCl and

8 mM Mg2+

for ~4 minutes, an incubation that produces ~30% native ribozyme. A Kratky

plot revealed that there was no systematic increase in peak height as a function of time

(Figure 5.7A), which would have indicated compaction during the monitored incubation

time. Rather, even the earliest time point after mixing (~3 sec) showed significant

compaction of the ribozyme compared to the unfolded state. Analogously, both the

Mss116p –ATP and Mss116p +ATP reactions indicated that no significant compaction

occurs after the initial compaction that takes place in the dead time of the experiment

(Figure 5.7B,C). It appears that the compactness possessed by these three states after 20-

30 minutes is essentially complete ~3 seconds after mixing with the folding solution

conditions (Figure 5.7, compare thick lines to thin lines).

5.4 DISCUSSION

Both the DMS footprinting data and SAXS data reveal significant formation of

structure for aI5γ constructs soon after the addition of Mg2+

. The footprinting data

illustrate that large amounts of secondary structure and some tertiary structure have

formed while the SAXS data indicate that significant global compaction of the RNA also

occurs within minutes of Mg2+

addition. Moreover, the SAXS data show that the majority

of compaction appears to be complete after 3 sec, as no significant further compaction

was observed beyond the initial compaction seen after the dead time of the experiment.

While this study did not uncover any specific misfolded conformations of aI5γ,

misfolded RNA conformations do not necessarily strictly differ in overall appearance or

Page 154: Copyright by Jeffrey Philip Potratz 2012

138

structural features when compared to native states. Rather, they can mimic the native

state quite closely in some cases (17). Additionally, it is still possible that a multitude of

misfolded intermediate structures exist. These structures would not strictly give rise to a

distinct DMS footprinting profile because of their inherent heterogeneity but would

almost certainly be more compact than the unfolded conformation. As for the conclusions

about the group II intron aI5γ folding pathway derived from this study, it appears that

significant structural formation occurs early in the folding pathway upon the addition of

Mg2+

.

5.5 FOOTNOTE

* The unfolded state of the D135 ribozyme does not contain KCl, as opposed to

the unfolded state of aI5γ, which contains 100 mM KCl. The unfolded state for D135

ribozyme lacks KCl because a discontinuous activity assay revealed that a significant

portion of the ribozyme folds productively in the presence of KCl (data not shown).

‡ A control reaction using the discontinuous assay revealed that the same amount

of D135 ribozyme folds to the native state whether 50 or 100 mM Mg2+

is in the folding

reaction (data not shown).

Page 155: Copyright by Jeffrey Philip Potratz 2012

139

Figure 5.1: Representative DMS footprinting gel

A DMS footprinting gel is shown above. The red box indicates the two different full-

length cDNA products resulting from reactions in which the intron was spliced with

Mss116p. The shorter cDNA product being more prevalent than the longer cDNA

product indicates that the majority of the intron is spliced and the reverse transcription

reaction terminates at the 5’ splice site most often. Lanes 1 and 2 are probing the native

state while lanes 3,4, and 5 are probing the unfolded state. The gel shows from

approximately nucleotide number 600 (bottom of gel) through the 5’ end.

Page 156: Copyright by Jeffrey Philip Potratz 2012

140

Figure 5.2: Many DMS accessible nucleotides in the unfolded state

Full-length aI5γ was subjected to DMS footprinting. Shown on the secondary structure

diagram are the A and C nucleotides that were accessible to DMS modification (black) or

were not accessible (gray) when the RNA was in an unfolded state. Note that even many

nucleotides involved in secondary structures in the native state are accessible here,

indicating that they lack defined secondary structure in these conditions. Additionally, as

expected, many nucleotides in loops or involved in tertiary contacts in the native state are

accessible to DMS in these conditions. Known tertiary contacts are noted with Greek

letters.

Page 157: Copyright by Jeffrey Philip Potratz 2012

141

Figure 5.3: Nucleotides protected and exposed in the native and early folding time

point states

Nucleotides that were protected from DMS modification or made more accessible, as

compared to the unfolded state, are shown. Nucleotides that are protected to a similar

degree in both the early time point and native states are green while those that are

protected in both states but are protected significantly more in the native state are shown

in orange. Protections only observed in the native state are red while those only protected

in the early time point state are blue. Therefore, together the green, orange, and blue

nucleotides represent the protections observed in the early time point state. In addition,

nucleotides that were more accessible to DMS in the native state compared to the

unfolded state are shown in magenta. Nucleotides shown in black were not protected in

any state. Domains V and VI were not probed in the native state due to a primer issue and

were not included in the analysis for the native state (gray box). See Figure 5.5A,B and

Table 5.1 for the raw footprinting data and a list of all the nucleotides modified by DMS.

Page 158: Copyright by Jeffrey Philip Potratz 2012

142

Figure 5.4: DMS footprinting results for domain IV

Panel A shows the DMS accessible nucleotides in black while panel B shows protections

and enhancements to DMS modification by comparing the unfolded state to the other

states. Color scheme is the same as in Figure 5.3. The gray box indicates nucleotides that

were not probed in the native state.

Page 159: Copyright by Jeffrey Philip Potratz 2012

143

Nucleotide #

Page 160: Copyright by Jeffrey Philip Potratz 2012

144

Nucleotide #

Figure 5.5: DMS footprinting profiles of full-length aI5γ

The normalized signal intensity values for A and C residues for the unfolded, early

folding time point, and native states are shown. Panel A shows the profiles in the

presence of DMS and panel B shows the profiles in the absence of DMS.

Page 161: Copyright by Jeffrey Philip Potratz 2012

145

Figure 5.6: Kratky plots of static samples reveal different compaction peak heights

D135 ribozyme samples (0.5 µM) were folded for approximately twenty minutes under

specified conditions and then put on ice until exposed to X-rays at room temperature. The

states probed are native (black), Mss116p +ATP (green), Mss116p –ATP (red), early

time point (blue), and unfolded (gray). These states correspond to 100%, 80%, 40%,

30%, and 15% native ribozyme, according to a discontinuous activity assay (142).

Page 162: Copyright by Jeffrey Philip Potratz 2012

146

Figure 5.7: Time-resolved SAXS data indicate an early compaction event

Time-resolved data (thin lines in all panels) are compared to static data (thick lines in all

panels) acquired on the (A) early time point state, (B) Mss116p –ATP state, and (C)

Mss116p +ATP state. The time-resolved data are collected beginning ~3 seconds after

mixing D135 ribozyme with folding conditions and continue for ~4 minutes. Time-

resolved data are grouped together in colors from earliest to latest time points. Thin red

lines represent times from ~3-15 seconds after initiating folding. Thin green lines

represent times from ~15-30 seconds. Thin yellow lines represent times from ~40-130

Page 163: Copyright by Jeffrey Philip Potratz 2012

147

seconds. Thin blue lines represent times from ~130-220 seconds. Note that while the time

dependent signal fluctuates, there is no systematic increase in peak height as a function of

time. By comparing the first time points of the time-resolved data (red thin lines) to the

static data (thick lines), it appears that the RNA is as compact after ~3 seconds as it is

after twenty minutes. All the data were collected using 1 µM D135 ribozyme except for

the unfolded state, which used 0.5 µM ribozyme with the signal normalized to account

for this difference.

(Time-dependent data acquired by Woongsoon Choi.)

Page 164: Copyright by Jeffrey Philip Potratz 2012

148

Nucleotide reaction with DMS Total Number Color Coding in Figures

accessible 156 Black

27, 38, 41, 51, 63, 66, 67, 71, 77, 82, 92, 93, 94, 100, 101, 105, 114, 115, 117, 121, 125, 126, 127, 128,

129, 130, 133, 148, 152, 157, 159, 169, 176, 177, 180, 181, 187, 188, 191, 193, 195, 196, 198, 204, 205,

208, 209, 210, 214, 215, 217, 218, 219, 220, 225, 227, 228, 229, 230, 233, 237, 238, 240, 241, 242, 245,

251, 252, 253, 254, 259, 262, 264, 269, 272, 289, 290, 293, 307, 309, 310, 331, 332, 333, 341, 348, 355,

362, 365, 368, 382, 383, 389, 395, 403, 406, 410, 411, 412, 413, 415, 416, 418, 419, 441, 442, 444, 445,

447, 449, 463, 482, 496, 561, 585, 586, 589, 594, 598, 599, 600, 604, 605, 617, 627, 629, 630, 639, 644,

645, 646, 648, 651, 652, 653, 661, 668, 729, 745, 753, 757, 758, 762, 802, 803, 804, 806, 830, 832, 851,

860, 861, 868, 869, 876, 886

early folding time point protections 71 Blue, Green

129, 169, 180, 181, 187, 188, 191, 193, 195, 204, 205, 209, 210, 218, 219, 220, 229, 230, 233, 237, 238,

240, 241, 242, 307, 309, 310, 332, 348, 368, 389, 395, 403, 406, 410, 411, 413, 415, 416, 441, 442, 444,

445, 463, 482, 561, 585, 586, 589, 594, 599, 600, 604, 617, 627, 644, 645, 646, 652, 668, 729, 753, 757,

758, 762, 802, 803, 804, 806, 876, 886

native protections 96 Red, Green, Orange

27, 51, 63, 66, 67, 71, 77, 92, 100, 101, 105, 117, 121, 130, 133, 148, 159, 169, 180, 181, 187, 188, 191,

193, 195, 196, 198, 204, 205, 208, 209, 210, 214, 215, 217, 218, 219, 220, 227, 228, 229, 230, 233, 237,

242, 245, 259, 264, 307, 309, 310, 333, 341, 348, 382, 389, 395, 403, 406, 410, 411, 413, 415, 416, 418,

419, 441, 463, 482, 585, 586, 589, 594, 598, 599, 600, 604, 605, 617, 627, 630, 639, 644, 645, 646, 648,

651, 652, 653, 661, 668, 729, 745, 757, 758, 762

native greater protections 23 Orange

169, 187, 191, 195, 204, 205, 218, 219, 220, 229, 230, 233, 242, 348, 395, 413, 585, 586, 599, 600, 604,

617, 762

native more accessible 2 Magenta

420, 734

Table 5.1: Nucleotides modified by DMS.

All the nucleotides that were modified by DMS are included. (See Section 5.2.2 for how

these assignments were determined.)

Page 165: Copyright by Jeffrey Philip Potratz 2012

149

Appendix

A.1 PRELIMINARY DMS FOOTPRINTING OF D135 RIBOZYME USING CAPILLARY

ELECTROPHORESIS

DMS footprinting reactions were carried out using 5 pmol of D135 ribozyme. The

RNA was denatured at 92°C for 1 minute in the absence of Mg2+

and then transferred to

reaction conditions. The states of the RNA probed were ‘unfolded’ (50 mM Na-MOPS

pH 7.0), and an ‘early folding time point’ that was initiated from the unfolded conditions

with the addition of Mg2+

(50 mM Na-MOPS pH 7.0, 100 mM KCl, 8 mM MgCl2, ~30

°C for < 2.5 min). After the RNA was folded, it was incubated with 1 µL of 420 mM

DMS for 30 seconds at ~30 °C. A published protocol was followed using Cy5 labeled

primers to obtain fluorescent cDNA fragments that were separated on a Beckman CEQ

instrument (154). The cDNA profiles were analyzed using CAFA software to quantify

the peak area for each nucleotide (203). Nucleotide signals were normalized by the

average peak area.

In order for a nucleotide to be considered accessible, the unfolded value +DMS

must be ≥ 0.2 larger than the value –DMS. Similarly, in order for a nucleotide to be

considered protected the unfolded+DMS value must be ≥ 0.2 larger than the early folding

time point value+DMS.

Despite a number of differences between the footprinting studies conducted on

the full-length aI5γ and the D135 ribozyme (different constructs, different DMS

incubation times, different concentrations of RNA probed, different normalization

procedures for nucleotide signal, and different methods of determining protections), the

general conclusion of structure formation soon after the addition of Mg2+

is maintained.

Page 166: Copyright by Jeffrey Philip Potratz 2012

150

~Nucleotide #

Figure A.1: DMS footprinting profile of the D135 ribozyme

The normalized signal intensities for all the nucleotides of the D135 ribozyme are shown

for the unfolded (black) and early folding time point (blue) states in the presence of

DMS. Highlighted are nucleotides that were accessible in the unfolded state, protected in

Page 167: Copyright by Jeffrey Philip Potratz 2012

151

the early folding time point state, and were also deemed protected in the study using the

full-length aI5γ construct (see Figure 5.3).

Page 168: Copyright by Jeffrey Philip Potratz 2012

152

Bibliography

1. Roth A & Breaker RR (2009) The structural and functional diversity of

metabolite-binding riboswitches. Annu Rev Biochem 78:305-334.

2. Noller HF (2005) RNA structure: reading the ribosome. Science 309(5740):1508-

1514.

3. Collins K (2006) The biogenesis and regulation of telomerase holoenzymes. Nat

Rev Mol Cell Biol 7(7):484-494.

4. Egea PF, Stroud RM, & Walter P (2005) Targeting proteins to membranes:

structure of the signal recognition particle. Curr Opin Struct Biol 15(2):213-220.

5. Serganov A, Polonskaia A, Phan AT, Breaker RR, & Patel DJ (2006) Structural

basis for gene regulation by a thiamine pyrophosphate-sensing riboswitch. Nature

441(7097):1167-1171.

6. Winkler W, Nahvi A, & Breaker RR (2002) Thiamine derivatives bind messenger

RNAs directly to regulate bacterial gene expression. Nature 419(6910):952-956.

7. Kruger K, et al. (1982) Self-splicing RNA: autoexcision and autocyclization of

the ribosomal RNA intervening sequence of Tetrahymena. Cell 31(1):147-157.

8. Guerrier-Takada C, Gardiner K, Marsh T, Pace N, & Altman S (1983) The RNA

moiety of ribonuclease P is the catalytic subunit of the enzyme. Cell 35(3 Pt

2):849-857.

9. Greenleaf WJ, Frieda KL, Foster DA, Woodside MT, & Block SM (2008) Direct

observation of hierarchical folding in single riboswitch aptamers. Science

319(5863):630-633.

10. Brion P & Westhof E (1997) Hierarchy and dynamics of RNA folding. Annu Rev

Biophys Biomol Struct 26:113-137.

11. Tinoco I, Jr. & Bustamante C (1999) How RNA folds. J Mol Biol 293(2):271-

281.

12. Pan C & Russell R (2010) Roles of DEAD-box proteins in RNA and RNP

Folding. RNA Biol 7(6):667-676.

13. Solomatin SV, Greenfeld M, Chu S, & Herschlag D (2010) Multiple native states

reveal persistent ruggedness of an RNA folding landscape. Nature

463(7281):681-684.

14. Treiber DK & Williamson JR (1999) Exposing the kinetic traps in RNA folding.

Curr Opin Struct Biol 9(3):339-345.

15. Herschlag D (1995) RNA chaperones and the RNA folding problem. J Biol Chem

270(36):20871-20874.

16. Treiber DK, Rook MS, Zarrinkar PP, & Williamson JR (1998) Kinetic

intermediates trapped by native interactions in RNA folding. Science

279(5358):1943-1946.

17. Russell R, et al. (2006) The paradoxical behavior of a highly structured misfolded

intermediate in RNA folding. J Mol Biol 363(2):531-544.

18. Schroeder R, Grossberger R, Pichler A, & Waldsich C (2002) RNA folding in

vivo. Curr Opin Struct Biol 12(3):296-300.

Page 169: Copyright by Jeffrey Philip Potratz 2012

153

19. Doetsch M, Schroeder R, & Fürtig B (2011) Transient RNA-protein interactions

in RNA folding. FEBS J 278(10):1634-1642.

20. Russell R (2008) RNA misfolding and the action of chaperones. Front Biosci

13:1-20.

21. Linder P (2006) Dead-box proteins: a family affair--active and passive players in

RNP-remodeling. Nucleic Acids Res 34(15):4168-4180.

22. Jankowsky E (2011) RNA helicases at work: binding and rearranging. Trends

Biochem Sci 36(1):19-29.

23. Linder P (2008) mRNA export: RNP remodeling by DEAD-box proteins. Curr

Biol 18(7):R297-299.

24. Jankowsky E & Bowers H (2006) Remodeling of ribonucleoprotein complexes

with DExH/D RNA helicases. Nucleic Acids Res 34(15):4181-4188.

25. Lorsch JR (2002) RNA chaperones exist and DEAD box proteins get a life. Cell

109(7):797-800.

26. Bhaskaran H & Russell R (2007) Kinetic redistribution of native and misfolded

RNAs by a DEAD-box chaperone. Nature 449(7165):1014-1018.

27. Coetzee T, Herschlag D, & Belfort M (1994) Escherichia coli proteins, including

ribosomal protein S12, facilitate in vitro splicing of phage T4 introns by acting as

RNA chaperones. Genes Dev 8(13):1575-1588.

28. Mohr S, Stryker JM, & Lambowitz AM (2002) A DEAD-box protein functions as

an ATP-dependent RNA chaperone in group I intron splicing. Cell 109(6):769-

779.

29. Zemora G & Waldsich C (2010) RNA folding in living cells. RNA Biol 7(6):8-15.

30. Mohr S, Matsuura M, Perlman PS, & Lambowitz AM (2006) A DEAD-box

protein alone promotes group II intron splicing and reverse splicing by acting as

an RNA chaperone. Proc Natl Acad Sci U S A 103(10):3569-3574.

31. Tijerina P, Bhaskaran H, & Russell R (2006) Nonspecific binding to structured

RNA and preferential unwinding of an exposed helix by the CYT-19 protein, a

DEAD-box RNA chaperone. Proc Natl Acad Sci U S A 103(45):16698-16703.

32. Raghunathan PL & Guthrie C (1998) RNA unwinding in U4/U6 snRNPs requires

ATP hydrolysis and the DEIH-box splicing factor Brr2. Curr Biol 8(15):847-855.

33. Staley JP & Guthrie C (1999) An RNA switch at the 5' splice site requires ATP

and the DEAD box protein Prp28p. Mol Cell 3(1):55-64.

34. Fairman-Williams ME, Guenther UP, & Jankowsky E (2010) SF1 and SF2

helicases: family matters. Curr Opin Struct Biol.

35. Jankowsky E & Fairman ME (2007) RNA helicases--one fold for many functions.

Curr Opin Struct Biol 17(3):316-324.

36. Cordin O, Banroques J, Tanner NK, & Linder P (2006) The DEAD-box protein

family of RNA helicases. Gene 367:17-37.

37. Del Campo M & Lambowitz AM (2009) Structure of the Yeast DEAD box

protein Mss116p reveals two wedges that crimp RNA. Mol Cell 35(5):598-609.

Page 170: Copyright by Jeffrey Philip Potratz 2012

154

38. Sengoku T, Nureki O, Nakamura A, Kobayashi S, & Yokoyama S (2006)

Structural basis for RNA unwinding by the DEAD-box protein Drosophila Vasa.

Cell 125(2):287-300.

39. Bono F, Ebert J, Lorentzen E, & Conti E (2006) The crystal structure of the exon

junction complex reveals how it maintains a stable grip on mRNA. Cell

126(4):713-725.

40. von Moeller H, Basquin C, & Conti E (2009) The mRNA export protein DBP5

binds RNA and the cytoplasmic nucleoporin NUP214 in a mutually exclusive

manner. Nat Struct Mol Biol 16(3):247-254.

41. Andersen CB, et al. (2006) Structure of the exon junction core complex with a

trapped DEAD-box ATPase bound to RNA. Science 313(5795):1968-1972.

42. Collins R, et al. (2009) The DEXD/H-box RNA helicase DDX19 is regulated by

an {alpha}-helical switch. J Biol Chem 284(16):10296-10300.

43. Banroques J, Doere M, Dreyfus M, Linder P, & Tanner NK (2010) Motif III in

superfamily 2 "helicases" helps convert the binding energy of ATP into a high-

affinity RNA binding site in the yeast DEAD-box protein Ded1. J Mol Biol

396(4):949-966.

44. Lorsch JR & Herschlag D (1998) The DEAD box protein eIF4A. 1. A minimal

kinetic and thermodynamic framework reveals coupled binding of RNA and

nucleotide. Biochemistry 37(8):2180-2193.

45. Polach KJ & Uhlenbeck OC (2002) Cooperative binding of ATP and RNA

substrates to the DEAD/H protein DbpA. Biochemistry 41(11):3693-3702.

46. Theissen B, Karow AR, Kohler J, Gubaev A, & Klostermeier D (2008)

Cooperative binding of ATP and RNA induces a closed conformation in a DEAD

box RNA helicase. Proc Natl Acad Sci U S A 105(2):548-553.

47. Karow AR & Klostermeier D (2009) A conformational change in the helicase

core is necessary but not sufficient for RNA unwinding by the DEAD box

helicase YxiN. Nucleic Acids Res 37(13):4464-4471.

48. Lohman TM, Tomko EJ, & Wu CG (2008) Non-hexameric DNA helicases and

translocases: mechanisms and regulation. Nat Rev Mol Cell Biol 9(5):391-401.

49. Singleton MR, Dillingham MS, & Wigley DB (2007) Structure and mechanism of

helicases and nucleic acid translocases. Annu Rev Biochem 76:23-50.

50. Rossler OG, Straka A, & Stahl H (2001) Rearrangement of structured RNA via

branch migration structures catalysed by the highly related DEAD-box proteins

p68 and p72. Nucleic Acids Res 29(10):2088-2096.

51. Huang Y & Liu ZR (2002) The ATPase, RNA unwinding, and RNA binding

activities of recombinant p68 RNA helicase. J Biol Chem 277(15):12810-12815.

52. Bizebard T, Ferlenghi I, Iost I, & Dreyfus M (2004) Studies on three E. coli

DEAD-box helicases point to an unwinding mechanism different from that of

model DNA helicases. Biochemistry 43(24):7857-7866.

53. Yang Q & Jankowsky E (2006) The DEAD-box protein Ded1 unwinds RNA

duplexes by a mode distinct from translocating helicases. Nat Struct Mol Biol

13(11):981-986.

Page 171: Copyright by Jeffrey Philip Potratz 2012

155

54. Chamot D, Colvin KR, Kujat-Choy SL, & Owttrim GW (2005) RNA structural

rearrangement via unwinding and annealing by the cyanobacterial RNA helicase,

CrhR. J Biol Chem 280(3):2036-2044.

55. Halls C, et al. (2007) Involvement of DEAD-box proteins in group I and group II

intron splicing. Biochemical characterization of Mss116p, ATP hydrolysis-

dependent and -independent mechanisms, and general RNA chaperone activity. J

Mol Biol 365(3):835-855.

56. Rozen F, et al. (1990) Bidirectional RNA helicase activity of eucaryotic

translation initiation factors 4A and 4F. Mol Cell Biol 10(3):1134-1144.

57. Rogers GW, Jr., Lima WF, & Merrick WC (2001) Further characterization of the

helicase activity of eIF4A. Substrate specificity. J Biol Chem 276(16):12598-

12608.

58. Yang Q, Del Campo M, Lambowitz AM, & Jankowsky E (2007) DEAD-box

proteins unwind duplexes by local strand separation. Mol Cell 28(2):253-263.

59. Rogers GW, Jr., Richter NJ, & Merrick WC (1999) Biochemical and kinetic

characterization of the RNA helicase activity of eukaryotic initiation factor 4A. J

Biol Chem 274(18):12236-12244.

60. Del Campo M, et al. (2007) Do DEAD-box proteins promote group II intron

splicing without unwinding RNA? Mol Cell 28(1):159-166.

61. Henn A, Cao W, Hackney DD, & De La Cruz EM (2008) The ATPase cycle

mechanism of the DEAD-box rRNA helicase, DbpA. J Mol Biol 377(1):193-205.

62. Cao W, et al. (2011) Mechanism of Mss116 ATPase Reveals Functional Diversity

of DEAD-Box Proteins. J Mol Biol 409(3):399-414.

63. Kistler AL & Guthrie C (2001) Deletion of MUD2, the yeast homolog of

U2AF65, can bypass the requirement for sub2, an essential spliceosomal ATPase.

Genes Dev. 15(1):42-49.

64. Chen JY, et al. (2001) Specific alterations of U1-C protein or U1 small nuclear

RNA can eliminate the requirement of Prp28p, an essential DEAD box splicing

factor. Mol Cell 7(1):227-232.

65. Perriman R & Ares M, Jr. (2000) ATP can be dispensable for prespliceosome

formation in yeast. Genes Dev 14(1):97-107.

66. Lund MK & Guthrie C (2005) The DEAD-box protein Dbp5p is required to

dissociate Mex67p from exported mRNPs at the nuclear rim. Mol Cell 20(4):645-

651.

67. Tran EJ, Zhou Y, Corbett AH, & Wente SR (2007) The DEAD-box protein Dbp5

controls mRNA export by triggering specific RNA:protein remodeling events.

Mol Cell 28(5):850-859.

68. Fairman ME, et al. (2004) Protein displacement by DExH/D "RNA helicases"

without duplex unwinding. Science 304(5671):730-734.

69. Bowers HA, et al. (2006) Discriminatory RNP remodeling by the DEAD-box

protein DED1. RNA 12(5):903-912.

70. Nielsen KH, et al. (2009) Mechanism of ATP turnover inhibition in the EJC. RNA

15(1):67-75.

Page 172: Copyright by Jeffrey Philip Potratz 2012

156

71. Staley JP & Guthrie C (1998) Mechanical devices of the spliceosome: motors,

clocks, springs, and things. Cell 92(3):315-326.

72. Gibson TJ & Thompson JD (1994) Detection of dsRNA-binding domains in RNA

helicase A and Drosophila maleless: implications for monomeric RNA helicases.

Nucleic Acids Res. 22(13):2552-2556.

73. Dossani ZY, Weirich CS, Erzberger JP, Berger JM, & Weis K (2009) Structure of

the C-terminus of the mRNA export factor Dbp5 reveals the interaction surface

for the ATPase activator Gle1. Proc Natl Acad Sci U S A 106(38):16251-16256.

74. Napetschnig J, et al. (2009) Structural and functional analysis of the interaction

between the nucleoporin Nup214 and the DEAD-box helicase Ddx19. Proc Natl

Acad Sci U S A 106(9):3089-3094.

75. Oberer M, Marintchev A, & Wagner G (2005) Structural basis for the

enhancement of eIF4A helicase activity by eIF4G. Genes Dev 19(18):2212-2223.

76. Fukao A, et al. (2009) The ELAV protein HuD stimulates cap-dependent

translation in a Poly(A)- and eIF4A-dependent manner. Mol Cell 36(6):1007-

1017.

77. Kossen K, Karginov FV, & Uhlenbeck OC (2002) The carboxy-terminal domain

of the DExDH protein YxiN is sufficient to confer specificity for 23S rRNA. J

Mol Biol 324(4):625-636.

78. Karginov FV, Caruthers JM, Hu Y, McKay DB, & Uhlenbeck OC (2005) YxiN is

a modular protein combining a DEx(D/H) core and a specific RNA-binding

domain. J Biol Chem 280(42):35499-35505.

79. Diges CM & Uhlenbeck OC (2001) Escherichia coli DbpA is an RNA helicase

that requires hairpin 92 of 23S rRNA. EMBO J 20(19):5503-5512.

80. Tsu CA, Kossen K, & Uhlenbeck OC (2001) The Escherichia coli DEAD protein

DbpA recognizes a small RNA hairpin in 23S rRNA. RNA 7(5):702-709.

81. Del Campo M, et al. (2009) Unwinding by local strand separation is critical for

the function of DEAD-box proteins as RNA chaperones. J Mol Biol 389(4):674-

693.

82. Potratz JP, Tijerina P, & Russell R (2010) Mechanisms of DEAD-box proteins in

ATP-dependent processes. RNA helicases, ed Jankowsky E (Royal Society of

Chemistry), Vol RSC Biomolecular Sciences No. 19, pp 61-98.

83. Grohman JK, et al. (2007) Probing the mechanisms of DEAD-box proteins as

general RNA chaperones: The C-terminal domain of CYT-19 mediates general

recognition of RNA. Biochemistry 46:3013-3022.

84. Mallam AL, et al. (2011) Solution structures of DEAD-box RNA chaperones

reveal conformational changes and nucleic acid tethering by a basic tail. Proc

Natl Acad Sci U S A 108(30):12254-12259.

85. Chuang RY, Weaver PL, Liu Z, & Chang TH (1997) Requirement of the DEAD-

Box protein Ded1p for messenger RNA translation. Science 275(5305):1468-

1471.

Page 173: Copyright by Jeffrey Philip Potratz 2012

157

86. de la Cruz J, Iost I, Kressler D, & Linder P (1997) The p20 and Ded1 proteins

have antagonistic roles in eIF4E-dependent translation in Saccharomyces

cerevisiae. Proc. Natl. Acad. Sci. U.S.A. 94(10):5201-5206.

87. Linder P (2003) Yeast RNA helicases of the DEAD-box family involved in

translation initiation. Biol Cell 95(3-4):157-167.

88. Marsden S, Nardelli M, Linder P, & McCarthy JE (2006) Unwinding single RNA

molecules using helicases involved in eukaryotic translation initiation. J Mol Biol

361(2):327-335.

89. Tarn WY & Chang TH (2009) The current understanding of Ded1p/DDX3

homologs from yeast to human. RNA Biol 6(1):17-20.

90. Clodi E, Semrad K, & Schroeder R (1999) Assaying RNA chaperone activity in

vivo using a novel RNA folding trap. EMBO J 18(13):3776-3782.

91. Huang HR, et al. (2005) The splicing of yeast mitochondrial group I and group II

introns requires a DEAD-box protein with RNA chaperone function. Proc Natl

Acad Sci U S A 102(1):163-168.

92. Mohr G, et al. (2008) Function of the C-terminal domain of the DEAD-box

protein Mss116p analyzed in vivo and in vitro. J Mol Biol 375(5):1344-1364.

93. Grohman JK, et al. (2007) Probing the mechanisms of DEAD-box proteins as

general RNA chaperones: the C-terminal domain of CYT-19 mediates general

recognition of RNA. Biochemistry 46(11):3013-3022.

94. Russell R, Millett IS, Doniach S, & Herschlag D (2000) Small angle X-ray

scattering reveals a compact intermediate in RNA folding. Nat Struct Biol

7(5):367-370.

95. Tijerina P, Bhaskaran H, & Russell R (2006) Non-specific binding to structured

RNA and preferential unwinding of an exposed helix by the CYT-19 protein, a

DEAD-box RNA chaperone. Proc. Natl. Acad. Sci. U.S.A. 103:16698-16703.

96. Jarmoskaite I & Russell R (2011) DEAD-box proteins as RNA helicases and

chaperones. WIREs RNA 2(1):135-152.

97. Rogers GW, Jr., Richter NJ, & Merrick WC (1999) Biochemical and kinetic

characterization of the RNA helicase activity of eukaryotic initiation factor 4A. J.

Biol. Chem. 274(18):12236-12244.

98. Yang Q, Del Campo M, Lambowitz AM, & Jankowsky E (2007) DEAD-box

proteins unwind duplexes by local strand separation. Mol. Cell 28:253-263.

99. Zaug AJ, Grosshans CA, & Cech TR (1988) Sequence-specific endoribonuclease

activity of the Tetrahymena ribozyme: enhanced cleavage of certain

oligonucleotide substrates that form mismatched ribozyme-substrate complexes.

Biochemistry 27(25):8924-8931.

100. Pörschke D & Eigen M (1971) Co-operative non-enzymic base recognition. 3.

Kinetics of the helix-coil transition of the oligoribouridylic--oligoriboadenylic

acid system and of oligoriboadenylic acid alone at acidic pH. J Mol Biol

62(2):361-381.

101. Craig ME, Crothers DM, & Doty P (1971) Relaxation kinetics of dimer formation

by self complementary oligonucleotides. J Mol Biol 62(2):383-401.

Page 174: Copyright by Jeffrey Philip Potratz 2012

158

102. Serra MJ & Turner DH (1995) Predicting thermodynamic properties of RNA.

Methods Enzymol 259:242-261.

103. Dillingham MS, Wigley DB, & Webb MR (2000) Demonstration of

unidirectional single-stranded DNA translocation by PcrA helicase: measurement

of step size and translocation speed. Biochemistry 39(1):205-212.

104. Martinez-Senac MM & Webb MR (2005) Mechanism of translocation and

kinetics of DNA unwinding by the helicase RecG. Biochemistry 44(51):16967-

16976.

105. Mackintosh SG & Raney KD (2006) DNA unwinding and protein displacement

by superfamily 1 and superfamily 2 helicases. Nucleic Acids Res 34(15):4154-

4159.

106. Grifo JA, Abramson RD, Satler CA, & Merrick WC (1984) RNA-stimulated

ATPase activity of eukaryotic initiation factors. J Biol Chem 259(13):8648-8654.

107. Peck ML & Herschlag D (1999) Effects of oligonucleotide length and atomic

composition on stimulation of the ATPase activity of translation initiation factor

elF4A. RNA 5(9):1210-1221.

108. Herschlag D (1995) RNA chaperones and the RNA folding problem. J. Biol.

Chem. 270(36):20871-20874.

109. Schroeder R, Barta A, & Semrad K (2004) Strategies for RNA folding and

assembly. Nat Rev Mol Cell Biol 5(11):908-919.

110. Liu F, Putnam A, & Jankowsky E (2008) ATP hydrolysis is required for DEAD-

box protein recycling but not for duplex unwinding. Proc Natl Acad Sci U S A

105(51):20209-20214.

111. Karbstein K, Lee J, & Herschlag D (2007) Probing the role of a secondary

structure element at the 5'- and 3'-splice sites in group I intron self-splicing: the

tetrahymena L-16 ScaI ribozyme reveals a new role of the G.U pair in self-

splicing. Biochemistry 46(16):4861-4875.

112. Allain FH & Varani G (1995) Structure of the P1 helix from group I self-splicing

introns. J Mol Biol 250(3):333-353.

113. Patel SS & Donmez I (2006) Mechanisms of helicases. J. Biol. Chem.

281(27):18265-18268.

114. Ray BK, et al. (1985) ATP-dependent unwinding of messenger RNA structure by

eukaryotic initiation factors. J. Biol. Chem. 260(12):7651-7658.

115. Yang Q & Jankowsky E (2005) ATP- and ADP-dependent modulation of RNA

unwinding and strand annealing activities by the DEAD-box protein DED1.

Biochemistry 44(41):13591-13601.

116. Yang Q & Jankowsky E (2006) The DEAD-box protein Ded1 unwinds RNA

duplexes by a mode distinct from translocating helicases. Nat. Struct. Mol. Biol.

13(11):981-986.

117. Konarska MM, Vilardell J, & Query CC (2006) Repositioning of the reaction

intermediate within the catalytic center of the spliceosome. Mol Cell 21(4):543-

553.

Page 175: Copyright by Jeffrey Philip Potratz 2012

159

118. Wan Y, Suh H, Russell R, & Herschlag D (2010) Multiple unfolding events

during native folding of the Tetrahymena group I ribozyme. J Mol Biol.

119. Russell R & Herschlag D (2001) Probing the folding landscape of the

Tetrahymena ribozyme: commitment to form the native conformation is late in

the folding pathway. J Mol Biol 308(5):839-851.

120. Chadee AB, Bhaskaran H, & Russell R (2010) Protein roles in group I intron

RNA folding: the tyrosyl-tRNA synthetase CYT-18 stabilizes the native state

relative to a long-lived misfolded structure without compromising folding

kinetics. J Mol Biol 395(3):656-670.

121. Wan Y, Suh H, Russell R, & Herschlag D (2010) Multiple unfolding events

during native folding of the Tetrahymena group I ribozyme. J Mol Biol

400(5):1067-1077.

122. Herschlag D, Khosla M, Tsuchihashi Z, & Karpel RL (1994) An RNA chaperone

activity of non-specific RNA binding proteins in hammerhead ribozyme catalysis.

EMBO J 13(12):2913-2924.

123. Tsuchihashi Z, Khosla M, & Herschlag D (1993) Protein enhancement of

hammerhead ribozyme catalysis. Science 262(5130):99-102.

124. Semrad K & Schroeder R (1998) A ribosomal function is necessary for efficient

splicing of the T4 phage thymidylate synthase intron in vivo. Genes Dev

12(9):1327-1337.

125. Semrad K, Green R, & Schroeder R (2004) RNA chaperone activity of large

ribosomal subunit proteins from Escherichia coli. RNA 10(12):1855-1860.

126. Rajkowitsch L, Semrad K, Mayer O, & Schroeder R (2005) Assays for the RNA

chaperone activity of proteins. Biochem Soc Trans 33(Pt 3):450-456.

127. Zaug AJ & Cech TR (1986) The intervening sequence RNA of Tetrahymena is an

enzyme. Science 231(4737):470-475.

128. Tanner M & Cech T (1996) Activity and thermostability of the small self-splicing

group I intron in the pre-tRNA(lle) of the purple bacterium Azoarcus. RNA

2(1):74-83.

129. Kuo LY, Davidson LA, & Pico S (1999) Characterization of the Azoarcus

ribozyme: tight binding to guanosine and substrate by an unusually small group I

ribozyme. Biochim Biophys Acta 1489(2-3):281-292.

130. Qin PZ & Pyle AM (1997) Stopped-flow fluorescence spectroscopy of a group II

intron ribozyme reveals that domain 1 is an independent folding unit with a

requirement for specific Mg2+ ions in the tertiary structure. Biochemistry

36(16):4718-4730.

131. Swisher J, Duarte CM, Su LJ, & Pyle AM (2001) Visualizing the solvent-

inaccessible core of a group II intron ribozyme. EMBO J 20(8):2051-2061.

132. Walstrum SA & Uhlenbeck OC (1990) The self-splicing RNA of Tetrahymena is

trapped in a less active conformation by gel purification. Biochemistry

29(46):10573-10576.

Page 176: Copyright by Jeffrey Philip Potratz 2012

160

133. Emerick VL, Pan J, & Woodson SA (1996) Analysis of rate-determining

conformational changes during self-splicing of the Tetrahymena intron.

Biochemistry 35(41):13469-13477.

134. Russell R & Herschlag D (1999) New pathways in folding of the Tetrahymena

group I RNA enzyme. J Mol Biol 291(5):1155-1167.

135. Zhang L, Xiao M, Lu C, & Zhang Y (2005) Fast formation of the P3-P7

pseudoknot: a strategy for efficient folding of the catalytically active ribozyme.

RNA 11(1):59-69.

136. Sinan S, Yuan X, & Russell R (2011) The Azoarcus Group I Intron Ribozyme

Misfolds and Is Accelerated for Refolding by ATP-dependent RNA Chaperone

Proteins. J Biol Chem 286(43):37304-37312.

137. Fedorova O, Waldsich C, & Pyle AM (2007) Group II intron folding under near-

physiological conditions: collapsing to the near-native state. J Mol Biol

366(4):1099-1114.

138. Fedorova O, Solem A, & Pyle AM (2010) Protein-facilitated folding of group II

intron ribozymes. J Mol Biol.

139. Steiner M, Karunatilaka KS, Sigel RK, & Rueda D (2008) Single-molecule

studies of group II intron ribozymes. Proc Natl Acad Sci U S A 105(37):13853-

13858.

140. Karunatilaka KS, Solem A, Pyle AM, & Rueda D (2010) Single-molecule

analysis of Mss116-mediated group II intron folding. Nature 467:935-939.

141. Pyle AM & Green JB (1994) Building a kinetic framework for group II intron

ribozyme activity: quantitation of interdomain binding and reaction rate.

Biochemistry 33(9):2716-2725.

142. Potratz JP, Del Campo M, Wolf RZ, Lambowitz AM, & Russell R (2011) ATP-

Dependent Roles of the DEAD-Box Protein Mss116p in Group II Intron Splicing

In Vitro and In Vivo. J Mol Biol 411(3):661-679.

143. Pan C & Russell R (2010) Roles of DEAD-box proteins in RNA and RNP

Folding. RNA Biol 7(6):667-676.

144. Shcherbakova I, Mitra S, Laederach A, & Brenowitz M (2008) Energy barriers,

pathways, and dynamics during folding of large, multidomain RNAs. Curr Opin

Chem Biol 12(6):655-666.

145. Chin K & Pyle AM (1995) Branch-point attack in group II introns is a highly

reversible transesterification, providing a potential proofreading mechanism for

5'-splice site selection. RNA 1(4):391-406.

146. Golden BL & Cech TR (1996) Conformational switches involved in orchestrating

the successive steps of group I RNA splicing. Biochemistry 35(12):3754-3763.

147. Woodson SA & Cech TR (1989) Reverse self-splicing of the tetrahymena group I

intron: implication for the directionality of splicing and for intron transposition.

Cell 57(2):335-345.

148. Karbstein K, Carroll KS, & Herschlag D (2002) Probing the Tetrahymena group I

ribozyme reaction in both directions. Biochemistry 41(37):11171-11183.

Page 177: Copyright by Jeffrey Philip Potratz 2012

161

149. Zingler N, Solem A, & Pyle AM (2010) Dual roles for the Mss116 cofactor

during splicing of the ai5g group II intron. Nucleic Acids Res.

150. Nolte A, Chanfreau G, & Jacquier A (1998) Influence of substrate structure on in

vitro ribozyme activity of a group II intron. RNA 4(6):694-708.

151. Wan Y, Mitchell D, & Russell R (2010) Catalytic activity as a probe of native

RNA folding. Methods in Enzymology 468:195-218.

152. Pan J & Woodson SA (1998) Folding intermediates of a self-splicing RNA:

mispairing of the catalytic core. J Mol Biol 280(4):597-609.

153. Sclavi B, Woodson S, Sullivan M, Chance MR, & Brenowitz M (1997) Time-

resolved synchrotron X-ray "footprinting", a new approach to the study of nucleic

acid structure and function: application to protein-DNA interactions and RNA

folding. J Mol Biol 266(1):144-159.

154. Tijerina P, Mohr S, & Russell R (2007) DMS footprinting of structured RNAs

and RNA-protein complexes. Nat Protoc 2(10):2608-2623.

155. King PA, Jamison E, Strahs D, Anderson VE, & Brenowitz M (1993)

'Footprinting' proteins on DNA with peroxonitrous acid. Nucleic Acids Res

21(10):2473-2478.

156. Mortimer SA & Weeks KM (2007) A fast-acting reagent for accurate analysis of

RNA secondary and tertiary structure by SHAPE chemistry. J Am Chem Soc

129(14):4144-4145.

157. Mortimer SA & Weeks KM (2008) Time-resolved RNA SHAPE chemistry. J Am

Chem Soc 130(48):16178-16180.

158. Mortimer SA & Weeks KM (2009) Time-resolved RNA SHAPE chemistry:

quantitative RNA structure analysis in one-second snapshots and at single-

nucleotide resolution. Nat Protoc 4(10):1413-1421.

159. Sclavi B, Sullivan M, Chance MR, Brenowitz M, & Woodson SA (1998) RNA

folding at millisecond intervals by synchrotron hydroxyl radical footprinting.

Science 279(5358):1940-1943.

160. Laederach A, Shcherbakova I, Liang MP, Brenowitz M, & Altman RB (2006)

Local kinetic measures of macromolecular structure reveal partitioning among

multiple parallel pathways from the earliest steps in the folding of a large RNA

molecule. J Mol Biol 358(4):1179-1190.

161. Laederach A, Shcherbakova I, Jonikas MA, Altman RB, & Brenowitz M (2007)

Distinct contribution of electrostatics, initial conformational ensemble, and

macromolecular stability in RNA folding. Proc Natl Acad Sci U S A

104(17):7045-7050.

162. Mitra S, Laederach A, Golden BL, Altman RB, & Brenowitz M (2011) RNA

molecules with conserved catalytic cores but variable peripheries fold along

unique energetically optimized pathways. RNA 17(8):1589-1603.

163. Kwok LW, et al. (2006) Concordant exploration of the kinetics of RNA folding

from global and local perspectives. J Mol Biol 355(2):282-293.

Page 178: Copyright by Jeffrey Philip Potratz 2012

162

164. Schlatterer JC & Brenowitz M (2009) Complementing global measures of RNA

folding with local reports of backbone solvent accessibility by time resolved

hydroxyl radical footprinting. Methods 49(2):142-147.

165. Roh JH, et al. (2010) Multistage collapse of a bacterial ribozyme observed by

time-resolved small-angle x-ray scattering. J Am Chem Soc.

166. Bokinsky G & Zhuang X (2005) Single-molecule RNA folding. Acc Chem Res

38(7):566-573.

167. Zhuang X, et al. (2000) A single-molecule study of RNA catalysis and folding.

Science 288(5473):2048-2051.

168. Bokinsky G, et al. (2003) Single-molecule transition-state analysis of RNA

folding. Proc Natl Acad Sci U S A 100(16):9302-9307.

169. Rueda D, et al. (2004) Single-molecule enzymology of RNA: essential functional

groups impact catalysis from a distance. Proc Natl Acad Sci U S A

101(27):10066-10071.

170. Zhuang X (2005) Single-molecule RNA science. Annu Rev Biophys Biomol Struct

34:399-414.

171. Liu S, Bokinsky G, Walter NG, & Zhuang X (2007) Dissecting the multistep

reaction pathway of an RNA enzyme by single-molecule kinetic "fingerprinting".

Proc Natl Acad Sci U S A 104(31):12634-12639.

172. Cech TR (1986) The generality of self-splicing RNA: relationship to nuclear

mRNA splicing. Cell 44(2):207-210.

173. Narlikar GJ & Herschlag D (1997) Mechanistic aspects of enzymatic catalysis:

lessons from comparison of RNA and protein enzymes. Annu Rev Biochem 66:19-

59.

174. Pyle AM, Fedorova O, & Waldsich C (2007) Folding of group II introns: a model

system for large, multidomain RNAs? Trends Biochem Sci 32(3):138-145.

175. Lambowitz AM & Zimmerly S (2010) Group II Introns: Mobile Ribozymes that

Invade DNA. Cold Spring Harb Perspect Biol.

176. Solem A, Zingler N, & Pyle AM (2006) A DEAD protein that activates intron

self-splicing without unwinding RNA. Mol Cell 24(4):611-617.

177. Bifano AL & Caprara MG (2008) A DExH/D-box protein coordinates the two

steps of splicing in a group I intron. J Mol Biol 383(3):667-682.

178. Duncan CD & Weeks KM (2008) SHAPE analysis of long-range interactions

reveals extensive and thermodynamically preferred misfolding in a fragile group I

intron RNA. Biochemistry 47(33):8504-8513.

179. Toor N, Keating KS, Taylor SD, & Pyle AM (2008) Crystal structure of a self-

spliced group II intron. Science 320(5872):77-82.

180. Dai L, et al. (2008) A three-dimensional model of a group II intron RNA and its

interaction with the intron-encoded reverse transcriptase. Mol Cell 30(4):472-485.

181. Michel F, Costa M, & Westhof E (2009) The ribozyme core of group II introns: a

structure in want of partners. Trends Biochem Sci 34(4):189-199.

182. Peebles CL, et al. (1986) A self-splicing RNA excises an intron lariat. Cell

44(2):213-223.

Page 179: Copyright by Jeffrey Philip Potratz 2012

163

183. Swisher JF, Su LJ, Brenowitz M, Anderson VE, & Pyle AM (2002) Productive

folding to the native state by a group II intron ribozyme. J Mol Biol 315(3):297-

310.

184. Su LJ, Brenowitz M, & Pyle AM (2003) An alternative route for the folding of

large RNAs: apparent two-state folding by a group II intron ribozyme. J Mol Biol

334(4):639-652.

185. Huang HR (2004) Functional studies of intron- and nuclear-encoded splicing

factors in the mitochondria of Saccharomyces cerevisiae, PhD thesis, The

University of Texas Southwestern Medical Center, Dallas.

186. Jarrell KA, Dietrich RC, & Perlman PS (1988) Group II intron domain 5

facilitates a trans-splicing reaction. Mol Cell Biol 8(6):2361-2366.

187. Amberg DC, Burke D, & Strathern JN (2005) Methods in yeast genetics: a Cold

Spring Harbor Laboratory course manual (CSHL Press).

188. Yaffe MP & Schatz G (1984) Two nuclear mutations that block mitochondrial

protein import in yeast. Proc Natl Acad Sci U S A 81(15):4819-4823.

189. Wan Y, Mitchell D, & Russell R (2009) Catalytic activity as a probe of native

RNA folding. Methods in Enzymology 468:195-218.

190. Su LJ, Waldsich C, & Pyle AM (2005) An obligate intermediate along the slow

folding pathway of a group II intron ribozyme. Nucleic Acids Res 33(21):6674-

6687.

191. Bifano AL, Turk EM, & Caprara MG (2010) Structure-guided mutational analysis

of a yeast DEAD-box protein involved in mitochondrial RNA splicing. J Mol Biol

398(3):429-443.

192. Woodson SA & Cech TR (1991) Alternative secondary structures in the 5' exon

affect both forward and reverse self-splicing of the Tetrahymena intervening

sequence RNA. Biochemistry 30(8):2042-2050.

193. Woodson SA & Emerick VL (1993) An alternative helix in the 26S rRNA

promotes excision and integration of the Tetrahymena intervening sequence. Mol

Cell Biol 13(2):1137-1145.

194. Pause A & Sonenberg N (1992) Mutational analysis of a DEAD box RNA

helicase: the mammalian translation initiation factor eIF-4A. EMBO J 11(7):2643-

2654.

195. Waldsich C & Pyle AM (2007) A folding control element for tertiary collapse of a

group II intron ribozyme. Nat Struct Mol Biol 14(1):37-44.

196. Chen Y, et al. (2008) DEAD-box proteins can completely separate an RNA

duplex using a single ATP. Proc Natl Acad Sci U S A 105(51):20203-20208.

197. Cannone JJ, et al. (2002) The comparative RNA web (CRW) site: an online

database of comparative sequence and structure information for ribosomal, intron,

and other RNAs. BMC Bioinformatics 3:2.

198. Mohr G, Caprara MG, Guo Q, & Lambowitz AM (1994) A tyrosyl-tRNA

synthetase can function similarly to an RNA structure in the Tetrahymena

ribozyme. Nature 370(6485):147-150.

Page 180: Copyright by Jeffrey Philip Potratz 2012

164

199. Weeks KM & Cech TR (1995) Efficient protein-facilitated splicing of the yeast

mitochondrial bI5 intron. Biochemistry 34(23):7728-7738.

200. Das R, Laederach A, Pearlman SM, Herschlag D, & Altman RB (2005) SAFA:

semi-automated footprinting analysis software for high-throughput quantification

of nucleic acid footprinting experiments. RNA 11(3):344-354.

201. Laederach A, et al. (2008) Semiautomated and rapid quantification of nucleic acid

footprinting and structure mapping experiments. Nat Protoc 3(9):1395-1401.

202. Waldsich C & Pyle AM (2008) A kinetic intermediate that regulates proper

folding of a group II intron RNA. J Mol Biol 375(2):572-580.

203. Mitra S, Shcherbakova IV, Altman RB, Brenowitz M, & Laederach A (2008)

High-throughput single-nucleotide structural mapping by capillary automated

footprinting analysis. Nucleic Acids Res 36(11):e63.