189
RNA TERMINUS CHEMISTRY POTENTIATES DECAY EVENTS THAT TARGET HAC1 mRNA DURING THE UNFOLDED PROTEIN RESPONSE. by Patrick Cherry B.A. Hendrix College, 2013 A thesis submitted to the Graduate School of the University of Colorado in partial fulfilment of the requirements for the degree of Doctor of Philosophy Molecular Biology Program 2019

RNA TERMINUS CHEMISTRY POTENTIATES DECAY

Embed Size (px)

Citation preview

RNA TERMINUS CHEMISTRY POTENTIATES DECAY EVENTS THAT

TARGET HAC1 mRNA DURING THE UNFOLDED PROTEIN RESPONSE.

by

Patrick Cherry

B.A. Hendrix College, 2013

A thesis submitted to the

Graduate School of the

University of Colorado in partial fulfilment

of the requirements for the degree of

Doctor of Philosophy

Molecular Biology Program

2019

ii

This dissertation for the Doctor of Philosophy degree of

Patrick Cherry

has been approved for the

Molecular Biology Program

by

Sandy Martin, Chair

Richard Davis

Jeffrey Kieft

Michael McMurray

James Costello

Jay Hesselberth, Advisor

Date: May 17, 2019

iii

Cherry, Patrick (PhD, Molecular Biology Program)

RNA-terminus chemistry potentiates decay events that target HAC1 mRNA during splicing to

regulate the unfolded protein response.

Thesis directed by Associate Professor Jay R. Hesselberth.

ABSTRACT

RNA repair enzymes catalyze rejoining of an RNA molecule after cleavage of

phosphodiester linkages. RNA repair in budding yeast is catalyzed by two separate enzymes

that process tRNA exons during their splicing and HAC1 mRNA exons during activation of the

unfolded protein response (UPR). The RNA ligase Trl1 joins 2′,3′-cyclic phosphate and 5′-

hydroxyl RNA fragments, creating a new phosphodiester linkage with a 2′-phosphate at the

junction. The 2′-phosphate is subsequently removed by the 2′-phosphotransferase Tpt1, which

catalyzes phosphate transfer to NAD+, producing nicotinamide and a unique ADP-ribose

metabolite. I bypassed the essential functions of TRL1 and TPT1 in budding yeast by expressing

“pre-spliced”/intronless versions of the ten normally intron-containing tRNAs, indicating this

repair pathway does not have additional essential functions. Consistent with previous studies,

expression of intronless tRNAs failed to rescue the growth of cells with deletions in components

of the SEN complex, implying an additional essential role for the splicing endonuclease. Finally,

I optimized a PCR-based method to detect RNA 2′-phosphate modifications and show that they

are present on ligated HAC1 mRNA.

In the unfolded protein response, stress in the endoplasmic reticulum (ER) activates a large

transcriptional program to increase ER folding capacity. During the budding yeast UPR, Ire1

excises an intron from the HAC1 mRNA and the exon products of cleavage are ligated, and the

translated protein induces dozens of stress-response genes. The trl1∆ and tpt1∆ mutants

accumulate tRNA and HAC1 splicing intermediates indicative of specific RNA repair defects. As

iv

expected, failure to induce the unfolded protein response in trl1∆ cells grown with tunicamycin

is lethal owing to their inability to ligate HAC1 after its cleavage by Ire1. In contrast, tpt1∆ mutants

grow in the presence of tunicamycin despite reduced accumulation of spliced HAC1 mRNA. I

show that phosphorylation of two different HAC1 splicing intermediates is required for their

degradation by the 5′→3′ exonuclease Xrn1 to enact opposing effects on the UPR. I also found

that ligated but 2′-phosphorylated HAC1 mRNA is cleaved, yielding a decay intermediate with

both 5′- and 2′-phosphates at its 5′-end that inhibit 5′→3′ decay. These decay events expand

the scope of RNA-based regulation in the budding yeast UPR, and these RNA repair mutants

enable new studies of the role of RNA repair in cellular physiology.

The form and content of this abstract are approved. I recommend its publication.

Approved: Jay R. Hesselberth

v

DEDICATION

I dedicate this dissertation to my parents. Absolutely none of this would be possible

without your unconditional love and encouragement. I dedicate this work to my dad, who shares

with me the same sentiments of hard work and fairness, who supports me always, and who is

full of wise advice for the good times and the bad times. I dedicate this work to Mar, who came

into my life later than most parents, but who accepted instant parenthood and did an outstanding

job. Mar modeled values I now hold close, like equality in opportunity, acceptance, and honesty

(and continues to do so). This is for you.

vi

ACKNOWLEDGEMENTS

Firstly, I must acknowledge and thank my advisor, Jay Hesselberth, for his tireless

encouragement and positivity. Jay’s intense curiosity is contagious. Solely through

inquisitiveness have we discovered some of the most interesting behaviors of RNA. Jay

accepted me into the lab from a cold email I sent him the summer before starting graduate

school, and since then it has been a great fit. I thank Jay for his mentorship in my growth as a

scientist. Thank you, Jay.

I also thank my committee, Sandy Martin, Dick Davis, Michael McMurray, Jeff Kieft, and

Jim Costello. Thank you for taking the time to mentor me on the most perplexing problems in

my project, as well as for the encouragement to reach higher.

Many mentors have guided me to this point from being just a freshman in undergrad

looking for something more challenging than coursework to work on: Joy Sturtevant, my first

science mentor, introduced me to the amazing power of yeast genetics—in Candida albicans no

less—and to the workings of academic labs at the Louisiana State University Health Sciences

Center in New Orleans, LA. Michael Shiloh at University of Texas Southwestern Medical Center

in Dallas, TX thought me perseverance in the face of recalcitrant cloning problems. Andres Caro

at Hendrix College in Conway, AR taught both my Biochemistry I and II courses with Oscar-

worthy acting of “so, you are running from a bear,” as the recurring stimulus for epinephrine

release. Andrew Schurko, also at Hendrix, taught a fantastic genetics class with Nasonia

vitripennis wasps as genetic models and let my science presentations get into the weeds in

Advanced Genetics. Liz Gron, also at Hendrix, trained me in the methods and reasoning of

analytical chemistry, and got in some future grad school and career mentoring between lessons

and labs.

All members of the Hesselberth lab have helped me succeed in graduate school. Laura

White has been my partner-in-science for virtually all endeavors, from “paired science” (not

vii

typically recommended) to maintaining lab equipment to inviting speakers. I co-first-authored

my very first paper with Laura, and in doing so I discovered a special bond that comes from the

mutual appreciation of scientific talent and thought. Additionally, Sally Peach advanced so much

of this project before I took over and was a mentor and teacher of programming to me and many

others. I arrived in the lab my first year to Kerri York, Monica Ransom, and Suzi Brian, all of whom

taught me many yeast, nucleic acid, and protein manipulation techniques, along with some

graduate school survival mentoring. Mandy Richer’s balancing of outdoor excursions with lab

responsibilities inspired me to keep up my other hobbies while remaining dedicated to my thesis

project. Also, Mandy’s ongoing fight for fair treatment against the RNA Mafia is working;

evidence of this hangs outside Jay’s office door to this day. I appreciate Rachel Ancar’s presence

in the lab, her patient listening skills, her fearless approach to trying new things in science, and

her slightly absurd sense of humor. Chris Snowden, Jackie Pierce, Maggie Balas, Alexis

Zukowski are all “next door lab neighbors,” and so received the brunt of my questions about

methods and ideas that Hesselberth lab members were not familiar. For all of you, I am so

thankful.

Sometimes there is an easier way to get to the good, fun science in a project, and cloning-

by-phone can save time and frustration. Strains, samples, and knowledge have been dropped

into my life from some fantastic people, whom I would like to thank now. Roy Parker, Stewart

Shuman, Beate Schwer, Anita Hopper, Eric Phizicky, Jane Jackman, Erich Chapman, Jeff Moore,

Dave Barton, and Volker Thiel.

I am lucky to have so many mentors and PIs looking out for me at the University. Bob

Sclafani, Rytis Prekeris, David Engelke, Lori Sussel, Suja Jagannathan, Chad Pearson, and Tânia

Reis all lent me support and advice. Sabrena Heilman, Caitlin Moloney, and Michele Hwozdyk-

Parsons helped me manage my responsibilities to the Molecular Biology Program and the

viii

Graduate School. Thank you to the Molecular Biology Program, the Department of Biochemistry

and Molecular Genetics, The RNA Bioscience Initiative, and Victor W. & Earleen D. Bolie.

Among the ranks of the RNA Mafia are lab mates Rachel Ancar and Laura White, as well

as Maggie Balas, Ryan Sheridan, Erik Hartwick, Lena Steckelberg, Ben Akiyama, Marissa Rhule,

and Laura Hudish. I am thankful for my friends in Molecular Biology program: Molly Kingsley,

Jenn Rabe, Cassi Estrem, Sara Flubacher, Sara Espinosa, James Till, Ryan Sheridan, as well as

my graduate school cohort outside of the program: Mike Shaffer, Kyle Smith, Nicolle Witte, Tania

Eliseeva, Brooke Sinnen, Joe Bednarek, Alex Barret, Katie Mishall Barret, and Travis Nemkov.

Having friends outside of science is healthy, and even fun. My non-science friends

arguably enrich my life so much more than the science friends on my quest to be an instrument

of appreciation, and I share with them bonds formed in the struggles of climbing, biking,

camping, hiking, gaming, pop-culture-critiquing [drag], and occasionally free-soloing (sorry,

Mom and Dad). So I am thankful for Monty Prekeris, Jake Beck, Nick Goedecke, Jack Rugile,

Mike Gorodinsky, Alex Hart, Alex Imhof, Mark Rohr, Nick Turner, Fred Gravagna, Denise Meyer,

Evan Robertson, Ben Parefsky, Gina Loftus, Mike Slater, Camila Restrepo, and Lena Gerber.

I am so thankful to my peers who believed in me in undergrad: Cheryl Mathis, Joana Ortiz-

Baca, Johnny Tran, Ina Agee, Lauren Irby, Gabe Gonzales, Rima AbiSamra, Sloan Zimmerman,

and Olivia Urbanoitz. I thank Emily Winters for inspiring me to chase my dreams, one skype

session at a time. Thank you to my high school science buddies Justin Shapiro and Matt

McTernan for getting through grad school alongside me; science wouldn’t be as fun without your

A-day-ing around.

Thank you to my parents, for so much love, understanding, and encouragement.

The science and learning have been fantastic, but the memories and connections to my

favorite people in life are what make grind of graduate school feel worthwhile. Thank you all, so

much.

ix

TABLE OF CONTENTS

CHAPTER

I INTRODUCTION ............................................................................................................... 1

Endoribonucleases impose terminus chemistry ......................................................... 2

Intrinsic RNA cleavage .................................................................................... 5

Self-cleaving ribozymes .................................................................................. 6

Sen2 & Sen34 .................................................................................................. 7

Las1, Grc3, & 35S rRNA processing ............................................................... 9

Stress-induced tRNA cleavage ..................................................................... 11

Ribotoxins in interspecies conflict ................................................................ 12

RNA decay in co-translational mRNA surveillance ....................................... 14

Ire1, the “splicing” endoribonuclease of the unfolded protein response (UPR) ............................................................................................. 17

RNA terminus modification ....................................................................................... 19

Clp1 RNA 5′-kinase ....................................................................................... 20

Grc3 RNA 5′-kinase ...................................................................................... 20

CNP (cyclic nucleotide phosphodiesterase) ................................................. 21

RtcA (3′-terminal phosphate cyclase) ........................................................... 21

RNA ligases combine RNA terminus modification domains with a ligase domain. .. 22

Trl1, the ligase of fungi & plants .................................................................... 22

RtcB, the ligase of prokaryotes & animals .................................................... 23

Fidelity of RNA ligases .................................................................................. 25

RNA repair in cellular physiology .............................................................................. 26

The splicing of tRNAs .................................................................................... 26

The unfolded protein response (UPR) ........................................................... 28

x

RNA repair in bacterial stress responses and interspecies conflict .............. 35

Terminus chemistry and exonucleolytic RNA decay ................................................ 37

5′→3′ RNA decay enzymes require substrates be 5′-phosphorylated ........ 38

3′→5′ RNA degradation ................................................................................ 42

Kinase-Mediated Decay ............................................................................................ 44

KMD of tRNA introns ..................................................................................... 44

KMD of rRNA processing intermediates ....................................................... 46

KMD of 3′-fragments of no-go mRNA decay (NGD) ..................................... 47

KMD of T4 bacteriophage mRNA .................................................................. 47

Toward mutants of RNA repair ................................................................................. 48

II GENETIC BYPASS OF ESSENTIAL RNA REPAIR ENZYMES IN BUDDING YEAST ...... 50

Abstract ..................................................................................................................... 50

Introduction ............................................................................................................... 51

Materials & Methods ................................................................................................. 53

General Methods ........................................................................................... 53

Northern blotting ........................................................................................... 54

Detection of 2′-Phosphate linkages by RT-PCR .......................................... 54

Expression vector for intronless tRNAs ........................................................ 55

HAC1 epitope tagging and western blotting ................................................. 55

Results & Discussion ................................................................................................. 60

Genetic bypass of essential RNA repair genes in budding yeast ................. 60

RNA repair mutants have defects in translation ............................................ 63

RNA repair mutants accumulate intermediates and products of tRNA splicing .......................................................................................................... 68

RNA repair mutants have defects in unfolded protein response activation ....................................................................................................... 72

xi

Summary ................................................................................................................... 78

III MULTIPLE DECAY EVENTS TARGET HAC1 mRNA DURING SPLICING TO REGULATE

THE UNFOLDED PROTEIN RESPONSE ........................................................................ 80

Abstract ..................................................................................................................... 80

Introduction ............................................................................................................... 80

Materials & Methods ................................................................................................. 83

Cell culture and RNA preparation ................................................................. 83

RT-PCR/qPCR .............................................................................................. 83

Primer Extension ........................................................................................... 83

Northern blotting ........................................................................................... 84

Yeast strains and plasmids ........................................................................... 84

Results ...................................................................................................................... 89

RNA repair mutants have unique HAC1 mRNA processing defects. ............ 89

Kinase-mediated decay of cleaved HAC1 3′-exon competes with its ligation. .......................................................................................................... 92

Kinase-mediated decay of excised intron is required for HAC1 translation. ..................................................................................................... 96

Incompletely processed HAC1 mRNA is endonucleolytically cleaved and degraded .............................................................................................. 101

Discussion ............................................................................................................... 110

Summary ................................................................................................................. 115

IV FUTURE DIRECTIONS .................................................................................................. 117

Does kinase-mediated decay also regulate Xbp1 splicing in animals? .................. 117

Identify precise HAC1s secondary cleavage location and nuclease. ...................... 120

Applications of 2′-phosphorylated RNA to enhance stability in vivo ...................... 123

xii

Transcriptome-wide sequencing of products of RNA repair via enrichment of 2′-PO4

RNAs from tpt1∆ yeast ........................................................................................... 124

“Fungification” of metazoan cells to replace RtcB with Trl1 for marking repaired

transcripts with 2′-PO4. ........................................................................................... 126

REFERENCES ........................................................................................................................... 128

xiii

LIST OF FIGURES

Figure 1.1: Mechanisms of endoribonuclease cleavage. ............................................................. 4

Figure 1.2: Model of RNA processing in tRNA splicing ................................................................ 8

Figure 1.3: Mechanisms of ligation. ............................................................................................ 24

Figure 1.4: Model of S. cerevisiae unfolded protein response activation ................................... 33

Figure 1.5: Three key mechanisms that couple HAC1 splicing to its translation. ...................... 34

Figure 1.6: Typical mRNA decay in the cytoplasm by Xrn1 and the exosome. ......................... 41

Figure 1.7: Examples of Kinase-Mediated Decay (KMD) ............................................................ 45

Figure 2.1: Genetic bypass of essential components of tRNA splicing with intronless tRNAs. . 61

Figure 2.2: Growth phenotype of RNA repair mutants. .............................................................. 64

Figure 2.3: tRNA processing phenotypes of RNA repair mutants. ............................................. 69

Figure 2.4: UPR-related phenotypes of RNA repair mutants. .................................................... 74

Figure 3.1. HAC1 mRNA processing defects in RNA repair and decay mutants. ...................... 90

Figure 3.2. Kinase-mediated decay of HAC1 3′-exon competes with its ligation. ..................... 93

Figure 3.3: Kinase-mediated decay of excised HAC1 intron is required to activate the unfolded

protein response. ............................................................................................................ 97

Figure 3.4: Incompletely processed HAC1s mRNA is cleaved and degraded. ........................ 102

Figure 3.5: A 5′- and 2′-phosphorylated HAC1 decay intermediate inhibits Xrn1. .................. 105

Figure 3.6: Kinetic analysis of HAC1 mRNA processing in cells lacking Tpt1. ........................ 108

Figure 3.7: Decay of HAC1 splicing intermediates regulates UPR activation, suppression, and

attenuation. ................................................................................................................... 111

Figure 4.1: Hypothesis of competition between decay and ligation in animals. ...................... 119

xiv

LIST OF TABLES

Table 1.1: Exoribonucleases and their terminus requirements .................................................. 39

Table 2.1: Oligonucleotide sequences. ...................................................................................... 57

Table 2.2: Strain numbers and genotypes. ................................................................................. 58

Table 2.3: Intron-containing tRNA copy number and intron-dependent modifications in S.

cerevisiae. ....................................................................................................................... 67

Table 3.1: HAC1 processing intermediates ................................................................................ 85

Table 3.2: Oligonucleotide sequences ....................................................................................... 86

Table 3.3: Strain numbers and genotypes .................................................................................. 87

1

CHAPTER I

I INTRODUCTION

RNA is a versatile molecule in the cell, functioning as a messenger encoding a protein

(mRNA), a decoder of the genetic code (tRNA), a structural and catalytic component of the

spliceosome (snRNA) and the ribosome (rRNA), and a plethora of other known (snoRNA, ncRNA,

siRNA, etc.) and undiscovered functions. The lifecycle of RNA can be complex, with many

intervening processing events between birth and death (transcription and decay). Emerging

evidence suggests that some of these processing and decay pathways compete, and so this

dissertation seeks to address how processing and decay compete, and what effect these

competition(s) have on the cell.

To address the question of how processing and decay compete, as well as how this

competition affects cellular physiology, I focused on the unfolded protein response (the UPR), a

cell-autonomous, homeostatic signaling pathway that requires the cleavage, processing, and

ligation of an mRNA. To study the effects of RNA terminus modification and decay, I developed,

characterized, and deployed tools to add and subtract various terminus modification activity to

and from the cells, notably the 10x tRNA Block, a genetic bypass of the essential yeast

(Saccharomyces cerevisiae) tRNA ligase and 2′-phosphotransferase genes, as well as a

heterologous RNA ligase, and exonucleolytic decay pathways. In Chapter II, I demonstrate that

expression of intronless forms of the ten intron-containing tRNAs encoded in the yeast genome

is sufficient to bypass the essential genes TRL1 and TPT1. I remark that the 10x tRNA Block

construct failed to bypass any genes in the tRNA splicing endonuclease (SEN) complex,

indicating that they have an essential function beyond splicing tRNAs. Lastly, I introduce the

molecular and growth phenotype those RNA repair mutants have with respect to the UPR.

Chapter III further describes all my findings on how RNA ligation and decay interact at the stage

2

of 5′-phosphorylated products: having HAC1 3′-exon decay compete with ligation suppresses

the UPR in the absence of stimulus, and how kinase-mediated decay (KMD) of HAC1 intron is

required for robust induction of the UPR upon stimulus. In Chapter IV, I summarize these

findings, place them in context to show how they advance the field, and propose future directions

the project could take. Together, these findings support a generalized RNA decay mechanism,

kinase-mediated decay, and provide examples of how KMD and its competition with ligation

affect cellular physiology.

Endoribonucleases impose terminus chemistry

The termini of RNA are determined by their mode of synthesis: prior to modification, RNA

polymerases produce an RNA molecule with a 5′-triphosphate at the beginning, left behind from

the nucleotide triphosphate used to initiate synthesis, and with a 2′,3′-cis-diol at the end, left

behind by the last ribonucleotide triphosphate incorporated. However, many processing events

impinge on the RNA termini: in eukaryotes, RNA polymerase II (polII)-transcribed RNAs (mRNAs,

lncRNAs) undergo capping to add a 7-methylguanosine 5′→5′-triphosphate adduct (Shatkin

1976), effectively hiding the 5′-end from 5′→3′ exoribonucleases and marking the transcript as

a valid substrate for translation by the ribosome. Additionally, pol II transcripts are

polyadenylated at the polyadenylation site on the primary transcript, which Ysh1 (CPSF-73)

cleaves (Jenny et al. 1996; Ryan et al. 2004) and onto which Pap1 transfers non-templated

adenosine ribonucleotides (Lingner et al. 1991). After synthesis, many RNAs are post-

transcriptionally modified either by “trimming” (exonucleolytically) (see rRNA below) or by

cleavage (endonucleolytically) (see Sen2, Sen34, Las1 below).

The biochemistry of an endoribonuclease active site determines the RNA termini that are

produced from the incision: metal-ion-dependent endoribonucleases (e.g. RNase H, P, Z, MRP)

coordinate the scissile phosphodiester bond and a water molecule to yield “canonical” RNA

3

termini: a 5′-phosphate (5′-PO4) and a 2′,3′-cis-diol (Fig. 1.1A); in contrast, endoribonucleases

that do not coordinate a metal ion exploit the 2′-hydroxyl of the ribonucleotide to attack the

adjacent phosphodiester, yielding two RNA molecules, one with a 5′-hydroxyl (5′-OH) and the

other with a 2′,3′-cyclic phosphate (2′,3′-cP) (Fig. 1.1B). The consequence to the cell of which

variety of endoribonuclease catalyzes the cleavage is that metal-ion-dependent endonucleases

yield ‘clean ends,’ or termini of canonical chemistry that can enter directly into downstream

biochemical reactions, like ligation or exonucleolytic decay. In contrast, metal-ion-independent

endoribonucleases yield products with ‘dirty ends’ that need modification before they are

compatible for downstream reactions.

Metal-ion-dependent endoribonucleases generally employ aspartic acid residues in the

active site to coordinate two divalent metal ions (e.g. Mg2+) to stabilize additional negative charge

on the non-bridging oxygens of the phosphate group and to favor dissociation of water to form

a basic hydroxide that will serve as the nucleophile. By nucleophilic attack, the hydroxide forms

a bond with the phosphorous, releasing the leaving group, the 3′-oxygen of the nucleotide 5′ of

the phosphate (Yang 2011) (see electron-pushing curved arrows in Fig. 1.1A). This class of

endoribonucleases pertains little to this dissertation and is relegated to this brief mention in this

section.

Metal-ion-independent endoribonucleases (e.g. Sen2, Sen34, Rny1/Angiogenin, Las1,

PrrC, Ire1) use a variety of basic active site amino acids to catalyze cleavage. Generally, a basic

residue will deprotonate the 2′-hydroxyl (2′-OH) of a ribonucleotide, allowing the electronegative

2′-oxygen to perform nucleophilic attack on the neighboring phosphorous, forming a 2′,3′-cyclic

phosphate. Then, generally, an acidic residue will induce the 5′-oxygen of the downstream

ribonucleotide to serve as the leaving group, yielding a 5′-hydroxyl product (Fig. 1.1B). Metal-

ion independent endoribonucleases fulfill a plethora functions for the cell, which are reviewed

below.

4

Figure 1.1: Mechanisms of endoribonuclease cleavage.

A: Metal-ion-dependent endoribonucleases coordinate metal ions to prepare the scissile phosphodiester for nucleophilic attack by a hydroxide nucleophile obtained by abstracting a proton from a water molecule. Small curved arrows represent electron pushing; small dotted lines represent coordination or hydrogen bonding. The products of such cleavage are a 5′-fragment with a 2′,3′-cis-diol and a 3′-fragment with a 5′-phosphate.

B: Metal-ion-independent endoribonucleases use acidic (AH+) and basic (B-) amino acid residues to induce reactivity of the 2′-O, causing it to perform nucleophilic attack (curved arrows) on the scissile phosphate, resulting in a 5′-fragment with a 2′,3′-cyclic phosphate and a 3′-fragment with a 5′-hydroxyl.

5

Intrinsic RNA cleavage

Any phosphodiester bond in an RNA molecule is susceptible to intrinsic cleavage because

a 2′-hydroxyl is ‘intrinsically’ present adjacent to the phosphate backbone. Intrinsic cleavage—

the non-enzymatic, spontaneous cleavage of RNA—releases 5′-hydroxyl and 2′,3′-cyclic

phosphate products. Intrinsic cleavage is an ever-present threat to RNA in aqueous solution,

and its kinetics can be accelerated in the presence of heat, base, and divalent cations (e.g. Mg2+,

Mn2+, Zn2+, Pb2+). Heat increases the dissociation constant of water (KW), increasing the

concentration of hydroxide anions that deprotonate the 2′-OH, the first step of the cleavage

reaction. Alkaline pHs (pH ≥ 8) also increase hydroxide concentration, similarly increasing the

frequency of deprotonation of 2′-OH in RNA. Metal ions function as Brønsted–Lowry bases by

abstracting a proton from the 2′-OH of the ribose, yielding a more nucleophilic 2′-oxyanion

capable of attacking the adjacent phosphate (Brown et al. 1985). These factors are the

motivation for typically keeping RNA cold when resuspended in liquid solution, for why RNA

buffers are neutral-to-slightly-acidic in pH, and for why RNA storage buffers often contain EDTA.

Intrinsic cleavage occurs on a spectrum of three-dimensional conformations, with the

optimal being an “in line” conformation that facilitates an SN2-like transesterification reaction

mechanism, in which the attacking 2′-oxygen is 180° opposed to [“in line” with] the downstream

ribonucleotide’s 5′-oxygen, which is the leaving group in the reaction (Usher 1969; Soukup and

Breaker 1999). Single-stranded RNA is more susceptible to adopting the in-line conformation

than double-stranded RNA, meaning single-stranded RNA is more vulnerable to intrinsic

cleavage. Intrinsic cleavage is not sequence-specific, making it a useful step in kits and protocols

that require unbiased fragmentation of RNA for downstream applications (such as high-

throughput sequencing).

6

Self-cleaving ribozymes

Some of the most striking examples of the versatility of RNA are self-cleaving ribozymes.

Six known ribozymes carry out such catalysis, each yielding 5′-hydroxyl and 2′,3′-cyclic

phosphate products. The hepatitis delta virus (HDV) (Sharmeen et al. 1988) and glmS (Winkler et

al. 2004) ribozymes can cleave their own RNA chain, and the hammerhead (Prody et al. 1986),

hairpin (Buzayan et al. 1986), Varkud satellite (VS) virus (Saville and Collins 1990) and twister

ribozymes (Roth et al. 2014) can even ligate compatible RNA termini after cleavage. These self-

cleaving ribozymes catalyze cleavage using general acid-base proton transfer (Das and Piccirilli

2005), similar to the Brønsted–Lowry proton transfer at work in intrinsic RNA cleavage.

The most easily comprehensible example of self-cleaving ribozymes is the hairpin

ribozyme because its active site closely mimics that of RNase A. In the hairpin ribozyme, two

purine ribonucleotides act as a general base and acid, analogous to the two histidines in the

RNase A active site. Furthermore, a conserved adenine base stabilizes the excess negative

charge on the phosphate during the transition state of the cleavage reaction, closely mimicking

the role a lysine’s positively-charged ε-amine plays in the active site of RNase A. The ribozyme

catalyzes cleavage without the use of metal-ion co-factors in the active site but does use metal

ions to fold into a catalytically-active conformation (Nesbitt et al. 1997; Young et al. 1997; Hampel

and Cowan 1997). Cleavage proceeds via the pathways depicted in Fig. 1.1B, yielding 5′-OH

and 2′,3′-cP products.

The cleavage reaction mechanism depicted in Fig 1.1B is reversible, meaning that self-

cleaving ribozymes capable of positioning a 2′,3′-cyclic phosphate RNA and a 5′-hydroxyl RNA

in the active site can catalyze ligation of the non-canonical termini natively. In fact, the only

distinction between self-cleaving ribozymes that can or cannot re-ligate RNA is having base-

pairing interactions on both the 5′-side and the 3′-side of the incision event, providing the

necessary mechanism to position the termini for the reverse reaction, re-ligation. The

7

hammerhead, hairpin, and VS ribozymes are all capable of the ligation reaction, ultimately using

it to self-circularize (Buzayan et al. 1986; Canny et al. 2007; Jones et al. 2001).

Sen2 & Sen34

The yeast genome codes for an estimated 275 tRNAs (transfer RNAs), 61 of which contain

introns (Lowe and Eddy 1997; Chan and Lowe 2016). Among those, ten isodecoding tRNAs are

supplied exclusively by genes interrupted with introns (Chan and Lowe 2009). In 1977, Howard

Goodman, Maynard Olson, and Benjamin Hall discovered the first tRNA intron in the DNA

sequence encoding SUP4, a tRNATyr gene with a 14 base-pair (bp) intron downstream of the

anticodon (Goodman et al. 1977), and many additional tRNA introns have been found since

(Valenzuela et al. 1978; van Tol and Beier 1988). Soon after, Anita Hopper, Fred Banks, and Vicky

Evangelidis first published evidence that a nuclease was required for complete tRNA

biosynthesis (Hopper et al. 1978). The evolutionary origin and physiological purpose of tRNA

introns remain unknown, however their mechanism of removal is well-studied.

Cleaving introns out of tRNA transcripts is an essential function of the Splicing

EndoNuclease (SEN), a complex composed of four essential proteins (Dhungel and Hopper

2012). The primary transcripts of intron-containing tRNAs undergo primary nuclear export into

the cytoplasm, where they meet up with the SEN complex on the outer surface of the

mitochondria (in budding yeast (Peebles et al. 1983; Yoshihisa et al. 2007; Yoshihisa 2014)). The

catalytic subunits of the SEN complex, Sen2 and Sen34 (Trotta et al. 1997), recognize their

intron-containing tRNA substrates on a structural basis and cleave at the exon-intron junctions

(Fig. 1.2). Transfer RNA introns are typically small (14 to 106 nt) and share no sequence similarity,

even within the same organism (Belfort and Weiner 1997). The introns are recognized primarily

by structure by virtue of being located at the end of the anti-codon stem and not interrupting the

structure of the rest of the tRNA (Swerdlow and Guthrie 1984).

8

Figure 1.2: Model of RNA processing in tRNA splicing

Intron-encoding genes for tRNAs are transcribed by RNA polymerase III, trimmed at the 5′- and 3′-ends, and exported into the cytoplasm. The SEN complex cleaves the intron out of the tRNA, leaving two tRNA “halves.” Trl1 ligates the halves together, leaving a 2′-phosphate at the ligation junction. Tpt1 removes the 2′-phosphate left behind by Trl1. Depending on the tRNA, it may return to the nucleus for further processing, or remain in the cytoplasm. The intron cleaved out of the tRNA originates with a 5′-hydroxyl, an RNA terminus on which Xrn1 does not initiate decay. The RNA kinase domain of Trl1 phosphorylates the 5′-end of the intron, transforming it into a substrate of Xrn1 (Wu and Hopper 2014). The 5′-PO4 introns are robustly degraded.

9

Specifically, two conserved positions, called “cardinal positions” are always located the same

number of nucleotides away from the beginning and end of the intron, giving the SEN complex

the “measurements” for where to cut (Bufardeci et al. 1993; Di Nicola Negri et al. 1997).

The catalytic subunits of the SEN complex, Sen2 and Sen34, release the 5′- and 3′-tRNA

exons as 2′,3′-cP and 5′-OH termini, respectively, and the exons remain associated with each

other via extensive secondary structure. Similarly, the intron is released with 5′-OH and 2′,3′-cP

termini. Exoribonuclease 1 (Xrn1) degrades the discarded tRNA intron in the 5′→3′ direction (Wu

and Hopper 2014). The tRNA halves are ligated in healing and sealing reactions reviewed in the

“RNA repair in cellular physiology” section of this chapter. Upon ligation, some tRNAs return to

the nucleus via “tRNA retrograde nuclear import” for additional modifications (Shaheen and

Hopper 2005; Takano et al. 2005).

In addition to tRNAs, the SEN complex also cleaves the mRNA CBP1, a nuclear-encoded

gene for a mitochondrial protein (Tsuboi et al. 2015). Cytochrome b mRNA processing 1

promotes the stability and translation of the mitochondrial mRNA for cytochrome b (Krause et

al. 2004). It’s interesting that a non-tRNA substrate of the SEN complex would function at the

mitochondria given that the SEN complex localizes to the outer surface of the mitochondria in

yeast (Yoshihisa et al. 2003; 2007). However, CBP1 is not an essential gene and therefore likely

does not contribute to the essential function(s) of the SEN complex (Dieckmann et al. 1982).

Nevertheless, while tRNA splicing can be accomplished by moving the SEN complex to the

nucleus, nuclear SEN fails to complement deletion of mitochondrial SEN complex (Dhungel and

Hopper 2012), suggesting that SEN has yet another essential function in the cytoplasm of yeast.

Las1, Grc3, & 35S rRNA processing

Ribosomes are the ribonucleoprotein ribozymes that catalyze the mRNA-dependent

synthesis of proteins. Ribosomes have a complex synthesis pathway, coalescing four ribosomal

RNAs (rRNA) and approximately 80 proteins (Fromm et al. 2017). A major step of ribosome

10

assembly is the production and processing of the 35S pre-rRNA, synthesized as a tandem

transcript of the 18S, 5.8S, and 25S rRNAs, along with intervening and surrounding sequence

called Internal Transcribed Spacers (ITS) and External Transcribed Spacers (ETS) that are

removed through the pathway of pre-rRNA maturation (Woolford and Baserga 2013). This long

transcript was discovered in the 45S fraction of a sucrose gradient and called the “giant RNA”

(Scherrer 2003). Eukaryotic rRNA processing was first evinced in the 1960s, when Klaus

Scherrer, Jim Darnell, and Robert Perry demonstrated that 14C-uridine-labeled giant RNA rapidly

shifted to the 35S, and then 28S and 16S fractions (Scherrer and Darnell 1962; Scherrer et al.

1963; Perry 1962; Perry et al. 1964). They concluded that the “giant RNA” was the “pre-RNA”

that was transcribed and then processed into smaller RNAs. The field of pre-rRNA processing

has advanced, but to this day, some details of the precise molecular mechanisms of ribosomal

RNA processing remain undiscovered (Scherrer 2003).

The rRNA of budding yeast (S. cerevisiae), is encoded in the genome (as “rDNA”) at ~150

copies in a tandem repeat array located on chromosome 12 (Fernández-Pevida et al. 2015). RNA

polymerase III (polIII) transcribes the 5S rRNA, at which point Rex1/Rnh70 hydrolyzes 7 to 13

nucleotides off the 3′-terminus (van Hoof et al. 2000). In contrast to the former, simpler rRNA

maturation pathway, RNA polymerase I (polI) transcribes the 25S, 18S, and 5.8S rRNAs as a

tandem 35S rRNA transcript, which requires eight endonucleolytic cleavages (A0, A1, D, A2, A3,

B1L, C2, and B0) and eight exonucleolytic trimming steps (B1s, B2, C1′, C1, and four trimming steps

from 7S/L to 5.8S/L) to form rRNAs of the mature length (Woolford and Baserga 2013; Fernández-

Pevida et al. 2015). Despite 57 years passing between the first publication of pre-rRNA

processing and the date of this dissertation’s writing, the agents that catalyze three of the

endonucleolytic cleavages (A0, A1, and B1L) remain unidentified (Scherrer 2003). The

endonuclease that cleaves at C2, however, was identified as Las1 in 2015 (Gasse et al. 2015).

Internal Transcribed Spacer 2 (ITS2) lies between the 5.8S and 25S rRNAs of the 27SBS pre-

11

rRNA. Las1 cleaves ITS2 to produce the 7SS and 26S pre-rRNAs (Veldman et al. 1981). In a

downstream step, Rat1 (Xrn2) trims the 26S pre-rRNA to yield a mature-length 25S rRNA

(Geerlings et al. 2000). Las1 is a metal-ion-independent endoribonuclease (Gasse et al. 2015), a

topic that will be revisited in the “Kinase-Mediated Decay” section of this chapter.

Stress-induced tRNA cleavage

Cleavage of tRNAs at or near their anticodon for emergency regulation of translation is a

theme in biology. In both bacteria and eukaryotes, cellular stresses can result in the cleavage of

tRNAs in their anticodon loops to produce 3′ and 5′ halves called “tiRNAs.” The Escherichia coli

anticodon nuclease, PrrC, mounts an antiviral response against bacteriophage infection by

cleaving tRNALys in the anticodon loop, preventing the tRNA from participating in translation

(Levitz et al. 1990; Kaufmann 2000). In addition, colicin E5 causes cell death by translational

arrest by cleaving 3′ of the queuine base modification at the wobble position in the anticodon

loop in tyrosine, histidine, asparagine, and aspartic acid tRNAs (Ogawa et al. 1999; Masaki and

Ogawa 2002). Eukaryotes also cleave tRNAs during times of stress. The ciliate Tetrahymena and

the parasitic protozoan Giardia lamblia both cleave tRNAs under starvation conditions (Lee and

Collins 2005; Li et al. 2008), and budding yeast, plants, and humans cleave tRNAs in response

to oxidative stress specifically (Thompson et al. 2008; Thompson and Parker 2009b). However,

a major difference between prokaryotic and eukaryotic tRNA cleavage is that prokaryotes tend

to cleave a high proportion of susceptible tRNAs, whereas eukaryotic cleavage is ever only a

minority of cleavable tRNAs (Thompson and Parker 2009a).

The regulation of the stress-inducible endoribonucleases Rny1 in yeast and angiogenin in

humans occurs through remarkably similar pathways despite the nucleases sharing no

evolutionary history. Rny1 is a yeast vacuolar endoribonuclease belonging to the RNase T2 family

and serves to harden yeast against osmotic and heat-shock stresses (MacIntosh et al. 2001;

Thompson and Parker 2009b). Rny1 is normally trafficked through the yeast endomembrane

12

system and secreted, reducing its chances of encountering cytoplasmic RNA. However, upon

oxidative stress, Rny1 is released into the cytoplasm, where it cleaves tRNAs, causing

widespread translational pausing (Pelechano et al. 2015).

In a similar mode of regulation, human angiogenin, a member of the RNase A family, is a

secreted protein that is stored in the lysosome but also held in the nucleolus with its inhibitory

binding partner Rnh1. While the anticodon nuclease function of angiogenin is similar to that of

Rny1 from yeast, the two are not homologs; curiously, Rny1 has an ortholog in humans,

RNASET2, but it appears not to cleave tRNAs (Yamasaki et al. 2009). Angiogenin was first

discovered by purification from the HT-29 adenocarcinoma cell line (Fett et al. 1985) and was

shown to produce tiRNAs in in vivo (Yamasaki et al. 2009; Fu et al. 2009). Angiogenin is activated

by stress and translocates from the nucleolus to the cytoplasm where it dissociates from Rnh1

and cleaves tRNAs to produce tiRNAs. In addition to slowing translation by depleting available

tRNAs, the tiRNAs themselves may also promote apoptosis (Emara et al. 2010). Transfection of

5′-tiRNAs (but not corresponding 3′-tiRNAs) promotes stress granule formation and inhibition of

translation independent of eIF2α phosphorylation. 5′-tiRNAs can also inhibit translation initiation

by displacing eIF4G from mRNAs in vitro (Ivanov et al. 2011). Interestingly, tiRNAs can

competitively bind cytochrome c over the apoptosis pathway protein APAF1, protecting cells

from apoptosis during osmotic stress (Saikia et al. 2014).

Ribotoxins in interspecies conflict

Under intense resource competition, organisms can benefit from sabotaging a competing

species. Microbes especially compete for scarce resources at close quarters, making the

nutrient intake of one microbe a great loss for a competing microbe. So many organisms deploy

toxins at the expense of a competing or predatorial species. These toxins may take the form of

small molecules, peptides, or fully-functioning proteins that enzymatically catalyze some toxic

reaction. Ribotoxins are the class of toxic enzymes whose substrates are RNA. Ribotoxins have

13

a variety of RNA substrates, but typically target RNA molecules central to the metabolism of their

victim, such as ribosomal RNA or transfer RNA for their roles in protein synthesis. Cleavage of

these RNA targets leads to malfunction of translation machinery and, via an unknown

mechanism, cell death due to double-strand breaks of genomic DNA during the S phase of the

cell cycle (Klassen et al. 2004; Klassen and Meinhardt 2005).

Colicins are a group of secreted proteins belonging to plasmid-encoded toxin-antitoxin

systems that kill neighboring E. coli cells that do not harbor the col plasmid (James et al. 1996;

Cramer et al. 1999; Masaki and Ohta 1985). The toxic mechanism of these proteins varies, but a

family of colicins harm their victims via their nuclease activity. Colicin E3 cleaves the 3′ 49

nucleotides off the 16S rRNA in E. coli, removing a rRNA helix of the 30S ribosomal subunit

critical for its interaction with the 50S subunit and the A-site tRNA (Bowman et al. 1971a; 1971b;

Dahlberg and Dahlberg 1975). The colicin E3 cleavage event stops translation of protein,

ultimately leading to cell death. Colicin E5 and D inhibit translation in E. coli by specifically

cleaving tRNAs, leading to cell death (Tomita et al. 2000). The particular queuine post-

transcriptional modification to a guanine residue of the wobble base in the anticodon of some

tRNAs is sufficient for colicin E5 and D cleavage (Ogawa et al. 1999). However, colicin E5 and D

are capable of cleaving tRNAs lacking the queuine so long as they contain a specific anticodon

sequence (Ogawa et al. 2006).

There are also secreted ribotoxins deployed among eukaryotes. Kluyveromyces lactis

Zymocin and Pichia acaciae Killer (PaT) are secreted fungal defense toxins that penetrate S.

cerevisiae cells and, upon arrival in the cytoplasm, cleave the anticodon loop of tRNAGlu(GAA)

and tRNAGln(UUG), respectively. Consequent depletion of the specific tRNAs arrests yeast

growth by translational inhibition, cell cycle arrest, and DNA damage response activation (the

latter by an unknown mechanism) (Lu et al. 2005; Klassen et al. 2004; Klassen and Meinhardt

2005; Klassen et al. 2008).

14

Both Zymocin and Killer toxins represent examples of tRNA ribotoxins that rely on an RNA

base modification for targeting. Zymocin specifically recognizes the U34-wobble base

modification (5-methoxycarbonylmethyl-2-thiouridine, or mcm5s2U) of tRNAGlu(GAA) as carried

out by a modification pathway requiring tRNA methyltransferase Trm9. Consequently, trm9∆

cells and cells overexpressing tRNAGlu(GAA) to the point of saturation of Trm9 pathway enzyme

kinetics exhibit resistance to the Zymocin ribotoxin (Jablonowski et al. 2006). Pichia acaciae Killer

toxin (PaT) is a linear-plasmid encoded toxin-antitoxin system that confers toxin secretion again

competing yeast and simultaneous resistance to production of the toxin (Worsham and Bolen

1990; Chakravarty et al. 2014). PaT targeting also relies on the mcm5s2U modification, but this

time of tRNAGln(UUG) (Klassen et al. 2008). Furthermore, PaT can make two incisions in the

tRNAGln(UUG), releasing a dinucleotide the eliminates the possibility of recovery from the toxin

via RNA repair (Meineke et al. 2012).

RNA decay in co-translational mRNA surveillance

Messenger RNAs encode protein sequences that are translated into amino acid sequences

by the ribosome, and this process of translation is monitored for irregularities in the mRNA. The

coding quality of mRNA is controlled through a family of processes called co-translational mRNA

surveillance, an umbrella term for a series of processes that monitor ribosome progress along

the open reading frame to detect mRNAs that are defective in translation. Adaptor proteins

interact with the ribosome to detect aberrant mRNAs and direct them toward RNA decay

pathways. Principally, three types of co-translational mRNA surveillance are known: Nonsense-

Mediated Decay (NMD), Non-Stop Decay (NSD), and No-Go mRNA Decay (NGD) (Reviewed in

(Shoemaker and Green 2012)). Together, these processes ensure that a message accurately

encodes a peptide sequence.

No-go mRNA decay (NGD) is one of the co-translational mRNA surveillance processes and

detects stalled ribosomes on messages (Doma and Parker 2006). When ribosomes encounter

15

strong stalls in translation elongation, as caused by stem-loops, rare codons, or poly-positively-

charged amino acids, NGD is activated and the mRNA is cleaved, releasing 2′,3′-cP and 5′-OH

RNA products (Navickas et al. 2018). The cytoplasmic exosome degrades the 5′-product of

endonucleolytic cleavage via recruitment through Ski7, and Xrn1 degrades the 3′-product. No

specific mRNAs have been identified as being consistently subject to NGD, indicating that NGD

is likely a general-purpose quality control pathway that removes stochastically corrupted mRNAs

from the pool of translating messages.

While the nuclease of NGD is not known (and is speculated to be the ribosome itself), the

release-factor-like proteins Dom34 and Hbs1 recognize these aberrantly translating ribosomes

and remove the ribosome from the message in a stop-codon-independent manner. Dom34 and

Hbs1 appeared central in the discovery of no-go decay (Doma and Parker 2006), and the two

proteins were suspected of working as a ribosome recycling factors because of Dom34 is

paralogous to eukaryotic Release Factor 1 (eRF1), and Hbs1 to eRF3. Dom34 and Hbs1 were

later shown to form a heterodimer that mimics the conformation of a eukaryotic translation

termination complex (Chen et al. 2010; Becker et al. 2011) and even uses energy from the

hydrolysis of GTP to separate the 60S and 40S ribosomal subunits (Becker et al. 2011).

Conflicting reports indicate that Dom34 and Hbs1 are required for endonucleolytic cleavage of

the no-go mRNA (Passos et al. 2009) and some indicate no requirement (Ikeuchi and Inada 2016).

Either way, the factor that induces cleavage of the mRNA just upstream of the stalled ribosome

remains unidentified.

Non-Stop mRNA Decay (NSD) is rapid degradation of mRNAs that lack a translation

termination codon (Frischmeyer et al. 2002; van Hoof et al. 2002). Non-stop mRNAs can occur

naturally by alternative polyadenylation within the Open Reading Frame (ORF), by mutation, and

by endonuclease cleavage. NSD is activated by the ribosome reaching the 3′-terminus of the

mRNA, causing it to stall. Because the ribosome is stalled without a stop codon in the A-site,

16

the stop-codon-independent release factors Dom34 and Hbs1 recycle the ribosome, and Ski7

couples the process to rapid 3′→5′ degradation of the mRNA by the exosome (Horikawa et al.

2016). NSD is related to no-go decay by the need for stop-codon-independent ribosome release

factors, but NSD is different from other forms of exosomal degradation because the GTPase

domain of Ski7 is required for the process (van Hoof et al. 2002). Decapping followed by 5′→3′

decay, as carried out by Dcp1 and Xrn1, plays a minor role in removal of non-stop mRNAs from

circulation (Frischmeyer et al. 2002; Inada and Aiba 2005). Furthermore, the peptide resulting

from the non-stop mRNA is rapidly degraded by the ubiquitin-proteasome pathway via Ltn1 or

Not4 ubiquitinylating the peptide from the stalled ribosome (Wilson et al. 2007; Dimitrova et al.

2009; Bengtson and Joazeiro 2010).

Lastly, Nonsense-Mediated Decay (NSD) is the quality-control mechanism that degrades

mRNAs with aberrant stop codons. Such a quality control system was first suggested in 1979

when Regine Losson and Francois Lacroute introduced amber nonsense codons into the yeast

URA3 gene, causing destabilization of the URA3 mRNA, otherwise known for its legendary

stability (Losson and Lacroute 1979). Other causes of NMD include long 3′-untranslated regions

that alter the relationship of the poly(A) tail to the ORF (Muhlrad and Parker 1999a; Kebaara and

Atkin 2009), alternative translation start codons out of frame with the larger ORF that lead to

premature termination (Welch and Jacobson 1999), pre-mRNAs with introns that contain stop

codons (He et al. 1993; Sayani et al. 2008), mRNAs that contain frame shift sequences or

mutations (Belew et al. 2011), and mRNAs that contain upstream ORFs (Gaba et al. 2005; Guan

et al. 2006). These aberrant mRNAs become targets of Upf1, Upf2, and Upf3, which associate

with the translation termination complex and cause deadenylation-independent decapping or

shortening of the poly(A) tail, ultimately leading to Xrn1 degrading the mRNA (Muhlrad et al. 1994;

Cao and Parker 2003; Mitchell and Tollervey 2003; Baker and Parker 2004). Additionally, Upf1

17

alone can act to repress translation of an aberrant mRNA (Muhlrad and Parker 1999b) and can

target nonsense mRNAs to P-bodies (Sheth and Parker 2006).

Ire1, the “splicing” endoribonuclease of the unfolded protein response (UPR)

Ire1 (Inositol-Requiring Enzyme 1, alias: Ern1) (Nikawa and Yamashita 1992; Mori et al.

1993) is the transmembrane sensor protein that conveys the signal of unfolded proteins from

within the lumen of the endoplasmic reticulum (ER) to the cytoplasm (Walter and Ron 2011).

When unfolded proteins accumulate in the ER, the unfolded protein response (UPR) is triggered,

leading to corrective changes in gene expression that increase the size and protein-folding

capacity of the ER (covered in-depth in the “RNA repair in cellular physiology” section)

(Kozutsumi et al. 1988; Schuck et al. 2009). Ire1 conveys this signal via two enzymatic activities

present on its cytoplasmic domain: (i) it autophosphorylates other Ire1 molecules

(autophosphorylation in trans) upon oligomerization caused by accumulating unfolded proteins

in the ER lumen; and (ii) Ire1 activates its endoribonuclease domain, which is capable of cleaving

cytoplasmic RNAs in the vicinity of the ER membrane (Moore and Hollien 2012).

A structural examination of Ire1 explains much of its function in sensing unfolded proteins

and activating the UPR (Korennykh and Walter 2012). The primary/domain structure of Ire1 is

composed of an N-terminal sensing domain in the lumen of the ER, then a transmembrane helix,

then a cytoplasmic CDK2-like serine/threonine kinase, and finally a C-terminal ribonuclease

domain (Sidrauski and Walter 1997). A crystal structure of the luminal sensing domain revealed

abundant β-strands that coalesce to form a β-sheet platform (Credle et al. 2005) with the

surprising finding of a deep pocket with architectural similarity to the disordered-peptide-binding

pocket of major histocompatibility complex (MHC) (Achour et al. 1998; Olson et al. 2006). This

finding supports the model that the sensor domain of Ire1 can directly bind disordered/unfolded

proteins in the lumen of the ER to support oligomerization/activation (Credle et al. 2005; Gardner

and Walter 2011). In contrast, a competing model of activation of Ire1 is based on the

18

interpretation of a different crystal structure with an occluded groove on the β-sheet platform

(Zhou et al. 2006) to support a model in which unfolded proteins sequester BiP/Kar2 away from

direct Ire1 binding, allowing Ire1 the opportunity to dimerize and oligomerize (Bertolotti et al.

2000). However, unfolded proteins are likely the direct ligand of luminal domain of Ire1, as shown

by in vitro binding studies with hydrophobic and basic peptides, as well as in vivo

immunoprecipitation of Ire1 to a constitutively-unfolded protein (Gardner and Walter 2011).

What the conflicting models have in common is that the luminal sensor domain of Ire1 must

undergo a conformation change upon protein folding stress (independent of its ligand) in order

to self-associate and form oligomers on the surface of the ER (Gardner and Walter 2011).

Quantitatively, the amount of Ire1 oligomerization is directly proportional to RNase activity,

strongly implying that Ire1 self-association is the mechanism of RNase domain activation

(Korennykh et al. 2011a). The higher-order assembly of Ire1 stabilizes a region of the protein

called the helix-loop element (Korennykh et al. 2009), whose residues are critical for cleavage of

substrate RNA (Lee et al. 2008). Important to this interaction are the phosphoserine and

phosphothreonine residues formed from the specific autophosphorylation in-trans activity. The

phospho-amino acids form salt-bridges across monomers of the higher-order assembly, further

stabilizing the interaction. The interface of two Ire1 proteins (dimer) stabilizes the helix-loop

element, completing the active site of the RNase domain. This dimer structure of the RNase

domain is further supported by the 1:2 stoichiometry of binding: in vitro, one stem-loop RNA

substrate binds per two molecules of Ire1 (Korennykh et al. 2011b).

Across phyla, Ire1 has only three well-supported substrates that all function similarly as

transcription factors in vivo: HAC1 in yeast, bZIP60 in plants, and Xbp1 in animals. All three

mRNAs have stem-loop structures in common: a short stem with a seven- or eight-nucleotide

loop bearing the consensus sequence 5′-CNGNNGN-3′ (Gonzalez et al. 1999; Sidrauski and

Walter 1997; Yoshida et al. 2001). How these substrates of Ire1 function once they have been

19

cleaved is reviewed in the “Unfolded protein response” area of the “RNA repair in cellular

physiology” section below.

Ire1 has additional functions in the UPR beyond cleaving RNA for splicing. Hollien and

Weissman first identified ER-associated mRNAs that Ire1 could cleave (other than Xbp1) using a

microarray on total RNA isolated from Drosophila S2 cells (Hollien and Weissman 2006). A subset

of RNAs downregulated during protein folding stress returned to normal expression levels when

Ire1 was knocked down with RNAi, but not under Xbp1 knock-down. Northern blots showed the

RNAs were endonucleolytically cleaved. Later named Regulated Ire1-Dependent Decay (RIDD)

(Hollien et al. 2009), it was shown to be a general phenomenon of Ire1 across many phyla,

including mammals, plans, and fission yeast (Kimmig et al. 2012) (Tam et al. 2014; Guydosh et

al. 2017), but not budding yeast (Niwa et al. 2005). In mammals, RIDD cleaves RNAs that maintain

Xbp1-like stem-loop structural motifs near the ER membrane (Moore and Hollien 2015). In fission

yeast, where no UPR-associated transcription factor is present, Ire1 cleaves RNAs with some

degree of site-specificity, enacting RIDD on some substrates, but also increasing the stability of

other mRNA substrates (Kimmig et al. 2012). The lack of significant sequence identity (29%)

between the Ire1 protein of S. cerevisiae and S. pombe may confer the functional divergence (Li

et al. 2018), possibly explaining the differences in substrates of the endoribonuclease domain of

Ire1.

RNA terminus modification

Further processing steps or downstream functions sometimes require the modification of

the terminus of an RNA molecule. Cells often express discrete enzymes with RNA terminus

modification activity, like kinase, cyclic-phosphodiesterase, or cyclase activities, to assist in

processing RNA molecules for maturation or decay. 5′-termini can be phosphorylated

specifically by RNA 5′-kinases. Examples of RNA 5′-kinases include bacteriophage T4

20

polynucleotide kinase (PNK) (Wang et al. 2002) and the kinase domain of fungal or plant RNA

ligase (Trl1) (Wang and Shuman 2005; Wang et al. 2006). These kinases prepare the terminus of

the substrate RNA for ligation because the respective ligase will need a 5′-PO4 for transferring

an NTP onto the terminus, effectively “charging” the terminus as a high-energy 5′-Npp-RNA

intermediate for the ligation reaction. By changing the biochemistry of these RNA termini, the

RNA ligases/kinases expose the RNA to decay enzymes that specifically recognize the 5′-

phosphate of RNA (Wu and Hopper 2014). Notably, an absence of 2′- or 3′- direct kinases has

implications for the fidelity of ligation of RNA fragments, which is described in the next section

on RNA Ligases.

Clp1 RNA 5′-kinase

Clp1 (5′-RNA kinase) is an essential structural component of the yeast co-transcriptional

RNA processing complex (CLeavage and Polyadenylation factor 1a subunit) and a 5′-RNA kinase

in metazoans. Stefan Weitzer and Javier Martinez were motivated to identify the kinase that

converts 5′-OH synthetic siRNAs to the required 5′-PO4 terminus for incorporation into the RNA-

induced silencing complex (RISC) (Nykänen et al. 2001; Weitzer and Martinez 2007). The

homologous S. cerevisiae Clp1 was found not to have ATP-hydrolyzing activity, and so appears

to be a structural component of the cleavage and polyadenylation complex in yeast rather than

a catalytic subunit (Noble et al. 2007). Weitzer and Martinez also found that tRNA 5′- and 3′-

exons and introns are in vivo substrates of the kinase in animal cells. Clp1 is the candidate RNA

kinase for degradation of introns cleaved from spliced tRNAs in humans. In mice and humans,

mutations in Clp1 have been linked to neurodegenerative diseases, possibly caused by defects

in tRNA intron decay (Hanada et al. 2013; Karaca et al. 2014; Schaffer et al. 2014).

Grc3 RNA 5′-kinase

Grc3 (name etymology not known) is an essential 5′-RNA kinase in yeast first

bioinformatically identified as a triple-A ATP/GTPase by protein sequence homology (El-

21

Moghazy et al. 2000). Grc3 was identified in a microarray screen for non-coding RNA processing

and individually confirmed to participate in 27S pre-rRNA processing (a process that ultimately

produces mature 5.8S and 25S rRNAs) (Peng et al. 2003). Then Grc3 was demonstrated to be a

polynucleotide kinase associated with rRNA transcription centers, where it assists RNA

polymerase I terminate transcription (likely by the torpedo model) (Braglia et al. 2010). Another

study showed that an important function of Grc3 kinase is to phosphorylate the downstream

pre-rRNA fragment generated from Las1 cleavage, allowing the 5′-PO4-dependent Xrn2/Rat1

exonuclease to partially trim the pre-rRNA (Gasse et al. 2015).

CNP (cyclic nucleotide phosphodiesterase)

CNP (Cyclic Nucleotide Phosphodiesterase) is a 2′,3′-cyclic nucleotide phosphodiesterase

that catalyzes the irreversible hydrolysis of 3′-phosphodiester bonds in 2′,3′-cyclic nucleotides

to produce 2′-PO4/3′-OH termini (Sprinkle 1989). CNP is found extensively throughout cells of

the nervous system of mammals, making up ~4% of protein in myelinating glial cells (Vogel and

Thompson 1988), and is required in oligodendrocytes for survival of neuronal axons (Lappe-

Siefke et al. 2003). Surprisingly, only one in vivo substrate of CNP has been identified: Xbp1, the

mRNA of the unfolded protein response of animals (Unlu et al. 2018). Nonetheless, CNP is

frequently employed in biochemical manipulations of RNA termini in vitro and in vivo (Schwer et

al. 2008).

RtcA (3′-terminal phosphate cyclase)

RtcA (2′,3′-cyclase) uses ATP to cyclize 3′-phosphate RNA termini to 2′,3′-cyclic

phosphates. RtcA is widely distributed across the genomes and proteomes of eukaryotes and

prokaryotes (Genschik et al. 1997). 2′,3′-cyclase activity was first discovered in the lysate of

HeLa cells (Filipowicz et al. 1983) and in Xenopus cell nuclei (Laski et al. 1983). 2′,3′-cyclic

phosphate RNAs, like the kind that RtcA yields, are the substrates of the RNA ligases discussed

in the next section, conferring RtcA with a possible function of preparing substrates for ligases.

22

However, cyclic phosphates are stable once formed, so it is unclear what the physiological

substrates of RtcA could be across the domains of life. Potential physiological roles for RtcA

have been proposed after discovering additional products that RtcA is capable of generating,

including 5′-adenylated RNA (from 5′-PO4 substrates) and 5′-adenylated DNA in a nicked double

strand, owing to much overlap in the catalytic mechanism of RtcA with RNA/DNA ligases

(Chakravarty and Shuman 2011). Notably, yeast have no 2′,3′-cyclase activity in cell lysate (Billy

et al. 2000) and RtcA is not essential in E. coli (Genschik et al. 1998).

RNA ligases combine RNA terminus modification domains with a ligase domain.

Cells deploy a variety of strategies to “clean up”—or “heal”—the ends of RNA released

from cleavage events. Some of the RNA terminus modifications discussed in the previous

section resurface in multifunctional RNA ligases. The modification strategy used by the cell

largely depends on the RNA ligase expressed in the organism: fungi and plants use a “5′→3′”

mechanism of ligation, whereas animals, bacteria, and archaea use a “3′→5′” mechanism,

named so because of the terminus on which the phosphate becomes incorporated into the

phosphodiester backbone of the RNA molecule (Fig. 1.3) (Unlu et al. 2018).

Trl1, the ligase of fungi & plants

The tRNA Ligase in plants and fungi, Trl1 (alias: Rnl1, for RNA Ligase) was at first only

detected via its activity in yeast and in wheat germ extract (Knapp et al. 1978; Abelson 1979;

Konarska et al. 1981). In 1983, Greer et al. determined the enzymatic steps of the ligase from

enzyme isolated by activity purification (Greer et al. 1983), the same technique used a decade

earlier on T4 Bacteriophage RNA ligase (Silber et al. 1972). In 1986, Eric Phizicky, John Abelson

et al. showed that Trl1 can perform ligation in purified form on tRNA halves and an arbitrary

oligoribonucleotide substrates, implying that Trl1 is a general RNA ligase (Phizicky et al. 1986).

Trl1 has three separable domains (Fig. 1.3A): (i) a cyclic phosphodiesterase domain that

23

hydrolyzes the 2′,3′-cyclic phosphate to a 2′-phosphate/3′-hydroxyl; (ii) a RNA 5′-kinase

domain; and (iii) a 5′-adenylation domain that forms the RNA-5′-adenylate intermediate that

undergoes nucleophilic attack by the 3′-OH terminus to form a phosphodiester bond (Greer et

al. 1983; Schwartz et al. 1983; Xu et al. 1990; Apostol and Greer 1991). The termini-modifying

activities constitute “healing” of the RNA ends, and the adenylyl-transferase/ligase activity

represents “sealing” of the two molecules of RNA together. All three domains/functions of Trl1

are essential in yeast, but they can be expressed as separate proteins (Sawaya et al. 2003).

The completed ligation reaction heals the termini, and seals together the RNA molecules

with a phosphodiester bond, but a 2′-phosphate “scar” remains at the ligation junction due to

the cyclic phosphodiesterase specificity of Trl1. Plants and fungi have an enzyme, 2′-

PhosphoTransferase 1 (Tpt1) to remove the 2′-phosphate (Fig. 1.3A) (McCraith and Phizicky

1990). Tpt1 transfers the 2′-phosphate to a molecule of NAD+, displacing the nicotinamide and

yielding an unusual product, ADP-ribose 1″,2″-cyclic phosphate (Culver et al. 1993).

RtcB, the ligase of prokaryotes & animals

RtcB, RNA ligase present in archaea, bacteria, and animals, uses a different strategy to

ligate RNA. The locus of rtcB in the E. coli genome (in an operon alongside rtcA, an RNA 2′,3′-

terminal cyclase, and rtcR, a σ54 transcription factor) and its sequence conservation with other

bacterial ligases suggested it could be an RNA ligase (Galperin and Koonin 2004). RtcB and

HSPC117 were nearly simultaneously discovered as the RNA ligases responsible for tRNA

splicing in prokaryotes and animals, respectively (Tanaka and Shuman 2011; Popow et al. 2011).

Soon after, RtcB was expressed in yeast to show that it could complement deletion of the

essential gene TRL1 and that it also functioned as the ligase in the unfolded protein response

(Tanaka et al. 2011b). (The unfolded protein response, and the effect of RtcB on it, is discussed

in greater detail below and in Chapter III.) Furthermore, the biochemical activity of RtcB was

24

characterized as distinct from that of plant and fungal ligases, involving a “two-step,” 3′→5′

reaction mechanism (Zofallova et al. 2000; Tanaka et al. 2011a; Chakravarty et al. 2012).

Figure 1.3: Mechanisms of ligation.

A: In fungi and plants, the tRNA Ligase 1, Trl1 (also called RNA Ligase 1, Rnl1), is a multifunctional enzyme with distributive cyclic phosphodiesterase (CPDase), 5′-kinase, and ligase (involving transient 5′-andelylation) activity. The CPDase hydrolyzes the 2′,3′-cP to a 2′-PO4 and a 3′-OH, which serves as the nucleophile during ligation to displace the high-energy adenylate group. The ligation product has a 2′-phosphate “scar” (in blue, left over from the CPDase reaction), which is transferred to a molecule of NAD+, releasing nicotinamide (not depicted) and ADP-ribose 1″,2″-cyclic phosphate (depicted). The phosphate highlighted in red (5′) is incorporated into the phosphodiester backbone, hence the name “5′→3′ ligation.”

B: In animals, bacteria, and archaea, the enzyme RtcB (HSPC117) also works as a multifunctional enzyme with cyclic phosphodiesterase (CPDase) and ligase (with transient 3′-gyanelation) activity. The CPDase hydrolyzes the 2′,3′-cP to a 2′-OH and a 3′-PO4. The 5′-OH serves as the nucleophile during ligation to displace the high-energy guanylate group. The 3′-PO4 highlighted in red ultimately becomes incorporated into the phosphodiester backbone, giving it the name “3′→5′ ligation.”

25

RtcB/HSPC117 hydrolyzes the 2′,3′-cyclic phosphate to a 3′-phosphate, which it then

guanylates to form a 3′-ppG intermediate; finally, the 5′-OH, so-far unmodified throughout this

process, displaces the guanylate group, sealing the RNAs together with a phosphodiester bond

(Fig 1.3B).

Meanwhile, the metazoan ligase RtcB (alias: HSPC117), already known to carry out tRNA

ligation, was demonstrated to be a ligase essential for activation of the unfolded protein response

(UPR) in the nematode C. elegans, as shown via deletion mutants expressing a pre-spliced tRNA

construct (Kosmaczewski et al. 2014). In mammals, HSPC117 was identified using a synthetic

biology approach coupled with a siRNA screen (Lu et al. 2014) and with HeLa whole-cell extracts

coupled with RNAi knock-down of RtcB (and its partner protein, Archease) (Jurkin et al. 2014).

Fidelity of RNA ligases

The biochemistry of RNA termini encodes information about the fidelity of the molecule,

allowing ligases to enforce fidelity requirements on their substrates. Generating a seamless

ligation of two RNA ends is essential for producing tRNAs with correct anticodon loop structure

and preserving the coding potential of the 3′-exon of HAC1/Xbp1/bZIP60 in the unfolded protein

response. But this observation raises the question: how do ligases ensure fidelity? There are no

known terminal 2′- or 3′-RNA kinases, meaning that any RNA with a 2′,3′-cyclic phosphate, 2′-

phosphate, or 3′-phosphate was cleaved at that exact site, and so is a faithful substrate for

ligation. Supporting this idea are three examples of ligase specificities for such substrates, thus

enforcing fidelity of their ligation products. Firstly, RtcB, the ligase of prokaryotes and animals,

requires a 3′-PO4 to transfer the guanylate group in preparation for ligation. In the case where

the cyclic phosphate is hydrolyzed to a 2′-PO4, the RNA cyclase RtcA can restore the substrate

to a 2′,3′-cyclic phosphate, preserving the fidelity of the terminus. Secondly, in an interesting

twist, S. cerevisiae ligase, Trl1, requires a 2′-PO4 of its ligation substrates despite the 2′-PO4 not

being involved in the biochemical reaction itself (Schwer et al. 2004). Similarly, the RNA ligase of

26

the plant Arabidopsis thaliana requires a 2′-PO4 on its tRNA substrates for ligation (Wang et al.

2006). Thirdly, a [counter]example of a low-fidelity ligase is found in T4 RNA Ligase 1 (T4 Rnl1).

The previously discussed enforcement of fidelity via the 2′ or 3′ terminus of the RNA is not

present in bacteriophage T4 Rnl1, and consistent with that observation, the ligation products of

T4 Rnl1 occasionally lack nucleotides from what were the 3′-ends of its substrates (Schwer et

al. 2004). Thus, the RNA termini contain information about the fidelity of the molecule, and ligases

can read this information to enforce fidelity requirements on their substrates. In the next section,

the biochemistry of the 5′-terminus of RNA is explored for its effect on ligation and other

processes.

RNA repair in cellular physiology

RNA repair is distributed widely throughout biology. Therefore, it is not surprising that it

has been discovered in a variety of contexts, performing a variety of tasks for the cell, or

sometimes not for the cell, reviewed in the T4 bacteriophage RNA repair section below. RNA

repair ligates tRNA exons back together once their introns have been excised, and similarly, RNA

repair enables the unfolded protein response to activate by ligating the exons of an inducibly-

cleaved mRNA. Many stress responses use RNA repair to reverse the response, like after stress-

induced cleavage of tRNAs in prokaryotes or restoring ribosome function after MazF cleavage.

RNA repair is also a direct response during interspecies conflict with ribotoxins, examples of

which include the PnkP/Hen1 RNA repair and immunization system (Chan et al. 2009) and

defense against the Killer and Zymocin anti-fungal toxins (Klassen et al. 2008; Lu et al. 2005).

The splicing of tRNAs

The genomes of organisms encode some fraction of tRNA genes with introns, first

discovered as intervening sequence (IVS) in the SUP4 gene (Heinemann et al. 2010; Goodman

et al. 1977). The budding yeast genome contains an estimated 275 tRNA genes, 61 of which

27

contain introns (Lowe and Eddy 1997; Chan and Lowe 2016). Among those, ten isodecoding

tRNAs are supplied exclusively by genes interrupted with introns (Chan and Lowe 2009). The

splicing endonuclease complex (SEN) recognizes the structure of the intron-containing tRNA and

precisely cleaves at each side of the intron, releasing the intron and cleaved tRNA molecule.

Because the catalytic subunits of the SEN complex catalyze cleavage independent of metal ions,

the resulting tRNA fragments have 2′,3′-cyclic phosphate (2′,3′-cP) and a 5′-hydroxyl (5′-OH)

termini (Trotta et al. 1997).

In 1986, Eric Phizicky and John Abelson purified the ligase responsible for putting the two

halves of the tRNA together (Phizicky et al. 1986). Subsequent studies in John Abelson’s and

Chris Greer’s labs revealed the mechanistic steps tRNA Ligase (Trl1, alias: Rlg1 for RNA ligase

1) uses to ligate tRNAs (Greer et al. 1983). Trl1 performs RNA repair by preparing the termini for

ligation and then forming the phosphodiester bond in three successive steps (Figs 1.2 & 1.3A):

(i) the cyclic phosphodiesterase domain (CPDase) of Trl1 hydrolyzes the 2′,3′-cP to form a 2′-

phosphate, 3′-hydroxyl terminus; (ii) the RNA 5′-kinase domain phosphorylates the 5′-OH to

form a 5′-PO4; (iii) the ligase domain adenylates the 5′-PO4 to form a 5′-App intermediate, and

uses the 3′-OH to displace the adenylate, forming a ligated product with a 3′→5′ phosphodiester

bond and a residual 2′-phosphate at the junction of ligation.

The ligated tRNA is released from Trl1 with a residual 2′-PO4, which has to be removed for

the tRNA to undergo further modification and function in translation (Culver et al. 1997; Spinelli

et al. 1997). The 2′-phosphotransferase enzyme, Tpt1, specifically catalyzes the removal of the

2′-PO4 via a transesterification reaction of the phosphate onto a molecule of NAD+, yielding a

canonical 5′-3′-phosphodiester with an adjacent 2′-OH and an unusual product: ADP-ribose-

1″,2″-cyclic phosphate (McCraith and Phizicky 1990; 1991; Culver et al. 1993). Sherry Spinelli

and Eric Phizicky isolated the TPT1 gene, a highly-specific 2′-dephosphorylation enzyme, and

28

showed, using the conditional mutant tpt1-1, that spliced tRNAs are under-modified at

nucleotides near the splice site (Spinelli et al. 1997).

In humans, some tRNAs also require splicing by the conserved TSEN complex (tRNA

splicing endonuclease), with the catalytic subunits conservatively named TSEN2 and TSEN34

(Paushkin et al. 2004). However, in contrast to yeast, human tRNA splicing uses a different ligase

with a more direct mechanism (Abelson et al. 1998). In 2011 Johannes Popow and Javier

Martinez discovered the ligase, HSPC117 in humans, using an activity-guided purification

procedure on HeLa cell extracts, and they confirmed the requirement of the enzyme in cells via

RNAi knock-down (Popow et al. 2011). Popow et al. noted that the high degree of sequence

conservation between human HSCP117 and E. coli RtcB suggests conservation of the ligase

function(s), even between such distantly related organisms. Given that the RNA ligase in animals

RtcB, which does not yield ligation products with 2′-phoshphates (Chakravarty et al. 2012), is

able to activate the UPR (Lu et al. 2014), it is surprising that animal cells also express a 2′-

phosphotransferase, Trpt1. However, Trpt1 is dispensable in mice (Harding et al. 2008), calling

into question why this conserved enzyme remains present in an organism so evolutionarily

distant from the last common ancestor that used an RNA ligase that generated 2′-

phosphorylated products.

Once spliced out of the tRNAs, the introns are degraded by Xrn1, the primary means of

degrading these RNAs (Wu and Hopper 2014), but an intervening phosphorylation step would

be required to make the 5′-OH introns (as released by the Sen2 in the SEN complex) into a 5′-

PO4 substrate of Xrn1 (see kinase-mediated decay section).

The unfolded protein response (UPR)

The unfolded protein response (UPR) is a homeostatic intracellular signaling pathway,

conserved throughout eukaryotes, that signals to the nucleus during Endoplasmic Reticulum

(ER) stress and ultimately increases the protein folding capacity of the ER. Misfolding proteins,

29

protein processing overload, oxidative stress, calcium efflux, and viral infection each constitute

a stress to the ER and can cause UPR activation (Kozutsumi et al. 1988). Activation of the UPR

increases the protein folding capacity of the cell by upregulating expression of dozens of genes

like chaperones, membrane lipid biosynthesis genes, and ER-resident protein degradation

genes, ultimately resulting in an enlargement of the organelle and an increase of the ER’s protein

folding capacity (Schuck et al. 2009; Cox et al. 1997; Scheuner et al. 2001; Travers et al. 2000;

Casagrande et al. 2000).

As a homeostatic process, the UPR can be physiologically activated (i.e. adaptive/healthy)

in cells with high protein-processing requirements, like pancreatic β cells secreting insulin,

activated B-cells (plasma cells) secreting antibody, differentiating T-cells dividing rapidly, muscle

cells, and in active neurons synthesizing large amounts of protein (Okabayashi et al. 1985;

Reimold et al. 2000; Pramanik et al. 2018; Reimold et al. 2001; Clauss et al. 1993; Wang et al.

2010; Tan et al. 2018). However, UPR activation can be pathological (i.e. maladaptive/unhealthy)

in genetic diseases that cause protein misfolding (e.g. retinitis pigmentosa, cystic fibrosis) and

in type II diabetes (Ozcan et al. 2004). Additionally, and for the purposes of study, the UPR can

be activated by small molecules that put stress on the endoplasmic reticulum. Tunicamycin is a

potent inhibitor of N-linked glycosylation of proteins, thereby reducing protein stability in the ER

(Duksin and Mahoney 1982). Thapsigargin is often chosen for inducing the unfolded protein

response in mammalian cells, where it inhibits the sarcoplasmic Ca2+ ATPase, disrupting ionic

homeostasis in the ER (Rogers et al. 1995; DuRose et al. 2006). Dithiothreitol (DTT) can also be

used to induce the UPR by breaking the disulfide bonds formed in the ER, destabilizing those

proteins (Lee 1992).

While the UPR is present in all eukaryotes, different phyla possess different sets of

pathways of transmitting the stress signal, with the most conserved arm of being the Ire1

signaling pathway. Ire1 is the transmembrane sensor protein that conveys the signal of unfolded

30

proteins from within the lumen of the ER to the cytoplasm. Ire1 conveys this signal via two

enzymatic activities of its cytoplasmic domain: (i) it autophosphorylates other Ire1 molecules (in

trans) upon oligomerization caused by accumulating unfolded proteins; and (ii) Ire1 activates its

endoribonuclease domain, which can cleave cytoplasmic RNAs near the ER surface.

In budding yeast, Saccharomyces cerevisiae, Ire1 is the only known mechanism of

signaling ER stress to the nucleus. The discovery of the unprecedented mechanism of signaling

(and mRNA splicing) was demonstrated by a series of experiments published nearly

simultaneously by Peter Walter’s lab at UCSF and Kazutoshi Mori, first at UT Southwestern and

then at Kyoto University. Walter and Mori jointly received the Lasker Award in 2014 for their

discovery. But the idea that cells tailor their protein processing capacity on the basis of stress

was supported by initial observations of glucose-regulated proteins (GRPs) increasing in

expression not just under glucose starvation stress (Lee et al. 1983), but also under treatment

with amino acid analogues (Kelley and Schlesinger 1978). The connection between protein

metabolism and chaperones GRP78 (BiP, Kar2 in yeast) and GRP94 was inferred by expressing

influenza hemagglutinin gene alleles known to misfold and accumulate in the ER, which in turn

caused GRP78 and GRP94 expression to rise (Munro and Pelham 1986; Normington et al. 1989;

Kozutsumi et al. 1988).

Studies of the BiP gene of yeast, KAR2 (Normington et al. 1989; Rose et al. 1989;

Nicholson et al. 1990), set the stage for the discovery of Ire1 and Hac1 as signal transducers of

the unfolded protein response. A 22-bp conserved sequence in the promoter of KAR2 appeared

in the promoters of many other UPR-inducible genes and was sufficient to promote transcription

of a reporter gene (lacZ) upon induction of protein folding stress (Mori et al. 1992). Building on

that result, Mori then showed that the 22-bp element (named the UPRE for “UPR Response

Element”) is required for KAR2 induction under protein folding stress (Kohno et al. 1993). The

Walter and Mori labs cloned the UPRE upstream of a lacZ gene expressed in yeast, effectively

31

making a colorimetric screen for deletion mutants with defective UPR signaling. Using this

approach, Walter and Mori independently identified Ire1 (Inositol-requiring enzyme 1, alias: Ern1)

as a required component of the signaling pathway (Cox et al. 1993; Mori et al. 1993). Because

Mori also showed that Ire1 is a transmembrane protein with its N-terminus in the lumen and that

the cytoplasmic C-terminal kinase is required for signaling, Mori et al. hypothesized that Ire1

directly phosphorylated a latent transcription factor, causing it to activate, localize to the nucleus,

and drive expression of UPRE genes.

Consistent with that hypothesis, Mori’s group identified the transcription factor of the UPR,

HAC1 (aliases: Ern4 & Ire15), using a yeast one-hybrid screen that showed that Hac1 binds the

UPRE in vivo, activates transcription of UPRE genes, and is required for UPR signaling (Nojima

et al. 1994; Mori et al. 1996). Soon after, Peter Walter’s group used a synthetic overexpression

assay to determine that HAC1 is the transcription factor of the UPR, but that its expression is

regulated by a splicing event (Cox and Walter 1996).

The discovery that UPR signaling was transduced through splicing, not phosphorylation,

was compounded by the even more unexpected finding that yeast’s tRNA ligase was required

for splicing (and therefore signaling) (Cox and Walter 1996; Sidrauski et al. 1996). Walter and

colleagues found that Hac1 protein was only present in UPR-induced cells (as assayed by

immunofluorescence and western blot) and not in uninduced cells. This finding is inconsistent

with the typical mode of regulating a transcription factor’s activity, where the protein is present

at all times but inactive without stimulation. Northern blot analysis showed that the mRNA for

HAC1 decreased in length by about 250 nucleotides (nt) upon induction of the UPR in an Ire1-

dependent manner. Sequencing of the shortened HAC1 RNA species revealed that a 252 nt

“intervening sequence” (IVS) was missing from the shortened species (Cox and Walter 1996).

Furthermore, a random-mutagenesis genetic screen (sectoring screen) for mutants unable to

lose a KAR2 plasmid, indicating a defect in the UPR, uncovered an allele, rlg1-100, of the

32

essential tRNA ligase, TRL1 (alias RLG1) (Sidrauski et al. 1996). TRL1 is essential in yeast

because it ligates together tRNAs after the splicing endonuclease cleaves out the intron (Phizicky

et al. 1992), and so the rlg1-100 allele was discovered to be a separation-of-function mutant

allele, still enabling the yeast to ligate tRNA haves together, as required for viability, but not able

to ligate the halves of the HAC1 mRNA. Soon after, Walter’s group found Ire1 to be a site-specific

endoribonuclease that cleaves the HAC1 mRNA in a stress-dependent manner, providing a

mechanism for how information in the endoplasmic reticulum can be communicated to the

nucleus (Sidrauski and Walter 1997; Kawahara et al. 1998).

The aforementioned experiments revealed the only known signaling pathway entailing the

cytoplasmic splicing of an mRNA (Fig. 1.4): Normally bound by Kar2 under healthy conditions,

Ire1 instead binds unfolded or misprocessed proteins under conditions of ER stress, causing it

to oligomerize and activate the endoribonuclease domain (Gardner and Walter 2011; Lee et al.

2008; Korennykh et al. 2009). Ire1 recognizes HAC1 mRNA via its characteristic stem-loops,

which Ire1 cleaves to release the 5′-exon with a 2′,3′-cyclic phosphate terminus, the 3′-exon

with a 5′-hydroxyl terminus, and the intron with like termini (Gonzalez et al. 1999). The 5′- and

3′-exons are held together by extensive base-pairing, and so the termini are robustly ligated

together by Trl1 (as described previously). This spliced mRNA encodes a potent transcription

factor that localizes to the nucleus and upregulates dozens of genes, about 7% of the yeast

genome (Travers et al. 2000).

Hints that HAC1u mRNA processing was both necessary and sufficient for robust UPR

signaling came from expression of manipulated HAC1 ORFs and mRNAs, with the 5′-exon/UTR,

intron, and 3′-exon/UTR each responsible for important regulatory functions (Fig. 1.5).

33

Figure 1.4: Model of S. cerevisiae unfolded protein response activation

Unfolded proteins in the endoplasmic reticulum (ER) cause Ire1 to oligomerize and activate its cytoplasmic endoribonuclease domain, which site-specifically recognizes the cleavage site flanking the intron of HAC1 mRNA. Ire1 cleaves the intron out from the mRNA, at which point Trl1 ligates the two exons together, leaving a 2′-phosphate at the ligation junction due to its catalytic mechanism. Tpt1 removes the 2′-phosphate from the HAC1 mRNA, leaving the ligation event seamless. HAC1 mRNA is then translated to produce Hac1, a potent transcription factor that is trafficked to the nucleus to upregulate dozens of genes that increase protein folding capacity of the cell, thus relieving ER stress and re-establishing homeostasis. Notably, Hac1 promotes transcription of its own mRNA, making HAC1 splicing a positive feedback loop. Cleavage of HAC1 mRNA is reported to be the rate-limiting step in the pathway.

34

Figure 1.5: Three key mechanisms that couple HAC1 splicing to its translation.

HAC1 mRNA has two key features that ensure its translation is inhibited by the presence of the intron, and one feature that makes the signaling cascade quick to respond to ER stress. Firstly, HAC1 intron forms a strong base-pairing interaction with the 5′-untranslated region (5′-UTR) of the message, effectively preventing scanning ribosomes from reaching the open reading frame to begin translation (Rüegsegger et al. 2001). Secondly, in the event that a ribosome escapes inhibition and initiates translation on unspliced HAC1, a failsafe ORF encoded in the intron, in frame with the ORF of the 5′–exon, codes for a potent degron, resulting in rapid ubiquitination and degradation of the resulting protein (Di Santo et al. 2016). Thirdly and lastly, HAC1 3′-UTR folds into the bipartite element (BE) that anchors the mRNA to the cytoplasmic surface of the ER, near Ire1, on standby for rapid response to ER stress (Aragón et al. 2009).

35

HAC1u was found co-sedimenting with polyribosomes (Chapman and Walter 1997) despite there

being very little Hac1u (Unspliced Hac1) protein produced in unstressed cells (Cox and Walter

1996); this paradox was resolved by results indicating that the intron of HAC1u acts to prevent

scanning ribosomes from initiating translation on the unspliced mRNA (Rüegsegger et al. 2001).

In 2000, Kazutoshi Mori and Peter Walter both showed that swapping of the C-terminal domain

of Hac1 via the splicing of its mRNA was required for robust accumulation of the protein

transcription factor, but the C-terminal domain of Hac1s (spliced Hac1) was dispensable for

transcriptional activation (Mori et al. 2000). The Hac1 transcription factor upregulates

transcription of its own locus, HAC1, and that positive feedback loop is required for sustained

UPR activation (Ogawa and Mori 2004). HAC1u mRNA is tethered to the cytoplasmic surface of

the ER by the bipartite element (BE) in its 3′-UTR, a requirement for its rapid processing in the

event of protein folding stress (Aragón et al. 2009). And as recently as 2016, studies from David

Weinberg’s lab showed that the ORF of the intron of HAC1u in frame with that of the 5′-exon

encodes a potent degron, causing the protein resulting from spurious translation of HAC1u to be

rapidly degraded by the ubiquitin-proteasome pathway, as mediated by the ubiquitin ligase

DUH1 (alias: DAS1) (Di Santo et al. 2016). Importantly, both the translation initiation block and

the intron-encoded degron work together to maintain tight suppression of Hac1 protein

production when the cell is not experiencing ER stress, and thus not splicing HAC1.

RNA repair in bacterial stress responses and interspecies conflict

Microbes are subjected to quickly changing environmental conditions and to direct conflict

with microbes of other species. Two interesting cases of RNA breakage and repair occur in

microbes in response to stress and to interspecies conflict. In both instances, a situation of

broken RNA is rectified by repair. The RNA cleavage can be caused either by the microbe itself

or a competing microbe, but the repair is carried out to aid in recovery. Both examples support

the idea that RNA repair constitutes an adaptive function across many phyla.

36

E. coli under stress adapt their protein synthesis program by cleaving their rRNA and

mRNA, causing the ribosomes to favor translating a specialized subset of mRNAs. The

endoribonuclease MazF is part of a the mazEF toxin-antitoxin system (Aizenman et al. 1996;

Engelberg-Kulka et al. 2006) and was identified as the endoribonuclease that enacts this

translational program. MazF cleaves the 3′ 43 nucleotides off of the 16S rRNA, introducing

70S∆43 specialized ribosomes in the cell (Vesper et al. 2011). These ribosomes lack helix 45 and

the anti-Shine-Dalgarno sequence (aSD), both of which are essential for the translation of typical

mRNAs in E. coli (Shine and Dalgarno 1974). The specialized ribosomes preferentially translate

“leaderless” mRNAs that were similarly processed by MazF (Zhang et al. 2005), thus imposing

an alternative translation program.

Given the immense energetic costs of manufacturing a ribosome de novo, Temmel et al.

hypothesized that E. coli may ligate the rRNA back together to restore the ribosomes to their

pre-stress state once the stress passes. They identified RtcB as the ligase that reverses MazF

cleavage of rRNA to restore normal translation after stress (Temmel et al. 2016). Additionally, the

rtcB mRNA is itself processed by MazF, including it in the subpopulation of mRNAs selectively

translated during stress. This discovery provided a biological role for RtcB in E. coli and

demonstrated how RNA cleavage and ligation can be used to regulate an important bacterial

stress response.

PnkP/Hen1 RNA repair and immunization system is a cooperating 2′-O-methyltransferase

and RNA ligase system that repairs RNAs (usually tRNAs) damaged by ribotoxin endonucleases

and installs a methyl group at the 2′-position to block that oxygen from participating in cleavage

reactions in the future, effectively immunizing the RNA from cleavage again by the same

ribotoxin. Hen1, the methyltransferase, was first discovered in plants as a factor required for the

processing of miRNAs (3′-terminal 2′-O-methylation) (Kishi et al. 2005). Based on peptide

sequence similarity, homologous Hen1 proteins are also found in the genomes of various

37

bacteria, but their function in bacteria was unknown (Tkaczuk et al. 2006). Unexpectedly, the

Clostridium thermocellum operon coding for Hen1 contains a downstream gene possessing

kinase, phosphatase, and adenylyltransferase activities, hallmarks of an RNA ligase (Martins and

Shuman 2005). Chan and colleagues cloned and purified the gene from Anabaena variabilis and

showed that, in complex with the Pnkp ligase, the proteins carry out RNA repair on ribotoxin-

cleaved tRNAs and install a methyl group at the 2′-O position of the 5′-cleavage fragment prior

to ligating the fragments together (Chan et al. 2009).

Given that most ribotoxins use a metal-ion-independent mechanism to cleave their target

RNAs in their victims (Ogawa et al. 1999; Soelaiman et al. 2001; Morad et al. 1993; Nariya and

Inouye 2008; Tomita et al. 2000; Pedersen et al. 2003), the 2′-O-methylation is an effective

strategy to overcome the effects of the ribotoxins by blocking further cleavage. Quantitatively,

Wang et al. found a 31% decrease in cleavage susceptibility in vitro, a much-preferred outcome

over creating a futile cycle of cleavage and ligation in response to the ribotoxin (Chan et al. 2009).

In the Pnkp-Rnl-Hen1 complex, methylation occurs before ligation, and methylation may even

be conducive to the ligase component because methylated substrates are ligated 10-fold faster

(Wang et al. 2015). Furthermore, from a structural perspective, the methyltransferase active site

is situated between the 3′-phosphatase and the ligase active sites, maximizing the likelihood of

methylation of substrates prior to ligation.

Terminus chemistry and exonucleolytic RNA decay

The chemistry of the end of a molecule of RNA can have major impacts on which

exonucleases will engage with it and at what rates they will catalyze degradation. The three

known 5′→3′ exoribonucleases all require 5′-PO4 termini to catalyze decay. 3′→5′

exoribonucleases have a more heterogeneous set of requirements. These biochemical signals

sent from the RNA termini to the exonucleases operating on them represent a form of regulation

38

of their activity in decay and processing. Salient exoribonucleases are discussed here and

summarized in Table 1.1.

5′→3′ RNA decay enzymes require substrates be 5′-phosphorylated

All RNA is subject to decay, and cells require regulated RNA decay machinery to adapt to

their environment by changes in behavior through new signals sent from the nucleus. Messenger

RNAs are capped at their 5′-ends for protection from RNA degradation machinery in the cell

(Furuichi et al. 1977; Shimotohno et al. 1977). One of the primary decay factors against which

caps protect their RNAs is a 5′-phosphate dependent 5′→3′ exoribonuclease discovered by

Audrey Stevens by purifying it from yeast (Stevens 1980). Later named XRN1 (for

exoribonuclease 1) (Larimer and Stevens 1990), Xrn1 was shown to be abundant in cells

(Ghaemmaghami et al. 2003) and to processively degrade its substrates (Stevens 2001).

In budding yeast, the poly-adenylate tail (poly(A) tail) is a master regulator of mRNA

stability. Typically, RNA decay is initiated via deadenylation (Decker and Parker 1993) by the

Ccr4/Not complex (Tucker et al. 2001) or the Pan2/Pan3 complex (Boeck et al. 1996; Brown and

Sachs 1998). Deadenylation is a highly regulated process, with several RNA-binding proteins

known as actors in this regulation (Parker 2012). The principal regulator of deadenylation is the

Poly(A)-Binding protein, Pab1, which inhibits deadenylation by the Ccr4 complex when bound

to poly(A) tails (Caponigro and Parker 1995).

Once an RNA is deadenylated, the most common outcome is that the decapping complex

(Dcp1/Dcp2) hydrolyzes the 7-methylguanosine cap of the RNA to produce a 5′-phosphate,

exposing the 5′-terminus to degradation by Xrn1 (Hsu and Stevens 1993). Alternatively,

decapped RNAs can also be degraded in a 3′→5′ scheme by the cytoplasmic RNA exosome

(Anderson and Parker 1998). Comparatively, the 5′→3′ decay pathway, as executed by Xrn1,

appears to be more consequential to exponentially-growing cells.

39

Table 1.1: Exoribonucleases and their terminus requirements

Name Direction Species Active Termini

Inhibitory Termini

Reference

Xrn1/Xrn4 5′→3′ Eukaryotes 5′-PO4 5′-OH (Stevens 2001)

(Nagarajan et al. 2013)

Xrn2/Rat1 5′→3′ Eukaryotes 5′-PO4 5′-OH (Johnson 1997) (Geerlings et al. 2000)

Dxo1 5′→3′ Yeast 5′-PO4 5′-OH (Chang et al. 2012)

Dis3/Rrp44 3′→5′ Eukaryotes 2′,3′-OH,

2′,3′-cP N/A

(Meaux and van Hoof 2006)

Rrp6 3′→5′ Eukaryotes 2′,3′-OH 3′-PO4 (Burkard and Butler 2000)

Usb1 3′→5′ Yeast 2′,3′-OH

2′,3′-cP 3′-PO4 (Didychuk et al. 2017)

PNPase 3′→5′ Bacteria 2′,3′-OH 2′-, 3′-PO4 (Munir et al. 2018a)

40

Cells defective for cytoplasmic 5′→3′ RNA decay (e.g. xrn1∆, dcp2∆) exhibit slow growth or are

inviable (depending on genetic background), whereas yeast with mutations in 3′→5′ decay

factors (e.g. ski2∆, ski3∆) (discussed further below) exhibit a very modest growth phenotype

(Giaever et al. 2002). Interestingly, cells lacking any combination of both 5′→3′ decay and 3′→5′

decay are inviable (Anderson and Parker 1998), indicating that the decay pathways can

compensate for one another, leading to a synthetic lethal genetic interaction when ablated

simultaneously. A small proportion of the RNA in a yeast cells may decapped independent of

deadenylation (Muhlrad et al. 1995), and some RNAs may infrequently be subject to

endonucleolytic cleavage as a means of regulated decay (unpublished data, Y. Harigaya & R.

Parker) (Parker 2012), but the overarching trend is that decay is regulated primarily at the

deadenylation stage, followed by decapping, and then finally by exonucleolytic degradation (Fig.

1.6).

Beyond Xrn1, two alternative 5′-phosphate-dependent 5′→3′ decay pathways exist:

Rat1/Xrn2 and Dxo1. Rat1 (alias: Xrn2) is an essential nuclear exonuclease found by a screen for

mutants unable to export poly(A) RNA into the cytoplasm (giving it the name Ribonucleic Acid

Trafficking 1) (Amberg et al. 1992). Overexpression of wild-type Rat1 does not completely rescue

cells from the growth and RNA decay defects imparted by deleting Xrn1, but Rat1 mutants that

increase localization to the cytoplasm fully rescue the xrn1∆ phenotype (Poole and Stevens

1995). Conversely, Xrn1 tagged with a nuclear localization signal can complement the rat1-1

conditional mutant at non-permissive temperatures (Johnson 1997). Thus, Xrn1 and Rat1 are

functionally similar exoribonucleases that have become compartmentalized in eukaryotes.

41

Figure 1.6: Typical mRNA decay in the cytoplasm by Xrn1 and the exosome.

(1) The Ccr4/Not complex or Pan2/Pan3 complex deadenylate the poly(A) tail of the mRNA. (2) Dcp1 and Dcp2 are induced by to short poly(A) tail to de-cap the mRNA, producing a 5′-PO4 terminus. (3) Xrn1 is the primary decay factor in the cytoplasm, and it rapidly and processively degrades 5′-PO4 RNAs in the 5′→3′ direction. (4) The exosome is the secondary decay factor of the cytoplasm and is composed of a complex of structural and regulatory proteins, as well as the exoribonuclease Rrp44. The exosome degrades deadenylated mRNAs in the 3′→5′ direction. (Parker 2012)

42

A third enzyme was discovered to have distributive 5′→3′ exoribonuclease activity during

structural studies that were focused on decapping (Chang et al. 2012). Thus, Decapping

eXOnuclease 1 (Dxo1) was named for its combined nuclease activities on methylated and

unmethylated 7-methylguanosine caps and 5′-phosphorylated RNA. Dxo1 is structurally similar

to Rai1 (Rat1 Interacting protein 1), a nuclear decapping enzyme with affinity toward

unmethylated 7-methylguanosine caps, and Dom3Z, the mammalian homolog of Rai1 (Xue et al.

2000).

3′→5′ RNA degradation

A second RNA decay pathway executed by the cytoplasmic exosome in yeast can degrade

RNA from the opposite terminus, proceeding in a 3′→5′ direction (Fig. 1.6). A generalized 3′→5

decay pathway was proposed because mutants of components required for 5′→3′ decay are

viable and continue to degrade most RNAs, albeit at slower rates (Hsu and Stevens 1993;

Muhlrad et al. 1994; 1995; Beelman et al. 1996), and mutations of both the 5′→3′ and 3′→5′

pathways are synthetic lethal (Johnson and Kolodner 1995). The exosome was identified as a

multi-protein complex via an immunoprecipitation experiment performed on Rrp4 (Ribosomal

RNA Processing 4), a protein involved in 3′→5′ trimming of the 5.8S rRNA (Mitchell et al. 1996).

Mass spectrometry identified proteins associated with Rrp4, Rrp41 (Ski6), Rrp42, Rrp43, and

Rrp44 (Dis3), all required for 3′→5′ processing of the 5.8S rRNA, and the complex was dubbed

the exosome. In vitro characterization demonstrated that: Rrp4 is a distributive RNA hydrolase;

Rrp44 is a processive RNA hydrolase; and Ski6 (Rrp41) is a processive phosphorolytic enzyme

(Mitchell et al. 1997). Further structural studies revealed the exosome complex to be composed

of 10 key proteins: the exonuclease and endonuclease domain-containing Dis3 (Rrp44) protein,

homologous to E. coli RNase R (vacB), and a nine-member ring structure formed by six RNase

PH-like proteins (Rrp41, Rrp45, Rrp46, Rrp43, Mtr3, and Rrp42) and three small RNA-binding

43

proteins (Rrp4, Rrp40, and Csl4) analogous to bacterial PNPase (Mian 1997; Cheng et al. 1998;

Liu et al. 2006).

The activity of the exosome is largely determined by its association with regulatory

substrate recognition complexes (Araki et al. 2001). Beyond rRNA processing, subsequent

studies indicated that the exosome is present in the cytoplasm and functions as a general

pathway for degradation of mRNAs. In addition to Ski6, Anderson and Parker demonstrated that

Ski2, Ski8, and the adaptor protein Ski7 function as part of the cytoplasmic exosome, performing

both routine turnover of RNA, as well as degradation of aberrant mRNA, like in non-stop and no-

go mRNA decay (Anderson and Parker 1998; Wang et al. 2005). In contrast to the principal

enzymes of 5′→3′ decay, there is mixed evidence for terminus requirements for the RNA

exosome. Stacie Meaux and Ambro van Hoof showed, using self-cleaving ribozymes, that the

yeast cytoplasmic exosome degraded 2′,3′-cP just as well as 2′,3′-cis-diol terminated RNAs

(Meaux and van Hoof 2006). However, 3′-PO4 RNAs appear to be inhibitory to Rrp6, the catalytic

subunit of the nuclear exosome, similar to how DNA (with its 2′-H) may be inhibitory (Burkard

and Butler 2000). Usb1 (U Six Biogenesis 1), the 3′→5′ exoribonuclease in yeast essential for

U6 snRNA biogenesis, is also inhibited by 3′-PO4 (Didychuk et al. 2017). As Usb1 degrades the

pre-U6 snRNA, which starts off with a 2′,3′-cis-diol, Usb1 generates 2′,3′-cP with each

exonucleolytic step. Interestingly, Usb1 has a second activity of cyclic-phosphate hydrolysis to

a 3′-PO4, which inhibits further exonuclease activity, thereby preventing Usb1 from degrading

too much RNA off the 3′-end of U6 snRNA.

In addition to terminus chemistry requirements, a recent discovery provided an example

of a 3′→5′ exoribonuclease (phosphorylase) that is blocked by a non-canonical internal linkage.

Dissatisfied with the ability to resolve 2′-phosphorylated RNA 15-mers on polyacrylamide gels,

Stewart Shuman discovered that PNPase (PolyNucleotide Phosphorylase) from Mycobacterium

smegmatis is halted site-specifically by internal 2′-PO4 (Munir et al. 2018a).

44

Kinase-Mediated Decay

A major caveat of 5′→3′ RNA decay is that all three known exonucleases that degrade

5′→3′, Xrn1, Rat1, and Dxo1, require 5′-phosphates of their substrates. This phenomenon is

consequential because many metal-ion-independent RNA incision events produce 5′-hydroxyl

RNAs: Sen2/Sen34 cleavage of tRNAs (Knapp et al. 1979); Ire1 cleavage of HAC1 mRNA

(Gonzalez et al. 1999); Rny1 and angiogenin (Thompson and Parker 2009a); no-go mRNA decay

(Navickas et al. 2018); intrinsic cleavage of RNA. These products of cleavage are effectively

immune to 5′→3′ decay because the enzymes of 5′→3′ decay do not recognize the 5′–OH RNAs

as substrates. Any evidence of their 5′→3′ decay would necessarily imply a 5′-phosphorylation

event, as would be provided by a 5′-RNA kinase, hence Kinase-Mediated Decay (KMD).

Examples of KMD already exist: tRNA introns, pre-rRNA, no-go mRNA decay 3′-fragments, and

T4 bacteriophage mRNA.

KMD of tRNA introns

Introns from tRNAs, once cleaved out from transcripts, undergo kinase-mediated decay.

The catalytic subunits of the SEN complex, Sen2 and Sen34, catalyze metal-ion-independent

cleavage, and release 5′-OH and 2′,3′-cP products (Knapp et al. 1979). Anita Hopper uncovered

an intron retention defect for tRNAIle(UAU) in xrn1∆ yeast via an unbiased reverse-genetic screen,

indicating that the 5′-OH tRNA introns are converted to 5′-PO4 in order to be degraded by Xrn1

(Fig. 1.7A) (Wu et al. 2015), A follow-up paper showed that, in addition to tRNAIle(UAU), liberated

introns from tRNALeu(CAA), tRNALys(UUU), tRNATrp(CCA), and tRNAPro(UGG) accumulated in the

absence of Xrn1 (Wu and Hopper 2014). Furthermore, using the previously described conditional

mutants of the tRNA ligase rlg1-4 and rlg1-10 (Phizicky et al. 1992), Hopper and Wu showed that

ligase-mutant yeast grown at the non-permissive temperature also accumulated liberated tRNA

introns.

45

Figure 1.7: Examples of Kinase-Mediated Decay (KMD)

A: The splicing endonuclease (SEN) complex cleaves the intron out of the pre-tRNA, releasing it as a 5′-OH. The kinase domain of Trl1 phosphorylates the 5′-terminus, yielding a 5′-PO4, which is a substrate or Xrn1. (Wu et al. 2015)

B: Las1 cleaves at the C2 site in internal transcribed spacer 2 (ITS2) of the 27SB pre-rRNA, releasing a 26S pre-rRNA with a 5′-OH terminus. Grc3 RNA 5′-kinase phosphorylates the 5′-end, yielding a 5′-PO4 substrate of Rat1/Xrn2. (Gasse et al. 2015)

C: No-go mRNA decay (NGD) releases cleavage fragments with a 5′-OH by an unknown nuclease/mechanism. Trl1 phosphorylates the 5′-terminus of the 3′-cleavage fragment, yielding a 5′-PO4, which is a substrate or Xrn1. (Navickas et al. 2018)

D: During T4 phage infection of E. coli, the regulatory nuclease RegB cleaves the Shine-Dalgarno sequence of “early” RNAs during the transition to “late” gene expression. The cleaved mRNAs are released with 5′-OH termini and require phosphorylation by T4 polynucleotide kinase (PNK) for recognition and decay by host RNase E and RNase G. (Durand et al. 2012)

46

The experiment elegantly demonstrated that the tRNA introns have to be phosphorylated by the

5′-RNA kinase domain of yeast tRNA ligase (Trl1) prior to degradation by Xrn1, constituting an

example of kinase-mediated decay (Fig. 1.2).

KMD of rRNA processing intermediates

A second example of the kinase-mediated decay phenomenon was discovered at work

during ribosomal RNA (rRNA) processing. During the production of 25S and 5.8S rRNAs, their

tandem polI transcript has to be cleaved and trimmed to form rRNAs of the mature length. Lisa

Gasse and Ed Hurt identified the endoribonuclease that cleaves at the C2 position in ITS2 of the

27SBS pre-rRNA as Las1, a metal-ion-independent endonuclease. Once Las1 cleaves ITS2 at

the C2 site, yielding a 26S rRNA with a 5′-OH, the 5′–P exonuclease, Rat1, “trims” the 26S pre-

rRNA. Gasse et al. showed that the polynucleotide kinase Grc3 phosphorylates the 5′-terminus

of the 26S pre-rRNA to allow Rat1 to trim ribonucleotides off the 5′-end (Fig. 1.7B) (Gasse et al.

2015). This sequential generation of 5′-OH RNA, phosphorylation to a 5′-PO4, and

degradation/chew-back by a 5′-PO4 RNase (Rat1/Xrn2) constitutes kinase mediated decay as a

mechanism of rRNA maturation.

The factors that carry out this example of KMD are conserved in human pre-rRNA

processing. The LAS1L endonuclease cleaves both the C1′ site and 30 nt 3′ of the E′ site of the

ITS2 spacer in 45S pre-rRNA, producing the 6SS and 26S pre-rRNAs (Schillewaert et al. 2012).

Similar to Grc3, human Nol9 is the nucleolar RNA 5′-kinase (Heindl and Martinez 2010) and likely

phosphorylates the 5′-terminus of the 26S pre-rRNA, preparing it for trimming by XRN2, the

human 5′-phosohate-dependent 5′→3′ exoribonuclease homologous to Rat1 from budding

yeast. Thus, kinase-mediated decay is likely a conserved mechanism of pre-rRNA processing in

humans.

47

KMD of 3′-fragments of no-go mRNA decay (NGD)

The coding quality of mRNA is controlled through a family of processes called co-

translational mRNA surveillance, an umbrella term for a series of processes that monitor

ribosome progress along the open reading frame to detect mRNAs that are defective in

translation. Adaptor proteins interact with the ribosome to detect aberrant mRNAs and direct

them toward RNA decay pathways. No-go mRNA decay (NGD) is one of these co-translational

mRNA surveillance processes that detects and rectifies stalled ribosomes on messages (Doma

and Parker 2006). When ribosomes encounter strong stalls in translation elongation, such as

stem-loops, rare codons, poly-positively-charged amino acids, NGD is activated and the mRNA

is cleaved, releasing 5′ and 3′ RNA products. The cytoplasmic exosome degrades the 5′-product

of endonucleolytic cleavage, and Xrn1 degrades the 3′-product.

While the endoribonuclease that catalyzes the cleavage of no-go decay mRNAs remains

unknown, the cleavage site reliably occurs 5′ of the elements that causes the ribosome(s) to

stall. In a 2018 bioRχiv pre-print, Navickas, et al. demonstrate that the products of cleavage are

generated with 5′-OH termini (Navickas et al. 2018). With Xrn1 as the required decay factor for

degradation of 3′-NGD fragments (and Dxo1 to a much lesser degree), a 5′-phosphorylation step

would be a required intermediate step. Using the temperature-sensitive allele of Trl1 (rlg1-4),

Navickas, et al. showed that a phosphorylation event was indeed required prior to Xrn1-mediated

decay of the no-go decay product (Fig. 1.7C).

KMD of T4 bacteriophage mRNA

Bacteria have 3′→5′ and 5′→3′ decay programs for RNA, but these exonucleases typically

do not initiate decay (Belasco 2010). Instead, RNA decay typically starts with an endonucleolytic

cleavage event that generates substrates for exonucleases in the cell. One example of this decay

mode occurs via a cascading series of endonucleolytic cleavages that generate short substrates

for 3′→5′ exoribonucleases. To initiate decay, the bacterial RNA pyrophosphohydrolase (RppH)

48

converts the 5′-triphosphate terminus to a 5′-phosphate (Celesnik et al. 2007). Next, RNases E

and G recognize the 5′-PO4 terminus and cleave downstream, generating yet another (shorter)

5′-PO4 RNA that serves as a substrate for RNases E and G until the remaining RNA is sufficiently

short (Mackie 1998; Spickler et al. 2001). The fragmented RNA products of RNases E and G are

degraded by the bacterial 3′→5′ exoribonucleases RNase R, RNase II, and polynucleotide

phosphorylase (PNPase) (Andrade et al. 2009).

During infection of E. coil, T4 Bacteriophage must coordinate the expression of early and

late genes. The phage uses a virally-encoded endoribonuclease, RegB, to site-specifically cleave

at the Shine-Dalgarno sequence of early genes (Uzan et al. 1988; Sanson et al. 2000). RegB is a

metal-ion-independent endoribonuclease and so generates 5′-OH products of cleavage (Saïda

et al. 2003). However, with RNases E and G being 5′-phosphate-dependent decay factors, an

intervening phosphorylation step would be necessary for them to degrade early mRNAs on the

phage’s behalf. T4 phage supply a polynucleotide kinase (PNK) that phosphorylates these 5′-

OH termini and permits RNases E and G to degrade the early genes after RegB cleavage (Fig.

1.7D) (Durand et al. 2012). This RegB-directed KMD mechanism presents an interesting example

of KMD in prokaryotes, the only example known to date. The example is also remarkable

because of the cooperation the T4 phage exhibits by adapting its viral mRNA regulatory system

to the endonuclease and kinase of the general RNA decay machinery of the bacterium.

Toward mutants of RNA repair

To date, investigations of the role of RNA repair have occurred through temperature-

sensitive (conditional) mutants of the tRNA ligase, TRL1/RNL1, requiring shifting the yeast to the

non-permissive temperature for experiments in the absence of ligase functions (Phizicky et al.

1992), complicating the study of substrates of RNA repair. Furthermore, conditional mutants of

49

the 2′-phosphotransferase, TPT1, obscure detection of the products of ligation, as they are

marked by a residual 2′-phosphate at the ligation junction (Spinelli et al. 1997).

An outright genetic deletion of the genes TRL1 or TPT1 is not possible because they are

essential, but those genotypes would be useful for studying the substrates and products of RNA

repair, as well as any additional contributions the enzymes make to RNA processing, such as

kinase mediated decay. In Chapter II, I present my studies on performing a genetic bypass of

the essential genes TRL1 and TPT1 using a construct to express “pre-spliced” tRNAs, showing

that the only essential function of Trl1 and Tpt1 is to splice intron-containing tRNAs. I go on to

use the genetic bypass, called the “10x tRNA Block,” in Chapter III, to describe how RNA repair

regulates the unfolded protein response (UPR) downstream of Ire1-mediated cleavage, using

KMD to effectively suppress the UPR under normal conditions and to facilitate activation of the

UPR upon endoplasmic reticulum stress. In Chapter IV, I propose using the findings of these

studies to discover substrates and products of RNA repair in both fungi and metazoans and

stabilize RNAs with 5′-terminal 2′-phosphates.

50

CHAPTER II

II GENETIC BYPASS OF ESSENTIAL RNA REPAIR ENZYMES IN BUDDING YEAST1

Abstract

RNA repair enzymes catalyze rejoining of an RNA molecule after cleavage of

phosphodiester linkages. RNA repair in budding yeast is catalyzed by two separate enzymes

that process tRNA exons during their splicing and HAC1 mRNA exons during activation of the

unfolded protein response. The RNA ligase Trl1 joins 2′,3′-cyclic phosphate and 5′-hydroxyl

RNA fragments, creating a new phosphodiester linkage with a 2′-phosphate at the junction. The

2′-phosphate is subsequently removed by the 2′-phosphotransferase Tpt1, which catalyzes

phosphate transfer to NAD+, producing nicotinamide and a unique ADP-ribose metabolite. I

bypassed the essential functions of TRL1 and TPT1 in budding yeast by expressing “pre-

spliced”/intronless versions of the ten normally intron-containing tRNAs, indicating this repair

pathway does not have additional essential functions. Consistent with previous studies,

expression of intronless tRNAs failed to rescue the growth of cells with deletions in components

of the SEN complex, implying an additional essential role for the splicing endonuclease. The

trl1∆ and tpt1∆ mutants accumulate tRNA and HAC1 splicing intermediates indicative of specific

RNA repair defects and are hypersensitive to drugs that inhibit translation. As expected, failure

to induce the unfolded protein response in trl1∆ cells grown with tunicamycin is lethal owing to

their inability to ligate HAC1 after its cleavage by Ire1. In contrast, tpt1∆ mutants grow in the

presence of tunicamycin despite reduced accumulation of spliced HAC1 mRNA. Finally, I

optimized a PCR-based method to detect RNA 2′-phosphate modifications and show that they

1Published with permission under a Creative Commons License (Attribution-NonCommercial 4.0 International) from RNA: PD Cherry, LK White, K York, and JR Hesselberth. Genetic bypass of essential RNA repair enzymes in budding yeast. 2018. 24: 313-323. doi: 10.1261/rna.061788.117

51

are present on ligated HAC1 mRNA. These RNA repair mutants enable new studies of the role

of RNA repair in cellular physiology.

Introduction

RNA repair enzymes catalyze rejoining of an RNA molecule after cleavage of

phosphodiester linkages. RNA repair is carried out by enzymes that prepare the RNA termini for

ligation, ligate the termini, and “clean-up” the ligated product when required. RNA repair systems

are present in all domains of life and catalyze rejoining of the 5′-hydroxyl and 2′,3′-cyclic

phosphate products of cleavage. RNA repair catalyzes the maturation of endogenous tRNAs and

mRNAs (Schwer et al. 2004), and can also counteract the action of endonucleases during inter-

organismal conflicts (Burroughs and Aravind 2016; Amitsur et al. 1987; Nandakumar et al. 2008).

E. coli RtcB can repair ribosomal RNA cleaved during stress by the endonuclease MazF, thereby

reversing ribosomal heterogeneity and restoring translational activity to MazF-processed

ribosomes (Temmel et al. 2016). The Pnkp–Hen1 RNA repair complex, which is present in more

than 250 bacterial species, combines the enzymatic activities of the bacteriophage T4 RNA

repair system with the Hen1 methyltransferase, which installs a 2′-O-methyl group to “immunize”

repaired RNA molecules against future cleavage by the same ribotoxin (Wang et al. 2015). In

light of these and other recent developments in our understanding of the functional and genetic

diversity of RNA ligases (Burroughs and Aravind 2016; Kosmaczewski et al. 2015), I developed

a genetic bypass of the Saccharomyces cerevisiae RNA repair system, enabling direct study of

the roles of RNA repair proteins in budding yeast.

In budding yeast, S. cerevisiae, after tRNA intron excision, ligation of the exons is carried

out in successive steps (Phizicky et al. 1986). First, the splicing endonuclease (SEN) complex

cleaves the intron at two sites, creating two exon products, one with a 2′,3′-cyclic phosphate

terminus and one with a 5′-hydroxyl terminus, and an intron product with 5′-hydroxyl and 2′,3′-

52

cyclic phosphate termini that is degraded (Wu and Hopper 2014). In a second step, the two

exons are ligated by the RNA ligase Trl1. Trl1 joins substrates with 2′,3′-cyclic phosphate and

5′-hydroxyl termini via: (i) conversion of the 2′,3′-cyclic phosphate to a 2′-PO4/3′-hydroxyl; (ii)

phosphorylation of the 5′-OH; (iii) adenylylation of the 5′-PO4 and iv) nucleophilic attack on the

adenylate by the 3′-OH, producing AMP and a new 5′→3′ phosphodiester linkage with a 2′-PO4

(Greer et al. 1983). The 2′-PO4 left at the ligation junction is removed by the NAD+-dependent 2′-

phosphotransferase Tpt1, creating a canonical 5′→3′ phosphodiester linkage (McCraith and

Phizicky 1990; Culver et al. 1993) (Fig. 2.1A).

Trl1 has a second known role in the cell: activating the unfolded protein response (UPR).

Trl1 ligates the two exons of the HAC1 mRNA after cleavage by Ire1, activating the UPR

(Gonzalez et al. 1999). Ire1 excises an intron from the HAC1 pre-mRNA, and Trl1 subsequently

ligates the HAC1 exons together, enabling its translation into a transcription factor that localizes

to the nucleus and drives transcription of hundreds of stress response genes (Sidrauski et al.

1996). Yeast cells that lack Trl1 and Tpt1 and that express RNA repair enzymes from T4

bacteriophage are viable, but they exhibit low-fidelity HAC1 mRNA cleavage and ligation,

suggesting the 2′- PO4/3′-hydroxyl terminus produced by the cyclic phosphodiesterase domain

of Trl1 directs precise ligation (Schwer et al. 2004). Trpt1, the mammalian 2′-

phosphotransferase, was shown to be dispensable for UPR activation in mammals (Harding et

al. 2008). However, subsequent studies showed that the HSPC117/RtcB RNA ligase—which

does not create 2′-PO4 ligation products (Chakravarty et al. 2012)—activates the mammalian

UPR (Lu et al. 2014), explaining why Trpt1 2′-phosphotransferase activity is dispensable

(Harding et al. 2008). The role of Tpt1 in budding yeast in the UPR has not been previously

explored.

53

Using a genetic bypass strategy, RNA repair was previously shown to be essential for

growth only for fulfilling the need of C. elegans (Kosmaczewski et al. 2014) and trypanosomes

(Lopes et al. 2016) to splice tRNAs. Using a similar strategy, I designed and tested a genetic

bypass for deletion of the essential RNA repair enzymes Trl1 and Tpt1 in budding yeast and

show that rescued trl1∆ and tpt1∆ cells (the “RNA repair mutants”) have unique phenotypes for

both tRNA splicing and HAC1 mRNA splicing during the unfolded protein response.

Materials & Methods

General Methods

Saccharomyces cerevisiae W303 strains were cultured in YPD and synthetic “drop-out”

media (Sherman 2002) for experimental cultures and for plasmid selection, respectively. Single-

copy URA3 SEN plasmids were created by recombining SEN-containing genomic fragments

from the yeast tiling collection (Jones et al. 2008) into Advanced Gateway plasmids (Jones et al.

2008) via Gateway LR reactions (Life Technologies). Plasmid counterselection was performed on

synthetic complete solid media containing 5-fluoroorotic acid (FOA, US Biologicals) (1 mg/mL).

Yeast transformations were performed using a Lithium Acetate PEG-3350 Sheared Salmon

Sperm protocol (Gietz and Schiestl 2007). Genotypes of yeast were confirmed by PCR using

forward and reverse primers flanking the disrupted locus and outward-facing primers within the

disruption cassette. Success of FOA plasmid counterselections was confirmed using RT-PCR

for the shuffled gene in DNased total RNA isolated from shuffled strains and positive controls.

Experiments with tunicamycin (Sigma) treatment were performed at final concentrations of

0.08 µg/mL in solid media and at 2.5 µg/mL in liquid media, with negative controls using DMSO

(solvent). Cultures were incubated at 30°C unless otherwise indicated. UPR inductions ±

tunicamycin were carried out for 2 hours during exponential phase growth. For growth assays,

cells were grown in YPD overnight and liquid cultures were normalized to OD600 = 0.2 before plating

54

tenfold serial dilutions on indicated media. For translation inhibition assays, dilutions were

spotted onto YPD agar supplemented with 2 µg/mL anisomycin (Sigma), 100 ng/mL hygromycin

B (Invivogen), or 100 ng/mL cycloheximide (Sigma).

Northern blotting

Oligonucleotide probes (IDT) were designed to hybridize with RNA species to be analyzed

(Table 2.1). Oligonucleotide probes were 5′-radiolabeled with [γ-32P]-ATP (Perkin Elmer) using

T4 PNK (Enzymatics), and excess unincorporated label was removed with G-25 Sephadex (GE)

spin columns. Probes were heated to 100°C and diluted into 10 mL ULTRAhyb Oligo

hybridization buffer (Life Technologies). Total RNA (normalized to SCR1 loading) was

electrophoresed, along with RiboRuler Low Range RNA Ladder (Thermo) and ss10 ssDNA ladder

(Simplex Sciences) on pre-cast 10% polyacrylamide urea TBE Novex gels (Life Technologies) at

160 V for 90 minutes. Gels were stained with SybrGold Nucleic Acid Stain (Life Technologies),

imaged, and electrotransferred with a Genie electroblotter (Idea Scientific) onto Hybond N+ nylon

membranes (GE) at 19 V for 60 minutes submerged in 1x TBE. RNA was crosslinked to the

membranes with 0.12 joules of 254 nm UV light (Stratalinker 1800, Stratagene). Nylon

membranes were blocked with 10 mL ULTRAhyb Oligo for 1 hour at 42°C and hybridized with

probe at 42°C for 18 hours. Blots were washed at 42°C for 30 min each with buffer (2X SSC,

0.1% SDS) and were developed by phosphorimaging (GE, Molecular Dynamics).

Detection of 2′-Phosphate linkages by RT-PCR

A reverse primer that anneals directly downstream of the 2′-phosphate site (Table 2.1,

HAC1-R0 2-P) was designed so that the first deoxyribonucleotide incorporated by the reverse

transcriptase is complementary to the 2′-phosphorylated ribonucleotide. Total RNA (2 µg) was

DNased (Turbo DNase Kit, Ambion), treated ± with calf intestinal phosphatase (NEB), and acid-

phenol extracted. DNased and ±CIP-ed total RNA was denatured at 65°C for 5 minutes with 10

pmol of the reverse primer and then transferred to ice. The concentration of dNTPs in reverse

55

transcription reactions was reduced from 500 µM to 1 µM to inhibit reverse transcriptase

(SuperScript III, Life Technologies, lot #1826824) at sites of 2′-phosphorylation (Fig. 4E,

schematic). PCR was performed with the same reverse primer and a forward primer that anneals

to the 5′-exon of HAC1; 28 cycles were performed on low dNTP reactions and 22 cycles were

performed on RT reactions with normal dNTP concentrations. Amplified DNA was

electrophoresed in a 2% agarose TBE 1X SybrGold (Life Technologies) gel and imaged.

Expression vector for intronless tRNAs

I designed a sequence to express ten pre-spliced tRNAs using the SUP4 promoter and

RPR1 terminator, respectively (Good and Engelke 1994). The full sequence is available at

Addgene, record 70125. Two restriction sites at the 5′- and 3′-ends of the sequence for SacI

and ClaI are underlined. Ten tRNA genes are encoded with an intervening spacer (“CTTTGT”)

derived from a dicistronic tRNA (Engelke et al. 1985). From 5′ to 3′, the tRNAs are: Phe(GAA),

Leu(CAA), Lys(UUU), Ser(GCU), Ile(UAU), Trp(CCA), Tyr(GUA), Pro(UGG), Ser(CGA) and

Leu(UAG). The pol III promoter contains sequence elements within the SUP4 Tyr(GUA) tRNA

(Good and Engelke 1994), and thus tRNATyr(GUA) occurs twice in the construct. The sequence

was synthesized as a gBlock from IDT and recombined into pDONR221 in a Gateway BP reaction

(Life Technologies). The product plasmid (pDONR221-10X-tRNA) was recombined by a Gateway

LR reaction into pAG424-ccdB (created by removing the GPD promoter from pAG424-GPD-

ccdB (Alberti et al. 2007), which has the TRP1 selectable marker and a 2µ origin for high-copy

number per cell (Chan et al. 2013). The pDONR221-10X-tRNA, pAG424-ccdB (empty vector) and

pAG424-10X-tRNA plasmids are available from Addgene (plasmids 70125, 70124, and 70123).

HAC1 epitope tagging and western blotting

Oligonucleotide probes HAC1-pML104-plus and HAC1-pML104-minus (IDT) (Table 1) form

an insert that was ligated into the Cas9 and sgRNA yeast expression plasmid pML104 (Addgene

entry 67638) (Laughery et al. 2015). Yeast strains were simultaneously transformed with pML104-

56

HAC1-ct and FLAG donor DNA with homology to the HAC1 gene for homology-directed repair.

Co-transformed yeast were selected on uracil drop-out media (U.S. Biologicals) and confirmed

to have an in-frame FLAG epitope tag by Sanger sequencing. Cell lysates were analyzed by

immunoblotting using an anti-FLAG M2 antibody (Sigma) and secondary anti-mouse HRP-

conjugated antibodies (ThermoFisher Scientific) with Enhanced Chemiluminescence detection

(Promega). Loading controls included Ponceau staining of the membrane and anti-GAPDH

antibody (UBP-Bio).

57

Table 2.1: Oligonucleotide sequences.

Oligonucleotide Name Oligonucleotide sequence (5′→3′)

tRNA Ile UAU 5-Exon N Probe TATAAGCACGAAGCTCTAACCACTGAGCTACACGAGC

tRNA Ile UAU Intron N Probe CGTTGCTTTTAAAGGCCTGTTTGAAAGGTCTTTGGCACAGAAACTT

Phe-GAA-5-Exon-Probe CTTCAGTCTGGCGCTCTCCCAACTGAGCTAAATCCGC

Leu-CAA-5-Exon-Probe CTTGAATCAGGCGCCTTAGACCGCTCGGCCAAACAACC

Pro-UGG-5-Exon-Probe CCCAAAGCGAGAATCATACCACTAGACCACACGCCC

Tyr-GUA-5-Exon-Probe TTACAGTCTTGCGCCTTAAACCAACTTGGCTACCGAGAG

HAC1-F RT-PCR ACCTGCCGTAGACAACAACAAT

HAC1-R RT-PCR AAAACCCACCAACAGCGATAAT

HAC1-F 2-P RT-PCR ATGGGAGCTGCAGATGTTTAAG

HAC1-R0 2-P GAATTCAAACCTGACTGCGCTT

TPT1-F RT-PCR TGTTCAGGTCGCTCAATAATGT

TPT1-R RT-PCR TCTTTTCGAGCGGTATGTTTCT

TRL1-F2 RT-PCR GTGGCAGAATATTGCGATGA

TRL1-R2 RT-PCR ATCCTCCAAGGTGTTCGATG

KAR2-QPCR-F AAGACAAGCCACCAAGGATG

KAR2-QPCR-R AGTGGCTTGGACTTCGAAAA

PGK1-QPCR-F TCTTAGGTGGTGCCAAAGGTT

PGK1-QPCR-R GCCTTGTCGAAGATGGAGTC

HAC1-pML104-plus GATCTTCATGAAGACAATCGCAAGGTTTTAGAGCTAG

HAC1-pML104-minus CTAGCTCTAAAACCTTGCGATTGTCTTCATGAA

58

Table 2.2: Strain numbers and genotypes.

All strains are background W303 (MATa {leu2-3,112 trp1-1 can1-100 ura3-1 ade2-1 his3-11,15}).

Strain ID Genotype Source

YJH829 tpt1∆::LEU2 (TPT1 CEN ARS URA3) B. Schwer

YJH830 tpt1∆::LEU2 (TPT1 CEN ARS URA3) (pAG424-ccdB)

YJH832 tpt1∆::LEU2 (TPT1 CEN ARS URA3) (pAG424-10x-tRNA)

YJH834 tpt1∆::LEU2 (pAG424-10x-tRNA)

YJH681 trl1∆::KanMX (TRL1 CEN URA3) B. Schwer

YJH708 trl1∆::KanMX (TRL1 CEN URA3) (pAG424)

YJH709 trl1∆::KanMX (TRL1 CEN URA3) (pAG424-10x-tRNA)

YJH835 trl1∆::KanMX (pAG424-10x-tRNA)

YJH836 sen2∆::KanMX (SEN2 tiling block CEN ARS URA3)

YJH837 sen2∆::KanMX (SEN2 tiling block CEN ARS URA3) (SEN2 LEU2 2µ)

YJH838 sen2∆::KanMX (SEN2 tiling block CEN ARS URA3) (SEN15 LEU2 2µ)

YJH839 sen2∆::KanMX (SEN2 tiling block CEN ARS URA3) (SEN34 LEU2 2µ)

YJH840 sen2∆::KanMX (SEN2 tiling block CEN ARS URA3) (SEN54 LEU2 2µ)

YJH841 sen15∆::KanMX (SEN15 tiling block CEN ARS URA3)

YJH842 sen15∆::KanMX (SEN15 tiling block CEN ARS URA3) (SEN2 LEU2 2µ)

YJH843 sen15∆::KanMX (SEN15 tiling block CEN ARS URA3) (SEN15 LEU2 2µ)

YJH844 sen15∆::KanMX (SEN15 tiling block CEN ARS URA3) (SEN34 LEU2 2µ)

YJH845 sen15∆::KanMX (SEN15 tiling block CEN ARS URA3) (SEN54 LEU2 2µ)

YJH846 sen34∆::KanMX (SEN34 tiling block CEN ARS URA3)

YJH847 sen34∆::KanMX (SEN34 tiling block CEN ARS URA3) (SEN2 LEU2 2µ)

YJH848 sen34∆::KanMX (SEN34 tiling block CEN ARS URA3) (SEN15 LEU2 2µ)

YJH849 sen34∆::KanMX (SEN34 tiling block CEN ARS URA3) (SEN34 LEU2 2µ)

YJH850 sen34∆::KanMX (SEN34 tiling block CEN ARS URA3) (SEN54 LEU2 2µ)

59

YJH851 sen54∆::KanMX (SEN54 tiling block CEN ARS URA3)

YJH852 sen54∆::KanMX (SEN54 tiling block CEN ARS URA3) (SEN2 LEU2 2µ)

YJH853 sen54∆::KanMX (SEN54 tiling block CEN ARS URA3) (SEN15 LEU2 2µ)

YJH854 sen54∆::KanMX (SEN54 tiling block CEN ARS URA3) (SEN34 LEU2 2µ)

YJH855 sen54∆::KanMX (SEN54 tiling block CEN ARS URA3) (SEN54 LEU2 2µ)

YJH856 sen2∆::KanMX (SEN2 tiling block CEN ARS URA3) (pAG424-ccdB)

YJH857 sen2∆::KanMX (SEN2 tiling block CEN ARS URA3) (pAG424-10x-tRNA)

YJH858 sen15∆::KanMX (SEN15 tiling block CEN ARS URA3) (pAG424-ccdB)

YJH859 sen15∆::KanMX (SEN15 tiling block CEN ARS URA3) (pAG424-10x-tRNA)

YJH860 sen34∆::KanMX (SEN34 tiling block CEN ARS URA3) (pAG424-ccdB)

YJH861 sen34∆::KanMX (SEN34 tiling block CEN ARS URA3) (pAG424-10x-tRNA)

YJH862 sen54∆::KanMX (SEN54 tiling block CEN ARS URA3) (pAG424-ccdB)

YJH863 sen54∆::KanMX (SEN54 tiling block CEN ARS URA3) (pAG424-10x-tRNA)

YJH864 HAC1-C-terminal-3x-FLAG

YJH865 trl1∆::KanMX HAC1-C-terminal-3x-FLAG (pAG424-10x-tRNA)

YJH866 tpt1∆::LEU2 HAC1-C-terminal-3x-FLAG (pAG424-10x-tRNA)

60

Results & Discussion

Genetic bypass of essential RNA repair genes in budding yeast

Ten S. cerevisiae tRNA isodecoders are encoded with introns (Chan and Lowe 2009),

which must be accurately processed for cells to faithfully translate messenger RNA (Hopper

2013). I adapted a strategy first documented in C. elegans (Kosmaczewski et al. 2014) to express

these ten tRNAs in “pre-spliced” form (Fig. 1B, the 10x-tRNA plasmid) and found that expression

of these intronless tRNAs rescues the growth of cells with deletions in the essential genes TRL1

and TPT1 (Fig. 2.1C, 2.2A). This result is consistent with previous findings that TPT1 is essential

only in the context of the generation of 2′-phosphorylated tRNAs by Trl1 (Schwer et al. 2004)

and that a growth defect caused by TRL1 knockdown in the trypanosome Trypanosoma brucei

is rescued by expressing intronless tRNATyr (Lopes et al. 2016).

The SEN complex cleaves introns from pre-tRNAs (Fig. 2.1A), and each component of the

heterotetrameric endonuclease (Sen2, Sen15, Sen34 and Sen54) is essential for growth (Trotta

et al. 1997). I created cells with single genomic deletions in each of the SEN2, SEN15, SEN34,

and SEN54 genes and found that in each case cells were only viable when complemented by

plasmid-mediated expression of the cognate SEN gene (Fig. 2.1D). I expressed the ten pre-

spliced tRNAs in cells containing deletions of each of the SEN genes and found that pre-spliced

tRNAs failed to rescue deletion of any component of the SEN complex (Fig. 2.1C), consistent

with this complex having an essential function beyond splicing of pre-tRNA (Dhungel and Hopper

2012). The gene encoding the only other known substrate of the SEN complex, CBP1, is non-

essential (Tsuboi et al. 2015), suggesting that the SEN complex has another essential function,

possibly to process unknown RNA substrates.

61

Figure 2.1: Genetic bypass of essential components of tRNA splicing with intronless

tRNAs.

A. Functions of tRNA splicing enzymes. Introns from pre-tRNAs are removed by the SEN complex, which contains the Sen2 and Sen34 endonucleases that cleave the 5′ and 3′ splice sites, respectively, as well as Sen15 and Sen54. Following cleavage, the tRNA exons are ligated by the multifunctional RNA ligase Trl1, producing a ligated tRNA with a 2′-PO4 at the splice

junction. Tpt1 removes the 2′-PO4 in an NAD++-dependent reaction, producing ADP-ribose-

1″,2″-cyclic phosphate and nicotinamide.

62

Figure 2.1 (continued): Genetic bypass of essential components of tRNA splicing with

intronless tRNAs.

B. Schematic of a plasmid encoding the ten S. cerevisiae tRNAs in intronless form. The tRNAs are expressed from a high-copy 2µ TRP1 plasmid containing a SUP4 promoter and terminated by the RPR1 terminator.

C. Genetic bypass of RNA repair mutants by expression of intronless tRNAs (“10x-tRNA” plasmid). Strains expressing a URA3 covering plasmid were transformed with an empty vector (TRP1 2µ) or a high-copy plasmid encoding intronless tRNAs (B). URA+ TRP+ colonies (top row) were selected and struck on FOA media (bottom row), which selects against cells with the covering plasmid, to assess intronless tRNA-mediated bypass. Plates were photographed after 5 to 7 days of incubation at 30°C. Intronless tRNAs complement deletion of TRL1 and TPT1 but do not rescue deletion of SEN components (bottom right).

D. Cells with genomic deletions of SEN genes (SEN2, SEN15, SEN34, SEN54) and a single-copy URA3 plasmid expressing the deleted gene were individually transformed with high-copy LEU2 plasmids containing the genomic locus of each of the SEN genes. LEU+ colonies were plated on –LEU media (top), and FOA media (bottom) to select against the covering plasmid. Only those plasmids that contain the cognate genomic locus of the deleted SEN gene were able to rescue growth on FOA (bottom).

63

RNA repair mutants have defects in translation

Genetically-bypassed trl1∆ and tpt1∆ cells share common growth phenotypes that may

reflect translational defects in the absence of RNA repair or incomplete rescue by intronless

tRNAs. I found that rescued trl1∆ and tpt1∆ cells grow slowly at 30°C and fail to grow at 37°C

on rich media (Fig. 2.2B). These growth defects could be due to differences in the abundance or

functionality of pre-spliced tRNAs or due to the accumulation of endogenous intermediates of

tRNA splicing. Alterations in tRNA pools and post-transcriptional modifications of specific tRNA

species have been implicated in regulating translational dynamics in both prokaryotes and

eukaryotes (Kirchner and Ignatova 2015), providing a possible explanation for the temperature

sensitivity. Rescued trl1∆ cells should accumulate unligated pre-tRNA half-molecules that have

been cut by the SEN complex, whereas tpt1∆ cells should accumulate spliced and ligated pre-

tRNAs that retain a 2′-PO4 at the ligation junction (Fig. 2.1A). Accumulation of these molecular

species could have distinct effects on translation. However, both RNA repair mutants exhibit

broad sensitivity to sub-lethal doses of translational inhibitors (Fig. 2.2C).

It is possible that the pre-spliced tRNAs used to rescue trl1∆ and tpt1∆ mutants are

suboptimal due to defects in their expression levels, amino acid charging, or post-transcriptional

modification. S. cerevisiae tRNAs have an average of 13 post-transcriptional modifications per

tRNA species (Phizicky and Hopper 2010), and the diverse set of modifications added to tRNAs

can impact maintenance of reading frame, translational fidelity, and tRNA stability (Hopper 2013).

The rapid tRNA decay pathway (RTD) comprises a quality control mechanism that degrades

aberrant tRNA molecules via the 5′→3′ exonucleases Rat1 and Xrn1. Structural stability is a

major determinant of RTD substrates: both the introduction of destabilizing mutations in tRNA

sequences and the deletion of modification enzymes that act to enhance tRNA stability cause

temperature sensitivity in budding yeast.

64

Figure 2.2: Growth phenotype of RNA repair mutants.

Laura K White performed the experiments displayed in this figure.

A. RNA repair mutants rescued by pre-spliced tRNAs expressed from the (10x-tRNA TRP1 2µ) plasmid grow on -TRP and FOA media, but not on -URA media. Plates were photographed after 3 days of growth at 30°C.

B. Growth of RNA repair mutants is temperature-sensitive. Cells were serially diluted, spotted on

YPD media and incubated at the indicated temperatures. Wild-type and complemented trl1∆

(TRL1) and tpt1∆ (TPT1) cells grow at all temperatures, whereas bypassed trl1∆ and tpt1∆ cells fail to grow at 37°C.

65

Figure 2.2 (continued): Growth phenotype of RNA repair mutants.

C. Sensitivity of RNA repair mutants to translational inhibitors. Cells were serially diluted and spotted on YPD media or YPD media supplemented with sub-lethal doses of the translational inhibitors anisomycin, hygromycin B and cycloheximide at the concentrations indicated. Wild-type cells are viable under all conditions, whereas RNA repair mutants fail to grow in the presence of inhibitors of translation.

D. Deletion of XRN1 does not suppress the temperature sensitive phenotype of trl1∆ rescued with pre-spliced tRNAs. All plates were photographed after 3 to 4 days of incubation at 30°C or 37°C.

66

However, growth at 37°C can be rescued by compensatory mutations that restore tRNA

structural stability or by deletion of the Xrn1 exonuclease (Whipple et al. 2011). Based on these

prior observations, I tested whether RTD is responsible for temperature sensitivity by deleting

XRN1 in a rescued trl1∆ mutant and found that the double deletion, trl1∆ xrn1∆, grows no better

at 37°C than the single trl1∆ mutant (Fig. 2.2D), suggesting that the temperature sensitivity of

RNA repair mutants is not a consequence of RTD.

Whereas our results suggest that pre-spliced tRNAs are not rapidly degraded due to

structural instability, these intronless tRNA species may still be hypomodified or otherwise

suboptimal. Previous studies of tRNA introns identified two types of post-transcriptional

modifications that are only added at the pre-tRNA stage, when an intron is still present. Site-

specific introduction of 5-methylcytidine (m5C) or pseudouridine in the anticodon or at position

40 is dependent on the presence of the intron, and these intron-dependent modifications occur

in four of the ten intron-containing tRNAs in S. cerevisiae (Grosjean et al. 1997) (Table 2.3). In

particular, incorporation of m5C at the wobble position in the pre-tRNALeu(CAA) is catalyzed by

the tRNA-specific methyltransferase Trm4, and disruption of TRM4 causes sensitivity to the

antibiotic paromomycin, an aminoglycoside that interferes with translational fidelity (Wu et al.

1998). This observation is consistent with our finding that RNA repair mutants rescued with pre-

spliced tRNAs fail to grow on rich media containing low doses of the translational inhibitors

anisomycin, hygromycin B, or cycloheximide (Fig. 2.2C). Thus, this phenotype could be

indicative of hypomodification of one or more of the intronless tRNAs from the construct.

Alternatively, it is possible that bypass of trl1∆ and tpt1∆ with intronless tRNAs provides a

complete rescue of tRNA splicing but causes reduced fitness due to an additional but

nonessential role of RNA repair related to translation.

67

Table 2.3: Intron-containing tRNA copy number and intron-dependent modifications in S.

cerevisiae.

Laura K White produced this table for our 2018 publication in RNA.

tRNA Species Copy Number Intron Length (nt) Intron-Dependent Base Modifications

tRNA-Ser(CGA) 1 19

tRNA-Ile(UAU) 2 60 Pseudouridine, positions 34 & 36

tRNA-Leu(UAG) 3 19

tRNA-Ser(GCU) 4 19

tRNA-Trp(CCA) 6 34

tRNA-Lys(UUU) 7 23

tRNA-Tyr(GUA) 8 14 Pseudouridine, position 35

tRNA-Leu(CAA) 10 32 5-methylcytidine, position 35

tRNA-Phe(GAA) 10 18-19 5-methylcytidine, position 40

tRNA-Pro(UGG) 10 30-33

Genomic copy number of tRNAs listed are reproduced here, reorganized, from (Chan and Lowe 2009); intron-dependent tRNA modifications are from (Grosjean et al. 1997).

68

RNA repair mutants accumulate intermediates and products of tRNA splicing

To evaluate the ability of cells lacking RNA repair to produce mature tRNAs, I analyzed the

processing of tRNAIle(UAU) by northern blot using total RNA from rescued trl1∆ and tpt1∆ cells,

unshuffled control strains, wild-type, and xrn1∆ mutants. Wild-type, trl1∆ (TRL1), and tpt1∆

(TPT1) cells have expected tRNA processing intermediates, including primary transcript (145 nt),

pre-tRNA with 5′- and 3′-processing by RNaseP, Rex1, RNase Z, and Lph1 (136 nt), mature

tRNA at 76 nt, excised intron at 60 nt, and 5′-exon at 38 nt (Fig. 3 A, B). Total RNA from trl1∆

mutants complemented with intronless tRNAs contains a band of the same size as mature tRNA

from wild-type controls, and this tRNA must arise from the intronless tRNA construct because

these cells cannot ligate exons processed from genomic intron-containing tRNAs (Fig. 2.3A, lane

4). Cells that contain the pre-spliced tRNA plasmid exhibit high molecular weight species to

which the northern probe hybridizes, which are likely tandem tRNA transcripts that are not

completely processed. This result is consistent with the construct functioning as a single

transcriptional unit, owing to its design with a single SUP4 pol III promoter and RPR1 pol III

terminator; additionally, the intronless tRNA genes may also be transcribed individually via

internal promoters (Galli et al. 1981).

The S. cerevisiae genome contains tRNA genes at varying copy number (Table 2.3) (Chan

and Lowe 2009), raising the possibility that rescue by a high-copy plasmid expressing intronless

tRNAs could alter the levels of specific tRNA isodecoders. To assess the ability of the genetic

bypass to produce quantities of tRNAs similar to those found in wild-type cells, I performed

northern blotting with probes hybridizing to the 5′-exons of tRNAs with varying genomic copy

number.

69

Figure 2.3: tRNA processing phenotypes of RNA repair mutants.

Diagrams of tRNA primary structure, annotations, and probe locations are depicted next to each northern blot. SCR1 loading control blots are below and were the basis of loading for each lane.

A. Northern blot using probe that hybridizes to the 5′-exon of tRNAIle(UAU) identifies pre-processed tRNA intermediates derived from primary transcripts of intronless tRNAs (lanes 4, 5, and 7, bracket annotation), as well as intermediates arising from the intron-containing endogenous

tRNAs (at ~150 nt). Each strain produces mature tRNAs (~76 nt). The trl1∆ mutant (lane 4) is unable to ligate tRNAIle(UAU) exons arising from chromosomal copies of the gene, and thus mature tRNAs in these cells occur from processing of the intronless tRNA transcript. A band at ~32 nt in lanes 4 and 7 is likely a product of SEN cleavage of chromosomally-encoded intron-containing tRNAs that are not re-ligated.

70

Figure 2.3 (continued): tRNA processing phenotypes of RNA repair mutants.

B. Northern blot with probe to tRNAIle(UAU) intron shows increased accumulation of the intron in

trl1∆ mutants (lanes 4 and 7), as well as in xrn1∆ and in double deletion trl1∆ xrn1∆ mutants (lanes 6 and 7, respectively). A band at ~90 nt putatively represents 5′-exon/intron (question mark). Densitometry quantifications of intron signal relative to wild-type (lane 1) and normalized to SCR1 signal are displayed below lane numbers.

C. Northern blot with probe to the 5′-exon of tRNATyr(GUA), a tRNA with a copy number of 8 in the budding yeast genome, reveals mature tRNA bands at approximately the same intensity across wild-type (lane 1) and repair mutant strains (lanes 4 & 5).

D. Northern blot with probe to the 5′-exon of tRNAPhe(GAA), a tRNA with a copy number of 10 in the genome, exhibits approximately equal density of mature tRNA band intensity in wild-type (lane 1) and deletions in RNA repair genes (lanes 4 & 5).

E. Northern blot with probe to the 5′-exon of tRNALeu(CAA), a tRNA with a copy number of 10 in

the genome, shows a decrease in mature tRNA band intensity in the trl1∆ mutant (lane 4), and

also in the tpt1∆ mutant (lane 5), as compared to wild-type and covered strains (lanes 1 through 3).

F. Northern blot with probe to the 5′-exon of tRNAPro(UGG), a tRNA with a copy number of 10 in

the genome, reveals a decrease in mature tRNA band intensity in both trl1∆ and tpt1∆ RNA repair mutants (lanes 4 & 5, respectively). All lanes show a doublet of bands, with the upper band being consistent in size with mature tRNAPro(UGG), and the lower band is annotated with an asterisk.

Lane 5 displays a strong tpt1∆-dependent band, annotated with a question mark.

71

In the case of low copy number, tRNAIle(UAU) is encoded at two copies per genome, and the

intronless tRNA plasmid produces quantities of mature tRNAIle(UAU) similar to wild-type cells

(Fig. 2.3A, lane 1 versus 4 & 5). In contrast, the isodecoders tRNATyr(GUA), tRNALeu(CAA),

tRNAPhe(GAA), and tRNAPro(UGG) are encoded in the genome at relatively high copy number (8,

10, 10, and 10 copies, respectively) (Table 2.3). The intronless tRNA plasmid produces

tRNATyr(GUA) and tRNAPhe(GAA) at abundances similar to wild-type (Fig. 2.3C & 2.3D, lane 1

versus 4 & 5), but the plasmid fails to produce equivalent amounts of tRNALeu(CAA) and

tRNAPro(UGG) (Fig. 2.3E & 2.3F, lanes 1 versus 4 & 5). These results show that there is not a

monotonic relationship between copy number in the genome and the ability of the intronless

tRNA plasmid to produce wild-type quantities of each tRNA. However, these results do not

comment on the modification status of intronless tRNAs. Furthermore, the discrepancy in

quantities of tRNALeu(CAA) and tRNAPro(UGG) produced by the intronless tRNA plasmid could

contribute to the slow growth phenotype of the RNA repair mutants (Fig. 2.2).

The northern blot for tRNAPro(UGG) (Fig. 2.3F) reveals several additional features of interest

that appear to be unique to this tRNA isodecoder. Doublet bands of spliced tRNAPro(UGG), which

I also detected in all strains tested, have been previously observed by denaturing polyacrylamide

gel (Winey et al. 1986). The upper of these two bands has a length consistent with the 75 nt

mature tRNAPro(UGG) (Chan and Lowe 2009). The lower band, approximately 10 nucleotides

shorter, cannot be explained as an intron-containing tRNA half or processing intermediate, as

the intron for tRNAPro(UGG) is 31 nt and a processing intermediate that contained the intron and

both exons would migrate at ~103 nt. Because these intermediates accumulate in the RNA ligase

mutant—and ligase is required for intron splicing—the pre-spliced tRNA construct is also

competent to produce both species (i.e., the putative processing intermediate marked with an

asterisk and mature tRNA). In addition, both tpt1∆ (TPT1) and tpt1∆ cells (Fig. 2.3F, lanes 3 and

72

5) contain a third tRNAPro(UGG) band (indicated by a question mark) that accumulates to high

levels in tpt1∆ cells and is not seen in other genotypes, which may correspond to an additional

tRNA species resolved via 2D PAGE analysis of tRNA from a tpt1 conditional mutant (Spinelli et

al. 1997). It is possible that this third band is derived from endogenous spliced tRNAPro(UGG) that

retains a 2′-phosphate at the splice junction in the absence of TPT1; however, the precise

identity of these three species of tRNAPro(UGG) remains unresolved.

In addition to analysis of mature tRNA production, I further investigated the fate of the

introns from endogenous tRNA genes in the context of the RNA repair mutants. Northern blot

analysis of the tRNAIle(UAU) intron recapitulated the finding that decay of excised tRNA introns

requires both Xrn1 and Trl1(Wu and Hopper 2014). Bypassed trl1∆ cells show an increase in

levels of tRNA intron (47-fold compared to wild-type, normalized to SCR1 levels), as do xrn1∆

cells (6.3-fold) and double mutant trl1∆ xrn1∆ cells (22-fold) (Fig. 2.3B, lanes 4, 6, and 7). These

observations are in line with kinase-mediated decay of cleaved tRNA introns, in which the 5′-

RNA kinase activity of Trl1 (Fig. 2.1A) is required to phosphorylate the 5′-hydroxyl intron

products of SEN cleavage to enable their 5′→3′ exonucleolytic decay by Xrn1, which specifically

degrades 5′-phosphorylated RNA substrates (Stevens 2001).

RNA repair mutants have defects in unfolded protein response activation

Induction of the unfolded protein response (UPR) requires the ligation of two exons after

intron excision from the HAC1 mRNA (Gonzalez et al. 1999). I tested cells lacking RNA repair for

their ability to activate the UPR. Bypassed trl1∆ mutants fail to grow on media containing the

UPR-inducing drug tunicamycin (Fig. 2.4A), confirming that Trl1 is required for UPR activation in

response to protein folding stress (Sidrauski et al. 1996). In contrast, tpt1∆ mutants grow equally

well in the presence and absence of tunicamycin (Fig. 2.4A), suggesting that the 2′-PO4

remaining on HAC1 mRNA after ligation in tpt1∆ cells does not interfere with its translation or

73

that partial HAC1 translation is sufficient for UPR activation (see below). Cells with Hac1-FLAG

(in wild-type, trl1∆, and tpt1∆ backgrounds) were also spotted onto tunicamycin containing

media and had similar growth to cells with untagged Hac1, confirming that Hac1-FLAG can

function in the UPR (Fig. 2.4A).

I corroborated the growth assay by analyzing HAC1 splicing in trl1∆ and tpt1∆ cells using RT-

PCR (Fig. 2.4B). The wild-type strain catalyzes cleavage and ligation of HAC1 upon UPR

stimulation with tunicamycin, whereas the trl1∆ mutant exhibits no detectable spliced HAC1 and

a reduction in unspliced HAC1. The decrease in unspliced HAC1 observed in lanes 4 and 6 (Fig.

2.4B) as compared to their tunicamycin-null control lanes can be explained by two facts: first,

both strains express Ire1, the endonuclease responsible for cleaving HAC1 mRNA prior to

ligation; second, the Hac1 protein is a transcription factor that activates transcription of the

HAC1 gene, creating a positive feedback loop to sustain UPR activation (Ogawa and Mori 2004).

When Ire1 cleaves HAC1 mRNA, but Trl1 is unavailable for ligation, Hac1 protein is not translated

and cells fail to activate the positive feedback loop. Then, without the newly-transcribed HAC1

mRNA, the remaining pool of HAC1 is efficiently cleaved so that the cDNAs synthesized do not

contain both PCR priming sites and thus fail to amplify in this PCR assay. These simultaneous

deficits could lead to a decrease in unspliced HAC1 in RNA repair mutants treated with

tunicamycin as compared to their null-treatment controls.

74

Figure 2.4: UPR-related phenotypes of RNA repair mutants.

A. Growth assay of RNA repair mutants on UPR-inducing media. Yeast cells (wild-type, trl1∆,

and tpt1∆) were serially diluted and spotted onto rich media (YPD) and tunicamycin-containing media (80 ng/mL tunicamycin) to induce the UPR. Plates were imaged after 3 days of growth at

30°C. Wild-type and tpt1∆ cell growth is unaffected by tunicamycin, whereas trl1∆ cells fail to grow on media containing tunicamycin. Serial dilution growth assays for cells of the same genetic background as above, but with C-terminal FLAG tags on Hac1, are shown below.

75

Figure 2.4 (continued): UPR-related phenotypes of RNA repair mutants.

B. Analysis of HAC1 splicing in RNA repair mutants. Total RNA from untreated and tunicamycin-

treated wild-type, trl1∆, and tpt1∆ cells was analyzed by RT-PCR using primers specific for HAC1, producing products at 499 bp (unspliced HAC1) and 247 bp (spliced HAC1). A no-template (NT) control is shown in lane 7. The proportion of spliced HAC1 upon tunicamycin

treatment increases in wild-type (lanes 1 and 2) and tpt1∆ cells (lanes 5 and 6), and spliced HAC1 is visible in RNA from both cells (lanes 2 and 6, 247 bp). Spliced HAC1 is undetectable in

trl1∆ cells (lane 4) upon tunicamycin treatment (lanes 3 and 4), owing to the inability of trl1∆ cells to ligate HAC1 exons. Asterisk marks an unknown PCR product dependent on tunicamycin treatment. Cells with C-terminal FLAG tags of Hac1 were also analyzed for splicing in the same manner (below) with a no-template (NT) control in lane 9. Reactions lacking reverse transcriptase (RT-) were negative for amplification (data not shown).

C. Hac1 protein levels in RNA repair mutants. Whole cell lysates were prepared from wild-type,

trl1∆, and tpt1∆ cells expressing C-terminal Hac1-FLAG and grown in the presence and absence of tunicamycin. Lysates were analyzed by SDS-PAGE and nitrocellulose transfer followed by Ponceau S staining and cross-reaction with anti-FLAG and anti-GAPDH antibodies. Scale to the left is nominal molecular mass of a protein ladder (kDa); the expected mass of Hac1-FLAG is 31 kDa. Hac1-FLAG is detected in wild-type cells upon tunicamycin addition but is undetectable in

trl1∆ and tpt1∆ cells.

D. Induction of the UPR-responsive KAR2 gene in RNA repair mutants. Amounts of KAR2 mRNA

(normalized to PGK1 mRNA abundance) were measured by RT-qPCR in wild-type, trl1∆, and

tpt1∆ cells in the presence and absence of tunicamycin. Error bars are 95% confidence intervals, n=3. Relative abundance of KAR2 mRNA increased 20-fold in wild-type cells treated with

tunicamycin, whereas the corresponding levels of KAR2 did not increase in trl1∆ and increased

1.4-fold in tpt1∆ cells.

E. Detection of ligated and 2′-phosphorylated HAC1 mRNA. Total RNA was treated with calf intestinal phosphatase (CIP) to remove 2′-phosphates (diagram) and reverse-transcribed using HAC1-specific primer under high (500 µM) concentrations of dNTPs. The cDNA products were PCR amplified, yielding products for unspliced (456 bp) and spliced (204 bp) HAC1 mRNA. Using high dNTP concentrations, I find that splicing of HAC1 in wild-type cells increases upon tunicamycin treatment (compare lanes 1 and 2 versus 3 and 4), similar to B, but is unaffected by CIP treatment (compare lanes 1 versus 2, and 3 versus 4). Likewise, I find that spliced HAC1

mRNA in tpt1∆ cells increases in response to tunicamycin (compare lanes 9 and 10 versus lanes 11 and 12), albeit to a lesser extent that wild-type cells and is unaffected by CIP treatment (compare lanes 9 versus 10 and lanes 11 versus 12). An asterisk marks an unknown PCR product dependent on tunicamycin treatment.

F. Detection of ligated and 2′-phosphorylated HAC1 mRNA. Using low (1 µM) dNTP concentrations, I find that spliced HAC1 is preferentially amplified in wild-type cells over unspliced HAC1 (lanes 1 through 4). The abundance of spliced HAC1 mRNA from wild-type cells increases in response to tunicamycin but is unaffected by CIP treatment (compare lanes 1 versus

2, and 3 versus 4). RT-PCR analysis of HAC1 mRNA from tpt1∆ cells reverse transcribed under low dNTP concentrations shows both unspliced and spliced forms of HAC1 mRNA, and spliced HAC1 mRNA increases in response to tunicamycin treatment (compare lanes 10 versus 12).

However, in contrast to wild-type, amplification of spliced HAC1 mRNA from tpt1∆ is strongly

76

Figure 2.4 (continued): UPR-related phenotypes of RNA repair mutants.

dependent on prior treatment with CIP. In the absence of tunicamycin and CIP treatment, spliced HAC1 mRNA is undetectable, whereas treatment with CIP enables reverse transcription (compare lane 9 to lane 11; see panel with enhanced contrast to the right). Similarly, the

abundance of spliced HAC1 mRNA from tpt1∆ cells increases in response to tunicamycin, and its abundance is further increased upon CIP treatment (compare lane 11 versus 12).

77

I found that tpt1∆ cells accumulate spliced HAC1, but to a lesser extent than wild-type,

suggesting that the 2′-phosphorylated products of ligation interfere with processing or stability

of the HAC1 mRNA (Fig. 2.4B, compare lanes 2 versus 6). Cells with Hac1-FLAG spliced HAC1

mRNA with the same pattern as their untagged controls (Fig. 2.4B). An immunoblot for Hac1-

FLAG (Fig. 2.4C) showed that wild-type cells showed no detectable Hac1 until the UPR was

induced with tunicamycin, but Hac1-FLAG could not be detected in lysate from trl1∆ or tpt1∆

cells.

In addition, trl1∆ cells failed to increase expression of the chaperone KAR2 (Fig. 2.4D), a

representative UPR-responsive gene strongly induced by Hac1 transcription factor protein (Mori

et al. 1992; Nikawa et al. 1996) indicating that trl1∆ cells fail to induce UPR-responsive genes.

Consistent with reduced spliced HAC1 mRNA accumulation (Fig. 2.4B), tpt1∆ mutants also

show reduced expression of KAR2 (Fig. 2.4D). However, this reduced degree of UPR induction,

as observed by HAC1 mRNA splicing and KAR2 expression, is nonetheless sufficient for growth

of tpt1∆ cells in the presence of tunicamycin (Fig. 2.4A). Despite accumulation of spliced HAC1

mRNA (Fig. 2.4B), accumulation of KAR2 mRNA (Fig. 2.4D) and growth on tunicamycin (Fig.

2.4A), I was unable to detect Hac1-FLAG by immunoblot from tpt1∆ cells. (Fig. 2.4C). It is

possible that the 2′-PO4 allows only partial translation of the spliced HAC1 mRNA 5′-exon,

leading to production of Hac1 N-terminal bZIP domain and low level UPR activation, but

precluding translation of the C-terminal FLAG epitope. Consistent with this possibility, translation

of unspliced HAC1 mRNA is sufficient to restore growth on tunicamycin-containing media in the

absence of the Duh1 ubiquitin ligase (Di Santo et al. 2016).

I found that tpt1∆ mutants accumulate less spliced HAC1, less Hac1 protein, and less

KAR2 mRNA than wild-type yeast despite having a functional RNA ligase. To determine whether

ligated HAC1 mRNA retains a 2′-PO4 at the ligation junction in tpt1∆ cells, I adapted a reverse-

78

transcriptase-based method from previous reports that show 2′-phosphates inhibit reverse

transcription (Dhungel and Hopper 2012; Schutz et al. 2010). I used calf intestine phosphatase

(CIP) treatment, which removes 2′-phosphates from RNA (Fig. 2.4E schematic) (McCraith and

Phizicky 1990), and RT-PCR to test whether 2′-phosphates were present in ligated HAC1 mRNA.

Using low concentrations of dNTPs (1 µM) (Fig. 2.4F) and a reverse transcription primer adjacent

to the expected site of 2′-phosphorylation, CIP treatment substantially enhanced detection of

spliced HAC1 mRNA in tpt1∆ cells, but levels of amplified HAC1 in RNA from wild-type cells

were unaffected, indicating that a 2′-PO4 remains at the ligation junction in the HAC1 mRNA

from tpt1∆ cells. Reverse transcription reactions with typical dNTP concentrations (500 µM) (Fig.

2.4E) restored amplification of unspliced HAC1 mRNA. I surmise that reduced dNTP

concentrations lower the processivity of reverse transcriptase, favoring synthesis of shorter

cDNA substrates. The presence of spliced HAC1 mRNA in tpt1∆ cells in the absence of CIP

treatment (Fig. 2.4F, lane 11) could indicate that the assay is not quantitatively sensitive to sites

of 2′-phosphorylation or that not all molecules of ligated HAC1 mRNA retain 2′-phosphates

despite the absence of Tpt1. In any case, 2′-phosphorylated RNA reduces cDNA synthesis under

these conditions, enabling their detection via this RT-PCR strategy.

Summary

I showed the one essential function of RNA repair in budding yeast is catalyzing the ligation

of tRNA halves resulting from splicing; however, the reduced-growth phenotypes of RNA repair

mutants caused by various translational inhibitors, a UPR-inducing drug, and elevated

temperature suggest either that the pre-spliced tRNAs do not function at the same

efficiency/abundance as endogenously-encoded tRNAs—or that RNA repair is generally helpful

to cells, albeit not essential. I demonstrated that “pre-spliced” tRNA genes are transcribed

79

and/or processed to a size consistent with wild-type mature tRNA. Furthermore, I provided

evidence of dysfunction of the UPR in both the ligase mutant and the 2′-phosphotransferase

mutant, suggesting that 2′-phosphorylated HAC1 mRNA contributes to UPR induction

dynamics. Lastly, I showed that RNA retaining 2′-PO4 residues is detectable when the enzyme

responsible for removing them is deleted.

These new genetic reagents enable studies to identify other targets of RNA repair. For

example, cells lacking the Tpt1 2′-phosphotransferase accumulate 2′-phosphates at sites of

RNA ligation, enabling their possible identification by methods to identify 2′-O-modifications by

RNA-seq (Birkedal et al. 2015) or by affinity purification by tagging with mutant Tpt1 enzyme and

biotin-NAD (Steiger et al. 2005). These approaches address the diversity of RNA repair in biology

and are the subject of further study.

80

CHAPTER III

III MULTIPLE DECAY EVENTS TARGET HAC1 mRNA DURING SPLICING TO

REGULATE THE UNFOLDED PROTEIN RESPONSE2

Abstract

In the unfolded protein response (UPR), stress in the endoplasmic reticulum (ER) activates

a large transcriptional program to increase ER folding capacity. During the budding yeast UPR,

Ire1 excises an intron from the HAC1 mRNA and the exon products of cleavage are ligated, and

the translated protein induces dozens of stress-response genes. Using cells with mutations in

RNA repair and decay enzymes, we show that phosphorylation of two different HAC1 splicing

intermediates is required for their degradation by the 5′→3′ exonuclease Xrn1 to enact opposing

effects on the UPR. We also found that ligated but 2′-phosphorylated HAC1 mRNA is cleaved,

yielding a decay intermediate with both 5′- and 2′-phosphates at its 5′-end that inhibit 5′→3′

decay and suggesting that Ire1 degrades incompletely processed HAC1. These decay events

expand the scope of RNA-based regulation in the budding yeast UPR and have implications for

the control of the metazoan UPR.

Introduction

During the unfolded protein response (UPR), protein folding stress in the lumen of the

endoplasmic reticulum leads to oligomerization of the transmembrane kinase/endoribonuclease

Ire1 and the processing of a cytoplasmic mRNA to yield splicing intermediates with 2′,3′-cyclic

phosphate (PO4) and 5′-hydroxyl (OH) termini (Gonzalez et al. 1999). In budding yeast, excision

2 Published with permission under a Creative Commons License (Attribution 4.0 International (CC BY 4.0)) from eLife: PD Cherry, SE Peach, and JR Hesselberth. Multiple decay events target HAC1 mRNA during splicing to regulate the unfolded protein response. 2019. doi: 10.7554/eLife.42262

81

of an intron from the HAC1u mRNA (“u” denoting the unspliced mRNA) by Ire1 is followed by

exon ligation by the multifunctional Trl1 RNA ligase (Sidrauski et al. 1996) involving 5′-

phosphorylation of the 5′-OH product, adenylylation of the 5′-PO4, and resolution of the 2′,3′-

cyclic PO4 to a 2′-PO4/ 3′-OH. The newly produced 3′-OH serves as the nucleophile to attack

the 5′-adenylate intermediate, yielding a ligated mRNA with an internal 2′-PO4. The 2′-PO4 is

assumed to be removed in a separate reaction by the 2′-phosphotransferase, Tpt1, in a NAD+ -

dependent reaction (Culver et al. 1997). The spliced mRNA, called HAC1s mRNA (“s” denoting

spliced mRNA) (Li et al. 2018), is translated into a transcription factor that activates the

expression of dozens of stress-response genes to mitigate protein-folding stress (Ron and

Walter 2007). In addition, Hac1 activates its own promoter in a positive feedback loop that

generates more HAC1u and permits sustained UPR activation (Ogawa and Mori 2004) (Fig. 3.1A).

Control of this positive feedback loop ensures UPR suppression during normal growth and

rapid activation upon stress exposure. To facilitate the control of UPR activation, HAC1u contains

cis- regulatory elements that suppress unintended translation and promote rapid processing. If

a ribosome initiates on HAC1u , translation through the 5′-exon/intron junction yields a truncated

protein with an intron-encoded C-terminal peptide “degron” that targets it for ubiquitylation and

degradation (Di Santo et al. 2016). A stem-loop in its 3′-untranslated region (the “3′-BE”) tethers

HAC1u to the ER membrane, ensuring rapid Ire1-mediated cleavage following ER stress (Aragón

et al. 2009). Finally, a long-range base-pairing interaction between the 5′-UTR and intron

prevents ribosome initiation to suppress translation of HAC1u mRNA (Chapman and Walter 1997;

Di Santo et al. 2016).

Previous work found unexpected roles for RNA decay and repair enzymes acting on HAC1

mRNA in the budding yeast unfolded protein response. Ire1 is a metal-ion-independent

endonuclease that produces RNA cleavage products with 5′-OH termini (Gonzalez et al. 1999).

82

In cells lacking the cytoplasmic 5′→3′ exonuclease Xrn1, HAC1 splicing intermediates

accumulate with 5′-PO4 termini, indicating that an RNA 5′-kinase phosphorylates HAC1

processing intermediates and that not all HAC1 splicing intermediates are productively ligated

(Harigaya and Parker 2012; Peach et al. 2015). In addition to its role in HAC1 exon ligation, Trl1

is required to relieve translational attenuation of HAC1s by an unknown mechanism (Mori et al.

2010). In cells expressing the T4 bacteriophage RNA repair enzymes PNK and RNL in lieu of

TRL1, ligated HAC1 molecules contained single nucleotide deletions from the 3′-terminus of the

5′-exon, indicating that a 3′→5′ exonucleolytic activity acts on the cleaved 5′-exon (Schwer et

al. 2004) and nuclear 3′→5′ decay of HAC1u liberates the 3′-BE, tuning the activation potential

of the UPR (Sarkar et al. 2018).

Recent studies showed that RNA decay also plays a role in the UPR in other organisms.

During UPR activation in the fission yeast, Ire1 incises specific mRNAs to promote their

stabilization or degradation (Kimmig et al. 2012; Guydosh et al. 2017). This mode of Ire1 cleavage

is similar to the metazoan Regulated Ire1-Dependent Decay (RIDD) pathway wherein Ire1 incises

some ER-localized mRNAs and the cleavage products are degraded by Xrn1 and the

cytoplasmic exosome (Hollien and Weissman 2006).

Here, we used budding yeast with mutations in RNA repair and decay enzymes to show

that HAC1 splicing intermediates are processed at multiple steps prior to ligation, limiting the

impact of spurious Ire1 activation and unintentional HAC1 cleavage. Our studies also show that

incompletely spliced HAC1s mRNA is targeted for degradation, which may be used to attenuate

the UPR.

83

Materials & Methods

Cell culture and RNA preparation

Single colonies were inoculated in drop-out media supplemented with relevant amino

acids and incubated at 30°C overnight with rotation. Cultures were diluted to an OD600 of 0.2 in

yeast-extract, peptone, dextrose (YPD) media, and UPR induction was carried out when yeast

were growing at mid-log phase with a 2-hour treatment (unless otherwise indicated) with

tunicamycin (final concentration of 2.5 μg/mL, Sigma-Aldrich) or DMSO mock treatment. Cells

were harvested by centrifugation, and total RNA was isolated by hot acid phenol extraction. For

RT-PCR and RT-qPCR experiments, total RNA was treated with TURBO DNase (2 U, Ambion)

to degrade contaminating genomic DNA.

RT-PCR/qPCR

DNase-treated RNA was reverse transcribed with 200 U of SuperScript III reverse

transcriptase (Invitrogen) using a gene-specific reverse primer (Table 3.2). Products analyzed on

a 1.5% agarose TBE gel, stained with 1x GelRed (Sigma) and imaged with a Bio Rad GelDoc.

Densitometry was performed with Bio-Rad Image Analysis software and splicing quantifications

were computed and visualized in R using ggplot2 and cowplot R Packages. Quantitative PCR

(qPCR) for KAR2 was also performed on cDNA as generated above and assayed for KAR2 and

PGK1 using Sso Advanced Universal SYBR Green Supermix (Bio Rad) and cycled on a Bio Rad

C1000 384-well thermal cycler and plate reader. Output Ct values were analyzed in Microsoft

Excel and plotted in R using ggplot2 and cowplot R Packages.

Primer Extension

Primers specific for HAC1 mRNA and U6 snRNA were PAGE-purified and ethanol

precipitated. Oligonucleotide primers were 5′-end-labeled with PNK (Enzymatics) and γ-3 2P-

ATP (Perkin Elmer) and purified with Sephadex G-25 spin columns (GE Healthcare resin, Thermo

84

empty columns). Radiolabeled primers and total RNA (15 μg) were heated to 65°C for 5 minutes

and cooled to 42°C. SuperScript III reverse transcriptase (200 U, Invitrogen) was added and

reverse transcription reactions were run with a final concentration of 500 μM dNTPs. Primers

were extended for 30 minutes at 42°C, 15 minutes at 45°C, and 15 minutes at 50°C. SuperScript

III RT was inactivated by heating for 20 minutes at 75°C. RNA was destroyed with in 10 mM

NaOH at 90°C for 3 minutes and neutralized with HCl. Formamide loading dye was added and

products were run on an 8% acrylamide TBE 7M Urea gel. Gels were dried (Bio Rad) and

exposed on a phosphor-imager screen and imaged on a Typhoon 9400 (GE Healthcare).

Northern blotting

Total RNA (3 μg) was electrophoresed on 6% acrylamide TBE 7M urea gels and transferred

to nylon membrane (Hybond N+, GE) by electroblotting. Membranes were UV-crosslinked (254

nm, 120 mJ dose), blocked in ULTRAhyb-Oligo Buffer (Ambion), and incubated with 5′-32P-

labeled oligonucleotide probes (Table 3.2) in ULTRAhyb-Oligo at 42°C for 18 hours. Membranes

were washed with 2X SSC/0.5% SDS washing buffer two time for 30 minutes each, exposed on

a phosphor-imager storage screen, and imaged on a Typhoon 9400 (GE Healthcare).

Membranes were stripped of original probe with 3 washes in stripping buffer (2% SDS) at 80°C

for 30 minutes per wash. Membranes were re-blocked and probed a second time for the loading

control, SCR1 (Table 3.2).

Yeast strains and plasmids

Yeast strains and sources used in this study are listed in Table 3.3.

85

Table 3.1: HAC1 processing intermediates

Name Size (nt) Visual Summary Description

HAC1u 1450* Full-length, genomic HAC1 transcript

HAC1s 1198* Spliced HAC1; intron removed

HAC1 5′-exon 728 Everything 5′ of the intron

Cleaved 5′-exon ~678 Fragment of 5′-Exon missing ~50 nt off its 3′-end

HAC1 intron 252 Liberated intron (alone)

circularized intron ~500 Circularized intron, visible in wild-type & RtcB cells

HAC1 3′-exon 474* Everything 3′ of the intron

Cleaved 3′-exon ~524* 3′-Exon with ~50 nt of 5′-Exon on its 5′-end

5′-exon + intron 980 5′-exon + Intron

Intron + 3′-exon 726* Intron + 3′-exon

*Size does not include poly(A) tail.

Sizes of HAC1 processing intermediates are predicted from strand-specific RNA sequencing data (Levin et al. 2010) mapped to the sacCer1 genome.

86

Table 3.2: Oligonucleotide sequences

Oligonucleotide Name Oligonucleotide sequence (5′→3′)

HAC1-F RT-PCR ACCTGCCGTAGACAACAACAAT

HAC1-R RT-PCR AAAACCCACCAACAGCGATAAT

KAR2 qPCR F AAGACAAGCCACCAAGGATG

KAR2 qPCR R AGTGGCTTGGACTTCGAAAA

PGK1 qPCR F TCTTAGGTGGTGCCAAAGGTT

PGK1 qPCR R GCCTTGTCGAAGATGGAGTC

HAC1 5′-exon probe 1 AAGTCTCTTGGTCCGACGCGGAATCGCGCA

HAC1 5′-exon probe 2 CTGGATTACGCCAATTGTCAAGATCAATTG

HAC1 intron probe 1 AACCGGCTCCTCCCCCATCAGAGAACCACGA

HAC1 intron probe 2 GGACAGTACAAGCAAGCCGTCCATTTCTTAGT

HAC1 3′-exon probe (primer

extension and northern)

ACCGGAGACAGAACAGTAGAAACCACTAAGCG

KAR2 probe ACCGTAGGCAATGGCGGCTGCGGTTGGTTC

SCR1 probe (oRP100) GTCTAGCCGCGAGGAAGG

87

Table 3.3: Strain numbers and genotypes

All strains are background W303 (MATa {leu2-3,112 trp1-1 can1-100 ura3-1 ade2-1 his3-11,15}).

Strain ID Genotype Source

YJH682 “wild-type” (CRY1) Mingxia Huang

YJH632 xrn1∆::HygMX

YJH745 ski2∆::NatMX

YJH898 dxo1∆::KanMX

YJH867 xrn1∆::HygMX dxo1∆::KanMX

YJH829 tpt1∆::LEU2 (TPT1 CEN ARS URA3) Schwer et al., 2004 PNAS

YJH830 tpt1∆::LEU2 (TPT1 CEN ARS URA3) (pAG424-ccdB)

YJH832 tpt1∆::LEU2 (TPT1 CEN ARS URA3) (pAG424-10x-tRNA)

YJH834 tpt1∆::LEU2 (pAG424-10x-tRNA)

YJH980 tpt1∆::LEU2 (pAG424-10x-tRNA) (pAG413-NPr-TPT1)

YJH891 tpt1∆::LEU2 (pAG424-10x-tRNA) (pAG413-NPr-tpt1-R138A)

YJH902 tpt1∆::LEU2 xrn1∆::HygMX (pAG424-10x-tRNA)

YJH901 tpt1∆::LEU2 ski2∆::NatMX (pAG424-10x-tRNA)

YJH681 trl1∆::KanMX (pRS416-TRL1) Schwer et al., 2004 PNAS

YJH708 trl1∆::KanMX (pRS416-TRL1) (pAG424)

YJH709 trl1∆::KanMX (pRS416-TRL1) (pAG424-10x-tRNA)

YJH835 trl1∆::KanMX (pAG424-10x-tRNA)

YJH887 trl1∆::KanMX (pAG424-10x-tRNA) (pRS413-TRL1)

YJH811 trl1∆::KanMX (pAG424-10x-tRNA) (pRS413-trl1-D425N)

YJH812 trl1∆::KanMX xrn1∆::HygMX (pAG424-10x-tRNA) (pRS413-trl1-D425N)

88

YJH912 trl1∆::KanMX (pAG424-10x-tRNA) (pRS413-trl1-K114A)

YJH913 trl1∆::KanMX (pAG424-10x-tRNA) (pRS413-trl1-K114A-D425N)

YJH808 trl1∆::KanMX (pRS423-TPI-RtcB)

YJH809 trl1∆::KanMX xrn1∆::HygMX (pRS423-TPI-RtcB)

YJH899 trl1∆::KanMX xrn1∆::HygMX (pAG424-10x-tRNA)

YJH900 trl1∆::KanMX ski2∆::NatMX (pAG424-10x-tRNA)

YJH903 trl1∆::KanMX hac1∆::HygMX (pAG424-10x-tRNA) (pAG413-GPD-HAC1u)

YJH904 trl1∆::KanMX hac1∆::HygMX (pAG424-10x-tRNA) (pAG413-GPD-HAC1s)

YJH920 hac1∆::NatMX (pAG413-GPD-HAC1u)

YJH921 hac1∆::NatMX (pAG413-GPD-HAC1s)

YJH923 tpt1∆::LEU2 hac1∆::NatMX (pAG424-10x-tRNA) (pAG413-GPD-HAC1u)

YJH924 tpt1∆::LEU2 hac1∆::NatMX (pAG424-10x-tRNA) (pAG413-GPD-HAC1s)

89

Results

RNA repair mutants have unique HAC1 mRNA processing defects.

I recently showed that the functions of the essential RNA repair enzymes Trl1 and Tpt1 in

budding yeast can be genetically bypassed by the expression of intronless tRNAs, which are

able to support translation in trl1∆ and tpt1∆ cells (Cherry et al. 2018). Because Trl1 is required

for HAC1s ligation and subsequent UPR activation (Sidrauski et al. 1996), trl1∆ cells are unable

to grow on media containing tunicamycin (Fig. 3.1B). In contrast, the general growth defect of

tpt1∆ cells is unaffected by tunicamycin (Cherry et al. 2018) (Fig. 3.1B), indicating that these cells

can activate the UPR. Combination of trl1∆ and tpt1∆ with mutations in 5′→3′ and 3′→5′ decay

factors xrn1∆ or ski2∆ led to more pronounced growth defects than the single deletions, but

removal of these decay factors did not affect the growth deficit of trl1∆ or tpt1∆ cells on

tunicamycin (Fig. 3.1B).

Given the multiple enzymatic roles of Trl1 and Tpt1 during RNA repair, I sought to

understand how the loss of these enzymes affected HAC1 mRNA splicing. I visualized HAC1

splicing intermediates by northern blotting with probes for the HAC1 3′-exon and intron (Fig.

3.1C, D) and found robust cleavage and ligation of HAC1u in wild-type cells in the presence of

tunicamycin, leading to high levels of HAC1s . As expected, trl1∆ cells lacking RNA ligase activity

did not produce HAC1s upon tunicamycin treatment. However, cleaved 3′-exon and intron

accumulated upon tunicamycin treatment in trl1∆ cells (Fig. 3.1C, D), indicating a defect in 3′-

exon decay. Cleaved HAC1 3′-exon often appears as a smear of products between ~450 nt and

~575 nt (Fig. 1C); I attribute this size heterogeneity to differences in poly(A) tail presence or

length, as the 5′-ends of these products occur uniformly at one site (Fig. 5A).

90

Figure 3.1: HAC1 mRNA processing defects in RNA repair and decay mutants.

A. Schematic of the budding yeast unfolded protein response. ER stress activates Ire1 (green),

which excises an intron (thin line) from HAC1u mRNA. The 5′ (black) and 3′ (grey) exons are

ligated by Trl1 yielding spliced HAC1 (HAC1s ) with a 2′-phosphate at the new ligation junction,

which is subsequently removed by the 2′-phosphotransferase Tpt1. HAC1s mRNA is translated into a transcription factor (Hac1, purple) that upregulates the HAC1 gene itself (a positive feedback loop), as well as several chaperones and heat shock proteins that resolve the stress.

B. Yeast cells with mutations in RNA repair and decay factors were serially diluted (5-fold) and spotted onto agar media (YPD and YPD containing tunicamycin (Tm; 0.16 μg/mL)), grown at 30°C for 2 days, and photographed. The “10x tRNA” plasmid encodes for 10 intronless tRNAs that bypass the lethality of trl1∆ and tpt1∆ (Cherry et al., 2018). The top panels depict cells with deletions of the RNA ligase TRL1, and the bottom panels depict growth of cells deletions of the 2′-phosphotransferase TPT1.

91

Figure 3.1 (continued): HAC1 mRNA processing defects in RNA repair and decay mutants.

C. HAC1 processing in RNA repair and decay mutants (3′-exon probe). HAC1u cleavage and ligation were analyzed in mutants of TRL1 RNA ligase and TPT1 2′-phosphotransferase by denaturing acrylamide gel northern blotting using a probe to the HAC1 3′-exon. Diagrams of

HAC1u, HAC1s, and HAC1 splicing intermediates are drawn next to predominant bands (see

Table 1 for descriptions and sizes of all annotations). HAC1u is cleaved and ligated to produce

HAC1s in wild-type cells (lanes 1 & 2). Intron/3′-exon and 3′-exon splicing intermediates

accumulate in trl1∆ cells, but HAC1s is not produced (lanes 3 & 4). In tpt1∆ cells, a small amount

of cleaved 3′-exon is present in the absence of tunicamycin (lane 5), whereas HAC1s and 3′-exon accumulate upon tunicamycin induction (lane 6). Cells lacking xrn1∆ grown in the absence

of tunicamycin produce HAC1s and Intron/3′-exon and 3′-exon splicing intermediates (lane 7),

and tunicamycin addition causes an increase in production of HAC1s (lane 8). The blot was stripped and reprobed using a probe for SCR1 as a loading control.

D. HAC1 processing in RNA repair and decay mutants (intron probe). Linear intron (252 nt) is

excised from HAC1u upon tunicamycin treatment (lanes 1 & 2) and linear intron is excised and accumulates in trl1∆ cells in the presence and absence of treatment (lanes 3 & 4). Excised intron is present a low level in tpt1∆ cells (lanes 5 & 6), whereas xrn1∆ cells accumulate high levels of full-length intron and shorter, intron-derived decay intermediates (lanes 7 & 8). A star denotes excised and circularized intron, which migrates at ~500 nt.

92

The 2′-phosphotransferase Tpt1 is essential in budding yeast to remove 2′-phosphate

groups from ligated tRNAs (Culver et al. 1997), but its role in HAC1 mRNA processing during the

UPR has not been defined. Although the growth of tpt1∆ cells is unaffected by tunicamycin (Fig.

3.1B), specific perturbations of HAC1 processing in tpt1∆ cells indicate that residual 2′-

phosphate groups on HAC1 mRNA cause defects in cleavage and ligation. Whereas tunicamycin

treatment led to cleavage of HAC1u and production of HAC1s in tpt1∆ cells, the levels of HAC1s

(Fig. 3.1C) and excised intron (Fig. 3.1D) are significantly lower than in wild-type cells. In

addition, despite the fact that tpt1∆ cells have functional RNA ligase, cleaved 3′-exon

accumulated to high levels upon tunicamycin treatment (Fig. 3.1C).

Kinase-mediated decay of cleaved HAC1 3′-exon competes with its ligation.

To further investigate the 3′-exon decay defect, I examined splicing of HAC1 in xrn1∆ cells.

I found that HAC1s accumulated in the absence of tunicamycin (Figs. 3.1C, 3.2A, C, D, and E).

This promiscuous processing was surprising given that HAC1s is undetectable in wild-type cells

under normal growth conditions, and it suggested that Xrn1 somehow limits production of

HAC1s. In xrn1∆ cells, 3′-exon accumulated to modest levels in both the absence and presence

of tunicamycin (Fig. 3.2A), whereas in trl1∆ cells, HAC1 3′-exon accumulated to higher levels

(Fig. 3.2A, 3.2B). Moreover, the abundance of 3′-exon was similar in trl1∆ and trl1∆ xrn1∆ cells

(Fig. 3.2A), indicating that Xrn1 requires Trl1 for 3′-exon degradation. Previous work showed that

Trl1 5′-kinase activity is required for the Xrn1-mediated degradation of excised tRNA introns in

budding yeast (Wu and Hopper 2014), and I considered whether this pathway also degraded

HAC1 3′-exon. Indeed, expression of a kinase-inactive version of Trl1 (Trl1-D425N) (Wang et al.

2006) did not restore Xrn1-mediated decay of the 3′-exon (Fig. 3.2B), affirming that Trl1 ligase

5′-kinase activity is required for Xrn1-mediated suppression of HAC1 splicing.

93

Figure 3.2: Kinase-mediated decay of HAC1 3′-exon competes with its ligation.

A. Decay of cleaved HAC1 3′-exon requires Trl1. A northern blot for HAC1 3′-exon reveals that xrn1∆ (lanes 3 & 4) and trl1∆ (lanes 5 & 6) are mutations sufficient to cause 3′-exon to accumulate, compared to wild-type. In xrn1∆ cells, the accumulations appears tunicamycin-treatment independent, whereas in trl1∆ cells, the accumulation increases upon treatment. The accumulation is also present in xrn1∆ trl1∆ cells (lanes 7 & 8).

B. Decay of cleaved HAC1 3′-exon requires the catalytic activity of Trl1 5′-kinase. We expressed a kinase-inactive missense mutant of Trl1 (Wang et al., 2006), trl1- D425N, to assess the contribution of RNA 5′-kinase activity to decay of HAC1 intermediates. Expression of trl1-D425N (lanes 5 & 6) caused similar accumulation of 3′-exon as in the trl1∆ deletion, and xrn1∆ trl1- D425N cells have levels of 3′-exon similar to trl1- D425N alone (compare lanes 6 & 8), indicating that Trl1 5′-kinase activity is required for Xrn1-mediated decay.

C. 5′-kinase and ligase domains contribute to the abundance of liberated 3′-exon. We expressed a ligase-inactive missense mutant of Trl1 (Sawaya et al. 2003), trl1-K114A, to assess the contribution of ligation activity to levels of HAC1 intermediates. Expression of trl1-K114A (lanes 5 & 6) lead to a moderate accumulation of 3′-exon. The double missense mutant, trl1-K114A-D425N (lanes 7 & 8), has 3′-exon accumulation much like that of trl1-D425N, albeit stronger.

94

Figure 3.2 (continued): Kinase-mediated decay of HAC1 3′-exon competes with its ligation.

D. Dxo1 activity does not affect HAC1 3′-exon abundance or promiscuous ligation. Northern blot analysis of 3′-exon shows that dxo1∆ cells phenocopy wild-type cells (2 & 6), whereas and xrn1∆ and dxo1∆ xrn1∆ cells accumulate similar levels of 3′-exon, indicating that Dxo1 does not contribute substantially to 3′-exon abundance.

E. Cells lacking Trl1 and expressing E. coli RtcB promiscuously splice HAC1s. Northern blot

analysis for HAC1 5′-exon shows that trl1∆ (RtcB) cells promiscuously splice HAC1s , similar to xrn1∆ cells (compare lanes 5 and 7 to lane 3), and the defects in HAC1 splicing in trl1∆ (RtcB) cells are unaffected by xrn1∆ (compare lanes 7 & 8 to 5 & 6).

F. RT-PCR assay to measure HAC1 splicing shows promiscuous splicing of HAC1s in RtcB xrn1∆ cells. Similar to E, here an endpoint RT-PCR assay using primers that flank the intron or splice

junction assesses and semi-quantifies HAC1 splicing. Tunicamycin induces HAC1s production

in wild-type cells, (lanes 1 & 2) but HAC1s is detected in in trl1∆ (RtcB) and xrn1∆ cells under normal growth conditions (without tunicamycin, lanes 3 & 5). An asterisk marks an unidentified PCR product.

G. Model for kinase-mediated decay of cleaved HAC1 3′-exon. Ire1 cleavage produces a 3′-exon with a 5′-OH that is phosphorylated by Trl1 5′-kinase. The 5′-phosphorylated product is then adenylated and ligated to the HAC1 5′-exon by Trl1 or degraded by the 5′-phosphate-dependent 5′→3′ exonuclease Xrn1. In xrn1∆ cells (top), the lack of robust 5′→3′ decay favors ligation, leading to promiscuous splicing under normal growth conditions. In trl1∆ cells expressing RtcB (bottom), RtcB directly ligates the 5′-OH products of Ire1 cleavage, and the lack of Trl1 5′-kinase activity renders Xrn1 decay irrelevant, causing promiscuous production of

HAC1s.

95

I also tested whether the ligase activity of Trl1 affected HAC1 3′-exon abundance using an

adenylyl-transferase/ligase defective allele (Trl1-K114A) (Sawaya et al. 2003) and found that

additional 3′-exon accumulates compared to wild-type (Fig. 2C), indicating that ligation also

contributes to processing of free 3′-exon.

I also examined the accumulation of 3′-exon in cells lacking Dxo1, a distributive, 5′-

phosphate-dependent 5′→3′ exonuclease (Chang et al. 2012) and found that HAC1 3′-exon

accumulation was unaffected in dxo1∆ cells. In addition, the levels of 3′-exon were similar in

xrn1∆ and dxo1∆ xrn1∆ cells (Fig. 3.2C), indicating that Xrn1 is the primary factor responsible for

5′→3′ decay of the 3′-exon.

Together these data indicate that ligation and Xrn1-mediated 5′→3′ decay compete for

the 5′-phosphorylated 3′-exon splicing intermediate (Fig. 3.2F, top). Examination of HAC1

splicing in trl1∆ cells expressing the E. coli RtcB RNA ligase (Tanaka et al. 2011b) provided

additional evidence of a competition between ligation and decay. RtcB catalyzes ligation of 2′,3′-

cyclic PO4 and 5′-OH RNA termini via a unique mechanism involving nucleophilic attack of the

5′-OH on a 3′-guanylate intermediate; accordingly, RtcB does not have 5′-kinase activity

(Chakravarty et al. 2012). I found that upon tunicamycin treatment, HAC1s was produced in trl1∆

(RtcB) cells (Fig. 3.2D), as shown previously (Tanaka et al. 2011b). However, under normal

growth conditions, trl1∆ (RtcB) cells also promiscuously spliced HAC1s at levels similar to xrn1∆

cells (Figs. 3.2D & E). I propose that because HAC1 ligation by RtcB does not involve a 5′-

phosphate intermediate, Xrn1 is unable to degrade the 5′-hydroxyl exon product of Ire1

cleavage, tipping the balance toward ligation and producing HAC1s under normal growth

conditions (Fig. 2F, bottom). Thus Xrn1-mediated decay of HAC1 3′-exon appears to counteract

a low rate of background Ire1 cleavage to ensure the UPR is only activated when legitimately

stressed.

96

Kinase-mediated decay of excised intron is required for HAC1 translation.

In several instances, cells with mutations in repair and decay factors can splice HAC1 but

fail to grow on media containing tunicamycin (Fig. 3.3A), indicating that HAC1s production is not

sufficient to activate the UPR. I assayed expression of KAR2, an ER chaperone and direct target

of the Hac1 transcription factor (Kohno et al. 1993) and found that all repair and decay mutants

express significantly less KAR2 mRNA upon tunicamycin treatment (Fig. 3.3B), consistent with

another layer of UPR regulation downstream of HAC1s production. Excised HAC1 intron

accumulates upon tunicamycin treatment in trl1∆, xrn1∆, and trl1∆ xrn1∆ cells (Fig. 3.3C).

Moreover, the intron decay products that accumulate in these cells are indicative of kinase-

mediated decay: in xrn1∆ cells that lack Xrn1 but have 5′-kinase activity, intron products appear

with a few distinct, smaller products below the full-length 252 nt excised intron (Fig. 3.3C). In

contrast, excised intron accumulated as a uniform, ~250 nt product in trl1∆ cells that lack 5′-

kinase activity (Fig. 3.3C), independent of XRN1 status. Moreover, production of shorter decay

products (like those present in xrn1∆ cells) was dependent on Trl1 5′-kinase catalytic activity

(Fig. 3.3D). Interestingly, expression of the adenylyl-transferase-dead/ligase-dead allele, trl1-

K114A, also led to accumulation of some free HAC1 intron (Fig. 3I), potentially indicating a role

for ligation in the processing of liberated HAC1 intron. Together, these data show that excised

HAC1 intron is a substrate for kinase-mediated decay with strict dependence on a 5′-

phosphorylation step to promote 5′→3′ decay.

A single deletion of Dxo1 had no effect on HAC1 intron degradation (Fig. 3.3E); however,

the sizes of smaller decay products in xrn1∆ dxo1∆ cells were subtly different than in xrn1∆ cells

(Fig. 3.3E), indicating that Dxo1 and other exonucleases partially degrade excised HAC1 intron,

but only when it accumulates in xrn1∆ cells.

97

Figure 3.3: Kinase-mediated decay of excised HAC1 intron is required to activate the

unfolded protein response.

A. A serial dilution (5-fold) yeast growth assay on rich media (YPD) and tunicamycin-containing media (+Tm) compares the growth of xrn1∆ , trl1∆ (10x tRNA), and trl1∆ (RtcB) cells to resist

98

Figure 3.3 (continued): Kinase-mediated decay of excised HAC1 intron is required to

activate the unfolded protein response.

protein-folding stress. Growth of wild-type cells is unaffected by tunicamycin, whereas growth of xrn1∆ cells is partially inhibited by tunicamycin. Cells that lack ligase (trl1∆), and cells expressing E. coli RtcB RNA ligase in lieu of TRL1 (Tanaka et al., 2011) both fail to grow on media containing tunicamycin, and this growth defect is not affected by xrn1∆.

B. Expression of a UPR-responsive gene is compromised in RNA repair and decay mutants. RT-qPCR of mRNA for KAR2 (BiP), a direct target of Hac1 (Kohno et al., 1993), performed on total RNA from the indicated genotypes shows that wild-type cells induce KAR2 expression by 16-fold upon tunicamycin treatment. (Error bars are 95% confidence interval, n=3; comparison bars represent p < 0.01, Student’s t-test). trl1∆ cells show an insignificant increase in UPR induction, whereas tpt1∆ cells have elevated KAR2 levels in the absence of tunicamycin, which does not change significantly after tunicamycin treatment. trl1∆ (RtcB) and xrn1∆ cells have a modest increase in expression, but not to the same degree as wild-type (p < 0.01).

C. Excised HAC1 intron is stabilized in xrn1∆ and trl1∆ cells. Northern blot analysis using a probe to HAC1 intron reveals that excised intron (252 nt) and partially-degraded intron intermediates accumulate in xrn1∆ cells. Ligase-delete cells (trl1∆ ) also accumulate intron as a uniformly sized 252 nt product. In xrn1∆ and trl1∆ cells, intron accumulates in the absence tunicamycin.

D. Catalytic activity of Trl1 5′-kinase is required for 5′→3′ decay of excised HAC1 intron. Northern blot analysis using a probe to HAC1 intron shows that a missense mutation in the 5′-kinase domain of Trl1 (trl1-D 425N) phenocopies the HAC1 accumulation of trl1∆ cells (lanes 4 & 6, also C).

E. The distributive 5′→3′ exonuclease Dxo1 can partially degrade HAC1 intron. Northern blot analysis for HAC1 intron on total RNA from dxo1∆ and dxo1∆ xrn1∆ cells shows that Dxo1 can partially degrade HAC1 intron when it accumulates in xrn1∆ cells (compare lanes 4, 6 & 8). F. A slow-migrating intron species accumulates in trl1∆ cells expressing RtcB. Northern blot analysis of RNA from wild-type and trl1∆ cells shows that wild-type cells accumulate linear, partially degraded intron (lanes 1-4), whereas trl1∆ (RtcB), and trl1∆ xrn1∆ (RtcB) cells accumulate a slower-migrating species (~500 nt; lanes 5-8).

G. Cells expressing RtcB accumulate circular HAC1 intron. To test whether the slower-migrating band was circularized, total RNA was treated with RNase R and analyzed by northern blot. The slower-migrating species is largely protected from degradation, indicating it is a circle. Linear HAC1 intron and SCR1 (bottom) are degraded upon RNase R treatment (lanes 2, 4, 6 & 8). A panel of enhanced contrast shows that the slower migrating species in wild-type cells are circular, excised HAC1 introns, resistant to RNase R. Circular intron only occurs in samples from cells expressing a ligase.

H. The cytoplasmic exosome degrades HAC1 intron when 5′→3′ decay is disabled. Northern blot analysis using a HAC1 intron probe on total RNA from ski2∆ (a component of the cytoplasmic exosome) cells showed that ski2∆ cells accumulate excised HAC1 intron (lane 4) at levels similar to wild-type (lane 3). In trl1∆ ski2∆ cells, a lack of both kinase-mediated decay (trl1∆) and 3′→5′ decay (ski2∆) causes accumulation of excised intron relative to trl1∆ cells (compare lanes 6 & 8).

99

Figure 3.3 (continued): Kinase-mediated decay of excised HAC1 intron is required to

activate the unfolded protein response.

I. Catalytic activity of Trl1 ligase domain contributes to processing of excised HAC1 intron. We expressed a ligase-inactive Trl1 allele, trl1- K114A, as in Fig. 3.2C. Expression of trl1-K114A (lanes 5 & 6) leads to a modest accumulation of intron, though not to the same extent as in the kinase-inactivated mutant (trl1-D425N) (lanes 3 & 4) . The double missense mutant, trl1-K114A-D425N (lanes 7 & 8), exhibits intron accumulation similar that of trl1-D425N.

100

Consistent with this notion, we found that the cytoplasmic exosome also contributes to HAC1

intron turnover (Fig. 3.3H), but this mode of decay is unlikely to regulate the UPR as the growth

of ski2∆ cells is unaffected by tunicamycin (Fig. 3.1B).

In trl1∆ (RtcB) cells, excised HAC1 intron accumulates as a circle, evinced by its altered

mobility and resistance to Xrn1-mediated decay in vivo (Fig. 3.3F), and its resistance to RNase

R degradation in vitro (Fig. 3.3G). Robust circularization of the HAC1 intron by RtcB in trl1∆ cells

is facilitated by 5′-OH and 2′,3′-cyclic PO4 termini created by Ire1 cleavage and the absence of

Trl1 end modification activities that could otherwise produce termini incompatible with RtcB

ligation (5′-PO4 or adenylylate; and 2′-PO4/3′-OH). It is noteworthy that circularized intron

accumulates to high levels in the absence of tunicamycin (Fig. 3.3F & 3.G), indicating that Ire1

catalyzes a low level of intron excision (and 3′-exon excision (Fig. 3.1C)) from HAC1u during

normal growth, leading to the accumulation of stable, circularized introns in the presence of

RtcB.

The HAC1 intron and 5′-UTR form an extensive base-pairing interaction that inhibits

ribosome initiation (Chapman and Walter 1997; Di Santo et al. 2016). Thus, together these data

evoke a model in which kinase-mediated decay of the excised HAC1 intron is required for HAC1s

translation, and a failure to degrade HAC1 intron—even when HAC1s is produced—prevents

HAC1s translation and subsequent expression of stress-responsive genes. Despite their ability

to make HAC1s, xrn1∆ and trl1∆ (RtcB) cells have growth defects on media containing

tunicamycin (Fig. 3.3A) and the relative severity of their defects parallels the accumulation of

excised intron in these cells. Cells lacking Xrn1 have a modest growth defect and accumulate

linear intron, which can be degraded by other exonucleases (Fig 3.3C-E, G). In contrast, trl1∆

(RtcB) cells have a severe growth defect and accumulate high levels of a stable, circularized

intron that is immune to exonucleolytic decay (Fig. 3.3F & G). I believe these findings resolve the

101

mystery of the previously identified “second function” of Trl1 required for UPR activation (Mori

et al. 2010), namely that—in addition to its ligase activity—Trl1 initiates kinase-mediated decay

of the excised HAC1 intron, relieving its repressive effect on HAC1s and activating translation.

Incompletely processed HAC1 mRNA is endonucleolytically cleaved and degraded

In addition to the canonical 5′-exon product of 5′-splice site cleavage, a second product

uniquely accumulates in tpt1∆ and tpt1∆ ski2∆ cells that is ~50 nt shorter than full-length 5′-exon

(Fig. 3.4A). Expression of a catalytically inactive form of Tpt1 (Tpt1-R138A, (Sawaya et al. 2005))

in tpt1∆ cells failed to rescue this defect (Fig. 3.4B), affirming that the catalytic activity of Tpt1 is

required to prevent accumulation of the shorter 5′-exon fragment. A corresponding elongated

3′-exon fragment accumulates in tpt1∆, and more intensely in tpt1∆ xrn1∆ cells, indicating it is

degraded by Xrn1 (Fig. 3.4C-E). The elongated 3′-exon is specifically detected using a northern

probe with a sequence complementary to the distal 3′-end of the 5′-exon (Fig. 3.4C & D),

indicating that a portion of the 5′-exon is responsible for the increased size of this decay

intermediate. Moreover, a fragment of similar size hybridizes to a probe for the 3′-exon,

suggesting that the sequence derived from 5′-exon is linked to the 3′-exon (Fig. 3.4E); together

these data indicate that HAC1s is cleaved upstream of the 2′-phosphorylated ligation junction in

tpt1∆ cells, and these products are degraded by both Xrn1 and the cytoplasmic exosome.

To determine whether HAC1s sequence is sufficient for cleavage, I expressed plasmid-

encoded HAC1u and HAC1s in hac1∆ cells to identify HAC1 splicing intermediates. I performed

the analysis on trl1∆ cells, reasoning that if HAC1s sequence were sufficient to cause cleavage,

we would expect an accumulation of 3′-exon in cells unable to ligate or degrade the products

by kinase-mediated decay. Tunicamycin treatment of trl1∆ hac1∆ cells expressing HAC1u caused

Ire1-mediated cleavage and accumulation of cleaved 3′-exon (Fig. 3.4F, lane 2).

102

Figure 3.4: Incompletely processed HAC1s mRNA is cleaved and degraded.

A. A shortened form of 5′-exon accumulates in tpt1∆ cells and is degraded by the cytoplasmic exosome. Northern blot analysis of tpt1∆ and tpt1∆ ski2∆ using a probe to 5′-exon identified a shortened form of liberated 5′-exon, about 50 nt smaller than full-length (red arrowheads). A region of enhanced contrast shows the specific accumulation of this product in tpt1∆ and tpt1∆

ski2∆ cells. The bottom diagram depicts the relative positions of probes 1 and 2 on the HAC1 5′-exon; a red arrowhead marks the putative site of cleavage.

B. Production of shortened 5′-exon requires the catalytic activity of Tpt1. Northern blot analysis of tpt1∆cells expressing either a plasmid-encoded wild-type copy of TPT1 or a catalytically-inactive missense mutant (tpt1-R138A) (Sawaya et al. 2005) using a probe for 5′-exon. Production of the cleaved 5′-exon seen in tpt1∆ cells (lane 4, red arrow marks the band) is rescued by plasmid-mediated expression of wild-type TPT1 (lane 6) but not by Tpt1-R138A (lane

103

Figure 3.4 (continued): Incompletely processed HAC1s mRNA is cleaved and degraded.

8, red arrow marks the band), confirming that Tpt1 catalytic activity is required to prevent HAC1s secondary cleavage.

C. The shortened 5′-exon is missing a portion of its 3′-end. A northern blot was probed with 5′-exon probe 2, which hybridizes to the 3′-most 30 nt of HAC1 5′-exon (see diagram in A). Compared to the blot in A, the shortened 5′-exon species is absent (lanes 4 and 8) and instead the probe detects smaller bands consistent with the length of the elongated 3′-exon (see region of enhanced contrast).

D. A lengthened form of 3′-exon accumulates in tpt1∆cells. Total RNA from tpt1∆ and xrn1∆ cells was analyzed by northern blot with 5′-exon probe 2, a probe that anneals to the 3′-most 30 nt of HAC1 5′-exon. D and E share the same band interpretation key, so dashed lines have been drawn across to illustrate the different positions of typical 3′-exon (474 nt) and elongated 3′-exon (524 nt).

E. The elongated 3′-exon from tpt1∆ cells co-migrates and co-hybridizes with 5′-exon probe 2 (see D). Stripping the blot from D and re-hybridizing it with HAC1 3′-exon probe detects the same, elongated band(s) as in the 5′-exon probe 2 blot (lane 8), as well as HAC1 3′-exon bands of typical length.

F. Expression of “pre-spliced” HAC1s is not sufficient to promote cleavage. Total RNA from trl1∆

hac1∆ cells expressing HAC1u and HAC1s from a plasmid was analyzed by northern blot with a

3′-exon probe. Full length HAC1u expressed from this construct is cleaved upon tunicamycin

addition (lane 2). Expression of “pre-spliced” HAC1s yields a single product with no decay

intermediates (lanes 3 and 4), indicating that “pre-spliced” HAC1s—which was not produced by ligation and therefore lacks a 2′-phosphate—is not sufficient to recapitulate the cleavage found in tpt1∆ cells.

G. Expression of “pre-spliced” HAC1s is not sufficient to promote secondary cleavage. Total RNA from hac1∆ and tpt1∆ hac1∆ cells expressing HAC1u and “pre-spliced” HAC1s from a plasmid was analyzed by northern blot with a 3′-exon probe. In hac1∆ cells, full length HAC1u expressed from this construct is cleaved and ligated upon tunicamycin addition (lane 2). Expression of “pre-spliced” HAC1s yields a single product with no processing intermediates (lanes 3 and 4), indicating that “pre-spliced” HAC1s is not further processed. The HAC1u construct expressed in cells of genotype tpt1∆ hac1∆ (10x tRNA) gets cleaved upon tunicamycin treatment (lane 6), yielding a secondary cleavage fragment. Expression of pre-spliced HAC1s in tpt1∆ hac1∆ cells does not produce any additional processing fragments, indicating that 2′-phosphorylated products from HAC1u are required for secondary cleavage in tpt1∆ cells.

104

However, I found that full length HAC1s was the only product produced in the trl1∆ hac1∆ cells

in the presence and absence of tunicamycin (Fig. 3.4F). Furthermore, expression of HAC1s in a

tpt1∆ background failed to produce additional fragments (Fig. 3.4G), whereas expression of

HAC1u is sufficient in the tpt1∆ background to produce free HAC1 3′-exon, consistent with its

secondary cleavage. Together, these results (Fig. 3.4F & 3.4G) indicate that the HAC1s transcript

is not sufficient to recapitulate the secondary cleavage, and that HAC1u must be cleaved and

ligated to produce the 2′-phosphorylated HAC1s secondary cleavage substrate.

To further characterize this secondary cleavage product, I determined the 5′ end of the

elongated 3′-exon. I analyzed the 5′-ends of cleaved 3′-exon by primer extension and found that

cleaved 3′-exon was barely detectable in wild-type cells upon tunicamycin addition, whereas 3′-

exon accumulated in trl1∆ cells due to a lack of kinase-mediated decay (Fig. 3.5A, lanes 2 and

8). In tpt1∆ cells treated with tunicamycin, two products accumulated upon tunicamycin addition:

a product consistent with canonical length 3′-exon and a small amount of elongated 3′-exon

(Fig. 3.5A, lane 4). The elongated product accumulated in tpt1∆ xrn1∆ cells, again indicating it is

degraded by Xrn1 (Fig. 3.5A, lane 6). To test this prediction (summarized in Fig. 3.5B), I measured

the susceptibility of 3′-exon fragments to treatment in vitro with recombinant Xrn1 (rXrn1/TEX).

As expected, fragments from xrn1∆ cells were degraded by rXrn1 (Fig. 3.5C), establishing that

they have 5′-PO4 termini, whereas 3′-exon fragments from trl1∆ cells were resistant to rXrn1

degradation (Fig. 3.5C), indicating that they have 5′-OH termini.

The instructive findings came from examining 3′-exon accumulation in tpt1∆ cells. The 3′-

exon fragment of canonical length that accumulates in tpt1∆ cells was resistant to rXrn1

treatment (Fig. 3.5C, lanes 11 and 12), while the elongated fragment of 3′-exon is susceptible to

rXrn1 treatment (Fig. 3.5C, lanes 15 and 16).

105

Figure 3.5: A 5′- and 2′-phosphorylated HAC1 decay intermediate inhibits Xrn1.

A. Primer extension analysis of HAC1 3′-exon cleavage products. Primer extension using a probe for 3′-exon was performed on total RNA from wild-type, tpt1∆ (10x tRNA) , tpt1∆ xrn1∆

(10x tRNA) , and trl1∆ (10x tRNA) cells treated with or without tunicamycin. A loading control for U6 snRNA is depicted in the bottom panel. The extension product of cleaved 3′-splice site is 174 nt, found in wild-type cells treated with tunicamycin (lane 2) and is present in untreated trl1∆ c ells but increases upon tunicamycin treatment (lanes 7 and 8). An extension product from tpt1∆

cells accumulates upon tunicamycin treatment (lanes 3 and 4) and co-migrates with the product from wild-type (lane 2) and trl1∆ (lane 8) cells. In addition, a faint elongated product at ~225 nt is present in tpt1∆ cells (lane 4). This elongated product accumulates to higher levels in tpt1∆ xrn1∆ cells treated with tunicamycin (lane 6).

B. Model of HAC1 3′-exon primer extension product lengths. A 5′-radiolabeled (yellow star) primer (HAC1 3′-exon probe, see Table 2) anneals to HAC1 3′-exon mRNA and primes cDNA synthesis by a reverse transcriptase. The reverse transcriptase stops synthesizing cDNA when it runs out of RNA template at the 5′-terminus of the 3′-exon. The model shows three situations as observed in A: 1) canonical 3′-exon cleavage fragments (lower right), as observed in trl1∆ cells; 2) extended HAC1 3′-exon (above reaction arrow), as caused by secondary cleavage of

HAC1s, observed in tpt1∆ xrn1∆ (10x tRNA) cells; and 3′-extended HAC1 3′-exon, on which Xrn1(cellular or in vitro) initiated decay, but failed to proceed past the 2′-P (beneath arrow).

106

Figure 3.5 (continued): A 5′- and 2′-phosphorylated HAC1 decay intermediate inhibits Xrn1.

C. Susceptibility of 3′-exon cleavage products to in vitro Xrn1 degradation. Total RNA from A

was treated with recombinant Xrn1 (rXrn1) and analyzed by primer extension for the 3′-exon. The loading control performed on U6 snRNA is depicted in the bottom panel. The 3′-exon from trl1∆ cells accumulates upon tunicamycin treatment (compare lanes 1 & 2 to 3 & 4) but the 3′-exon is resistant to rXrn1 (lanes 3 and 4) because it lacks a 5′-phosphate due to lack of Trl1 5′-kinase activity in these cells. In contrast, the 3′-exon products from xrn1∆ cells, which have Trl1 5′-kinase and thus 5′-phosphates, are degraded by rXrn1 (lanes 6 and 8). The 3′-exon product in tpt1∆ cells is resistant to rXrn1 treatment (compare lanes 9 to 10, and 11 to 12), despite the fact these cells have both Trl1 and Xrn1. The elongated 3′-exon product that accumulates in tpt1∆

xrn1∆ cells is partially degraded by rXrn1 (compare lane 15 to 16) and the decay intermediate co-migrates with cleaved 3′-exon.

D. Model depicting cleavage and decay of 2′-phosphorylated HAC1s. HAC1s is cleaved (likely by Ire1) ~50 nt upstream of the 2′-phosphorylated ligation junction, creating a 3′-product with ~50 nt of sequence of the 5′-exon (black) and an internal 2′-phosphate. Xrn1 initiates decay at the 5′-terminus, degrading the 5′-exon portion (black) up to the site of 2′-phosphorylation. The product of partial decay contains a 5′- and 2′-phosphate at its first position, which inhibits further degradation by Xrn1.

107

These observations indicate that 3′ product of secondary cleavage of HAC1s is a substrate of

Xrn1 in vivo and in vitro, raising the possibility that it may also be a substrate of kinase-mediated

decay, depending on the chemistry of the endoribonuclease that generates the secondary

cleavage. I propose these products are created via two steps: (i) HAC1s is cleaved ~50 nt

upstream of the 2′-phosphorylated ligation junction; (ii) Xrn1 partially degrades the intermediate

fragment to the site of 2′-phosphorylation, which inhibits further degradation (Fig. 3.5D). Under

this model, the 3′-exon fragment that accumulates in tpt1∆ cells has both 5′-PO4 and 2′-PO4

moieties at its first position, which inhibits Xrn1-mediated decay in vivo and in vitro (Fig. 3.1C;

3.4D; 3.5A & C).

The accumulation of HAC1 decay intermediates in tpt1∆ cells over time further supports a

model of HAC1s cleavage by Ire1. In tpt1∆ cells, the accumulation of secondary cleavage product

coincides with the increase in production of HAC1s at 20 minutes (Fig. 3.6). In tpt1∆ cells, cleaved

HAC1 3′-exon is present at low levels at steady state and accumulates over the course of two

hours upon tunicamycin treatment (Fig. 3.6A). HAC1s is also generated in tpt1∆ cells, but at

significantly reduced levels compared to wild-type (Fig. 3.6A). It is also notable that tpt1∆ xrn1∆

cells (Fig. 3.6B) and tpt1∆ ski2∆ cells (Fig. 3.6D) accumulate more spliced HAC1s than tpt1∆ cells

(Fig. 3.6E), suggesting that some HAC1s molecules or splicing intermediates in tpt1∆ cells are

degraded, possibly because they contain 2′-PO4 moieties. At all time points, tpt1∆ cells contain

more free 3′-exon than HAC1s, a ratio opposite to wild-type cells (Fig. 3.6A), indicating that that

3′-exon cleaved from HAC1u accumulates as a result of partial decay of 5′- and 2′-

phosphorylated HAC1s. The augmented accumulation of 3′-exon in tpt1∆ cells is also observed

in the primer extension analysis (Fig. 3.5A).

108

Figure 3.6: Kinetic analysis of HAC1 mRNA processing in cells lacking Tpt1.

A. Northern blot analysis of a time course of tunicamycin treatment (0 to 120 minutes) in tpt1∆

(10x tRNA) cells showing the dynamics of HAC1u splicing using a probe for 3′-exon. The SCR1

loading control is shown below the panel. RNA from wild-type cells treated for 0 and 120 minutes

tunicamycin was loaded in lanes 7 and 8. HAC1s accumulates slowly over 120 minutes and its

109

Figure 3.6 (continued): Kinetic analysis of HAC1 mRNA processing in cells lacking Tpt1.

abundance at the end of the time-course is 35-fold lower than wild-type (lanes 6 and 8). A 3′-exon cleavage product (~475 nt) is present at time 0 and accumulates over time; its increase is

coincident with the appearance of HAC1s (20-minute time point) and its abundance at 120

minutes exceeds the level of HAC1s . A 3′-exon cleavage product in wild-type cells (~500 nt) is apparent at 120 minutes (lane 8).

B. Northern blot analysis of HAC1 splicing in tpt1∆ xrn1∆ (10x tRNA) cells using a probe for 3′-exon. Conditions are the same as A. HAC1s increases at 20 minutes and its final level (lane 6) is

higher than HAC1s in tpt1∆ cells (A, lane 6), approaching wild-type levels (lane 8). Levels of the product of 5′-splice site cleavage (intron/3′-exon) accumulate over 120 minutes. The 3′-exon product (~475 nt) and cleaved 3′-exon (~500 nt) begin accumulating at 20 minutes, coincident

with increased levels of HAC1s . A 3′-exon cleavage product in wild-type cells (~500 nt) is apparent at 120 minutes (lane 8).

C. Northern blot analysis of HAC1 splicing in tpt1∆ (10x tRNA) cells using a probe for 5′-exon.

Conditions are the same as A. HAC1s increases at 20 minutes up to final level (lane 6) (A, lane 6), approaching wild-type levels. Canonical 5′-exon (~725 nt) and secondarily-cleaved 5′-exon (~675 nt; red arrowheads) begin accumulating at 20 minutes, coincident with increased levels of

HAC1s.

D. Northern blot analysis of HAC1 splicing in tpt1∆ ski2∆ (10x tRNA) cells using a probe for 5′-exon. Deletion of SKI2 in the context of tpt1∆ increases the abundance of nearly all intermediates relative to tpt1∆ alone (A). Shortened HAC1 5′-exon is marked with red arrowheads. The product of 3′-splice site cleavage (5′-exon/intron) is present at time 0 and accumulates over 120 minutes,

whereas HAC1s accumulates beginning at 20 minutes.

E. Northern blot analysis of HAC1 splicing using a probe for 5′-exon to compare the beginning and end time points of time course experiments performed on tpt1∆ , tpt1∆ xrn1∆, and tpt1∆ ski2∆. Results from densitometry of the spliced and unspliced bands are written beneath the blot. Less HAC1 splicing takes place in tpt1∆ cells compared to wild-type, and disabling decay factors ski2∆ and xrn1∆ in the tpt1∆ mutant lead to increased HAC1 splicing, compared to tpt1∆ alone.

110

Xrn1’s in vitro ability to only partially degrade elongated HAC1 3′-exon from tpt1∆ xrn1∆ cells,

and inability to degrade canonical 3′-exon from tpt1∆ cells, indicates that the accumulation of

free HAC1 3′-exon is likely caused primarily by blocked 5′→3′ degradation.

Discussion

Many different regulatory events impinge on HAC1 mRNA to control its localization and

processing. It has been assumed that cleavage of HAC1u by Ire1 is the rate-limiting step for UPR

activation. Counter to this view, we found that decay of HAC1 splicing intermediates is required

for both UPR activation and suppression. I found several examples wherein “kinase-mediated

decay” (KMD) degrades HAC1 splicing intermediates containing 5′-OH termini by sequential 5′-

phosphorylation and 5′-phosphate-dependent 5′→3′ exonucleolytic degradation (Fig. 3.7).

I propose that after 3′-splice site cleavage by Ire1, the Trl1 5′-kinase domain associates

with and phosphorylates the 5′-OH of the 3′-exon product (Fig. 3.7). Dissociation of the Trl1

kinase active site from the 5′-PO4 product then enables a competition between re-association

of Trl1 (now its adenylyltransferase/ligase domain) to catalyze ligation—or Xrn1 to catalyze

degradation. In some circumstances, this balance is tipped to favor ligation even in the absence

of overt UPR stimulation: a lack of decay in xrn1∆ cells favors ligation, whereas the lack of 5′-

kinase activity in trl1∆ (RtcB) cells renders Xrn1 decay irrelevant (Fig. 3.2F). Xrn1 is abundant in

budding yeast (Ghaemmaghami et al. 2003), which may efficiently suppress the UPR under

normal conditions by degrading spuriously cleaved HAC1 3′-exon intermediates. Regulation of

the ligation step of HAC1s splicing makes intuitive sense because it is the last opportunity to act

during splicing; once ligated, HAC1s is again protected by a 7-methylguanosine cap and a poly-

(A) tail.

111

Figure 3.7: Decay of HAC1 splicing intermediates regulates UPR activation, suppression,

and attenuation.

Activation of Ire1 by ER stress activates its endoribonuclease activity, leading to cleavage of the

5′- and 3′-splice sites of HAC1u. The cleaved 3′-exon (grey) is phosphorylated by the 5′-kinase activity of Trl1, permitting either its ligation or decay by Xrn1. The 5′-exon (black) and intron (thin line) form an extensive base-pairing interaction (grey squiggle). Kinase-mediated decay of the

intron is required to activate the translation of HAC1s . Ligated but 2′-phosphorylated HAC1s may also be cleaved by Ire1 upstream of the ligation junction, and the cleavage products are degraded by Xrn1 and the cytoplasmic exosome. In cells lacking Tpt1, a unique KMD intermediate accumulates with 5′- and 2′-phosphates, which inhibit Xrn1-mediated degradation.

112

Spurious HAC1 splicing was previously reported in trl1∆ cells expressing the tRNA ligase from

Arabidopsis thaliana (Mori et al. 2010). It is noteworthy that there are mechanistic differences

between the tRNA ligases of A. thaliana and another yeast (K. lactis) (Remus and Shuman 2014),

suggesting that these differences could impact the balance between kinase-mediated decay and

ligation in budding yeast expressing plant RNA ligase.

I also found that excised HAC1 intron is a substrate of kinase-mediated decay (Fig. 3.3).

Indeed, we believe that phosphorylation of HAC1 intron by Trl1 to promote kinase-mediated

decay is the previously proposed “second role” of Trl1 ligase in activating HAC1 translation

independent of ligation (Mori et al. 2010). In this previous study, excised and circularized HAC1

intron was found to remain associated with HAC1s, inhibiting translation. Kinase-mediated decay

of excised intron therefore likely relieves the long-range base-pairing interaction that prevents

HAC1s translation (Cox et al. 1997; Rüegsegger et al. 2001; Di Santo et al. 2016), explaining how

HAC1s can accumulate without concomitant UPR activation. This second layer of control over

HAC1s translation by KMD adds another failsafe mechanism to prevent its translation and

unintentional UPR activation.

Previous examples of kinase-mediated decay of bacterial mRNAs (Durand et al. 2012),

eukaryotic tRNA introns (Wu and Hopper 2014), ribosomal RNA processing intermediates (Gasse

et al. 2015), and no-go mRNA decay cleavage products (Navickas et al. 2018) suggest that this

mode of decay may be widespread. Coupling of RNA 5′-kinase and 5′→3′ exonucleolytic decay

activities in the context of kinase-mediated decay may regulate the UPR in other organisms.

Splicing of Xbp-1 mRNA in metazoans (the functional homolog of HAC1) is catalyzed by Ire1-

mediated removal of an intron and ligation by RtcB RNA ligase (Kosmaczewski et al. 2014).

Fundamental differences in the chemistry of RNA ligation between fungal and metazoan ligases

suggest that Xbp-1 may be subject to a biochemically distinct mode of regulation. Because RtcB

depends on 5′-OH and 3′-PO4 termini for catalysis (Chakravarty et al. 2012), activities that

113

remodel 5′-OH RNA termini could divert Ire1-generated 5′-OH splicing intermediates from

productive ligation. To that point, cyclic nucleotide phosphodiesterase (CNP) and RtcA (a 2′,3′-

RNA cyclase) were shown to “tune” the UPR in metazoans by competing with RtcB/HSCP117

for ligation substrates (Unlu et al. 2018). Specifically, CNP hydrolyzes the 2′,3′-cyclic phosphate

compatible RtcB to a 2′-PO4 incompatible with RtcB, decreasing ligation of Xbp1. Conversely,

RtcA, which converts 2′-PO4 RNA to 2′,3′-cyclic phosphate, makes the terminus compatible

with RtcB, thus enhancing Xbp1 splicing.

Additionally, an RNA 5′-kinase (e.g., Clp1) may phosphorylate the 3′-exon product of Xbp-

1 cleavage, simultaneously inhibiting ligation by RtcB and promoting its degradation by a 5′-

phosphate-dependent exoribonuclease to limit UPR activation. In this vein, it is noteworthy that

kinase-inactivating mutations in Clp1 cause neurodevelopmental defects and neuronal

dysfunction in humans, mice, and zebrafish (Hanada et al. 2013; Karaca et al. 2014; Schaffer et

al. 2014), possibly due to chronic UPR activation in neural tissues (Clayton and Popko 2016).

Xrn1 degrades mRNA fragments generated during metazoan Regulated Ire1-dependent

Degradation (RIDD) (Hollien and Weissman 2006); however, it is not known how these 5′-OH

cleavage products of Ire1 are phosphorylated for Xrn1-mediated decay. The RNA 5′-kinase

Clp1(Weitzer and Martinez 2007) and polynucleotide kinase Nol9 (Heindl and Martinez 2010) are

candidates for this activity, though neither are known to phosphorylate mRNA decay

intermediates.

While 5′→3′ decay plays a major role in UPR regulation, I found little evidence for UPR

regulation by 3′→5′ decay activity. As shown previously (Schwer et al., 2004), the exposed 3′-

end of cleaved HAC1 5′-exon is a substrate for 3′→5′ decay (Fig. 3.4) . But while excised HAC1

intron is stabilized in ski2∆ cells lacking cytosolic 3′→5′ decay (Fig. 3.3H), their growth is

unaffected by tunicamycin (Fig. 3.1B) , indicating that 3′→5′ decay of the excised intron does

not contribute substantially to intron-mediated HAC1s repression.

114

I also found evidence that Ire1 cleaves incompletely processed HAC1s mRNA, which is

ligated but contains an internal 2′-PO4 moiety. Cleavage of HAC1s leads to 5′- and 3′-fragments

that are degraded; however, the 3′-fragment is only partially degraded by kinase-mediated

decay, producing a 5′- and 2′-phosphorylated molecule that cannot be degraded by Xrn1.

Consistent with these findings, a recent study also showed that an RNA with an internal 2′-

phosphate group is protected from 3′→5′ decay by E. coli PNPase in vitro (Munir et al. 2018b);

those and our results together indicate that site-specific installation of a 2′-PO4 is an effective

strategy to protect an RNA from complete exonucleolytic degradation in vivo or in vitro.

Decay intermediates produced from incompletely processed, 2′-phosphorylated HAC1s

have not been previously observed and suggest a plausible regulatory role for Tpt1 in regulating

HAC1s fate. I have yet to determine the impact of 2′-phosphorylation on HAC1s translation, but

given that tpt1∆ cells grow on tunicamycin (Fig. 3.1B & (Cherry et al. 2018)) and activate KAR2

gene expression (Fig. 3.3B), it is likely that some Hac1 protein is produced from 2′-PO4 HAC1s

mRNA. It also remains to be determined how and why incompletely processed HAC1s is cleaved.

Insofar as the HAC1s cleavage substrate is initially produced by tunicamycin-dependent Ire1

cleavage and ligation, we conjecture that Ire1 incises ligated, 2′-phosphorylated HAC1s

upstream of the original ligation junction, yielding smaller 5′-exon and larger 3′-exon fragments.

Formally, I cannot currently rule out the possibility that another endonuclease catalyzes

secondary HAC1s cleavage; however, Ire1 is the only endoribonuclease known to site-

specifically incise HAC1 mRNA. I note that ire1∆ mutants are unable to initiate processing of

HAC1u and therefore do not make HAC1s in the first place (Sidrauski and Walter 1997),

precluding direct analysis of HAC1s cleavage in ire1∆ mutants.

I showed that the 2′-PO4 is required for cleavage, as expression of “pre-spliced” HAC1s

does not lead to cleavage (Fig. 3.4F & G). It is possible that the presence of a 2′-PO4 in the

115

context of a composite Ire1 splice site (i.e., formed from two halves of the original 5′- and 3′-

splice sites (Sidrauski and Walter 1997; Hooks and Griffiths-Jones 2011)) on HAC1s is recognized

by Ire1, but because a 2′-OH is the nucleophile for transesterification by metal-independent Ire1

(Gonzalez et al. 1999), the 2′-PO4 inhibits the chemical step of incision. This model also provides

a plausible mechanism to explain how Ire1 incises a neighboring, non-canonical site. We

propose that a failure of Ire1 to release 2′-phosphorylated HAC1s would enable the active site of

a nearby Ire1 molecule—in the context of its activated, oligomeric form (Korennykh et al. 2009)—

to catalyze site-specific incision at the second, upstream site. My results also raise the question

of why Ire1 would cut incompletely processed HAC1s. Here, cleavage of incompletely processed

(ligated, but 2′-phosphorylated) HAC1s could be a means to inactivate HAC1s after prolonged

stimulation to attenuate the UPR (Chawla et al. 2011; Rubio et al. 2011).

Summary

In the unfolded protein response (UPR), protein-folding stress in the endoplasmic reticulum

(ER) activates an extensive transcriptional program to increase ER folding capacity. During the

budding yeast UPR, the trans-ER-membrane kinase-endoribonuclease Ire1 excises an intron

from the HAC1 mRNA and the exon cleavage products are ligated and translated to a

transcription factor that induces dozens of stress-response genes. HAC1 cleavage by Ire1 is

thought to be the rate limiting step of its processing. Using cells with mutations in RNA repair

and decay enzymes, I showed that phosphorylation of two different HAC1 splicing intermediates

by Trl1 RNA 5′-kinase is required for their degradation by the 5′→3′ exonuclease Xrn1 to enact

opposing effects on the UPR. Kinase-mediated decay (KMD) of cleaved HAC1 3′-exon competes

with its ligation to limit productive splicing and suppress the UPR, whereas KMD of the excised

intron activates HAC1 translation, likely by relieving an inhibitory base-pairing interaction

116

between the intron and 5′-untranslated region. We also found that ligated but 2′-phosphorylated

HAC1 mRNA is endonucleolytically cleaved, yielding a decay intermediate with both 5′- and 2′-

phosphates at its 5′-end that inhibit 5′→3′ decay and suggesting that Ire1 initiates the

degradation of incompletely processed HAC1s to proofread ligation or attenuate the UPR. These

multiple decay events expand the scope of RNA-based regulation in the budding yeast UPR and

may have implications for the control of the metazoan UPR by mRNA processing.

117

CHAPTER IV

IV FUTURE DIRECTIONS

This work illustrates four contributions to the field: (i) the only essential functions of Trl1

and Tpt1 are tRNA splicing; (ii); the tRNA splicing endonuclease (SEN) has an additional essential

function beyond tRNA splicing; (iii) kinase-mediated decay (KMD) occurs on two substrates,

HAC1 intron to relieve unfolded protein response repression and on HAC1 3′-exon to suppress

spurious UPR activation; and (iv) 2′-phosphorylated RNA is stabilized from Xrn1 degradation.

Many unanswered questions remain, addressed in detail below. Kinase-mediated decay occurs

on 5′-OH substrates in yeast, and likely other organisms that use a Trl1/Tpt1 mechanism of

ligation. But what about in organisms, like humans, that use a RtcB mechanism? Also, beyond

tRNAs, does RtcB serve an additional essential function? Can this Trl1/Tpt1 genetic bypass in

yeast be applied to identify the products of RNA repair, transcriptome-wide? And can the 2′-PO4

stalling of Xrn1 degradation be put to use to stabilize RNAs of our choice? What utility could

these 2′-PO4 stabilized RNAs have?

Does kinase-mediated decay also regulate Xbp1 splicing in animals?

In Chapter III of this dissertation (Cherry et al. 2019) I unveiled the first examples of kinase-

mediated decay (KMD) on mRNA. For both the 3′-exon and intron of HAC1 mRNA, the

endoribonucleolytic cleavage releases them as 5′-hydroxyl fragments that can be 5′-

phosphorlated and subsequently degraded by Xrn1. In addition to these being the first examples

of KMD among mRNAs, they also regulate/tune activation of the unfolded protein response, and

without them UPR signaling functions improperly. In this section, I lay out how KMD may regulate

the metazoan UPR.

118

Animals express an RNA ligase with a different biochemical mechanism than that of yeast,

but that difference could still permit a competition between decay and ligation. In fungi and

animals, Ire1 (IRE1α) functions as a metal-ion-independent endoribonuclease, releasing

products with 5′-OH termini. Mammalian RtcB/HSPC117 can then take those 5′-OH termini and

use them directly in ligation (Fig. 1.3). Because RtcB guanylates the 3′-terminus of the other

ligation substrate, RtcB biochemically requires a 5′-OH RNA to ligate. Therefore, an RNA 5′-

kinase could interfere with the ligation reaction by phosphorylating the 5′-terminus before RtcB

encounters it (Fig. 4.1). Humans have two RNA 5′-kinases that could participate in this

competition: Clp1 and Nol9 (Weitzer and Martinez 2007; Heindl and Martinez 2010). Both

proteins catalyze the ATP-dependent phosphorylation of the 5′-terminus of RNA with little

sequence specificity. Competition experiments could be carried out in the trl1∆ (RtcB) yeast

backgrounds already on hand, but that would be rather indirect because it would test

competition for substrate between enzymes from two different organisms expressed in yet

another organism. Furthermore, there is the potential for a synthetic lethal interaction of

expression, which may occur if the chosen 5′-kinase vastly outcompetes RtcB (by activity or

sheer protein abundance) for cleaved tRNAs, which RtcB is ligating together to keep the cells

alive. That result itself would be interesting and informative about the relative activity of the two

enzymes, but it would be an indirect assessment of competition.

A more human-specific experiment would be to knock-down and overexpress Clp1 and

Nol9 RNA 5′-kinases and test for a difference in Xbp1 splicing without and with UPR stimulation

(with thapsigargin or tunicamycin). Clp1 is an important structural protein in the RNA polymerase

II cleavage and polyadenylation complex (Minvielle-Sebastia et al. 1997; de Vries et al. 2000),

which may necessitate relegating Clp1 exclusively to the nucleus with a Clp1-nuclear-

localization-sequence construct that would limits its contact with cytoplasmic Xbp1 mRNA.

119

Figure 4.1: Hypothesis of competition between decay and ligation in animals.

At left, the model depicts the 3′-exon of HAC1 released from Ire1 cleavage with a 5′-hydroxyl terminus. The multifunctional Trl1 phosphorylates the 5′-terminus. As a 5′-phosphate, the 3′-exon can either be ligated (Trl1) or be degraded (Xrn1). On the right, a proposed model of competition in animals shows the 3′-exon of Xbp1 immediately after cleavage by Ire1α with a 5′-OH. At this step, the exon can either be ligated (RtcB/HSPC117) or be phosphorylated (Clp1 kinase). The ligation pathway would commit the 3′-exon to activating the UPR via Xbp1 translation, whereas the 5′-PO4 3′-exon would be degraded by Xrn1.

120

To test the disease relevance of Clp1 and UPR signaling, one of the disease-linked missense

mutations (clp1-R140H or -G419A) (Karaca et al. 2014) could be expressed to test for spurious

Xbp1 splicing, as would be expected from compromised kinase activity if decay and ligation

compete.

However, there are several reasons that KMD-based regulation of the UPR is likely not

occurring in animals. Firstly, compartmentalization of RNA 5′-kinases may deprive the cytoplasm

of robust kinase activity. Clp1 is ~85% localized to the nucleus (Schaffer et al. 2014); Nol9 is a

nucleolar-localized protein that participates in ribosomal RNA maturation (Heindl and Martinez

2010). Secondly, RtcB is an efficient ligase, and competing with its rapid catalysis would require

a sufficiently rapid 5′-kinase. While human RtcB/HSPC117 has not been directly tested for its

ligation prowess, E. coli RtcB, which operates by the same biochemical mechanism, is a fast

ligase, especially for RNA termini held in proximity to one another (Tanaka and Shuman 2011).

Thirdly animals have two other signaling mechanisms for rectifying ER stress, ATF6 and PERK,

which can compensate for signaling (to some degree) and thus limit the evolutionary pressure to

develop intricate or nuanced mechanisms of regulation on the Ire1/Xbp1 arm of the response.

Furthermore, regulation of the UPR in animals via KMD on the intron of Xbp1 is unlikely because

the intron has not been shown to participate in regulation once cleaved out of the mRNA. (In

contrast, the “dual function” of Trl1 in yeast is to start KMD of HAC1 intron (Mori et al. 2010).) In

fact, the secondary structure of Xbp1 favors ejection of the intron upon cleavage at the exon-

intron junctions, zipping up to position the 5′- and 3′-splice sites in close proximity for rapid

ligation (Peschek et al. 2015).

Identify precise HAC1s secondary cleavage location and nuclease.

Products of secondary cleavage of HAC1s are observed in tpt1∆ cells. The 5′-fragment,

shortened by the cleavage, accumulates in tpt1∆ ski2∆ cells (Fig. 3.4A); the 3′-fragment,

121

lengthened by the cleavage, accumulates in tpt1∆ xrn1∆ cells (Fig. 3.4D & E). As the project

currently stands, it is not clear how these secondary cleavage fragments are generated or where

precisely the secondary cleavage occurs on the HAC1s mRNA. Below, I state two hypotheses

about the mechanism of secondary cleavage, and I propose two genetic experiments to test the

hypotheses. I also propose two methods to precisely identify the site of secondary cleavage of

HAC1s.

Given that expression of “pre-spliced” HAC1s from a plasmid is not sufficient for secondary

cleavage of the transcript, the 2′-PO4 appears to be a necessary feature of the HAC1s mRNA

from tpt1∆ cells for secondary cleavage. Thus, all experiments on this topic will have to occur

on tunicamycin-treated tpt1∆ cells so that 2′-PO4 HAC1s mRNA can be generated. Two

hypotheses I have for the generation of this fragment are: (1) Ire1 cleaves a different site on 2′-

PO4 HAC1s; or (2) no-go mRNA decay occurs because 2′-PO4 in an open reading frame can stall

ribosomes during translation elongation.

Testing whether Ire1 is cleaving HAC1s is complicated by the requirement for Ire1 to

generate 2′-PO4 HAC1s in tpt1∆ cells. Therefore, I would generate a degron-tagged Ire1 strain

(Johnston et al. 1995; Sheridan and Bentley 2016) in the tpt1∆ background so that Ire1 could be

stabilized in the presence of methotrexate to produce 2′-PO4 HAC1s, and then depleted by

removal of methotrexate from the media. With a starting amount of HAC1s in the cells, northern

blotting RNA harvested in a time course after degradation of Ire1 would show whether any

additional secondary cleavage products accumulate after removal of Ire1. A failure to accumulate

more secondary cleavage product would indicate that Ire1 is required for the secondary

cleavage. Given that Ire1 is an endonuclease already known to cleave HAC1, the finding would

debut a curious function of Ire1, where it cleaves a product (at a non-canonical site) that it

generated, meaning that it can antagonize the same signaling pathway that it activates. Such a

122

function is reminiscent of a quality-control pathway, where Ire1 would recognize defective

products and mark them for decay.

Alternatively, secondary cleavage HAC1s products could be caused by no-go mRNA

decay, wherein an elongating ribosome(s) stalls upon encountering the 2′-PO4 in the open

reading frame. It is unclear how much Hac1 protein is produced in tpt1∆ cells because the

epitope tag I chose to use in Chapter II was a C-terminal tag, and therefore would be synthesized

less than an N-terminal tag on an mRNA that causes stalls in translation. If no-go decay is

responsible for the secondary cleavage of HAC1, a simple genetic deletion of DOM34, the

release-factor-like protein that detects and rescues stalled ribosomes, would prevent the

secondary cleavage, and so the fragment would become undetectable on northern blots or

primer extension assays. A complementary approach would be to express a mutant HAC1

mRNA with no start codons and so would test whether translation is required for generating the

secondary cleavage.

The precise 5′-end of the extended 3′–fragment generated from secondary cleavage of

HAC1s remains undetermined. Northern blot and primer extension results suggest that the

secondary cleavage of HAC1s occurs ~40 nt upstream of the original ligated site, generating the

observed products. Simultaneously, northern results indicate that sequence from the

downstream region of the 5′-exon goes missing from the shortened form of 5′-exon, consistent

with cleavage at an alternative site. I propose to perform 5′-end capture, PCR amplification, and

Sanger sequencing of extended 3′-exon to determine precisely where the secondary cleavage

site is. A similar process could be done with 3′-end-capture of the shortened 5′-exon.

Alternatively, the sequence of the extended 3′-exon could be read of by dideoxy-sequencing the

RNA with ddNTPs spiked into primer extension reactions. Identifying the precise location of

secondary cleavage could provide insight into the nuclease that cleaves at that site and would

123

provide helpful information for performing in vitro experiments to recapitulate the 2′-phosphate-

dependent cleavage of HAC1 RNA.

Applications of 2′-phosphorylated RNA to enhance stability in vivo

Typical cellular RNAs resist 5′→3′ decay via the regulated process of capping and

decapping (Parker 2012). Typically, 5′-PO4 RNAs are destined for decay by the abundant and

processive exonuclease Xrn1. However, in the course of studying the effect of tpt1∆ on the

unfolded protein response, I found via northern blot that an elongated form of HAC1 3′-exon

was present in cells at the expense of a shortened 5′-exon (Fig. 3.4). Curiously, the stabilized

3′-exon accumulated at its typical length in tpt1∆ cells, but tpt1∆ xrn1∆ caused the band to shift

upward, suggesting additional sequence at the 5′-end, ruling out the previously hypothesized

reason for variability of the band: poly(A) tail length changes. Primer extension analysis revealed

that only those two products are recovered from cells, consistent with a processive decay factor.

Recombinant Xrn1 (rXrn1) incubated with RNA samples in vitro degraded extended 3′-exon to

the canonical 3′-exon site, where the 2′-PO4 resides in tpt1∆ cells. Thus, I concluded that internal

2′-phosphorylation of a substrate of Xrn1 is necessary for the stall in decay I observe (though it

may not be sufficient). Furthermore, I found that 3′-exon from tpt1∆ cells that has already been

partially degraded by cellular Xrn1 was resistant to further decay by rXrn1 in vitro, indicating that

a 5′-PO4, 2′-PO4 nucleotide at the beginning of an RNA molecule can prevent Xrn1 degradation.

Structures that stabilize RNAs from 5′→3′ decay are sought after by scientists and

biological agents alike. Scientists seek stable RNAs to maintain high levels of gene expression

in experiments and to ensure long half-lives of RNA-based drugs delivered to patients. Viruses

steal 5′-caps and fold their RNA genomes into three-dimensional structures to evade

degradation by 5′→3′ exoribonucleases (Plotch et al. 1979; 1981; Chapman et al. 2014). While

the 2′-PO4 left behind on the HAC1s mRNA (and on tRNAs) in tpt1∆ cells may not be

124

physiological, the modification nonetheless represents a means of stabilizing an RNA of one’s

choice. Much like the Xrn1-resistant RNA structures, 2′-PO4 could be chemically or enzymatically

installed on an RNA of one’s choice, either terminally or internally. A 5′-terminal 2′-PO4 would

stabilize the RNA in its entirety, whereas an internal 2′-PO4 would permit decay of the RNA up

to a point. Furthermore, the 5′,2′-phosphorylated RNA would be stable even in cells with

functioning Tpt1/Trpt1 activity because the 2′-phosphotransferase activity of Tpt1/Trpt1 appears

restricted to internal 2′-sites, whereas terminal 2′-PO4 would likely be resistant to Tpt1/Trpt1

activity (Zillmann et al. 1992; Culver et al. 1997).

The functions of RNAs stabilized with 5′-PO4, 2′-PO4 termini would likely be restricted to

non-coding roles in the cell, with the one exception of open reading frames translated with the

help of an Internal Ribosome Entry Site (IRES) (Jang et al. 1988). Because translation initiation

largely relies on 5′-caps and poly(A) tails (Dever et al. 2016), a 2′-PO4 stabilized RNA would likely

need another means of recruiting the translational machinery to the ORF. However, the plethora

of non-coding functions of RNAs (e.g. long non-coding RNAs, antisense RNAs, RNAi) would all

be compatible with the 2′-PO4 stabilization technique.

Transcriptome-wide sequencing of products of RNA repair via enrichment of 2′-PO4 RNAs

from tpt1∆ yeast

What RNAs get cleaved and put back together? The field of RNA repair is interested in

identifying the substrates and/or products of RNA repair, especially in eukaryotes. While the

mechanism of repair (i.e. the enzymes) are well studied, their substrates are less so.

Endonucleolytic cleavage of mRNA has a broad range of nucleases and substrates, only a

handful of which are known to be repaired. Studies that captured and sequenced the specific

termini of endoribonucleolytic cleavage (2′,3′-cP capture and 5′-OH capture) (Schutz et al. 2010;

Cooper et al. 2014; Peach et al. 2015) have detected some promising leads, but as many

125

examples show, not all cleaved RNAs are destined to be ligated back together. Currently, the

limited number of known RNA repair substrates include bacterial rRNA, tRNA, and HAC1/Xbp1

mRNA. But is RNA repair a broader phenomenon in biology? The tpt1∆ (10x tRNA) yeast may be

able to help us find out. If capturing the substrates of repair (e.g. products of cleavage) haven’t

answered this question, the next logical step is to capture products of ligation.

Cells with a Trl1/Rnl1 ligation system that are ablated for Tpt1/Trpt1 activity accumulate

products of ligation with a stable 2′-PO4 at the ligation junction. This 2′-PO4 is a site-specific

mark of ligation, and so provides evidence that RNA repair took place on that molecule at that

site. Enrichment of these RNAs could be accomplished by exploiting a characterized mutation

of Tpt1 that could transfer a biotin group to 2′-PO4 sites. Tpt1 uses a 2-step mechanism to

remove 2′-PO4: (i) use NAD+ to attach an ADP-ribose molecule to the 2′-PO4, displacing the

nicotinamide group, and then (ii) use the neighboring 2″-OH to displace the RNA away from the

ADP-ribose. The mutant Tpt1-K69A-R71S has a compromised second step, meaning that the

mutant enzyme will quickly transfer an ADP-ribose group onto a 2′-PO4, but then slowly remove

the ADP-ribose group (Steiger et al. 2005). This difference in rates would generate a population

of marked RNAs toward the beginning of incubation with the mutant Tpt1 in vitro. If, instead of

NAD+, 6-Biotin-17-NAD+ were given as a substrate for Tpt1-K69A-R71S in vitro reaction, the

RNAs would be marked with an affinity-resin compatible tag. To pull down the biotinylated RNAs,

Tpt1 would need to be inactivated, and the labeled RNA would be affinity-purified from the rest

of the RNA. After binding to the resin, either the RNA could be reverse transcribed into a cDNA

library, or the RNA could be fragmented followed by the library preparation, or the RNA even

eluted with wild-type Tpt1, which would complete the reaction and release the RNAs with a

seamless 2′-OH at the ligation junction. Enrichment of RNAs via Tpt1-assisted affinity-

purification could identify which RNAs undergo repair, but the method may only yield limited

information about the specific site of repair.

126

An alternative approach, in light of the Xrn1-stall observation, would be to digest the RNA

up to sites of 2′-PO4 in vitro, and then to perform 5′-end capture and sequencing on the RNAs

that survive incubation with Xrn1. The benefit of this method would be that the site of repair

could be identified at single-nucleotide resolution, as is indicated by the rXrn1 digestion and

primer extension results on 2′-PO4 HAC1 3′-exon (Fig. 3.5). However, Xrn1 can be halted by

elements other than 2′-PO4, so the assay could be made specific to 2′-PO4 by performing paired

Tpt1± pre-treatments in vitro on the RNA samples to be analyzed by subsequent rXrn1

degradation and library preparation. Site-specific information would not only identify the RNAs

on which repair occurs but would also identify the sequence motif at which the repair occurs,

obtaining data that could aid in determining which possible RNAs are regulated in a concerted

manner based on primary or secondary structure. These classes of recognition motifs could then

be associated with RNA cleavage or repair programs that activate under specific circumstances,

ultimately establishing the paradigm of regulation by RNA breakage and reunion.

“Fungification” of metazoan cells to replace RtcB with Trl1 for marking repaired transcripts

with 2′-PO4.

The capture and sequencing of 2′-PO4 RNAs could be extended to organisms that naturally

use a different, mutually-exclusive mechanism of ligating RNA (i.e. RtcB/HPSC117). Recently,

Trl1 was used to replace RtcB in mouse embryonic stem cells (Unlu et al. 2018). However, the

Trpt1 gene (Tpt1 homolog), previously shown to be dispensable in mouse cells (Harding et al.

2008), became essential upon switching the ligase to Trl1, consistent with Trpt1 complementing

tpt1∆ in yeast (Hu et al. 2003) and with mammalian cells using RtcB (Popow et al. 2011), a ligase

that does not yield 2′-phosphorylated products. That fact means that, after “yeast-ifying” the

RNA ligase of mammals, a “pre-spliced” tRNA bypass mechanism could be used to bypass the

essential function of Trpt1, allowing Trpt1 to be deleted and permitting the cell to accumulate

127

2′-PO4 products of ligation. As described previously, these could then be captured and

sequenced to determine how widespread RNA repair is, and on what RNAs it typically acts.

128

REFERENCES

Abelson J. 1979. RNA processing and the intervening sequence problem. Annu Rev Biochem 48: 1035–1069.

Abelson J, Trotta CR, Li H. 1998. tRNA Splicing. J Biol Chem 1–5.

Achour A, Persson K, Harris RA, Sundbäck J, Sentman CL, Lindqvist Y, Schneider G, Kärre K. 1998. The Crystal Structure of H-2Dd MHC Class I Complexed with the HIV-1-Derived Peptide P18-I10 at 2.4 Å Resolution. Immunity 9: 199–208.

Aizenman E, Engelberg-Kulka H, Glaser G. 1996. An Escherichia coli chromosomal “addiction module” regulated by guanosine 3′, 5′-bispyrophosphate: a model for programmed bacterial cell death. PNAS 93: 6059–6063.

Amberg DC, Goldstein AL, Cole CN. 1992. Isolation and characterization of RAT1: an essential gene of Saccharomyces cerevisiae required for the efficient nucleocytoplasmic trafficking of mRNA. Genes & Development 6: 1173–1189.

Amitsur M, Levitz R, Kaufmann G. 1987. Bacteriophage T4 anticodon nuclease, polynucleotide kinase, and RNA ligase reprocess the host lysine tRNA. The EMBO Journal 6: 2499–2503.

Anderson JSJ, Parker R. 1998. The 3′ to 5′ degradation of yeast mRNAs is a general mechanism for mRNA turnover that requires the SKI2 DEVH box protein and 3′ to 5′ exonucleases of the exosome complex. The EMBO Journal 17: 1497–1506.

Andrade JM, Pobre V, Silva IJ, Domingues S, Arraiano CM. 2009. The role of 3“-5” exoribonucleases in RNA degradation. Prog Mol Biol Transl Sci 85: 187–229.

Apostol BL, Greer CL. 1991. Preferential binding of yeast tRNA ligase to pre-tRNA substrates. Nucleic Acids Research 19: 1853–1860.

Aragón T, van Anken E, Pincus D, Serafimova IM, Korennykh AV, Rubio CA, Walter P. 2009. Messenger RNA targeting to endoplasmic reticulum stress signalling sites. Nature 457: 736–740.

Araki Y, Takahashi S, Kobayashi T, Kajiho H, Hoshino S-I, Katada T. 2001. Ski7p G protein interacts with the exosome and the Ski complex for 3′‐to‐5′ mRNA decay in yeast. The

EMBO Journal 20: 4684–4693.

Baker KE, Parker R. 2004. Nonsense-mediated mRNA decay: terminating erroneous gene expression. Curr Opin Cell Biol 16: 293–299.

Becker T, Armache J-P, Jarasch A, Anger AM, Villa E, Sieber H, Motaal BA, Mielke T, Berninghausen O, Beckmann R. 2011. Structure of the no-go mRNA decay complex Dom34–Hbs1 bound to a stalled 80S ribosome. Nat Struct Mol Biol 18: 715–720.

129

Beelman CA, Stevens A, Caponigro G, LaGrandeur TE, Hatfield L, Fortner DM, Parker R. 1996. An essential component of the decapping enzyme required for normal rates of mRNA turnover. Nature 382: 642–646.

Belew AT, Advani VM, Dinman JD. 2011. Endogenous ribosomal frameshift signals operate as mRNA destabilizing elements through at least two molecular pathways in yeast. Nucleic

Acids Research 39: 2799–2808.

Belfort M, Weiner A. 1997. Another Bridge between Kingdoms: tRNA Splicing in Archaea and Eukaryotes. Cell 89: 1003–1006.

Bengtson MH, Joazeiro CAP. 2010. Role of a ribosome-associated E3 ubiquitin ligase in protein quality control. Nature 467: 470–473.

Bertolotti A, Zhang Y, Hendershot LM, Harding HP, Ron D. 2000. Dynamic interaction of BiP and ER stress transducers in the unfolded-protein response. Nat Cell Biol 2: 326–332.

Billy E, Wegierski T, Nasr F, Filipowicz W. 2000. Rcl1p, the yeast protein similar to the RNA 3'-phosphate cyclase, associates with U3 snoRNP and is required for 18S rRNA biogenesis. The EMBO Journal 19: 2115–2126.

Birkedal U, Christensen-Dalsgaard M, Krogh N, Sabarinathan R, Gorodkin J, Nielsen H. 2015. Profiling of ribose methylations in RNA by high-throughput sequencing. Angew Chem Int

Ed Engl 54: 451–455.

Boeck R, Tarun S, Rieger M, Deardorff JA, Müller-Auer S, Sachs AB. 1996. The yeast Pan2 protein is required for poly(A)-binding protein-stimulated poly(A)-nuclease activity. J Biol

Chem 271: 432–438.

Bowman CM, Dahlberg JE, Ikemura T, Konisky J, Nomura M. 1971a. Specific inactivation of 16S ribosomal RNA induced by colicin E3 in vivo. PNAS 68: 964–968.

Bowman CM, SIDIKARO J, Nomura M. 1971b. Specific inactivation of ribosomes by colicin E3 in vitro and mechanism of immunity in colicinogenic cells. Nature New Biol 234: 133–137.

Braglia P, Heindl K, Schleiffer A, Martinez J, Proudfoot NJ. 2010. Role of the RNA/DNA kinase Grc3 in transcription termination by RNA polymerase I. EMBO reports 11: 758–764.

Brown CE, Sachs AB. 1998. Poly(A) tail length control in Saccharomyces cerevisiae occurs by message-specific deadenylation. Molecular and Cellular Biology 18: 6548–6559.

Brown RS, Dewan JC, Klug A. 1985. Crystallographic and biochemical investigation of the lead(II)-catalyzed hydrolysis of yeast phenylalanine tRNA. Biochemistry 24: 4785–4801.

Bufardeci E, Fabbri S, Baldi MI, Mattoccia E, Tocchini-Valentini GP. 1993. In vitro genetic analysis of the structural features of the pre-tRNA required for determination of the 3′ splice site in the intron excision reaction. The EMBO Journal 12: 4697–4704.

130

Burkard KT, Butler JS. 2000. A nuclear 3´-5´exonuclease involved in mRNA degradation interacts with Poly(A) polymerase and the hnRNA protein Npl3p. Molecular and Cellular

Biology 20: 604–616.

Burroughs AM, Aravind L. 2016. RNA damage in biological conflicts and the diversity of responding RNA repair systems. Nucleic Acids Research gkw722–31.

Buzayan JM, Gerlach WL, Bruening G. 1986. Non-enzymatic cleavage and ligation of RNAs complementary to a plant virus satellite RNA. Nature 323: 349–353.

Canny MD, Jucker FM, Pardi A. 2007. Efficient ligation of the Schistosoma hammerhead ribozyme. Biochemistry 46: 3826–3834.

Cao D, Parker R. 2003. Computational modeling and experimental analysis of nonsense-mediated decay in yeast. Cell 113: 533–545.

Caponigro G, Parker R. 1995. Multiple functions for the poly(A)-binding protein in mRNA decapping and deadenylation in yeast. Genes & Development 9: 2421–2432.

Casagrande R, Stern P, Diehn M, Shamu C, Osario M, Zúñiga M, Brown PO, Ploegh H. 2000. Degradation of proteins from the ER of S. cerevisiae requires an intact unfolded protein response pathway. Molecular Cell 5: 729–735.

Celesnik H, Deana A, Belasco JG. 2007. Initiation of RNA decay in Escherichia coli by 5' pyrophosphate removal. Molecular Cell 27: 79–90.

Chakravarty AK, Shuman S. 2011. RNA 3“-phosphate cyclase (RtcA) catalyzes ligase-like adenylylation of DNA and RNA 5-”monophosphate ends. J Biol Chem 286: 4117–4122.

Chakravarty AK, Smith P, Jalan R, Shuman S. 2014. Structure, mechanism, and specificity of a eukaryal tRNA restriction enzyme involved in self-nonself discrimination. Cell Rep 7: 339–347.

Chakravarty AK, Subbotin R, Chait BT, Shuman S. 2012. RNA ligase RtcB splices 3'-phosphate and 5'-OH ends via covalent RtcB-(histidinyl)-GMP and polynucleotide-(3“)pp(5”)G intermediates. Proc Natl Acad Sci USA 109: 6072–6077.

Chan CM, Zhou C, Huang RH. 2009. Reconstituting Bacterial RNA Repair and Modification in Vitro. Science 326: 247–247.

Chan PP, Lowe TM. 2016. GtRNAdb 2.0: an expanded database of transfer RNA genes identified in complete and draft genomes. Nucleic Acids Research 44: D184–9.

Chan PP, Lowe TM. 2009. GtRNAdb: a database of transfer RNA genes detected in genomic sequence. Nucleic Acids Research 37: D93–7.

Chang JH, Jiao X, Chiba K, Oh C, Martin CE, Kiledjian M, Tong L. 2012. Dxo1 is a new type of eukaryotic enzyme with both decapping and 5′-3′ exoribonuclease activity. Nature

Publishing Group 19: 1011–1017.

131

Chapman EG, Moon SL, Wilusz J, Kieft JS. 2014. RNA structures that resist degradation by Xrn1 produce a pathogenic Dengue virus RNA. eLife 3: e01892.

Chapman RE, Walter P. 1997. Translational attenuation mediated by an mRNA intron. Curr Biol 7: 850–859.

Chawla A, Chakrabarti S, Ghosh G, Niwa M. 2011. Attenuation of yeast UPR is essential for survival and is mediated by IRE1 kinase. J Cell Biol 193: 41–50.

Chen L, Muhlrad D, Hauryliuk V, Cheng Z, Lim MK, Shyp V, Parker R, Song H. 2010. Structure of the Dom34–Hbs1 complex and implications for no-go decay. Nat Struct Mol Biol 17: 1233–1240.

Cheng Z-F, Zuo Y, Li Z, Rudd KE, Deutscher MP. 1998. The vacB gene required for virulence in Shigella flexneri and Escherichia coli encodes the exoribonuclease RNase R. J Biol Chem 273: 14077–14080.

Cherry PD, Peach SE, Hesselberth JR. 2018a. Multiple decay events target HAC1 mRNA during splicing to regulate the unfolded protein response. bioRχiv.

Cherry PD, White LK, York K, Hesselberth JR. 2018b. Genetic bypass of essential RNA repair enzymes in budding yeast. RNA 24: 313–323.

Clauss IM, Gravallese EM, Darling JM, Shapiro F, Glimcher MJ, Glimcher LH. 1993. In situ hybridization studies suggest a role for the basic region-leucine zipper protein hXBP-1 in exocrine gland and skeletal development during mouse embryogenesis. Dev Dyn 197: 146–156.

Clayton BLL, Popko B. 2016. Endoplasmic reticulum stress and the unfolded protein response in disorders of myelinating glia. Brain Res 1648: 594–602.

Cooper DA, Jha BK, Silverman RH, Hesselberth JR, Barton DJ. 2014. Ribonuclease L and metal-ion-independent endoribonuclease cleavage sites in host and viral RNAs. Nucleic

Acids Research 42: 5202–5216.

Cox JS, Chapman RE, Walter P. 1997. The unfolded protein response coordinates the production of endoplasmic reticulum protein and endoplasmic reticulum membrane. Mol

Biol Cell 8: 1805–1814.

Cox JS, Shamu CE, Walter P. 1993. Transcriptional induction of genes encoding endoplasmic reticulum resident proteins requires a transmembrane protein kinase. Cell 73: 1197–1206.

Cox JS, Walter P. 1996. A novel mechanism for regulating activity of a transcription factor that controls the unfolded protein response. Cell 87: 391–404.

Cramer WA, Lindeberg M, Taylor R. 1999. The best offense is a good defense. Nat Struct Biol 6: 295–297.

Credle JJ, Finer-Moore JS, Papa FR, Stroud RM, Walter P. 2005. On the mechanism of sensing unfolded protein in the endoplasmic reticulum. PNAS 102: 18773–18784.

132

Culver GM, McCraith SM, Consaul SA, Stanford DR, Phizicky EM. 1997. A 2′-Phosphotransferase Implicated in tRNA Splicing Is Essential in Saccharomyces cerevisiae. J Biol Chem 272: 13203–13210.

Culver GM, McCraith SM, Zillmann M, Kierzek R, Michaud N, LaReau RD, Turner DH, Phizicky EM. 1993. An NAD derivative produced during transfer RNA splicing: ADP-ribose 1″-2″ cyclic phosphate. Science 261: 206–208.

Dahlberg AE, Dahlberg JE. 1975. Binding of ribosomal protein S1 of Escherichia coli to the 3' end of 16S rRNA. PNAS 72: 2940–2944.

Das SR, Piccirilli JA. 2005. General acid catalysis by the hepatitis delta virus ribozyme. Nat

Chem Biol 1: 45–52.

de Vries H, Rüegsegger U, Hübner W, Friedlein A, Langen H, Keller W. 2000. Human pre-mRNA cleavage factor II(m) contains homologs of yeast proteins and bridges two other cleavage factors. The EMBO Journal 19: 5895–5904.

Decker CJ, Parker R. 1993. A turnover pathway for both stable and unstable mRNAs in yeast: evidence for a requirement for deadenylation. Genes & Development 7: 1632–1643.

Dever TE, Kinzy TG, Pavitt GD. 2016. Mechanism and Regulation of Protein Synthesis in Saccharomyces cerevisiae. Genetics 203: 65–107.

Dhungel N, Hopper AK. 2012. Beyond tRNA cleavage: novel essential function for yeast tRNA splicing endonuclease unrelated to tRNA processing. Genes & Development 26: 503–514.

Di Nicola Negri E, Fabbri S, Bufardeci E, Baldi MI, Gandini Attardi D, Mattoccia E, Tocchini-Valentini GP. 1997. The eucaryal tRNA splicing endonuclease recognizes a tripartite set of RNA elements. Cell 89: 859–866.

Di Santo R, Aboulhouda S, Weinberg DE. 2016. The fail-safe mechanism of post-transcriptional silencing of unspliced HAC1 mRNA. eLife 5.

Didychuk AL, Montemayor EJ, Carrocci TJ, DeLaitsch AT, Lucarelli SE, Westler WM, Brow DA, Hoskins AA, Butcher SE. 2017. Usb1 controls U6 snRNP assembly through evolutionarily divergent cyclic phosphodiesterase activities. Nat Commun 8: 497.

Dieckmann CL, Pape LK, Tzagoloff A. 1982. Identification and cloning of a yeast nuclear gene (CBP1) involved in expression of mitochondrial cytochrome b. PNAS 79: 1805–1809.

Dimitrova LN, Kuroha K, Tatematsu T, Inada T. 2009. Nascent peptide-dependent translation arrest leads to Not4p-mediated protein degradation by the proteasome. J Biol Chem 284: 10343–10352.

Doma MK, Parker R. 2006. Endonucleolytic cleavage of eukaryotic mRNAs with stalls in translation elongation. Nature 440: 561–564.

Duksin D, Mahoney WC. 1982. Relationship of the structure and biological activity of the natural homologues of tunicamycin. J Biol Chem 257: 3105–3109.

133

Durand S, Richard G, Bontems F, Uzan M. 2012. Bacteriophage T4 polynucleotide kinase triggers degradation of mRNAs. PNAS 109: 7073–7078.

DuRose JB, Tam AB, Niwa M. 2006. Intrinsic capacities of molecular sensors of the unfolded protein response to sense alternate forms of endoplasmic reticulum stress. Mol Biol Cell 17: 3095–3107.

El-Moghazy AN, Zhang N, Ismail T, Wu J, Butt A, Ahmed Khan S, Merlotti C, Cara Woodwark K, Gardner DC, Gaskell SJ, et al. 2000. Functional analysis of six novel ORFs on the left arm of chromosome XII in Saccharomyces cerevisiae reveals two essential genes, one of which is under cell-cycle control. Yeast 16: 277–288.

Emara MM, Ivanov P, Hickman T, Dawra N, Tisdale S, Kedersha N, Hu G-F, Anderson P. 2010. Angiogenin-induced tRNA-derived stress-induced RNAs promote stress-induced stress granule assembly. J Biol Chem 285: 10959–10968.

Engelberg-Kulka H, Amitai S, Kolodkin-Gal I, Hazan R. 2006. Bacterial programmed cell death and multicellular behavior in bacteria. PLoS Genet 2: e135.

Engelke DR, Gegenheimer P, Abelson J. 1985. Nucleolytic Processing of a tRNAArg-tRNAAsp Dimeric Precursor by a Homologous Component from Saccharomyces cerevisiae. J Biol

Chem 260: 1271–1279.

Fernández-Pevida A, Kressler D, La Cruz De J. 2015. Processing of preribosomal RNA in Saccharomyces cerevisiae. WIREs RNA 6: 191–209.

Fett JW, Strydom DJ, Lobb RR, Alderman EM, Bethune JL, Riordan JF, Vallee BL. 1985. Isolation and characterization of angiogenin, an angiogenic protein from human carcinoma cells. Biochemistry 24: 5480–5486.

Filipowicz W, Konarska M, Gross HJ, Shatkin AJ. 1983. RNA 3'-terminal phosphate cyclase activity and RNA ligation in HeLa cell extract. Nucleic Acids Research 11: 1405–1418.

Frischmeyer PA, van Hoof A, O'Donnell K, Guerrerio AL, Parker R, Dietz HC. 2002. An mRNA surveillance mechanism that eliminates transcripts lacking termination codons. Science 295: 2258–2261.

Fromm L, Falk S, Flemming D, Schuller JM, Thoms M, Conti E, Hurt E. 2017. Reconstitution of the complete pathway of ITS2 processing at the pre-ribosome. Nat Commun 8: 1787.

Fu H, Feng J, Liu Q, Sun F, Tie Y, Zhu J, Xing R, Sun Z, Zheng X. 2009. Stress induces tRNA cleavage by angiogenin in mammalian cells. FEBS Lett 583: 437–442.

Furuichi Y, LaFiandra A, Shatkin AJ. 1977. 5′-Terminal structure and mRNA stability. Nature 266: 235–239.

Gaba A, Jacobson A, Sachs MS. 2005. Ribosome occupancy of the yeast CPA1 upstream open reading frame termination codon modulates nonsense-mediated mRNA decay. Molecular Cell 20: 449–460.

134

Galli G, Hofstetter H, Birnstiel ML. 1981. Two conserved sequence blocks within eukaryotic tRNA genes are major promoter elements. Nature 294: 626–631.

Galperin MY, Koonin EV. 2004. “Conserved hypothetical” proteins: prioritization of targets for experimental study. Nucleic Acids Research 32: 5452–5463.

Gardner BM, Walter P. 2011. Unfolded proteins are Ire1-activating ligands that directly induce the unfolded protein response. Science 333: 1891–1894.

Gasse L, Flemming D, Hurt E. 2015. Coordinated Ribosomal ITS2 RNA Processing by the Las1 Complex Integrating Endonuclease, Polynucleotide Kinase, and Exonuclease Activities. Molecular Cell 60: 808–815.

Geerlings TH, Vos JC, Raué HA. 2000. The final step in the formation of 25S rRNA in Saccharomyces cerevisiae is performed by 5′→3′ exonucleases. RNA 6: 1698–1703.

Genschik P, Billy E, Swianiewicz M, Filipowicz W. 1997. The human RNA 3'-terminal phosphate cyclase is a member of a new family of proteins conserved in Eucarya, Bacteria and Archaea. The EMBO Journal 16: 2955–2967.

Genschik P, Drabikowski K, Filipowicz W. 1998. Characterization of the Escherichia coli RNA 3′-Terminal Phosphate Cyclase and Its σ54-Regulated Operon. J Biol Chem 273: 25516–25526.

Ghaemmaghami S, Huh W-K, Bower K, Howson RW, Belle A, Dephoure N, O'Shea EK, Weissman JS. 2003. Global analysis of protein expression in yeast. Nature 425: 737–741.

Giaever G, Chu AM, Ni L, Connelly C, Riles L, Véronneau S, Dow S, Lucau-Danila A, Anderson K, André B, et al. 2002. Functional profiling of the Saccharomyces cerevisiae genome. Nature 418: 387–391.

Gietz RD, Schiestl RH. 2007. High-efficiency yeast transformation using the LiAc/SS carrier DNA/PEG method. Nature Protocols 2: 31–34.

Gonzalez TN, Sidrauski C, Dörfler S, Walter P. 1999. Mechanism of non-spliceosomal mRNA splicing in the unfolded protein response pathway. The EMBO Journal 18: 3119–3132.

Good PD, Engelke DR. 1994. Yeast expression vectors using RNA polymerase III promoters. Gene 151: 209–214.

Goodman HM, Olson MV, Hall BD. 1977. Nucleotide sequence of a mutant eukaryotic gene: the yeast tyrosine-inserting ochre suppressor SUP4-o. PNAS 74: 5453–5457.

Greer CL, Peebles CL, Gegenheimer P, Abelson J. 1983. Mechanism of action of a yeast RNA ligase in tRNA splicing. Cell 32: 537–546.

Grosjean H, Szweykowska-Kulinska Z, Motorin Y. 1997. Intron-dependent enzymatic formation of modified nucleosides in eukaryotic tRNAs: a review. Biochimie 79: 293–302.

135

Guan Q, Zheng W, Tang S, Liu X, Zinkel RA, Tsui K-W, Yandell BS, Culbertson MR. 2006. Impact of nonsense-mediated mRNA decay on the global expression profile of budding yeast. PLoS Genet 2: e203.

Guydosh NR, Kimmig P, Walter P, Green R. 2017. Regulated Ire1-dependent mRNA decay requires no-go mRNA degradation to maintain endoplasmic reticulum homeostasis in S. pombe. eLife 6: e29216.

Hampel A, Cowan JA. 1997. A unique mechanism for RNA catalysis: the role of metal cofactors in hairpin ribozyme cleavage. Chem Biol 4: 513–517.

Hanada T, Weitzer S, Mair B, Bernreuther C, Wainger BJ, Ichida J, Hanada R, Orthofer M, Cronin SJ, Komnenovic V, et al. 2013. CLP1 links tRNA metabolism to progressive motor-neuron loss. Nature 495: 474–480.

Harding HP, Lackey JG, Hsu H-C, Zhang Y, Deng J, Xu R-M, Damha MJ, Ron D. 2008. An intact unfolded protein response in Trpt1 knockout mice reveals phylogenic divergence in pathways for RNA ligation. RNA 14: 225–232.

Harigaya Y, Parker R. 2012. Global analysis of mRNA decay intermediates in Saccharomyces cerevisiae. PNAS 109: 11764–11769.

He F, Peltz SW, Donahue JL, Rosbash M, Jacobson A. 1993. Stabilization and ribosome association of unspliced pre-mRNAs in a yeast upf1- mutant. PNAS 90: 7034–7038.

Heindl K, Martinez J. 2010. Nol9 is a novel polynucleotide 5'-kinase involved in ribosomal RNA processing. The EMBO Journal 29: 4161–4171.

Hollien J, Lin JH, Li H, Stevens N, Walter P, Weissman JS. 2009. Regulated Ire1-dependent decay of messenger RNAs in mammalian cells. J Cell Biol 186: 323–331.

Hollien J, Weissman JS. 2006. Decay of Endoplasmic Reticulum-Localized mRNAs During the Unfolded Protein Response. Science 313: 104–107.

Hooks KB, Griffiths-Jones S. 2011. Conserved RNA structures in the non-canonical Hac1/Xbp1 intron. RNA Biol 8: 552–556.

Hopper AK. 2013. Transfer RNA Post-Transcriptional Processing, Turnover, and Subcellular Dynamics in the Yeast Saccharomyces cerevisiae. Genetics 194: 43–67.

Hopper AK, Banks F, Evangelidis V. 1978. A yeast mutant which accumulates precursor tRNAs. Cell 14: 211–219.

Horikawa W, Endo K, Wada M, Ito K. 2016. Mutations in the G-domain of Ski7 cause specific dysfunction in non-stop decay. Nature Publishing Group 6: 29295.

Hsu CL, Stevens A. 1993. Yeast cells lacking 5′→3′ exoribonuclease 1 contain mRNA species that are poly(A) deficient and partially lack the 5′ cap structure. Molecular and Cellular

Biology 13: 4826–4835.

136

Hu QD, Lu H, Huo K, Ying K, Li J, Xie Y, Mao Y, Li YY. 2003. A human homolog of the yeast gene encoding tRNA 2'-phosphotransferase: cloning, characterization and complementation analysis. Cell Mol Life Sci 60: 1725–1732.

Ikeuchi K, Inada T. 2016. Ribosome-associated Asc1/RACK1 is required for endonucleolytic cleavage induced by stalled ribosome at the 3' end of nonstop mRNA. Nature Publishing

Group 6: 28234.

Inada T, Aiba H. 2005. Translation of aberrant mRNAs lacking a termination codon or with a shortened 3'-UTR is repressed after initiation in yeast. The EMBO Journal 24: 1584–1595.

Ivanov P, Emara MM, Villen J, Gygi SP, Anderson P. 2011. Angiogenin-induced tRNA fragments inhibit translation initiation. Molecular Cell 43: 613–623.

Jablonowski D, Zink S, Mehlgarten C, Daum G, Schaffrath R. 2006. tRNAGlu wobble uridine methylation by Trm9 identifies Elongator's key role for zymocin-induced cell death in yeast. Molecular Microbiology 59: 677–688.

James R, Kleanthous C, Moore GR. 1996. The biology of E colicins: paradigms and paradoxes. Microbiology (Reading, Engl) 142 ( Pt 7): 1569–1580.

Jang SK, Kräusslich HG, Nicklin MJ, Duke GM, Palmenberg AC, Wimmer E. 1988. A segment of the 5' nontranslated region of encephalomyocarditis virus RNA directs internal entry of ribosomes during in vitro translation. J Virol 62: 2636–2643.

Jenny A, Minvielle-Sebastia L, Preker PJ, Keller W. 1996. Sequence similarity between the 73-kilodalton protein of mammalian CPSF and a subunit of yeast polyadenylation factor I. Science 274: 1514–1517.

Johnson AW. 1997. Rat1p and Xrn1p are functionally interchangeable exoribonucleases that are restricted to and required in the nucleus and cytoplasm, respectively. Molecular and

Cellular Biology 17: 6122–6130.

Johnson AW, Kolodner RD. 1995. Synthetic lethality of sep1 (xrn1) ski2 and sep1 (xrn1) ski3 mutants of Saccharomyces cerevisiae is independent of killer virus and suggests a general role for these genes in translation control. Molecular and Cellular Biology 15: 2719–2727.

Johnston JA, Johnson ES, Waller PR, Varshavsky A. 1995. Methotrexate inhibits proteolysis of dihydrofolate reductase by the N-end rule pathway. J Biol Chem 270: 8172–8178.

Jones FD, Ryder SP, Strobel SA. 2001. An efficient ligation reaction promoted by a Varkud Satellite ribozyme with extended 5′- and 3′-termini. Nucleic Acids Research 29: 5115–5120.

Jones GM, Stalker J, Humphray S, West A, Cox T, Rogers J, Dunham I, Prelich G. 2008. A systematic library for comprehensive overexpression screens in Saccharomyces cerevisiae. Nature Methods 5: 239–241.

137

Jurkin J, Henkel T, Nielsen AF, Minnich M, Popow J, Kaufmann T, Heindl K, Hoffmann T, Busslinger M, Martinez J. 2014. The mammalian tRNA ligase complex mediates splicing of XBP1 mRNA and controls antibody secretion in plasma cells. The EMBO Journal.

Karaca E, Weitzer S, Pehlivan D, Shiraishi H, Gogakos T, Hanada T, Jhangiani SN, Wiszniewski W, Withers M, Campbell IM, et al. 2014. Human CLP1 Mutations Alter tRNA Biogenesis, Affecting Both Peripheral and Central Nervous System Function. Cell 157: 636–650.

Kaufmann G. 2000. Anticodon nucleases. Trends in Biochemical Sciences 25: 70–74.

Kawahara T, Yanagi H, Yura T, Mori K. 1998. Unconventional splicing of HAC1/ERN4 mRNA required for the unfolded protein response. Sequence-specific and non-sequential cleavage of the splice sites. J Biol Chem 273: 1802–1807.

Kebaara BW, Atkin AL. 2009. Long 3'-UTRs target wild-type mRNAs for nonsense-mediated mRNA decay in Saccharomyces cerevisiae. Nucleic Acids Research 37: 2771–2778.

Kelley PM, Schlesinger MJ. 1978. The effect of amino acid analogues and heat shock on gene expression in chicken embryo fibroblasts. Cell 15: 1277–1286.

Kimmig P, Diaz M, Zheng J, Williams CC, Lang A, Aragón T, Li H, Walter P. 2012. The unfolded protein response in fission yeast modulates stability of select mRNAs to maintain protein homeostasis. eLife 1: e00048.

Kirchner S, Ignatova Z. 2015. Emerging roles of tRNA in adaptive translation, signalling dynamics and disease. Nat Rev Genet 16: 98–112.

Kishi M, Pan YA, Crump JG, Sanes JR. 2005. Mammalian SAD kinases are required for neuronal polarization. Science 307: 929–932.

Klassen R, Meinhardt F. 2005. Induction of DNA damage and apoptosis in Saccharomyces cerevisiae by a yeast killer toxin. Cellular Microbiology 7: 393–401.

Klassen R, Paluszynski JP, Wemhoff S, Pfeiffer A, Fricke J, Meinhardt F. 2008. The primary target of the killer toxin from Pichia acaciae is tRNA(Gln). Molecular Microbiology 69: 681–697.

Klassen R, Teichert S, Meinhardt F. 2004. Novel yeast killer toxins provoke S-phase arrest and DNA damage checkpoint activation. Molecular Microbiology 53: 263–273.

Knapp G, Beckmann JS, Johnson PF, Fuhrman SA, Abelson J. 1978. Transcription and processing of intervening sequences in yeast tRNA genes. Cell 14: 221–236.

Knapp G, Ogden RC, Peebles CL, Abelson J. 1979. Splicing of yeast tRNA precursors: structure of the reaction intermediates. Cell 18: 37–45.

Kohno K, Normington K, Sambrook J, Gething MJ, Mori K. 1993. The promoter region of the yeast KAR2 (BiP) gene contains a regulatory domain that responds to the presence of unfolded proteins in the endoplasmic reticulum. Molecular and Cellular Biology 13: 877–890.

138

Konarska M, Filipowicz W, Domdey H, Gross HJ. 1981. Formation of a 2′-phosphomonoester, 3′,5′-phosphodiester linkage by a novel RNA ligase in wheat germ. Nature 293: 112–116.

Korennykh A, Walter P. 2012. Structural basis of the unfolded protein response. Annu Rev Cell

Dev Biol 28: 251–277.

Korennykh AV, Egea PF, Korostelev AA, Finer-Moore J, Stroud RM, Zhang C, Shokat KM, Walter P. 2011a. Cofactor-mediated conformational control in the bifunctional kinase/RNase Ire1. BMC Biol 9: 48.

Korennykh AV, Egea PF, Korostelev AA, Finer-Moore J, Zhang C, Shokat KM, Stroud RM, Walter P. 2009. The unfolded protein response signals through high-order assembly of Ire1. Nature 457: 687–693.

Korennykh AV, Korostelev AA, Egea PF, Finer-Moore J, Stroud RM, Zhang C, Shokat KM, Walter P. 2011b. Structural and functional basis for RNA cleavage by Ire1. BMC Biol 9: 47.

Kosmaczewski SG, Edwards TJ, Han SM, Eckwahl MJ, Meyer BI, Peach SE, Hesselberth JR, Wolin SL, Hammarlund M. 2014. The RtcB RNA ligase is an essential component of the metazoan unfolded protein response. EMBO reports.

Kosmaczewski SG, Han SM, Han B, Irving Meyer B, Baig HS, Athar W, Lin-Moore AT, Koelle MR, Hammarlund M. 2015. RNA ligation in neurons by RtcB inhibits axon regeneration. Proc Natl Acad Sci USA 112: 8451–8456.

Kozutsumi Y, Segal M, Normington K, Gething M-J, Sambrook J. 1988. The presence of malfolded proteins in the endoplasmic reticulum signals the induction of glucose-regulated proteins. Nature 332: 462–464.

Krause K, Lopes de Souza R, Roberts DGW, Dieckmann CL. 2004. The mitochondrial message-specific mRNA protectors Cbp1 and Pet309 are associated in a high-molecular weight complex. Mol Biol Cell 15: 2674–2683.

Lappe-Siefke C, Goebbels S, Gravel M, Nicksch E, Lee J, Braun PE, Griffiths IR, Nave K-A. 2003. Disruption of Cnp1 uncouples oligodendroglial functions in axonal support and myelination. Nat Genet 33: 366–374.

Larimer FW, Stevens A. 1990. Disruption of the gene XRN1, coding for a 5′→3′ exoribonuclease, restricts yeast cell growth. Gene 95: 85–90.

Laski FA, Fire AZ, RajBhandary UL, Sharp PA. 1983. Characterization of tRNA precursor splicing in mammalian extracts. J Biol Chem 258: 11974–11980.

Laughery MF, Hunter T, Brown A, Hoopes J, Ostbye T, Shumaker T, Wyrick JJ. 2015. New vectors for simple and streamlined CRISPR-Cas9 genome editing in Saccharomyces cerevisiae. Yeast 32: 711–720.

Lee AS. 1992. Mammalian stress response: induction of the glucose-regulated protein family. Curr Opin Cell Biol 4: 267–273.

139

Lee AS, Delegeane AM, Baker V, Chow PC. 1983. Transcriptional regulation of two genes specifically induced by glucose starvation in a hamster mutant fibroblast cell line. J Biol

Chem 258: 597–603.

Lee KPK, Dey M, Neculai D, Cao C, Dever TE, Sicheri F. 2008. Structure of the dual enzyme Ire1 reveals the basis for catalysis and regulation in nonconventional RNA splicing. Cell 132: 89–100.

Lee SR, Collins K. 2005. Starvation-induced cleavage of the tRNA anticodon loop in Tetrahymena thermophila. J Biol Chem 280: 42744–42749.

Levin JZ, Yassour M, Adiconis X, Nusbaum C, Thompson DA, Friedman N, Gnirke A, Regev A. 2010. Comprehensive comparative analysis of strand-specific RNA sequencing methods. Nature Methods 7: 709–715.

Levitz R, Chapman D, Amitsur M, Green R, Snyder L, Kaufmann G. 1990. The optional E. coli prr locus encodes a latent form of phage T4-induced anticodon nuclease. The EMBO

Journal 9: 1383–1389.

Li W, Okreglak V, Peschek J, Kimmig P, Zubradt M, Weissman JS, Walter P. 2018. Engineering ER-stress dependent non-conventional mRNA splicing. eLife 7.

Li Y, Luo J, Zhou H, Liao J-Y, Ma L-M, Chen Y-Q, Qu LH. 2008. Stress-induced tRNA-derived RNAs: a novel class of small RNAs in the primitive eukaryote Giardia lamblia. Nucleic Acids

Research 36: 6048–6055.

Lingner J, Kellermann J, Keller W. 1991. Cloning and expression of the essential gene for poly(A) polymerase from S. cerevisiae. Nature 354: 496–498.

Liu Q, Greimann JC, Lima CD. 2006. Reconstitution, activities, and structure of the eukaryotic RNA exosome. Cell 127: 1223–1237.

Lopes RRS, Silveira G de O, Eitler R, Vidal RS, Kessler A, Hinger S, Paris Z, Alfonzo JD, Polycarpo C. 2016. The essential function of the Trypanosoma brucei Trl1 homolog in procyclic cells is maturation of the intron-containing tRNA Tyr. RNA 22: 1190–1199.

Losson R, Lacroute F. 1979. Interference of nonsense mutations with eukaryotic messenger RNA stability. PNAS 76: 5134–5137.

Lowe TM, Eddy SR. 1997. tRNAscan-SE: a program for improved detection of transfer RNA genes in genomic sequence. Nucleic Acids Research 25: 955–964.

Lu J, Huang B, Esberg A, Johansson MJO, Byström AS. 2005. The Kluyveromyces lactis γ-toxin targets tRNA anticodons. RNA 11: 1648–1654.

Lu Y, Liang F-X, Wang X. 2014. A Synthetic Biology Approach Identifies the Mammalian UPR RNA Ligase RtcB. Molecular Cell 1–13.

140

MacIntosh GC, Bariola PA, Newbigin E, Green PJ. 2001. Characterization of Rny1, the Saccharomyces cerevisiae member of the T2 RNase family of RNases: Unexpected functions for ancient enzymes? PNAS 98: 1018–1023.

Mackie GA. 1998. Ribonuclease E is a 5'-end-dependent endonuclease. Nature 395: 720–723.

Martins A, Shuman S. 2005. An end-healing enzyme from Clostridium thermocellum with 5′ kinase, 2′,3′ phosphatase, and adenylyltransferase activities. RNA 11: 1271–1280.

Masaki H, Ogawa T. 2002. The modes of action of colicins E5 and D, and related cytotoxic tRNases. Biochimie 84: 433–438.

Masaki H, Ohta T. 1985. Colicin E3 and its immunity genes. J Mol Biol 182: 217–227.

McCraith SM, Phizicky EM. 1990. A highly specific phosphatase from Saccharomyces

cerevisiae implicated in tRNA splicing. Molecular and Cellular Biology 10: 1049–1055.

McCraith SM, Phizicky EM. 1991. An enzyme from Saccharomyces cerevisiae uses NAD+ to transfer the splice junction 2'-phosphate from ligated tRNA to an acceptor molecule. J Biol

Chem 266: 11986–11992.

Meaux S, van Hoof A. 2006. Yeast transcripts cleaved by an internal ribozyme provide new insight into the role of the cap and poly(A) tail in translation and mRNA decay. RNA 12: 1323–1337.

Meineke B, Kast A, Schwer B, Meinhardt F, Shuman S, Klassen R. 2012. A fungal anticodon nuclease ribotoxin exploits a secondary cleavage site to evade tRNA repair. RNA 18: 1716–1724.

Mian IS. 1997. Comparative sequence analysis of ribonucleases HII, III, II PH and D. Nucleic

Acids Research 25: 3187–3195.

Minvielle-Sebastia L, Preker PJ, Wiederkehr T, Strahm Y, Keller W. 1997. The major yeast poly(A)-binding protein is associated with cleavage factor IA and functions in premessenger RNA 3'-end formation. PNAS 94: 7897–7902.

Mitchell P, Petfalski E, Shevchenko A, Mann M, Tollervey D. 1997. The exosome: a conserved eukaryotic RNA processing complex containing multiple 3′→5′ exoribonucleases. Cell 91: 457–466.

Mitchell P, Petfalski E, Tollervey D. 1996. The 3' end of yeast 5.8S rRNA is generated by an exonuclease processing mechanism. Genes & Development 10: 502–513.

Mitchell P, Tollervey D. 2003. An NMD pathway in yeast involving accelerated deadenylation and exosome-mediated 3′→ 5′ degradation. Molecular Cell 11: 1405–1413.

Moore K, Hollien J. 2015. Ire1-mediated decay in mammalian cells relies on mRNA sequence, structure, and translational status. ed. G. Voeltz. Mol Biol Cell 26: 2873–2884.

141

Moore KA, Hollien J. 2012. The unfolded protein response in secretory cell function. Annu Rev

Genet 46: 165–183.

Morad I, Chapman-Shimshoni D, Amitsur M, Kaufmann G. 1993. Functional expression and properties of the tRNA(Lys)-specific core anticodon nuclease encoded by Escherichia coli prrC. J Biol Chem 268: 26842–26849.

Mori K, Kawahara T, Yoshida H, Yanagi H, Yura T. 1996. Signalling from endoplasmic reticulum to nucleus: transcription factor with a basic-leucine zipper motif is required for the unfolded protein-response pathway. Genes Cells 1: 803–817.

Mori K, Ma W, Gething MJ, Sambrook J. 1993. A transmembrane protein with a cdc2+/CDC28-related kinase activity is required for signaling from the ER to the nucleus. Cell 74: 743–756.

Mori K, Ogawa N, Kawahara T, Yanagi H, Yura T. 2000. mRNA splicing-mediated C-terminal replacement of transcription factor Hac1p is required for efficient activation of the unfolded protein response. PNAS 97: 4660–4665.

Mori K, Sant A, Kohno K, Normington K, Gething MJ, Sambrook JF. 1992. A 22 bp cis-acting element is necessary and sufficient for the induction of the yeast KAR2 (BiP) gene by unfolded proteins. The EMBO Journal 11: 2583–2593.

Mori T, Ogasawara C, Inada T, Englert M, Beier H, Takezawa M, Endo T, Yoshihisa T. 2010. Dual functions of yeast tRNA ligase in the unfolded protein response: unconventional cytoplasmic splicing of HAC1 pre-mRNA is not sufficient to release translational attenuation. Mol Biol Cell 21: 3722–3734.

Muhlrad D, Decker CJ, Parker R. 1994. Deadenylation of the unstable mRNA encoded by the yeast MFA2 gene leads to decapping followed by 5“-->3” digestion of the transcript. Genes & Development 8: 855–866.

Muhlrad D, Decker CJ, Parker R. 1995. Turnover mechanisms of the stable yeast PGK1 mRNA. Molecular and Cellular Biology 15: 2145–2156.

Muhlrad D, Parker R. 1999a. Aberrant mRNAs with extended 3' UTRs are substrates for rapid degradation by mRNA surveillance. RNA 5: 1299–1307.

Muhlrad D, Parker R. 1999b. Recognition of yeast mRNAs as “nonsense containing” leads to both inhibition of mRNA translation and mRNA degradation: implications for the control of mRNA decapping. ed. T.D. Fox. Mol Biol Cell 10: 3971–3978.

Munir A, Abdullahu L, Damha M, Shuman S. 2018a. Two-step mechanism and step-arrest mutants of Runella slithyformis NAD+-dependent tRNA 2'-phosphotransferase Tpt1. RNA rna.067165.118.

Munir A, Banerjee A, Shuman S. 2018b. NAD+-dependent synthesis of a 5′-phospho-ADP-ribosylated RNA/DNA cap by RNA 2′-phosphotransferase Tpt1. Nucleic Acids Research.

142

Munro S, Pelham HRB. 1986. An Hsp70-like protein in the ER: Identity with the 78 kd glucose-regulated protein and immunoglobulin heavy chain binding protein. Cell 46: 291–300.

Nandakumar J, Schwer B, Schaffrath R, Shuman S. 2008. RNA repair: an antidote to cytotoxic eukaryal RNA damage. Molecular Cell 31: 278–286.

Nariya H, Inouye M. 2008. MazF, an mRNA interferase, mediates programmed cell death during multicellular Myxococcus development. Cell 132: 55–66.

Navickas A, Chamois S, Saint-Fort R, Henri J, Torchet C, Benard L. 2018. A unique No-Go Decay cleavage in mRNA exit-tunnel of ribosome produces 5′-OH ends phosphorylated by Rlg1. bioRχiv.

Nesbitt S, Hegg LA, Fedor MJ. 1997. An unusual pH-independent and metal-ion-independent mechanism for hairpin ribozyme catalysis. Chem Biol 4: 619–630.

Nicholson RC, Williams DB, Moran LA. 1990. An essential member of the HSP70 gene family of Saccharomyces cerevisiae is homologous to immunoglobulin heavy chain binding protein. PNAS 87: 1159–1163.

Nikawa J-I, Akiyoshi M, Hirata S, Fukuda T. 1996. Saccharomyces cerevisiae IRE2/HAC1 is involved in IRE1-mediated KAR2 expression. Nucleic Acids Research 24: 4222–4226.

Nikawa J-I, Yamashita S. 1992. IRE1 encodes a putative protein kinase containing a membrane-spanning domain and is required for inositol phototrophy in Saccharomyces

cerevisiae. Molecular Microbiology 6: 1441–1446.

Niwa M, Patil CK, DeRisi J, Walter P. 2005. Genome-scale approaches for discovering novel nonconventional splicing substrates of the Ire1 nuclease. Genome Biol 6: R3.

Noble CG, Beuth B, Taylor IA. 2007. Structure of a nucleotide-bound Clp1-Pcf11 polyadenylation factor. Nucleic Acids Research 35: 87–99.

Nojima H, Leem S-H, Araki H, Sakai A, Nakashima N, Kanaoka Y, Ono Y. 1994. Hac1: A novel yeast bZIP protein binding to the CRE motif is a multicopy suppressor for cdc10 mutant of Schizosaccharomyces pombe. Nucleic Acids Research 22: 5279–5288.

Normington K, Kohno K, Kozutsumi Y, Gething MJ, Sambrook J. 1989. S. cerevisiae encodes an essential protein homologous in sequence and function to mammalian BiP. Cell 57: 1223–1236.

Nykänen A, Haley B, Zamore PD. 2001. ATP requirements and small interfering RNA structure in the RNA interference pathway. Cell 107: 309–321.

Ogawa N, Mori K. 2004. Autoregulation of the HAC1 gene is required for sustained activation of the yeast unfolded protein response. Genes Cells 9: 95–104.

Ogawa T, Inoue S, Yajima S, Hidaka M, Masaki H. 2006. Sequence-specific recognition of colicin E5, a tRNA-targeting ribonuclease. Nucleic Acids Research 34: 6065–6073.

143

Ogawa T, Tomita K, Ueda T, Watanabe K, Uozumi T, Masaki H. 1999. A cytotoxic ribonuclease targeting specific transfer RNA anticodons. Science 283: 2097–2100.

Okabayashi Y, Ohki A, Sakamoto C, Otsuki M. 1985. Relationship between the severity of diabetes mellitus and pancreatic exocrine dysfunction in rats. Diabetes Res Clin Pract 1: 21–30.

Olson R, Dulac C, Bjorkman PJ. 2006. MHC homologs in the nervous system — they haven’t lost their groove. Current Opinion in Neurobiology 16: 351–357.

Ozcan U, Cao Q, Yilmaz E, Lee A-H, Iwakoshi NN, Ozdelen E, Tuncman G, Görgün C, Glimcher LH, Hotamisligil GS. 2004. Endoplasmic reticulum stress links obesity, insulin action, and type 2 diabetes. Science 306: 457–461.

Parker R. 2012. RNA degradation in Saccharomyces cerevisae. Genetics 191: 671–702.

Passos DO, Doma MK, Shoemaker CJ, Muhlrad D, Green R, Weissman J, Hollien J, Parker R. 2009. Analysis of Dom34 and its function in no-go decay. Mol Biol Cell 20: 3025–3032.

Paushkin SV, Patel M, Furia BS, Peltz SW, Trotta CR. 2004. Identification of a human endonuclease complex reveals a link between tRNA splicing and pre-mRNA 3' end formation. Cell 117: 311–321.

Peach SE, York K, Hesselberth JR. 2015. Global analysis of RNA cleavage by 5′-hydroxyl RNA sequencing. Nucleic Acids Research 43: e108–e108.

Pedersen K, Zavialov AV, Pavlov MY, Elf J, Gerdes K, Ehrenberg M. 2003. The bacterial toxin RelE displays codon-specific cleavage of mRNAs in the ribosomal A site. Cell 112: 131–140.

Peebles CL, Gegenheimer P, Abelson J. 1983. Precise excision of intervening sequences from precursor tRNAs by a membrane-associated yeast endonuclease. Cell 32: 525–536.

Pelechano V, Wei W, Steinmetz LM. 2015. Widespread Co-translational RNA Decay Reveals Ribosome Dynamics. Cell 161: 1400–1412.

Peng W-T, Robinson MD, Mnaimneh S, Krogan NJ, Cagney G, Morris Q, Davierwala AP, Grigull J, Yang X, Zhang W, et al. 2003. A panoramic view of yeast noncoding RNA processing. Cell 113: 919–933.

Perry RP. 1962. The Cellular Sites of Synthesis of Ribosomal and 4S RNA. PNAS 48: 2179–2186.

Perry RP, Srinivasan PR, Kelley DE. 1964. Hybridization of Rapidly Labeled Nuclear Ribonucleic Acids. Science 145: 504–507.

Peschek J, Acosta-Alvear D, Mendez AS, Walter P. 2015. A conformational RNA zipper promotes intron ejection during non-conventional XBP1 mRNA splicing. EMBO reports.

144

Phizicky EM, Consaul SA, Nehrke KW, Abelson J. 1992. Yeast tRNA ligase mutants are nonviable and accumulate tRNA splicing intermediates. J Biol Chem.

Phizicky EM, Hopper AK. 2010. tRNA biology charges to the front. Genes & Development 24: 1832–1860.

Phizicky EM, Schwartz RC, Abelson J. 1986. Saccharomyces cerevisiae tRNA ligase. J Biol

Chem 261: 2978–2986.

Plotch SJ, Bouloy M, Krug RM. 1979. Transfer of 5′-terminal cap of globin mRNA to influenza viral complementary RNA during transcription in vitro. PNAS 76: 1618–1622.

Plotch SJ, Bouloy M, Ulmanen I, Krug RM. 1981. A unique cap(m7GpppXm)-dependent influenza virion endonuclease cleaves capped RNAs to generate the primers that initiate viral RNA transcription. Cell 23: 847–858.

Poole TL, Stevens A. 1995. Comparison of features of the RNase activity of 5“-exonuclease-1 and 5-”exonuclease-2 of Saccharomyces cerevisiae. Nucleic Acids Symp Ser 79–81.

Popow J, Englert M, Weitzer S, Schleiffer A, Mierzwa B, Mechtler K, Trowitzsch S, Will CL, Lührmann R, Söll D, et al. 2011. HSPC117 is the essential subunit of a human tRNA splicing ligase complex. Science 331: 760–764.

Pramanik J, Chen X, Kar G, Henriksson J, Gomes T, Park J-E, Natarajan K, Meyer KB, Miao Z, McKenzie ANJ, et al. 2018. Genome-wide analyses reveal the IRE1a-XBP1 pathway promotes T helper cell differentiation by resolving secretory stress and accelerating proliferation. Genome Med 10: 76.

Prody GA, Bakos JT, Buzayan JM, Schneider IR, Bruening G. 1986. Autolytic processing of dimeric plant virus satellite RNA. Science 231: 1577–1580.

Reimold AM, Etkin A, Clauss I, Perkins A, Friend DS, Zhang J, Horton HF, Scott A, Orkin SH, Byrne MC, et al. 2000. An essential role in liver development for transcription factor XBP-1. Genes & Development 14: 152–157.

Reimold AM, Iwakoshi NN, Manis J, Vallabhajosyula P, Szomolanyi-Tsuda E, Gravallese EM, Friend D, Grusby MJ, Alt F, Glimcher LH. 2001. Plasma cell differentiation requires the transcription factor XBP-1. Nature 412: 300–307.

Remus BS, Shuman S. 2014. Distinctive kinetics and substrate specificities of plant and fungal tRNA ligases. RNA 20: 462–473.

Rogers TB, Inesi G, Wade R, Lederer WJ. 1995. Use of thapsigargin to study Ca2+ homeostasis in cardiac cells. Biosci Rep 15: 341–349.

Ron D, Walter P. 2007. Signal integration in the endoplasmic reticulum unfolded protein response. Nat Rev Mol Cell Biol 8: 519–529.

Rose MD, Misra LM, Vogel JP. 1989. KAR2, a karyogamy gene, is the yeast homolog of the mammalian BiP/GRP78 gene. Cell 57: 1211–1221.

145

Roth A, Weinberg Z, Chen AGY, Kim PB, Ames TD, Breaker RR. 2014. A widespread self-cleaving ribozyme class is revealed by bioinformatics. Nat Chem Biol 10: 56–60.

Rubio C, Pincus D, Korennykh A, Schuck S, El-Samad H, Walter P. 2011. Homeostatic adaptation to endoplasmic reticulum stress depends on Ire1 kinase activity. J Cell Biol 193: 171–184.

Rüegsegger U, Leber JH, Walter P. 2001. Block of HAC1 mRNA translation by long-range base pairing is released by cytoplasmic splicing upon induction of the unfolded protein response. Cell 107: 103–114.

Ryan K, Calvo O, Manley JL. 2004. Evidence that polyadenylation factor CPSF-73 is the mRNA 3' processing endonuclease. RNA 10: 565–573.

Saikia M, Jobava R, Parisien M, Putnam A, Krokowski D, Gao X-H, Guan B-J, Yuan Y, Jankowsky E, Feng Z, et al. 2014. Angiogenin-cleaved tRNA halves interact with cytochrome c, protecting cells from apoptosis during osmotic stress. Molecular and

Cellular Biology 34: 2450–2463.

Saïda F, Uzan M, Bontems F. 2003. The phage T4 restriction endoribonuclease RegB: a cyclizing enzyme that requires two histidines to be fully active. Nucleic Acids Research 31: 2751–2758.

Sanson B, Hu R-M, Troitskaya E, Mathy N, Uzan M. 2000. Endoribonuclease RegB from bacteriophage T4 is necessary for the degradation of early but not middle or late mRNAs. J

Mol Biol 297: 1063–1074.

Sarkar D, Paira S, Das B. 2018. Nuclear mRNA degradation tunes the gain of the unfolded protein response in Saccharomyces cerevisiae. Nucleic Acids Research 46: 1139–1156.

Saville BJ, Collins RA. 1990. A site-specific self-cleavage reaction performed by a novel RNA in Neurospora mitochondria. Cell 61: 685–696.

Sawaya R, Schwer B, Shuman S. 2003. Genetic and Biochemical Analysis of the Functional Domains of Yeast tRNA Ligase. J Biol Chem 278: 43928–43938.

Sawaya R, Schwer B, Shuman S. 2005. Structure-function analysis of the yeast NAD+-dependent tRNA 2´-phosphotransferase Tpt1. RNA 11: 107–113.

Sayani S, Janis M, Lee CY, Toesca I, Chanfreau GF. 2008. Widespread impact of nonsense-mediated mRNA decay on the yeast intronome. Molecular Cell 31: 360–370.

Schaffer AE, Eggens VRC, Caglayan AO, Reuter MS, Scott E, Coufal NG, Silhavy JL, Xue Y, Kayserili H, Yasuno K, et al. 2014. CLP1 founder mutation links tRNA splicing and maturation to cerebellar development and neurodegeneration. Cell 157: 651–663.

Scherrer K. 2003. Historical review: the discovery of “giant” RNA and RNA processing: 40 years of enigma. Trends in Biochemical Sciences 28: 566–571.

146

Scherrer K, Darnell JE. 1962. Sedimentation characteristics of rapidly labelled RNA from HeLa cells. Biochemical and Biophysical Research Communications 7: 486–490.

Scherrer K, Latham H, Darnell JE. 1963. Demonstration of an unstable RNA and of a precursor to ribosomal RNA in HeLa cells. PNAS 49: 240–248.

Scheuner D, Song B, McEwen E, Liu C, Laybutt R, Gillespie P, Saunders T, Bonner-Weir S, Kaufman RJ. 2001. Translational control is required for the unfolded protein response and in vivo glucose homeostasis. Molecular Cell 7: 1165–1176.

Schillewaert S, Wacheul L, Lhomme F, Lafontaine DLJ. 2012. The evolutionarily conserved protein Las1 is required for pre-rRNA processing at both ends of ITS2. Molecular and

Cellular Biology 32: 430–444.

Schuck S, Prinz WA, Thorn KS, Voss C, Walter P. 2009. Membrane expansion alleviates endoplasmic reticulum stress independently of the unfolded protein response. J Cell Biol 187: 525–536.

Schutz K, Hesselberth JR, Fields S. 2010. Capture and sequence analysis of RNAs with terminal 2′,3′-cyclic phosphates. RNA 16: 621–631.

Schwartz RC, Greer CL, Gegenheimer P, Abelson J. 1983. Enzymatic mechanism of an RNA ligase from wheat germ. J Biol Chem 258: 8374–8383.

Schwer B, Aronova A, Ramirez A, Braun P, Shuman S. 2008. Mammalian 2′,3′ cyclic nucleotide phosphodiesterase (CNP) can function as a tRNA splicing enzyme in vivo. RNA 14: 204–210.

Schwer B, Sawaya R, Ho CK, Shuman S. 2004. Portability and fidelity of RNA-repair systems. PNAS 101: 2788–2793.

Shaheen HH, Hopper AK. 2005. Retrograde movement of tRNAs from the cytoplasm to the nucleus in Saccharomyces cerevisiae. PNAS 102: 11290–11295.

Sharmeen L, Kuo MY, Dinter-Gottlieb G, Taylor J. 1988. Antigenomic RNA of human hepatitis delta virus can undergo self-cleavage. J Virol 62: 2674–2679.

Shatkin AJ. 1976. Capping of eucaryotic mRNAs. Cell 9: 645–653.

Sheridan RM, Bentley DL. 2016. Selectable one-step PCR-mediated integration of a degron for rapid depletion of endogenous human proteins. BioTechniques 60: 69–74.

Sheth U, Parker R. 2006. Targeting of aberrant mRNAs to cytoplasmic processing bodies. Cell 125: 1095–1109.

Shimotohno K, Kodama Y, Hashimoto J, Miura KI. 1977. Importance of 5′ -terminal blocking structure to stabilize mRNA in eukaryotic protein synthesis. PNAS 74: 2734–2738.

Shine J, Dalgarno L. 1974. The 3'-terminal sequence of Escherichia coli 16S ribosomal RNA: complementarity to nonsense triplets and ribosome binding sites. PNAS 71: 1342–1346.

147

Shoemaker CJ, Green R. 2012. Translation drives mRNA quality control. Nat Struct Mol Biol 19: 594–601.

Sidrauski C, Cox JS, Walter P. 1996. tRNA ligase is required for regulated mRNA splicing in the unfolded protein response. Cell 87: 405–413.

Sidrauski C, Walter P. 1997. The transmembrane kinase Ire1p is a site-specific endonuclease that initiates mRNA splicing in the unfolded protein response. Cell 90: 1031–1039.

Silber R, Malathi VG, Hurwitz J. 1972. Purification and properties of bacteriophage T4-induced RNA ligase. PNAS 69: 3009–3013.

Soelaiman S, Jakes K, Wu N, Li C, Shoham M. 2001. Crystal structure of colicin E3: implications for cell entry and ribosome inactivation. Molecular Cell 8: 1053–1062.

Soukup GA, Breaker RR. 1999. Relationship between internucleotide linkage geometry and the stability of RNA. RNA 5: 1308–1325.

Spickler C, Stronge V, Mackie GA. 2001. Preferential cleavage of degradative intermediates of rpsT mRNA by the Escherichia coli RNA degradosome. J Bacteriol 183: 1106–1109.

Spinelli SL, Consaul SA, Phizicky EM. 1997. A conditional lethal yeast phosphotransferase (tpt1) mutant accumulates tRNAs with a 2'-phosphate and an undermodified base at the splice junction. RNA 3: 1388–1400.

Sprinkle TJ. 1989. 2´,3´-cyclic nucleotide 3´-phosphodiesterase, an oligodendrocyte-Schwann cell and myelin-associated enzyme of the nervous system. Crit Rev Neurobiol 4: 235–301.

Steiger MA, Jackman JE, Phizicky EM. 2005. Analysis of 2'-phosphotransferase (Tpt1p) from Saccharomyces cerevisiae: evidence for a conserved two-step reaction mechanism. RNA 11: 99–106.

Stevens A. 2001. 5′-Exoribonuclease 1: Xrn1. Meth Enzymol 342: 251–259.

Stevens A. 1980. Purification and characterization of a Saccharomyces cerevisiae exoribonuclease which yields 5'-mononucleotides by a 5‘ leads to 3’ mode of hydrolysis. J

Biol Chem 255: 3080–3085.

Swerdlow H, Guthrie C. 1984. Structure of intron-containing tRNA precursors. Analysis of solution conformation using chemical and enzymatic probes. J Biol Chem 259: 5197–5207.

Takano A, Endo T, Yoshihisa T. 2005. tRNA actively shuttles between the nucleus and cytosol in yeast. Science 309: 140–142.

Tam AB, Koong AC, Niwa M. 2014. Ire1 has distinct catalytic mechanisms for XBP1/HAC1 splicing and RIDD. Cell Rep 9: 850–858.

Tan Z, Zhang W, Sun J, Fu Z, Ke X, Zheng C, Zhang Y, Li P, Liu Y, Hu Q, et al. 2018. ZIKV infection activates the IRE1-XBP1 and ATF6 pathways of unfolded protein response in neural cells. J Neuroinflammation 15: 275.

148

Tanaka N, Chakravarty AK, Maughan B, Shuman S. 2011a. Novel Mechanism of RNA Repair by RtcB via Sequential 2′,3′-Cyclic Phosphodiesterase and 3′-Phosphate/5′-Hydroxyl Ligation Reactions. J Biol Chem 286: 43134–43143.

Tanaka N, Meineke B, Shuman S. 2011b. RtcB, a novel RNA ligase, can catalyze tRNA splicing and HAC1 mRNA splicing in vivo. J Biol Chem 286: 30253–30257.

Tanaka N, Shuman S. 2011. RtcB Is the RNA Ligase Component of an Escherichia coli RNA Repair Operon. J Biol Chem 286: 7727–7731.

Temmel H, Müller C, Sauert M, Vesper O, Reiss A, Popow J, Martinez J, Moll I. 2016. The RNA ligase RtcB reverses MazF-induced ribosome heterogeneity in Escherichia coli. Nucleic

Acids Research gkw1018–14.

Thompson DM, Lu C, Green PJ, Parker R. 2008. tRNA cleavage is a conserved response to oxidative stress in eukaryotes. RNA 14: 2095–2103.

Thompson DM, Parker R. 2009a. Stressing out over tRNA cleavage. Cell 138: 215–219.

Thompson DM, Parker R. 2009b. The RNase Rny1p cleaves tRNAs and promotes cell death during oxidative stress in Saccharomyces cerevisiae. J Cell Biol 185: 43–50.

Tkaczuk KL, Obarska A, Bujnicki JM. 2006. Molecular phylogenetics and comparative modeling of HEN1, a methyltransferase involved in plant microRNA biogenesis. BMC Evol

Biol 6: 6.

Tomita K, Ogawa T, Uozumi T, Watanabe K, Masaki H. 2000. A cytotoxic ribonuclease which specifically cleaves four isoaccepting arginine tRNAs at their anticodon loops. PNAS 97: 8278–8283.

Travers KJ, Patil CK, Wodicka L, Lockhart DJ, Weissman JS, Walter P. 2000. Functional and genomic analyses reveal an essential coordination between the unfolded protein response and ER-associated degradation. Cell 101: 249–258.

Trotta CR, Miao F, Arn EA, Stevens SW, Ho CK, Rauhut R, Abelson JN. 1997. The yeast tRNA splicing endonuclease: a tetrameric enzyme with two active site subunits homologous to the archaeal tRNA endonucleases. Cell 89: 849–858.

Tsuboi T, Yamazaki R, Nobuta R, Ikeuchi K, Makino S, Ohtaki A, Suzuki Y, Yoshihisa T, Trotta C, Inada T. 2015. The tRNA Splicing Endonuclease Complex Cleaves the Mitochondria-localized CBP1 mRNA. J Biol Chem 290: 16021–16030.

Tucker M, Valencia-Sanchez MA, Staples RR, Chen J, Denis CL, Parker R. 2001. The transcription factor associated Ccr4 and Caf1 proteins are components of the major cytoplasmic mRNA deadenylase in Saccharomyces cerevisiae. Cell 104: 377–386.

Unlu I, Lu Y, Wang X. 2018. The cyclic phosphodiesterase CNP and RNA cyclase RtcA fine-tune noncanonical XBP1 splicing during ER stress. J Biol Chem.

Usher DA. 1969. On the mechanism of ribonuclease action. PNAS 62: 661–667.

149

Uzan M, Favre R, Brody E. 1988. A nuclease that cuts specifically in the ribosome binding site of some T4 mRNAs. PNAS 85: 8895–8899.

Valenzuela P, Venegas A, Weinberg F, Bishop R, Rutter WJ. 1978. Structure of yeast phenylalanine-tRNA genes: an intervening DNA segment within the region coding for the tRNA. PNAS 75: 190–194.

van Hoof A, Frischmeyer PA, Dietz HC, Parker R. 2002. Exosome-mediated recognition and degradation of mRNAs lacking a termination codon. Science 295: 2262–2264.

van Hoof A, Lennertz P, Parker R. 2000. Three conserved members of the RNase D family have unique and overlapping functions in the processing of 5S, 5.8S, U4, U5, RNase MRP and RNase P RNAs in yeast. The EMBO Journal 19: 1357–1365.

van Tol H, Beier H. 1988. All human tRNATyrgenes contain introns as a prerequisite for pseudouridine biosynthesis in the anticodon. Nucleic Acids Research 16: 1951–1966.

Veldman GM, Klootwijk J, van Heerikhuizen H, Planta RJ. 1981. The nucleotide sequence of the intergenic region between the 5.8S and 26S rRNA genes of the yeast ribosomal RNA operon. Possible implications for the interaction between 5.8S and 26S rRNA and the processing of the primary transcript. Nucleic Acids Research 9: 4847–4862.

Vesper O, Amitai S, Belitsky M, Byrgazov K, Kaberdina AC, Engelberg-Kulka H, Moll I. 2011. Selective translation of leaderless mRNAs by specialized ribosomes generated by MazF in Escherichia coli. Cell 147: 147–157.

Vogel US, Thompson RJ. 1988. Molecular Structure, Localization, and Possible Functions of the Myelin-Associated Enzyme 2′,3′-Cyclic Nucleotide 3′-Phosphodiesterase. J

Neurochem 50: 1667–1677.

Walter P, Ron D. 2011. The unfolded protein response: from stress pathway to homeostatic regulation. Science 334: 1081–1086.

Wang L, Lewis MS, Johnson AW. 2005. Domain interactions within the Ski2/3/8 complex and between the Ski complex and Ski7p. RNA 11: 1291–1302.

Wang LK, Lima CD, Shuman S. 2002. Structure and mechanism of T4 polynucleotide kinase: an RNA repair enzyme. The EMBO Journal 21: 3873–3880.

Wang LK, Schwer B, Englert M, Beier H, Shuman S. 2006. Structure-function analysis of the kinase-CPD domain of yeast tRNA ligase (Trl1) and requirements for complementation of tRNA splicing by a plant Trl1 homolog. Nucleic Acids Research 34: 517–527.

Wang LK, Shuman S. 2005. Structure-function analysis of yeast tRNA ligase. RNA 11: 966–975.

Wang M, Ye R, Barron E, Baumeister P, Mao C, Luo S, Fu Y, Luo B, Dubeau L, Hinton DR, et al. 2010. Essential role of the unfolded protein response regulator GRP78/BiP in protection from neuronal apoptosis. Cell Death and Differentiation 17: 488–498.

150

Wang P, Selvadurai K, Huang RH. 2015. Reconstitution and structure of a bacterial Pnkp1-Rnl-Hen1 RNA repair complex. Nat Commun 6: 6876.

Weitzer S, Martinez J. 2007. The human RNA kinase hClp1 is active on 3' transfer RNA exons and short interfering RNAs. Nature 447: 222–226.

Welch EM, Jacobson A. 1999. An internal open reading frame triggers nonsense-mediated decay of the yeast SPT10 mRNA. The EMBO Journal 18: 6134–6145.

Whipple JM, Lane EA, Chernyakov I, D'Silva S, Phizicky EM. 2011. The yeast rapid tRNA decay pathway primarily monitors the structural integrity of the acceptor and T-stems of mature tRNA. Genes & Development 25: 1173–1184.

Wilson MA, Meaux S, van Hoof A. 2007. A genomic screen in yeast reveals novel aspects of nonstop mRNA metabolism. Genetics 177: 773–784.

Winey M, Mendenhall MD, Cummins CM, Culbertson MR, Knapp G. 1986. Splicing of a yeast proline tRNA containing a novel suppressor mutation in the anticodon stem. J Mol Biol 192: 49–63.

Winkler WC, Nahvi A, Roth A, Collins JA, Breaker RR. 2004. Control of gene expression by a natural metabolite-responsive ribozyme. Nature 428: 281–286.

Woolford JL, Baserga SJ. 2013. Ribosome Biogenesis in the Yeast Saccharomyces cerevisiae. Genetics 195: 643–681.

Worsham PL, Bolen PL. 1990. Killer toxin production in Pichia acaciae is associated with linear DNA plasmids. Curr Genet 18: 77–80.

Wu J, Bao A, Chatterjee K, Wan Y, Hopper AK. 2015. Genome-wide screen uncovers novel pathways for tRNA processing and nuclear-cytoplasmic dynamics. Genes & Development 29: 2633–2644.

Wu J, Hopper AK. 2014. Healing for destruction: tRNA intron degradation in yeast is a two-step cytoplasmic process catalyzed by tRNA ligase Rlg1 and 5′-to-3′ exonuclease Xrn1. Genes

& Development 28: 1556–1561.

Wu P, Brockenbrough JS, Paddy MR, Aris JP. 1998. NCL1, a novel gene for a non-essential nuclear protein in Saccharomyces cerevisiae. Gene 220: 109–117.

Xu Q, Teplow D, Lee TD, Abelson J. 1990. Domain structure in yeast tRNA ligase. Biochemistry 29: 6132–6138.

Xue Y, Bai X, Lee I, Kallstrom G, Ho J, Brown J, Stevens A, Johnson AW. 2000. Saccharomyces cerevisiae RAI1 (YGL246c) is homologous to human DOM3Z and encodes a protein that binds the nuclear exoribonuclease Rat1p. Molecular and Cellular Biology 20: 4006–4015.

Yamasaki S, Ivanov P, Hu G-F, Anderson P. 2009. Angiogenin cleaves tRNA and promotes stress-induced translational repression. J Cell Biol 185: 35–42.

151

Yang W. 2011. Nucleases: diversity of structure, function and mechanism. Quart Rev Biophys 44: 1–93.

Yoshida H, Matsui T, Yamamoto A, Okada T, Mori K. 2001. XBP1 mRNA is induced by ATF6 and spliced by IRE1 in response to ER stress to produce a highly active transcription factor. Cell 107: 881–891.

Yoshihisa T. 2014. Handling tRNA introns, archaeal way and eukaryotic way. Front Genet 5: 213.

Yoshihisa T, Ohshima C, Yunoki-Esaki K, Endo T. 2007. Cytoplasmic splicing of tRNA in Saccharomyces cerevisiae. Genes to Cells 12: 285–297.

Yoshihisa T, Yunoki-Esaki K, Ohshima C, Tanaka N, Endo T. 2003. Possibility of cytoplasmic pre-tRNA splicing: the yeast tRNA splicing endonuclease mainly localizes on the mitochondria. Mol Biol Cell 14: 3266–3279.

Young KJ, Gill F, Grasby JA. 1997. Metal ions play a passive role in the hairpin ribozyme catalysed reaction. Nucleic Acids Research 25: 3760–3766.

Zhang Y, Zhang J, Hara H, Kato I, Inouye M. 2005. Insights into the mRNA cleavage mechanism by MazF, an mRNA interferase. J Biol Chem 280: 3143–3150.

Zhou J, Liu CY, Back SH, Clark RL, Peisach D, Xu Z, Kaufman RJ. 2006. The crystal structure of human IRE1 luminal domain reveals a conserved dimerization interface required for activation of the unfolded protein response. PNAS 103: 14343–14348.

Zillmann M, Gorovsky MA, Phizicky EM. 1992. HeLa cells contain a 2'-phosphate-specific phosphotransferase similar to a yeast enzyme implicated in tRNA splicing. J Biol Chem.

Zofallova L, Guo Y, Gupta R. 2000. Junction phosphate is derived from the precursor in the tRNA spliced by the archaeon Haloferax volcanii cell extract. RNA 6: 1019–1030.

Abelson J. 1979. RNA processing and the intervening sequence problem. Annu Rev Biochem 48: 1035–1069.

Abelson J, Trotta CR, Li H. 1998. tRNA Splicing. J Biol Chem 1–5.

Achour A, Persson K, Harris RA, Sundbäck J, Sentman CL, Lindqvist Y, Schneider G, Kärre K. 1998. The Crystal Structure of H-2Dd MHC Class I Complexed with the HIV-1-Derived Peptide P18-I10 at 2.4 Å Resolution. Immunity 9: 199–208.

Aizenman E, Engelberg-Kulka H, Glaser G. 1996. An Escherichia coli chromosomal “addiction module” regulated by guanosine 3′, 5′-bispyrophosphate: a model for programmed bacterial cell death. PNAS 93: 6059–6063.

152

Amberg DC, Goldstein AL, Cole CN. 1992. Isolation and characterization of RAT1: an essential gene of Saccharomyces cerevisiae required for the efficient nucleocytoplasmic trafficking of mRNA. Genes & Development 6: 1173–1189.

Amitsur M, Levitz R, Kaufmann G. 1987. Bacteriophage T4 anticodon nuclease, polynucleotide kinase, and RNA ligase reprocess the host lysine tRNA. The EMBO Journal 6: 2499–2503.

Anderson JSJ, Parker R. 1998. The 3′ to 5′ degradation of yeast mRNAs is a general mechanism for mRNA turnover that requires the SKI2 DEVH box protein and 3′ to 5′ exonucleases of the exosome complex. The EMBO Journal 17: 1497–1506.

Andrade JM, Pobre V, Silva IJ, Domingues S, Arraiano CM. 2009. The role of 3“-5” exoribonucleases in RNA degradation. Prog Mol Biol Transl Sci 85: 187–229.

Apostol BL, Greer CL. 1991. Preferential binding of yeast tRNA ligase to pre-tRNA substrates. Nucleic Acids Research 19: 1853–1860.

Aragón T, van Anken E, Pincus D, Serafimova IM, Korennykh AV, Rubio CA, Walter P. 2009. Messenger RNA targeting to endoplasmic reticulum stress signalling sites. Nature 457: 736–740.

Araki Y, Takahashi S, Kobayashi T, Kajiho H, Hoshino S-I, Katada T. 2001. Ski7p G protein interacts with the exosome and the Ski complex for 3′‐to‐5′ mRNA decay in yeast. The

EMBO Journal 20: 4684–4693.

Baker KE, Parker R. 2004. Nonsense-mediated mRNA decay: terminating erroneous gene expression. Curr Opin Cell Biol 16: 293–299.

Becker T, Armache J-P, Jarasch A, Anger AM, Villa E, Sieber H, Motaal BA, Mielke T, Berninghausen O, Beckmann R. 2011. Structure of the no-go mRNA decay complex Dom34–Hbs1 bound to a stalled 80S ribosome. Nat Struct Mol Biol 18: 715–720.

Beelman CA, Stevens A, Caponigro G, LaGrandeur TE, Hatfield L, Fortner DM, Parker R. 1996. An essential component of the decapping enzyme required for normal rates of mRNA turnover. Nature 382: 642–646.

Belasco JG. 2010. All things must pass: contrasts and commonalities in eukaryotic and bacterial mRNA decay. Nat Rev Mol Cell Biol 11: 467–478.

Belew AT, Advani VM, Dinman JD. 2011. Endogenous ribosomal frameshift signals operate as mRNA destabilizing elements through at least two molecular pathways in yeast. Nucleic

Acids Research 39: 2799–2808.

Belfort M, Weiner A. 1997. Another Bridge between Kingdoms: tRNA Splicing in Archaea and Eukaryotes. Cell 89: 1003–1006.

Bengtson MH, Joazeiro CAP. 2010. Role of a ribosome-associated E3 ubiquitin ligase in protein quality control. Nature 467: 470–473.

153

Bertolotti A, Zhang Y, Hendershot LM, Harding HP, Ron D. 2000. Dynamic interaction of BiP and ER stress transducers in the unfolded-protein response. Nat Cell Biol 2: 326–332.

Billy E, Wegierski T, Nasr F, Filipowicz W. 2000. Rcl1p, the yeast protein similar to the RNA 3'-phosphate cyclase, associates with U3 snoRNP and is required for 18S rRNA biogenesis. The EMBO Journal 19: 2115–2126.

Birkedal U, Christensen-Dalsgaard M, Krogh N, Sabarinathan R, Gorodkin J, Nielsen H. 2015. Profiling of ribose methylations in RNA by high-throughput sequencing. Angew Chem Int

Ed Engl 54: 451–455.

Boeck R, Tarun S, Rieger M, Deardorff JA, Müller-Auer S, Sachs AB. 1996. The yeast Pan2 protein is required for poly(A)-binding protein-stimulated poly(A)-nuclease activity. J Biol

Chem 271: 432–438.

Bowman CM, Dahlberg JE, Ikemura T, Konisky J, Nomura M. 1971a. Specific inactivation of 16S ribosomal RNA induced by colicin E3 in vivo. PNAS 68: 964–968.

Bowman CM, SIDIKARO J, Nomura M. 1971b. Specific inactivation of ribosomes by colicin E3 in vitro and mechanism of immunity in colicinogenic cells. Nature New Biol 234: 133–137.

Braglia P, Heindl K, Schleiffer A, Martinez J, Proudfoot NJ. 2010. Role of the RNA/DNA kinase Grc3 in transcription termination by RNA polymerase I. EMBO reports 11: 758–764.

Brown CE, Sachs AB. 1998. Poly(A) tail length control in Saccharomyces cerevisiae occurs by message-specific deadenylation. Molecular and Cellular Biology 18: 6548–6559.

Brown RS, Dewan JC, Klug A. 1985. Crystallographic and biochemical investigation of the lead(II)-catalyzed hydrolysis of yeast phenylalanine tRNA. Biochemistry 24: 4785–4801.

Bufardeci E, Fabbri S, Baldi MI, Mattoccia E, Tocchini-Valentini GP. 1993. In vitro genetic analysis of the structural features of the pre-tRNA required for determination of the 3′ splice site in the intron excision reaction. The EMBO Journal 12: 4697–4704.

Burkard KT, Butler JS. 2000. A nuclear 3´-5´exonuclease involved in mRNA degradation interacts with Poly(A) polymerase and the hnRNA protein Npl3p. Molecular and Cellular

Biology 20: 604–616.

Burroughs AM, Aravind L. 2016. RNA damage in biological conflicts and the diversity of responding RNA repair systems. Nucleic Acids Research gkw722–31.

Buzayan JM, Gerlach WL, Bruening G. 1986. Non-enzymatic cleavage and ligation of RNAs complementary to a plant virus satellite RNA. Nature 323: 349–353.

Canny MD, Jucker FM, Pardi A. 2007. Efficient ligation of the Schistosoma hammerhead ribozyme. Biochemistry 46: 3826–3834.

Cao D, Parker R. 2003. Computational modeling and experimental analysis of nonsense-mediated decay in yeast. Cell 113: 533–545.

154

Caponigro G, Parker R. 1995. Multiple functions for the poly(A)-binding protein in mRNA decapping and deadenylation in yeast. Genes & Development 9: 2421–2432.

Casagrande R, Stern P, Diehn M, Shamu C, Osario M, Zúñiga M, Brown PO, Ploegh H. 2000. Degradation of proteins from the ER of S. cerevisiae requires an intact unfolded protein response pathway. Molecular Cell 5: 729–735.

Celesnik H, Deana A, Belasco JG. 2007. Initiation of RNA decay in Escherichia coli by 5' pyrophosphate removal. Molecular Cell 27: 79–90.

Chakravarty AK, Shuman S. 2011. RNA 3“-phosphate cyclase (RtcA) catalyzes ligase-like adenylylation of DNA and RNA 5-”monophosphate ends. J Biol Chem 286: 4117–4122.

Chakravarty AK, Smith P, Jalan R, Shuman S. 2014. Structure, mechanism, and specificity of a eukaryal tRNA restriction enzyme involved in self-nonself discrimination. Cell Rep 7: 339–347.

Chakravarty AK, Subbotin R, Chait BT, Shuman S. 2012. RNA ligase RtcB splices 3'-phosphate and 5'-OH ends via covalent RtcB-(histidinyl)-GMP and polynucleotide-(3“)pp(5”)G intermediates. Proc Natl Acad Sci USA 109: 6072–6077.

Chan CM, Zhou C, Huang RH. 2009. Reconstituting Bacterial RNA Repair and Modification in Vitro. Science 326: 247–247.

Chan PP, Lowe TM. 2016. GtRNAdb 2.0: an expanded database of transfer RNA genes identified in complete and draft genomes. Nucleic Acids Research 44: D184–9.

Chan PP, Lowe TM. 2009. GtRNAdb: a database of transfer RNA genes detected in genomic sequence. Nucleic Acids Research 37: D93–7.

Chang JH, Jiao X, Chiba K, Oh C, Martin CE, Kiledjian M, Tong L. 2012. Dxo1 is a new type of eukaryotic enzyme with both decapping and 5′-3′ exoribonuclease activity. Nature

Publishing Group 19: 1011–1017.

Chapman EG, Moon SL, Wilusz J, Kieft JS. 2014. RNA structures that resist degradation by Xrn1 produce a pathogenic Dengue virus RNA. eLife 3: e01892.

Chapman RE, Walter P. 1997. Translational attenuation mediated by an mRNA intron. Curr Biol 7: 850–859.

Chawla A, Chakrabarti S, Ghosh G, Niwa M. 2011. Attenuation of yeast UPR is essential for survival and is mediated by IRE1 kinase. J Cell Biol 193: 41–50.

Chen L, Muhlrad D, Hauryliuk V, Cheng Z, Lim MK, Shyp V, Parker R, Song H. 2010. Structure of the Dom34–Hbs1 complex and implications for no-go decay. Nat Struct Mol Biol 17: 1233–1240.

Cheng Z-F, Zuo Y, Li Z, Rudd KE, Deutscher MP. 1998. The vacB gene required for virulence in Shigella flexneri and Escherichia coli encodes the exoribonuclease RNase R. J Biol Chem 273: 14077–14080.

155

Cherry PD, Peach SE, Hesselberth JR. 2019. Multiple decay events target HAC1 mRNA during splicing to regulate the unfolded protein response. eLife 8. https://elifesciences.org/articles/42262.

Cherry PD, White LK, York K, Hesselberth JR. 2018. Genetic bypass of essential RNA repair enzymes in budding yeast. RNA 24: 313–323.

Clauss IM, Gravallese EM, Darling JM, Shapiro F, Glimcher MJ, Glimcher LH. 1993. In situ hybridization studies suggest a role for the basic region-leucine zipper protein hXBP-1 in exocrine gland and skeletal development during mouse embryogenesis. Dev Dyn 197: 146–156.

Clayton BLL, Popko B. 2016. Endoplasmic reticulum stress and the unfolded protein response in disorders of myelinating glia. Brain Res 1648: 594–602.

Cooper DA, Jha BK, Silverman RH, Hesselberth JR, Barton DJ. 2014. Ribonuclease L and metal-ion-independent endoribonuclease cleavage sites in host and viral RNAs. Nucleic

Acids Research 42: 5202–5216.

Cox JS, Chapman RE, Walter P. 1997. The unfolded protein response coordinates the production of endoplasmic reticulum protein and endoplasmic reticulum membrane. Mol

Biol Cell 8: 1805–1814.

Cox JS, Shamu CE, Walter P. 1993. Transcriptional induction of genes encoding endoplasmic reticulum resident proteins requires a transmembrane protein kinase. Cell 73: 1197–1206.

Cox JS, Walter P. 1996. A novel mechanism for regulating activity of a transcription factor that controls the unfolded protein response. Cell 87: 391–404.

Cramer WA, Lindeberg M, Taylor R. 1999. The best offense is a good defense. Nat Struct Biol 6: 295–297.

Credle JJ, Finer-Moore JS, Papa FR, Stroud RM, Walter P. 2005. On the mechanism of sensing unfolded protein in the endoplasmic reticulum. PNAS 102: 18773–18784.

Culver GM, McCraith SM, Consaul SA, Stanford DR, Phizicky EM. 1997. A 2′-Phosphotransferase Implicated in tRNA Splicing Is Essential in Saccharomyces cerevisiae. J Biol Chem 272: 13203–13210.

Culver GM, McCraith SM, Zillmann M, Kierzek R, Michaud N, LaReau RD, Turner DH, Phizicky EM. 1993. An NAD derivative produced during transfer RNA splicing: ADP-ribose 1″-2″ cyclic phosphate. Science 261: 206–208.

Dahlberg AE, Dahlberg JE. 1975. Binding of ribosomal protein S1 of Escherichia coli to the 3' end of 16S rRNA. PNAS 72: 2940–2944.

Das SR, Piccirilli JA. 2005. General acid catalysis by the hepatitis delta virus ribozyme. Nat

Chem Biol 1: 45–52.

156

de Vries H, Rüegsegger U, Hübner W, Friedlein A, Langen H, Keller W. 2000. Human pre-mRNA cleavage factor II(m) contains homologs of yeast proteins and bridges two other cleavage factors. The EMBO Journal 19: 5895–5904.

Decker CJ, Parker R. 1993. A turnover pathway for both stable and unstable mRNAs in yeast: evidence for a requirement for deadenylation. Genes & Development 7: 1632–1643.

Dever TE, Kinzy TG, Pavitt GD. 2016. Mechanism and Regulation of Protein Synthesis in Saccharomyces cerevisiae. Genetics 203: 65–107.

Dhungel N, Hopper AK. 2012. Beyond tRNA cleavage: novel essential function for yeast tRNA splicing endonuclease unrelated to tRNA processing. Genes & Development 26: 503–514.

Di Nicola Negri E, Fabbri S, Bufardeci E, Baldi MI, Gandini Attardi D, Mattoccia E, Tocchini-Valentini GP. 1997. The eucaryal tRNA splicing endonuclease recognizes a tripartite set of RNA elements. Cell 89: 859–866.

Di Santo R, Aboulhouda S, Weinberg DE. 2016. The fail-safe mechanism of post-transcriptional silencing of unspliced HAC1 mRNA. eLife 5.

Didychuk AL, Montemayor EJ, Carrocci TJ, DeLaitsch AT, Lucarelli SE, Westler WM, Brow DA, Hoskins AA, Butcher SE. 2017. Usb1 controls U6 snRNP assembly through evolutionarily divergent cyclic phosphodiesterase activities. Nat Commun 8: 497.

Dieckmann CL, Pape LK, Tzagoloff A. 1982. Identification and cloning of a yeast nuclear gene (CBP1) involved in expression of mitochondrial cytochrome b. PNAS 79: 1805–1809.

Dimitrova LN, Kuroha K, Tatematsu T, Inada T. 2009. Nascent peptide-dependent translation arrest leads to Not4p-mediated protein degradation by the proteasome. J Biol Chem 284: 10343–10352.

Doma MK, Parker R. 2006. Endonucleolytic cleavage of eukaryotic mRNAs with stalls in translation elongation. Nature 440: 561–564.

Duksin D, Mahoney WC. 1982. Relationship of the structure and biological activity of the natural homologues of tunicamycin. J Biol Chem 257: 3105–3109.

Durand S, Richard G, Bontems F, Uzan M. 2012. Bacteriophage T4 polynucleotide kinase triggers degradation of mRNAs. PNAS 109: 7073–7078.

DuRose JB, Tam AB, Niwa M. 2006. Intrinsic capacities of molecular sensors of the unfolded protein response to sense alternate forms of endoplasmic reticulum stress. Mol Biol Cell 17: 3095–3107.

El-Moghazy AN, Zhang N, Ismail T, Wu J, Butt A, Ahmed Khan S, Merlotti C, Cara Woodwark K, Gardner DC, Gaskell SJ, et al. 2000. Functional analysis of six novel ORFs on the left arm of chromosome XII in Saccharomyces cerevisiae reveals two essential genes, one of which is under cell-cycle control. Yeast 16: 277–288.

157

Emara MM, Ivanov P, Hickman T, Dawra N, Tisdale S, Kedersha N, Hu G-F, Anderson P. 2010. Angiogenin-induced tRNA-derived stress-induced RNAs promote stress-induced stress granule assembly. J Biol Chem 285: 10959–10968.

Engelberg-Kulka H, Amitai S, Kolodkin-Gal I, Hazan R. 2006. Bacterial programmed cell death and multicellular behavior in bacteria. PLoS Genet 2: e135.

Engelke DR, Gegenheimer P, Abelson J. 1985. Nucleolytic Processing of a tRNAArg-tRNAAsp Dimeric Precursor by a Homologous Component from Saccharomyces cerevisiae. J Biol

Chem 260: 1271–1279.

Fernández-Pevida A, Kressler D, La Cruz De J. 2015. Processing of preribosomal RNA in Saccharomyces cerevisiae. WIREs RNA 6: 191–209.

Fett JW, Strydom DJ, Lobb RR, Alderman EM, Bethune JL, Riordan JF, Vallee BL. 1985. Isolation and characterization of angiogenin, an angiogenic protein from human carcinoma cells. Biochemistry 24: 5480–5486.

Filipowicz W, Konarska M, Gross HJ, Shatkin AJ. 1983. RNA 3'-terminal phosphate cyclase activity and RNA ligation in HeLa cell extract. Nucleic Acids Research 11: 1405–1418.

Frischmeyer PA, van Hoof A, O'Donnell K, Guerrerio AL, Parker R, Dietz HC. 2002. An mRNA surveillance mechanism that eliminates transcripts lacking termination codons. Science 295: 2258–2261.

Fromm L, Falk S, Flemming D, Schuller JM, Thoms M, Conti E, Hurt E. 2017. Reconstitution of the complete pathway of ITS2 processing at the pre-ribosome. Nat Commun 8: 1787.

Fu H, Feng J, Liu Q, Sun F, Tie Y, Zhu J, Xing R, Sun Z, Zheng X. 2009. Stress induces tRNA cleavage by angiogenin in mammalian cells. FEBS Lett 583: 437–442.

Furuichi Y, LaFiandra A, Shatkin AJ. 1977. 5′-Terminal structure and mRNA stability. Nature 266: 235–239.

Gaba A, Jacobson A, Sachs MS. 2005. Ribosome occupancy of the yeast CPA1 upstream open reading frame termination codon modulates nonsense-mediated mRNA decay. Molecular Cell 20: 449–460.

Galli G, Hofstetter H, Birnstiel ML. 1981. Two conserved sequence blocks within eukaryotic tRNA genes are major promoter elements. Nature 294: 626–631.

Galperin MY, Koonin EV. 2004. “Conserved hypothetical” proteins: prioritization of targets for experimental study. Nucleic Acids Research 32: 5452–5463.

Gardner BM, Walter P. 2011. Unfolded proteins are Ire1-activating ligands that directly induce the unfolded protein response. Science 333: 1891–1894.

Gasse L, Flemming D, Hurt E. 2015. Coordinated Ribosomal ITS2 RNA Processing by the Las1 Complex Integrating Endonuclease, Polynucleotide Kinase, and Exonuclease Activities. Molecular Cell 60: 808–815.

158

Geerlings TH, Vos JC, Raué HA. 2000. The final step in the formation of 25S rRNA in Saccharomyces cerevisiae is performed by 5′→3′ exonucleases. RNA 6: 1698–1703.

Genschik P, Billy E, Swianiewicz M, Filipowicz W. 1997. The human RNA 3'-terminal phosphate cyclase is a member of a new family of proteins conserved in Eucarya, Bacteria and Archaea. The EMBO Journal 16: 2955–2967.

Genschik P, Drabikowski K, Filipowicz W. 1998. Characterization of the Escherichia coli RNA 3′-Terminal Phosphate Cyclase and Its σ54-Regulated Operon. J Biol Chem 273: 25516–25526.

Ghaemmaghami S, Huh W-K, Bower K, Howson RW, Belle A, Dephoure N, O'Shea EK, Weissman JS. 2003. Global analysis of protein expression in yeast. Nature 425: 737–741.

Giaever G, Chu AM, Ni L, Connelly C, Riles L, Véronneau S, Dow S, Lucau-Danila A, Anderson K, André B, et al. 2002. Functional profiling of the Saccharomyces cerevisiae genome. Nature 418: 387–391.

Gietz RD, Schiestl RH. 2007. High-efficiency yeast transformation using the LiAc/SS carrier DNA/PEG method. Nature Protocols 2: 31–34.

Gonzalez TN, Sidrauski C, Dörfler S, Walter P. 1999. Mechanism of non-spliceosomal mRNA splicing in the unfolded protein response pathway. The EMBO Journal 18: 3119–3132.

Good PD, Engelke DR. 1994. Yeast expression vectors using RNA polymerase III promoters. Gene 151: 209–214.

Goodman HM, Olson MV, Hall BD. 1977. Nucleotide sequence of a mutant eukaryotic gene: the yeast tyrosine-inserting ochre suppressor SUP4-o. PNAS 74: 5453–5457.

Greer CL, Peebles CL, Gegenheimer P, Abelson J. 1983. Mechanism of action of a yeast RNA ligase in tRNA splicing. Cell 32: 537–546.

Grosjean H, Szweykowska-Kulinska Z, Motorin Y. 1997. Intron-dependent enzymatic formation of modified nucleosides in eukaryotic tRNAs: a review. Biochimie 79: 293–302.

Guan Q, Zheng W, Tang S, Liu X, Zinkel RA, Tsui K-W, Yandell BS, Culbertson MR. 2006. Impact of nonsense-mediated mRNA decay on the global expression profile of budding yeast. PLoS Genet 2: e203.

Guydosh NR, Kimmig P, Walter P, Green R. 2017. Regulated Ire1-dependent mRNA decay requires no-go mRNA degradation to maintain endoplasmic reticulum homeostasis in S. pombe. eLife 6: e29216.

Hampel A, Cowan JA. 1997. A unique mechanism for RNA catalysis: the role of metal cofactors in hairpin ribozyme cleavage. Chem Biol 4: 513–517.

Hanada T, Weitzer S, Mair B, Bernreuther C, Wainger BJ, Ichida J, Hanada R, Orthofer M, Cronin SJ, Komnenovic V, et al. 2013. CLP1 links tRNA metabolism to progressive motor-neuron loss. Nature 495: 474–480.

159

Harding HP, Lackey JG, Hsu H-C, Zhang Y, Deng J, Xu R-M, Damha MJ, Ron D. 2008. An intact unfolded protein response in Trpt1 knockout mice reveals phylogenic divergence in pathways for RNA ligation. RNA 14: 225–232.

Harigaya Y, Parker R. 2012. Global analysis of mRNA decay intermediates in Saccharomyces cerevisiae. PNAS 109: 11764–11769.

He F, Peltz SW, Donahue JL, Rosbash M, Jacobson A. 1993. Stabilization and ribosome association of unspliced pre-mRNAs in a yeast upf1- mutant. PNAS 90: 7034–7038.

Heindl K, Martinez J. 2010. Nol9 is a novel polynucleotide 5'-kinase involved in ribosomal RNA processing. The EMBO Journal 29: 4161–4171.

Heinemann IU, Söll D, Randau L. 2010. Transfer RNA processing in archaea: unusual pathways and enzymes. FEBS Lett 584: 303–309.

Hollien J, Lin JH, Li H, Stevens N, Walter P, Weissman JS. 2009. Regulated Ire1-dependent decay of messenger RNAs in mammalian cells. J Cell Biol 186: 323–331.

Hollien J, Weissman JS. 2006. Decay of Endoplasmic Reticulum-Localized mRNAs During the Unfolded Protein Response. Science 313: 104–107.

Hooks KB, Griffiths-Jones S. 2011. Conserved RNA structures in the non-canonical Hac1/Xbp1 intron. RNA Biol 8: 552–556.

Hopper AK. 2013. Transfer RNA Post-Transcriptional Processing, Turnover, and Subcellular Dynamics in the Yeast Saccharomyces cerevisiae. Genetics 194: 43–67.

Hopper AK, Banks F, Evangelidis V. 1978. A yeast mutant which accumulates precursor tRNAs. Cell 14: 211–219.

Horikawa W, Endo K, Wada M, Ito K. 2016. Mutations in the G-domain of Ski7 cause specific dysfunction in non-stop decay. Nature Publishing Group 6: 29295.

Hsu CL, Stevens A. 1993. Yeast cells lacking 5′→3′ exoribonuclease 1 contain mRNA species that are poly(A) deficient and partially lack the 5′ cap structure. Molecular and Cellular

Biology 13: 4826–4835.

Hu QD, Lu H, Huo K, Ying K, Li J, Xie Y, Mao Y, Li YY. 2003. A human homolog of the yeast gene encoding tRNA 2'-phosphotransferase: cloning, characterization and complementation analysis. Cell Mol Life Sci 60: 1725–1732.

Ikeuchi K, Inada T. 2016. Ribosome-associated Asc1/RACK1 is required for endonucleolytic cleavage induced by stalled ribosome at the 3' end of nonstop mRNA. Nature Publishing

Group 6: 28234.

Inada T, Aiba H. 2005. Translation of aberrant mRNAs lacking a termination codon or with a shortened 3'-UTR is repressed after initiation in yeast. The EMBO Journal 24: 1584–1595.

160

Ivanov P, Emara MM, Villen J, Gygi SP, Anderson P. 2011. Angiogenin-induced tRNA fragments inhibit translation initiation. Molecular Cell 43: 613–623.

Jablonowski D, Zink S, Mehlgarten C, Daum G, Schaffrath R. 2006. tRNAGlu wobble uridine methylation by Trm9 identifies Elongator's key role for zymocin-induced cell death in yeast. Molecular Microbiology 59: 677–688.

James R, Kleanthous C, Moore GR. 1996. The biology of E colicins: paradigms and paradoxes. Microbiology (Reading, Engl) 142 ( Pt 7): 1569–1580.

Jang SK, Kräusslich HG, Nicklin MJ, Duke GM, Palmenberg AC, Wimmer E. 1988. A segment of the 5' nontranslated region of encephalomyocarditis virus RNA directs internal entry of ribosomes during in vitro translation. J Virol 62: 2636–2643.

Jenny A, Minvielle-Sebastia L, Preker PJ, Keller W. 1996. Sequence similarity between the 73-kilodalton protein of mammalian CPSF and a subunit of yeast polyadenylation factor I. Science 274: 1514–1517.

Johnson AW. 1997. Rat1p and Xrn1p are functionally interchangeable exoribonucleases that are restricted to and required in the nucleus and cytoplasm, respectively. Molecular and

Cellular Biology 17: 6122–6130.

Johnson AW, Kolodner RD. 1995. Synthetic lethality of sep1 (xrn1) ski2 and sep1 (xrn1) ski3 mutants of Saccharomyces cerevisiae is independent of killer virus and suggests a general role for these genes in translation control. Molecular and Cellular Biology 15: 2719–2727.

Johnston JA, Johnson ES, Waller PR, Varshavsky A. 1995. Methotrexate inhibits proteolysis of dihydrofolate reductase by the N-end rule pathway. J Biol Chem 270: 8172–8178.

Jones FD, Ryder SP, Strobel SA. 2001. An efficient ligation reaction promoted by a Varkud Satellite ribozyme with extended 5′- and 3′-termini. Nucleic Acids Research 29: 5115–5120.

Jones GM, Stalker J, Humphray S, West A, Cox T, Rogers J, Dunham I, Prelich G. 2008. A systematic library for comprehensive overexpression screens in Saccharomyces cerevisiae. Nature Methods 5: 239–241.

Jurkin J, Henkel T, Nielsen AF, Minnich M, Popow J, Kaufmann T, Heindl K, Hoffmann T, Busslinger M, Martinez J. 2014. The mammalian tRNA ligase complex mediates splicing of XBP1 mRNA and controls antibody secretion in plasma cells. The EMBO Journal.

Karaca E, Weitzer S, Pehlivan D, Shiraishi H, Gogakos T, Hanada T, Jhangiani SN, Wiszniewski W, Withers M, Campbell IM, et al. 2014. Human CLP1 Mutations Alter tRNA Biogenesis, Affecting Both Peripheral and Central Nervous System Function. Cell 157: 636–650.

Kaufmann G. 2000. Anticodon nucleases. Trends in Biochemical Sciences 25: 70–74.

Kawahara T, Yanagi H, Yura T, Mori K. 1998. Unconventional splicing of HAC1/ERN4 mRNA required for the unfolded protein response. Sequence-specific and non-sequential cleavage of the splice sites. J Biol Chem 273: 1802–1807.

161

Kebaara BW, Atkin AL. 2009. Long 3'-UTRs target wild-type mRNAs for nonsense-mediated mRNA decay in Saccharomyces cerevisiae. Nucleic Acids Research 37: 2771–2778.

Kelley PM, Schlesinger MJ. 1978. The effect of amino acid analogues and heat shock on gene expression in chicken embryo fibroblasts. Cell 15: 1277–1286.

Kimmig P, Diaz M, Zheng J, Williams CC, Lang A, Aragón T, Li H, Walter P. 2012. The unfolded protein response in fission yeast modulates stability of select mRNAs to maintain protein homeostasis. eLife 1: e00048.

Kirchner S, Ignatova Z. 2015. Emerging roles of tRNA in adaptive translation, signalling dynamics and disease. Nat Rev Genet 16: 98–112.

Kishi M, Pan YA, Crump JG, Sanes JR. 2005. Mammalian SAD kinases are required for neuronal polarization. Science 307: 929–932.

Klassen R, Meinhardt F. 2005. Induction of DNA damage and apoptosis in Saccharomyces cerevisiae by a yeast killer toxin. Cellular Microbiology 7: 393–401.

Klassen R, Paluszynski JP, Wemhoff S, Pfeiffer A, Fricke J, Meinhardt F. 2008. The primary target of the killer toxin from Pichia acaciae is tRNA(Gln). Molecular Microbiology 69: 681–697.

Klassen R, Teichert S, Meinhardt F. 2004. Novel yeast killer toxins provoke S-phase arrest and DNA damage checkpoint activation. Molecular Microbiology 53: 263–273.

Knapp G, Beckmann JS, Johnson PF, Fuhrman SA, Abelson J. 1978. Transcription and processing of intervening sequences in yeast tRNA genes. Cell 14: 221–236.

Knapp G, Ogden RC, Peebles CL, Abelson J. 1979. Splicing of yeast tRNA precursors: structure of the reaction intermediates. Cell 18: 37–45.

Kohno K, Normington K, Sambrook J, Gething MJ, Mori K. 1993. The promoter region of the yeast KAR2 (BiP) gene contains a regulatory domain that responds to the presence of unfolded proteins in the endoplasmic reticulum. Molecular and Cellular Biology 13: 877–890.

Konarska M, Filipowicz W, Domdey H, Gross HJ. 1981. Formation of a 2′-phosphomonoester, 3′,5′-phosphodiester linkage by a novel RNA ligase in wheat germ. Nature 293: 112–116.

Korennykh A, Walter P. 2012. Structural basis of the unfolded protein response. Annu Rev Cell

Dev Biol 28: 251–277.

Korennykh AV, Egea PF, Korostelev AA, Finer-Moore J, Stroud RM, Zhang C, Shokat KM, Walter P. 2011a. Cofactor-mediated conformational control in the bifunctional kinase/RNase Ire1. BMC Biol 9: 48.

Korennykh AV, Egea PF, Korostelev AA, Finer-Moore J, Zhang C, Shokat KM, Stroud RM, Walter P. 2009. The unfolded protein response signals through high-order assembly of Ire1. Nature 457: 687–693.

162

Korennykh AV, Korostelev AA, Egea PF, Finer-Moore J, Stroud RM, Zhang C, Shokat KM, Walter P. 2011b. Structural and functional basis for RNA cleavage by Ire1. BMC Biol 9: 47.

Kosmaczewski SG, Edwards TJ, Han SM, Eckwahl MJ, Meyer BI, Peach SE, Hesselberth JR, Wolin SL, Hammarlund M. 2014. The RtcB RNA ligase is an essential component of the metazoan unfolded protein response. EMBO reports.

Kosmaczewski SG, Han SM, Han B, Irving Meyer B, Baig HS, Athar W, Lin-Moore AT, Koelle MR, Hammarlund M. 2015. RNA ligation in neurons by RtcB inhibits axon regeneration. Proc Natl Acad Sci USA 112: 8451–8456.

Kozutsumi Y, Segal M, Normington K, Gething M-J, Sambrook J. 1988. The presence of malfolded proteins in the endoplasmic reticulum signals the induction of glucose-regulated proteins. Nature 332: 462–464.

Krause K, Lopes de Souza R, Roberts DGW, Dieckmann CL. 2004. The mitochondrial message-specific mRNA protectors Cbp1 and Pet309 are associated in a high-molecular weight complex. Mol Biol Cell 15: 2674–2683.

Lappe-Siefke C, Goebbels S, Gravel M, Nicksch E, Lee J, Braun PE, Griffiths IR, Nave K-A. 2003. Disruption of Cnp1 uncouples oligodendroglial functions in axonal support and myelination. Nat Genet 33: 366–374.

Larimer FW, Stevens A. 1990. Disruption of the gene XRN1, coding for a 5′→3′ exoribonuclease, restricts yeast cell growth. Gene 95: 85–90.

Laski FA, Fire AZ, RajBhandary UL, Sharp PA. 1983. Characterization of tRNA precursor splicing in mammalian extracts. J Biol Chem 258: 11974–11980.

Laughery MF, Hunter T, Brown A, Hoopes J, Ostbye T, Shumaker T, Wyrick JJ. 2015. New vectors for simple and streamlined CRISPR-Cas9 genome editing in Saccharomyces cerevisiae. Yeast 32: 711–720.

Lee AS. 1992. Mammalian stress response: induction of the glucose-regulated protein family. Curr Opin Cell Biol 4: 267–273.

Lee AS, Delegeane AM, Baker V, Chow PC. 1983. Transcriptional regulation of two genes specifically induced by glucose starvation in a hamster mutant fibroblast cell line. J Biol

Chem 258: 597–603.

Lee KPK, Dey M, Neculai D, Cao C, Dever TE, Sicheri F. 2008. Structure of the dual enzyme Ire1 reveals the basis for catalysis and regulation in nonconventional RNA splicing. Cell 132: 89–100.

Lee SR, Collins K. 2005. Starvation-induced cleavage of the tRNA anticodon loop in Tetrahymena thermophila. J Biol Chem 280: 42744–42749.

Levin JZ, Yassour M, Adiconis X, Nusbaum C, Thompson DA, Friedman N, Gnirke A, Regev A. 2010. Comprehensive comparative analysis of strand-specific RNA sequencing methods. Nature Methods 7: 709–715.

163

Levitz R, Chapman D, Amitsur M, Green R, Snyder L, Kaufmann G. 1990. The optional E. coli prr locus encodes a latent form of phage T4-induced anticodon nuclease. The EMBO

Journal 9: 1383–1389.

Li W, Okreglak V, Peschek J, Kimmig P, Zubradt M, Weissman JS, Walter P. 2018. Engineering ER-stress dependent non-conventional mRNA splicing. eLife 7.

Li Y, Luo J, Zhou H, Liao J-Y, Ma L-M, Chen Y-Q, Qu LH. 2008. Stress-induced tRNA-derived RNAs: a novel class of small RNAs in the primitive eukaryote Giardia lamblia. Nucleic Acids

Research 36: 6048–6055.

Lingner J, Kellermann J, Keller W. 1991. Cloning and expression of the essential gene for poly(A) polymerase from S. cerevisiae. Nature 354: 496–498.

Liu Q, Greimann JC, Lima CD. 2006. Reconstitution, activities, and structure of the eukaryotic RNA exosome. Cell 127: 1223–1237.

Lopes RRS, Silveira G de O, Eitler R, Vidal RS, Kessler A, Hinger S, Paris Z, Alfonzo JD, Polycarpo C. 2016. The essential function of the Trypanosoma brucei Trl1 homolog in procyclic cells is maturation of the intron-containing tRNA Tyr. RNA 22: 1190–1199.

Losson R, Lacroute F. 1979. Interference of nonsense mutations with eukaryotic messenger RNA stability. PNAS 76: 5134–5137.

Lowe TM, Eddy SR. 1997. tRNAscan-SE: a program for improved detection of transfer RNA genes in genomic sequence. Nucleic Acids Research 25: 955–964.

Lu J, Huang B, Esberg A, Johansson MJO, Byström AS. 2005. The Kluyveromyces lactis γ-toxin targets tRNA anticodons. RNA 11: 1648–1654.

Lu Y, Liang F-X, Wang X. 2014. A Synthetic Biology Approach Identifies the Mammalian UPR RNA Ligase RtcB. Molecular Cell 1–13.

MacIntosh GC, Bariola PA, Newbigin E, Green PJ. 2001. Characterization of Rny1, the Saccharomyces cerevisiae member of the T2 RNase family of RNases: Unexpected functions for ancient enzymes? PNAS 98: 1018–1023.

Mackie GA. 1998. Ribonuclease E is a 5'-end-dependent endonuclease. Nature 395: 720–723.

Martins A, Shuman S. 2005. An end-healing enzyme from Clostridium thermocellum with 5′ kinase, 2′,3′ phosphatase, and adenylyltransferase activities. RNA 11: 1271–1280.

Masaki H, Ogawa T. 2002. The modes of action of colicins E5 and D, and related cytotoxic tRNases. Biochimie 84: 433–438.

Masaki H, Ohta T. 1985. Colicin E3 and its immunity genes. J Mol Biol 182: 217–227.

McCraith SM, Phizicky EM. 1990. A highly specific phosphatase from Saccharomyces

cerevisiae implicated in tRNA splicing. Molecular and Cellular Biology 10: 1049–1055.

164

McCraith SM, Phizicky EM. 1991. An enzyme from Saccharomyces cerevisiae uses NAD+ to transfer the splice junction 2'-phosphate from ligated tRNA to an acceptor molecule. J Biol

Chem 266: 11986–11992.

Meaux S, van Hoof A. 2006. Yeast transcripts cleaved by an internal ribozyme provide new insight into the role of the cap and poly(A) tail in translation and mRNA decay. RNA 12: 1323–1337.

Meineke B, Kast A, Schwer B, Meinhardt F, Shuman S, Klassen R. 2012. A fungal anticodon nuclease ribotoxin exploits a secondary cleavage site to evade tRNA repair. RNA 18: 1716–1724.

Mian IS. 1997. Comparative sequence analysis of ribonucleases HII, III, II PH and D. Nucleic

Acids Research 25: 3187–3195.

Minvielle-Sebastia L, Preker PJ, Wiederkehr T, Strahm Y, Keller W. 1997. The major yeast poly(A)-binding protein is associated with cleavage factor IA and functions in premessenger RNA 3'-end formation. PNAS 94: 7897–7902.

Mitchell P, Petfalski E, Shevchenko A, Mann M, Tollervey D. 1997. The exosome: a conserved eukaryotic RNA processing complex containing multiple 3′→5′ exoribonucleases. Cell 91: 457–466.

Mitchell P, Petfalski E, Tollervey D. 1996. The 3' end of yeast 5.8S rRNA is generated by an exonuclease processing mechanism. Genes & Development 10: 502–513.

Mitchell P, Tollervey D. 2003. An NMD pathway in yeast involving accelerated deadenylation and exosome-mediated 3′→ 5′ degradation. Molecular Cell 11: 1405–1413.

Moore K, Hollien J. 2015. Ire1-mediated decay in mammalian cells relies on mRNA sequence, structure, and translational status. ed. G. Voeltz. Mol Biol Cell 26: 2873–2884.

Moore KA, Hollien J. 2012. The unfolded protein response in secretory cell function. Annu Rev

Genet 46: 165–183.

Morad I, Chapman-Shimshoni D, Amitsur M, Kaufmann G. 1993. Functional expression and properties of the tRNA(Lys)-specific core anticodon nuclease encoded by Escherichia coli prrC. J Biol Chem 268: 26842–26849.

Mori K, Kawahara T, Yoshida H, Yanagi H, Yura T. 1996. Signalling from endoplasmic reticulum to nucleus: transcription factor with a basic-leucine zipper motif is required for the unfolded protein-response pathway. Genes Cells 1: 803–817.

Mori K, Ma W, Gething MJ, Sambrook J. 1993. A transmembrane protein with a cdc2+/CDC28-related kinase activity is required for signaling from the ER to the nucleus. Cell 74: 743–756.

Mori K, Ogawa N, Kawahara T, Yanagi H, Yura T. 2000. mRNA splicing-mediated C-terminal replacement of transcription factor Hac1p is required for efficient activation of the unfolded protein response. PNAS 97: 4660–4665.

165

Mori K, Sant A, Kohno K, Normington K, Gething MJ, Sambrook JF. 1992. A 22 bp cis-acting element is necessary and sufficient for the induction of the yeast KAR2 (BiP) gene by unfolded proteins. The EMBO Journal 11: 2583–2593.

Mori T, Ogasawara C, Inada T, Englert M, Beier H, Takezawa M, Endo T, Yoshihisa T. 2010. Dual functions of yeast tRNA ligase in the unfolded protein response: unconventional cytoplasmic splicing of HAC1 pre-mRNA is not sufficient to release translational attenuation. Mol Biol Cell 21: 3722–3734.

Muhlrad D, Decker CJ, Parker R. 1994. Deadenylation of the unstable mRNA encoded by the yeast MFA2 gene leads to decapping followed by 5“-->3” digestion of the transcript. Genes & Development 8: 855–866.

Muhlrad D, Decker CJ, Parker R. 1995. Turnover mechanisms of the stable yeast PGK1 mRNA. Molecular and Cellular Biology 15: 2145–2156.

Muhlrad D, Parker R. 1999a. Aberrant mRNAs with extended 3' UTRs are substrates for rapid degradation by mRNA surveillance. RNA 5: 1299–1307.

Muhlrad D, Parker R. 1999b. Recognition of yeast mRNAs as “nonsense containing” leads to both inhibition of mRNA translation and mRNA degradation: implications for the control of mRNA decapping. ed. T.D. Fox. Mol Biol Cell 10: 3971–3978.

Munir A, Abdullahu L, Damha M, Shuman S. 2018a. Two-step mechanism and step-arrest mutants of Runella slithyformis NAD+-dependent tRNA 2'-phosphotransferase Tpt1. RNA rna.067165.118.

Munir A, Banerjee A, Shuman S. 2018b. NAD+-dependent synthesis of a 5′-phospho-ADP-ribosylated RNA/DNA cap by RNA 2′-phosphotransferase Tpt1. Nucleic Acids Research.

Munro S, Pelham HRB. 1986. An Hsp70-like protein in the ER: Identity with the 78 kd glucose-regulated protein and immunoglobulin heavy chain binding protein. Cell 46: 291–300.

Nagarajan VK, Jones CI, Newbury SF, Green PJ. 2013. XRN 5′→3′ exoribonucleases: structure, mechanisms and functions. Biochim Biophys Acta 1829: 590–603.

Nandakumar J, Schwer B, Schaffrath R, Shuman S. 2008. RNA repair: an antidote to cytotoxic eukaryal RNA damage. Molecular Cell 31: 278–286.

Nariya H, Inouye M. 2008. MazF, an mRNA interferase, mediates programmed cell death during multicellular Myxococcus development. Cell 132: 55–66.

Navickas A, Chamois S, Saint-Fort R, Henri J, Torchet C, Benard L. 2018. A unique No-Go Decay cleavage in mRNA exit-tunnel of ribosome produces 5′-OH ends phosphorylated by Rlg1. bioRχiv.

Nesbitt S, Hegg LA, Fedor MJ. 1997. An unusual pH-independent and metal-ion-independent mechanism for hairpin ribozyme catalysis. Chem Biol 4: 619–630.

166

Nicholson RC, Williams DB, Moran LA. 1990. An essential member of the HSP70 gene family of Saccharomyces cerevisiae is homologous to immunoglobulin heavy chain binding protein. PNAS 87: 1159–1163.

Nikawa J-I, Akiyoshi M, Hirata S, Fukuda T. 1996. Saccharomyces cerevisiae IRE2/HAC1 is involved in IRE1-mediated KAR2 expression. Nucleic Acids Research 24: 4222–4226.

Nikawa J-I, Yamashita S. 1992. IRE1 encodes a putative protein kinase containing a membrane-spanning domain and is required for inositol phototrophy in Saccharomyces

cerevisiae. Molecular Microbiology 6: 1441–1446.

Niwa M, Patil CK, DeRisi J, Walter P. 2005. Genome-scale approaches for discovering novel nonconventional splicing substrates of the Ire1 nuclease. Genome Biol 6: R3.

Noble CG, Beuth B, Taylor IA. 2007. Structure of a nucleotide-bound Clp1-Pcf11 polyadenylation factor. Nucleic Acids Research 35: 87–99.

Nojima H, Leem S-H, Araki H, Sakai A, Nakashima N, Kanaoka Y, Ono Y. 1994. Hac1: A novel yeast bZIP protein binding to the CRE motif is a multicopy suppressor for cdc10 mutant of Schizosaccharomyces pombe. Nucleic Acids Research 22: 5279–5288.

Normington K, Kohno K, Kozutsumi Y, Gething MJ, Sambrook J. 1989. S. cerevisiae encodes an essential protein homologous in sequence and function to mammalian BiP. Cell 57: 1223–1236.

Nykänen A, Haley B, Zamore PD. 2001. ATP requirements and small interfering RNA structure in the RNA interference pathway. Cell 107: 309–321.

Ogawa N, Mori K. 2004. Autoregulation of the HAC1 gene is required for sustained activation of the yeast unfolded protein response. Genes Cells 9: 95–104.

Ogawa T, Inoue S, Yajima S, Hidaka M, Masaki H. 2006. Sequence-specific recognition of colicin E5, a tRNA-targeting ribonuclease. Nucleic Acids Research 34: 6065–6073.

Ogawa T, Tomita K, Ueda T, Watanabe K, Uozumi T, Masaki H. 1999. A cytotoxic ribonuclease targeting specific transfer RNA anticodons. Science 283: 2097–2100.

Okabayashi Y, Ohki A, Sakamoto C, Otsuki M. 1985. Relationship between the severity of diabetes mellitus and pancreatic exocrine dysfunction in rats. Diabetes Res Clin Pract 1: 21–30.

Olson R, Dulac C, Bjorkman PJ. 2006. MHC homologs in the nervous system — they haven’t lost their groove. Current Opinion in Neurobiology 16: 351–357.

Ozcan U, Cao Q, Yilmaz E, Lee A-H, Iwakoshi NN, Ozdelen E, Tuncman G, Görgün C, Glimcher LH, Hotamisligil GS. 2004. Endoplasmic reticulum stress links obesity, insulin action, and type 2 diabetes. Science 306: 457–461.

Parker R. 2012. RNA degradation in Saccharomyces cerevisae. Genetics 191: 671–702.

167

Passos DO, Doma MK, Shoemaker CJ, Muhlrad D, Green R, Weissman J, Hollien J, Parker R. 2009. Analysis of Dom34 and its function in no-go decay. Mol Biol Cell 20: 3025–3032.

Paushkin SV, Patel M, Furia BS, Peltz SW, Trotta CR. 2004. Identification of a human endonuclease complex reveals a link between tRNA splicing and pre-mRNA 3' end formation. Cell 117: 311–321.

Peach SE, York K, Hesselberth JR. 2015. Global analysis of RNA cleavage by 5′-hydroxyl RNA sequencing. Nucleic Acids Research 43: e108–e108.

Pedersen K, Zavialov AV, Pavlov MY, Elf J, Gerdes K, Ehrenberg M. 2003. The bacterial toxin RelE displays codon-specific cleavage of mRNAs in the ribosomal A site. Cell 112: 131–140.

Peebles CL, Gegenheimer P, Abelson J. 1983. Precise excision of intervening sequences from precursor tRNAs by a membrane-associated yeast endonuclease. Cell 32: 525–536.

Pelechano V, Wei W, Steinmetz LM. 2015. Widespread Co-translational RNA Decay Reveals Ribosome Dynamics. Cell 161: 1400–1412.

Peng W-T, Robinson MD, Mnaimneh S, Krogan NJ, Cagney G, Morris Q, Davierwala AP, Grigull J, Yang X, Zhang W, et al. 2003. A panoramic view of yeast noncoding RNA processing. Cell 113: 919–933.

Perry RP. 1962. The Cellular Sites of Synthesis of Ribosomal and 4S RNA. PNAS 48: 2179–2186.

Perry RP, Srinivasan PR, Kelley DE. 1964. Hybridization of Rapidly Labeled Nuclear Ribonucleic Acids. Science 145: 504–507.

Peschek J, Acosta-Alvear D, Mendez AS, Walter P. 2015. A conformational RNA zipper promotes intron ejection during non-conventional XBP1 mRNA splicing. EMBO reports.

Phizicky EM, Consaul SA, Nehrke KW, Abelson J. 1992. Yeast tRNA ligase mutants are nonviable and accumulate tRNA splicing intermediates. J Biol Chem.

Phizicky EM, Hopper AK. 2010. tRNA biology charges to the front. Genes & Development 24: 1832–1860.

Phizicky EM, Schwartz RC, Abelson J. 1986. Saccharomyces cerevisiae tRNA ligase. J Biol

Chem 261: 2978–2986.

Plotch SJ, Bouloy M, Krug RM. 1979. Transfer of 5′-terminal cap of globin mRNA to influenza viral complementary RNA during transcription in vitro. PNAS 76: 1618–1622.

Plotch SJ, Bouloy M, Ulmanen I, Krug RM. 1981. A unique cap(m7GpppXm)-dependent influenza virion endonuclease cleaves capped RNAs to generate the primers that initiate viral RNA transcription. Cell 23: 847–858.

168

Poole TL, Stevens A. 1995. Comparison of features of the RNase activity of 5“-exonuclease-1 and 5-”exonuclease-2 of Saccharomyces cerevisiae. Nucleic Acids Symp Ser 79–81.

Popow J, Englert M, Weitzer S, Schleiffer A, Mierzwa B, Mechtler K, Trowitzsch S, Will CL, Lührmann R, Söll D, et al. 2011. HSPC117 is the essential subunit of a human tRNA splicing ligase complex. Science 331: 760–764.

Pramanik J, Chen X, Kar G, Henriksson J, Gomes T, Park J-E, Natarajan K, Meyer KB, Miao Z, McKenzie ANJ, et al. 2018. Genome-wide analyses reveal the IRE1a-XBP1 pathway promotes T helper cell differentiation by resolving secretory stress and accelerating proliferation. Genome Med 10: 76.

Prody GA, Bakos JT, Buzayan JM, Schneider IR, Bruening G. 1986. Autolytic processing of dimeric plant virus satellite RNA. Science 231: 1577–1580.

Reimold AM, Etkin A, Clauss I, Perkins A, Friend DS, Zhang J, Horton HF, Scott A, Orkin SH, Byrne MC, et al. 2000. An essential role in liver development for transcription factor XBP-1. Genes & Development 14: 152–157.

Reimold AM, Iwakoshi NN, Manis J, Vallabhajosyula P, Szomolanyi-Tsuda E, Gravallese EM, Friend D, Grusby MJ, Alt F, Glimcher LH. 2001. Plasma cell differentiation requires the transcription factor XBP-1. Nature 412: 300–307.

Remus BS, Shuman S. 2014. Distinctive kinetics and substrate specificities of plant and fungal tRNA ligases. RNA 20: 462–473.

Rogers TB, Inesi G, Wade R, Lederer WJ. 1995. Use of thapsigargin to study Ca2+ homeostasis in cardiac cells. Biosci Rep 15: 341–349.

Ron D, Walter P. 2007. Signal integration in the endoplasmic reticulum unfolded protein response. Nat Rev Mol Cell Biol 8: 519–529.

Rose MD, Misra LM, Vogel JP. 1989. KAR2, a karyogamy gene, is the yeast homolog of the mammalian BiP/GRP78 gene. Cell 57: 1211–1221.

Roth A, Weinberg Z, Chen AGY, Kim PB, Ames TD, Breaker RR. 2014. A widespread self-cleaving ribozyme class is revealed by bioinformatics. Nat Chem Biol 10: 56–60.

Rubio C, Pincus D, Korennykh A, Schuck S, El-Samad H, Walter P. 2011. Homeostatic adaptation to endoplasmic reticulum stress depends on Ire1 kinase activity. J Cell Biol 193: 171–184.

Rüegsegger U, Leber JH, Walter P. 2001. Block of HAC1 mRNA translation by long-range base pairing is released by cytoplasmic splicing upon induction of the unfolded protein response. Cell 107: 103–114.

Ryan K, Calvo O, Manley JL. 2004. Evidence that polyadenylation factor CPSF-73 is the mRNA 3' processing endonuclease. RNA 10: 565–573.

169

Saikia M, Jobava R, Parisien M, Putnam A, Krokowski D, Gao X-H, Guan B-J, Yuan Y, Jankowsky E, Feng Z, et al. 2014. Angiogenin-cleaved tRNA halves interact with cytochrome c, protecting cells from apoptosis during osmotic stress. Molecular and

Cellular Biology 34: 2450–2463.

Saïda F, Uzan M, Bontems F. 2003. The phage T4 restriction endoribonuclease RegB: a cyclizing enzyme that requires two histidines to be fully active. Nucleic Acids Research 31: 2751–2758.

Sanson B, Hu R-M, Troitskaya E, Mathy N, Uzan M. 2000. Endoribonuclease RegB from bacteriophage T4 is necessary for the degradation of early but not middle or late mRNAs. J

Mol Biol 297: 1063–1074.

Sarkar D, Paira S, Das B. 2018. Nuclear mRNA degradation tunes the gain of the unfolded protein response in Saccharomyces cerevisiae. Nucleic Acids Research 46: 1139–1156.

Saville BJ, Collins RA. 1990. A site-specific self-cleavage reaction performed by a novel RNA in Neurospora mitochondria. Cell 61: 685–696.

Sawaya R, Schwer B, Shuman S. 2003. Genetic and Biochemical Analysis of the Functional Domains of Yeast tRNA Ligase. J Biol Chem 278: 43928–43938.

Sawaya R, Schwer B, Shuman S. 2005. Structure-function analysis of the yeast NAD+-dependent tRNA 2´-phosphotransferase Tpt1. RNA 11: 107–113.

Sayani S, Janis M, Lee CY, Toesca I, Chanfreau GF. 2008. Widespread impact of nonsense-mediated mRNA decay on the yeast intronome. Molecular Cell 31: 360–370.

Schaffer AE, Eggens VRC, Caglayan AO, Reuter MS, Scott E, Coufal NG, Silhavy JL, Xue Y, Kayserili H, Yasuno K, et al. 2014. CLP1 founder mutation links tRNA splicing and maturation to cerebellar development and neurodegeneration. Cell 157: 651–663.

Scherrer K. 2003. Historical review: the discovery of “giant” RNA and RNA processing: 40 years of enigma. Trends in Biochemical Sciences 28: 566–571.

Scherrer K, Darnell JE. 1962. Sedimentation characteristics of rapidly labelled RNA from HeLa cells. Biochemical and Biophysical Research Communications 7: 486–490.

Scherrer K, Latham H, Darnell JE. 1963. Demonstration of an unstable RNA and of a precursor to ribosomal RNA in HeLa cells. PNAS 49: 240–248.

Scheuner D, Song B, McEwen E, Liu C, Laybutt R, Gillespie P, Saunders T, Bonner-Weir S, Kaufman RJ. 2001. Translational control is required for the unfolded protein response and in vivo glucose homeostasis. Molecular Cell 7: 1165–1176.

Schillewaert S, Wacheul L, Lhomme F, Lafontaine DLJ. 2012. The evolutionarily conserved protein Las1 is required for pre-rRNA processing at both ends of ITS2. Molecular and

Cellular Biology 32: 430–444.

170

Schuck S, Prinz WA, Thorn KS, Voss C, Walter P. 2009. Membrane expansion alleviates endoplasmic reticulum stress independently of the unfolded protein response. J Cell Biol 187: 525–536.

Schutz K, Hesselberth JR, Fields S. 2010. Capture and sequence analysis of RNAs with terminal 2′,3′-cyclic phosphates. RNA 16: 621–631.

Schwartz RC, Greer CL, Gegenheimer P, Abelson J. 1983. Enzymatic mechanism of an RNA ligase from wheat germ. J Biol Chem 258: 8374–8383.

Schwer B, Aronova A, Ramirez A, Braun P, Shuman S. 2008. Mammalian 2′,3′ cyclic nucleotide phosphodiesterase (CNP) can function as a tRNA splicing enzyme in vivo. RNA 14: 204–210.

Schwer B, Sawaya R, Ho CK, Shuman S. 2004. Portability and fidelity of RNA-repair systems. PNAS 101: 2788–2793.

Shaheen HH, Hopper AK. 2005. Retrograde movement of tRNAs from the cytoplasm to the nucleus in Saccharomyces cerevisiae. PNAS 102: 11290–11295.

Sharmeen L, Kuo MY, Dinter-Gottlieb G, Taylor J. 1988. Antigenomic RNA of human hepatitis delta virus can undergo self-cleavage. J Virol 62: 2674–2679.

Shatkin AJ. 1976. Capping of eucaryotic mRNAs. Cell 9: 645–653.

Sheridan RM, Bentley DL. 2016. Selectable one-step PCR-mediated integration of a degron for rapid depletion of endogenous human proteins. BioTechniques 60: 69–74.

Sheth U, Parker R. 2006. Targeting of aberrant mRNAs to cytoplasmic processing bodies. Cell 125: 1095–1109.

Shimotohno K, Kodama Y, Hashimoto J, Miura KI. 1977. Importance of 5′ -terminal blocking structure to stabilize mRNA in eukaryotic protein synthesis. PNAS 74: 2734–2738.

Shine J, Dalgarno L. 1974. The 3'-terminal sequence of Escherichia coli 16S ribosomal RNA: complementarity to nonsense triplets and ribosome binding sites. PNAS 71: 1342–1346.

Shoemaker CJ, Green R. 2012. Translation drives mRNA quality control. Nat Struct Mol Biol 19: 594–601.

Sidrauski C, Cox JS, Walter P. 1996. tRNA ligase is required for regulated mRNA splicing in the unfolded protein response. Cell 87: 405–413.

Sidrauski C, Walter P. 1997. The transmembrane kinase Ire1p is a site-specific endonuclease that initiates mRNA splicing in the unfolded protein response. Cell 90: 1031–1039.

Silber R, Malathi VG, Hurwitz J. 1972. Purification and properties of bacteriophage T4-induced RNA ligase. PNAS 69: 3009–3013.

171

Soelaiman S, Jakes K, Wu N, Li C, Shoham M. 2001. Crystal structure of colicin E3: implications for cell entry and ribosome inactivation. Molecular Cell 8: 1053–1062.

Soukup GA, Breaker RR. 1999. Relationship between internucleotide linkage geometry and the stability of RNA. RNA 5: 1308–1325.

Spickler C, Stronge V, Mackie GA. 2001. Preferential cleavage of degradative intermediates of rpsT mRNA by the Escherichia coli RNA degradosome. J Bacteriol 183: 1106–1109.

Spinelli SL, Consaul SA, Phizicky EM. 1997. A conditional lethal yeast phosphotransferase (tpt1) mutant accumulates tRNAs with a 2'-phosphate and an undermodified base at the splice junction. RNA 3: 1388–1400.

Sprinkle TJ. 1989. 2´,3´-cyclic nucleotide 3´-phosphodiesterase, an oligodendrocyte-Schwann cell and myelin-associated enzyme of the nervous system. Crit Rev Neurobiol 4: 235–301.

Steiger MA, Jackman JE, Phizicky EM. 2005. Analysis of 2'-phosphotransferase (Tpt1p) from Saccharomyces cerevisiae: evidence for a conserved two-step reaction mechanism. RNA 11: 99–106.

Stevens A. 2001. 5′-Exoribonuclease 1: Xrn1. Meth Enzymol 342: 251–259.

Stevens A. 1980. Purification and characterization of a Saccharomyces cerevisiae exoribonuclease which yields 5'-mononucleotides by a 5‘ leads to 3’ mode of hydrolysis. J

Biol Chem 255: 3080–3085.

Swerdlow H, Guthrie C. 1984. Structure of intron-containing tRNA precursors. Analysis of solution conformation using chemical and enzymatic probes. J Biol Chem 259: 5197–5207.

Takano A, Endo T, Yoshihisa T. 2005. tRNA actively shuttles between the nucleus and cytosol in yeast. Science 309: 140–142.

Tam AB, Koong AC, Niwa M. 2014. Ire1 has distinct catalytic mechanisms for XBP1/HAC1 splicing and RIDD. Cell Rep 9: 850–858.

Tan Z, Zhang W, Sun J, Fu Z, Ke X, Zheng C, Zhang Y, Li P, Liu Y, Hu Q, et al. 2018. ZIKV infection activates the IRE1-XBP1 and ATF6 pathways of unfolded protein response in neural cells. J Neuroinflammation 15: 275.

Tanaka N, Chakravarty AK, Maughan B, Shuman S. 2011a. Novel Mechanism of RNA Repair by RtcB via Sequential 2′,3′-Cyclic Phosphodiesterase and 3′-Phosphate/5′-Hydroxyl Ligation Reactions. J Biol Chem 286: 43134–43143.

Tanaka N, Meineke B, Shuman S. 2011b. RtcB, a novel RNA ligase, can catalyze tRNA splicing and HAC1 mRNA splicing in vivo. J Biol Chem 286: 30253–30257.

Tanaka N, Shuman S. 2011. RtcB Is the RNA Ligase Component of an Escherichia coli RNA Repair Operon. J Biol Chem 286: 7727–7731.

172

Temmel H, Müller C, Sauert M, Vesper O, Reiss A, Popow J, Martinez J, Moll I. 2016. The RNA ligase RtcB reverses MazF-induced ribosome heterogeneity in Escherichia coli. Nucleic

Acids Research gkw1018–14.

Thompson DM, Lu C, Green PJ, Parker R. 2008. tRNA cleavage is a conserved response to oxidative stress in eukaryotes. RNA 14: 2095–2103.

Thompson DM, Parker R. 2009a. Stressing out over tRNA cleavage. Cell 138: 215–219.

Thompson DM, Parker R. 2009b. The RNase Rny1p cleaves tRNAs and promotes cell death during oxidative stress in Saccharomyces cerevisiae. J Cell Biol 185: 43–50.

Tkaczuk KL, Obarska A, Bujnicki JM. 2006. Molecular phylogenetics and comparative modeling of HEN1, a methyltransferase involved in plant microRNA biogenesis. BMC Evol

Biol 6: 6.

Tomita K, Ogawa T, Uozumi T, Watanabe K, Masaki H. 2000. A cytotoxic ribonuclease which specifically cleaves four isoaccepting arginine tRNAs at their anticodon loops. PNAS 97: 8278–8283.

Travers KJ, Patil CK, Wodicka L, Lockhart DJ, Weissman JS, Walter P. 2000. Functional and genomic analyses reveal an essential coordination between the unfolded protein response and ER-associated degradation. Cell 101: 249–258.

Trotta CR, Miao F, Arn EA, Stevens SW, Ho CK, Rauhut R, Abelson JN. 1997. The yeast tRNA splicing endonuclease: a tetrameric enzyme with two active site subunits homologous to the archaeal tRNA endonucleases. Cell 89: 849–858.

Tsuboi T, Yamazaki R, Nobuta R, Ikeuchi K, Makino S, Ohtaki A, Suzuki Y, Yoshihisa T, Trotta C, Inada T. 2015. The tRNA Splicing Endonuclease Complex Cleaves the Mitochondria-localized CBP1 mRNA. J Biol Chem 290: 16021–16030.

Tucker M, Valencia-Sanchez MA, Staples RR, Chen J, Denis CL, Parker R. 2001. The transcription factor associated Ccr4 and Caf1 proteins are components of the major cytoplasmic mRNA deadenylase in Saccharomyces cerevisiae. Cell 104: 377–386.

Unlu I, Lu Y, Wang X. 2018. The cyclic phosphodiesterase CNP and RNA cyclase RtcA fine-tune noncanonical XBP1 splicing during ER stress. J Biol Chem.

Usher DA. 1969. On the mechanism of ribonuclease action. PNAS 62: 661–667.

Uzan M, Favre R, Brody E. 1988. A nuclease that cuts specifically in the ribosome binding site of some T4 mRNAs. PNAS 85: 8895–8899.

Valenzuela P, Venegas A, Weinberg F, Bishop R, Rutter WJ. 1978. Structure of yeast phenylalanine-tRNA genes: an intervening DNA segment within the region coding for the tRNA. PNAS 75: 190–194.

van Hoof A, Frischmeyer PA, Dietz HC, Parker R. 2002. Exosome-mediated recognition and degradation of mRNAs lacking a termination codon. Science 295: 2262–2264.

173

van Hoof A, Lennertz P, Parker R. 2000. Three conserved members of the RNase D family have unique and overlapping functions in the processing of 5S, 5.8S, U4, U5, RNase MRP and RNase P RNAs in yeast. The EMBO Journal 19: 1357–1365.

van Tol H, Beier H. 1988. All human tRNATyrgenes contain introns as a prerequisite for pseudouridine biosynthesis in the anticodon. Nucleic Acids Research 16: 1951–1966.

Veldman GM, Klootwijk J, van Heerikhuizen H, Planta RJ. 1981. The nucleotide sequence of the intergenic region between the 5.8S and 26S rRNA genes of the yeast ribosomal RNA operon. Possible implications for the interaction between 5.8S and 26S rRNA and the processing of the primary transcript. Nucleic Acids Research 9: 4847–4862.

Vesper O, Amitai S, Belitsky M, Byrgazov K, Kaberdina AC, Engelberg-Kulka H, Moll I. 2011. Selective translation of leaderless mRNAs by specialized ribosomes generated by MazF in Escherichia coli. Cell 147: 147–157.

Vogel US, Thompson RJ. 1988. Molecular Structure, Localization, and Possible Functions of the Myelin-Associated Enzyme 2′,3′-Cyclic Nucleotide 3′-Phosphodiesterase. J

Neurochem 50: 1667–1677.

Walter P, Ron D. 2011. The unfolded protein response: from stress pathway to homeostatic regulation. Science 334: 1081–1086.

Wang L, Lewis MS, Johnson AW. 2005. Domain interactions within the Ski2/3/8 complex and between the Ski complex and Ski7p. RNA 11: 1291–1302.

Wang LK, Lima CD, Shuman S. 2002. Structure and mechanism of T4 polynucleotide kinase: an RNA repair enzyme. The EMBO Journal 21: 3873–3880.

Wang LK, Schwer B, Englert M, Beier H, Shuman S. 2006. Structure-function analysis of the kinase-CPD domain of yeast tRNA ligase (Trl1) and requirements for complementation of tRNA splicing by a plant Trl1 homolog. Nucleic Acids Research 34: 517–527.

Wang LK, Shuman S. 2005. Structure-function analysis of yeast tRNA ligase. RNA 11: 966–975.

Wang M, Ye R, Barron E, Baumeister P, Mao C, Luo S, Fu Y, Luo B, Dubeau L, Hinton DR, et al. 2010. Essential role of the unfolded protein response regulator GRP78/BiP in protection from neuronal apoptosis. Cell Death and Differentiation 17: 488–498.

Wang P, Selvadurai K, Huang RH. 2015. Reconstitution and structure of a bacterial Pnkp1-Rnl-Hen1 RNA repair complex. Nat Commun 6: 6876.

Weitzer S, Martinez J. 2007. The human RNA kinase hClp1 is active on 3' transfer RNA exons and short interfering RNAs. Nature 447: 222–226.

Welch EM, Jacobson A. 1999. An internal open reading frame triggers nonsense-mediated decay of the yeast SPT10 mRNA. The EMBO Journal 18: 6134–6145.

174

Whipple JM, Lane EA, Chernyakov I, D'Silva S, Phizicky EM. 2011. The yeast rapid tRNA decay pathway primarily monitors the structural integrity of the acceptor and T-stems of mature tRNA. Genes & Development 25: 1173–1184.

Wilson MA, Meaux S, van Hoof A. 2007. A genomic screen in yeast reveals novel aspects of nonstop mRNA metabolism. Genetics 177: 773–784.

Winey M, Mendenhall MD, Cummins CM, Culbertson MR, Knapp G. 1986. Splicing of a yeast proline tRNA containing a novel suppressor mutation in the anticodon stem. J Mol Biol 192: 49–63.

Winkler WC, Nahvi A, Roth A, Collins JA, Breaker RR. 2004. Control of gene expression by a natural metabolite-responsive ribozyme. Nature 428: 281–286.

Woolford JL, Baserga SJ. 2013. Ribosome Biogenesis in the Yeast Saccharomyces cerevisiae. Genetics 195: 643–681.

Worsham PL, Bolen PL. 1990. Killer toxin production in Pichia acaciae is associated with linear DNA plasmids. Curr Genet 18: 77–80.

Wu J, Bao A, Chatterjee K, Wan Y, Hopper AK. 2015. Genome-wide screen uncovers novel pathways for tRNA processing and nuclear-cytoplasmic dynamics. Genes & Development 29: 2633–2644.

Wu J, Hopper AK. 2014. Healing for destruction: tRNA intron degradation in yeast is a two-step cytoplasmic process catalyzed by tRNA ligase Rlg1 and 5′-to-3′ exonuclease Xrn1. Genes

& Development 28: 1556–1561.

Wu P, Brockenbrough JS, Paddy MR, Aris JP. 1998. NCL1, a novel gene for a non-essential nuclear protein in Saccharomyces cerevisiae. Gene 220: 109–117.

Xu Q, Teplow D, Lee TD, Abelson J. 1990. Domain structure in yeast tRNA ligase. Biochemistry 29: 6132–6138.

Xue Y, Bai X, Lee I, Kallstrom G, Ho J, Brown J, Stevens A, Johnson AW. 2000. Saccharomyces cerevisiae RAI1 (YGL246c) is homologous to human DOM3Z and encodes a protein that binds the nuclear exoribonuclease Rat1p. Molecular and Cellular Biology 20: 4006–4015.

Yamasaki S, Ivanov P, Hu G-F, Anderson P. 2009. Angiogenin cleaves tRNA and promotes stress-induced translational repression. J Cell Biol 185: 35–42.

Yang W. 2011. Nucleases: diversity of structure, function and mechanism. Quart Rev Biophys 44: 1–93.

Yoshida H, Matsui T, Yamamoto A, Okada T, Mori K. 2001. XBP1 mRNA is induced by ATF6 and spliced by IRE1 in response to ER stress to produce a highly active transcription factor. Cell 107: 881–891.

175

Yoshihisa T. 2014. Handling tRNA introns, archaeal way and eukaryotic way. Front Genet 5: 213.

Yoshihisa T, Ohshima C, Yunoki-Esaki K, Endo T. 2007. Cytoplasmic splicing of tRNA in Saccharomyces cerevisiae. Genes to Cells 12: 285–297.

Yoshihisa T, Yunoki-Esaki K, Ohshima C, Tanaka N, Endo T. 2003. Possibility of cytoplasmic pre-tRNA splicing: the yeast tRNA splicing endonuclease mainly localizes on the mitochondria. Mol Biol Cell 14: 3266–3279.

Young KJ, Gill F, Grasby JA. 1997. Metal ions play a passive role in the hairpin ribozyme catalysed reaction. Nucleic Acids Research 25: 3760–3766.

Zhang Y, Zhang J, Hara H, Kato I, Inouye M. 2005. Insights into the mRNA cleavage mechanism by MazF, an mRNA interferase. J Biol Chem 280: 3143–3150.

Zhou J, Liu CY, Back SH, Clark RL, Peisach D, Xu Z, Kaufman RJ. 2006. The crystal structure of human IRE1 luminal domain reveals a conserved dimerization interface required for activation of the unfolded protein response. PNAS 103: 14343–14348.

Zillmann M, Gorovsky MA, Phizicky EM. 1992. HeLa cells contain a 2'-phosphate-specific phosphotransferase similar to a yeast enzyme implicated in tRNA splicing. J Biol Chem.

Zofallova L, Guo Y, Gupta R. 2000. Junction phosphate is derived from the precursor in the tRNA spliced by the archaeon Haloferax volcanii cell extract. RNA 6: 1019–1030.