29.01.2013 Views

Mutagenesis Techniques - Science

Mutagenesis Techniques - Science

Mutagenesis Techniques - Science

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Sponsored by<br />

This booklet is brought to you by the AAAS/<strong>Science</strong> Business Office


Over 20 Years of Innovation<br />

As a global leader in the development of innovative technologies,<br />

our products simplify, accelerate, and improve drug discovery and<br />

drug development research. Since 1984, our reagents, kits, software,<br />

and instrumentation systems have been used in pharmaceutical and<br />

biotechnology laboratories worldwide to determine the molecular<br />

mechanisms of health and disease as well as to search for new drug<br />

therapies and diagnostic tests.<br />

To order a copy of our current product catalog, visit<br />

www.stratagene.com/catalog.<br />

AMPLIFICATION<br />

CELL BIOLOGY<br />

CLONING<br />

MICROARRAYS<br />

NUCLEIC ACID ANALYSIS<br />

PROTEIN FUNCTION & ANALYSIS<br />

QUANTITATIVE PCR<br />

SOFTWARE SOLUTIONS


Contents<br />

2 Introductions:<br />

Bringing About Change<br />

Sean Sanders<br />

The Power of <strong>Mutagenesis</strong><br />

Patricia Sardina<br />

4 Bypassing a Kinase Activity with an ATP-Competitive Drug<br />

Feroz R. Papa, Chao Zhang, Kevan Shokat, Peter Walter<br />

<strong>Science</strong> 28 November 2003 302: 1533-1537<br />

12 Human Catechol-O-Methyltransferase Haplotypes Modulate<br />

Protein Expression by Altering mRNA Secondary Structure<br />

A. G. Nackley, S.A. Shabalina, I. E. Tchivileva, K. Satterfield,<br />

O. Korchynskyi, S. S. Makarov, W. Maixner, L. Diatchenko<br />

<strong>Science</strong> 22 December 2006 314: 1930-1933<br />

17 An mRNA Surveillance Mechanism That Eliminates Transcripts<br />

Lacking Termination Codons<br />

Pamela A. Frischmeyer, Ambro van Hoof, Kathryn O’Donnell,<br />

Anthony L. Guerrerio, Roy Parker, Harry C. Dietz<br />

<strong>Science</strong> 22 March 2002 295: 2258–2261<br />

24 Direct Demonstration of an Adaptive Constraint<br />

Stephen P. Miller, Mark Lunzer, Antony M. Dean<br />

<strong>Science</strong> 20 October 2006 314: 458–461<br />

30 Technical Notes: Large Insertions: Two Simple Steps Using<br />

Quikchange® II Site-Directed <strong>Mutagenesis</strong> Kits<br />

About the Cover:<br />

γ -Exonuclease binds to an end of double-stranded<br />

DNA and degrades one of the strands in a<br />

highly processive manner. The structure of the<br />

exonuclease consists of a toroidal trimer (~94<br />

angstroms in diameter) and is presumed to enclose<br />

its substrate in the manner illustrated. [Image from<br />

the cover of <strong>Science</strong> 277 (19 September 1997)]<br />

Design and Layout: Amy Hardcastle<br />

Copy Editor: Robert Buck<br />

© 2007 by The American Association for the Advancement of <strong>Science</strong>.<br />

All rights reserved.<br />

1


Introductions<br />

Bringing About Change<br />

This new method of mutagenesis has considerable potential in genetic studies.… [S]pecific<br />

mutation within protein structural genes will allow the production of proteins with specific amino<br />

acid changes. This will allow precise studies of structure-function relationships, for example<br />

the role of specific amino acids in enzyme active sites, and the more convenient production of<br />

proteins whose action is species specific, for example the conversion of a cloned gene coding for<br />

a lower vertebrate hormone to that for a human.<br />

— closing paragraph of the seminal paper from Michael Smith’s lab first describing<br />

site-directed mutagenesis 1<br />

The art of life lies in a constant readjustment to our surroundings.<br />

— Okakura Kakuzo, Japanese scholar<br />

It is a truism that, to survive, we must be able to change. This, too, is the essence of<br />

evolution on the most fundamental biological level: without the innate ability for<br />

DNA to undergo mutation, it would not be possible for organisms to adapt and survive<br />

changes in their environment. Since humankind achieved an understanding of this mutagenesis<br />

process and its power, we have been trying to control it and bend it to our will.<br />

First described in 1978 by Michael Smith at the University of British Columbia in<br />

Vancouver, 1 site-directed mutagenesis (SDM) has become an invaluable tool for molecular<br />

biologists. Originally named oligonucleotide-directed mutagenesis, the first experiments<br />

made use of short, 12-nucleotide strands of synthetic DNA containing a mismatch<br />

to an intrinsic reporter gene of the bacteriophage ΦX174. Using these oligonucleotides,<br />

together with DNA polymerase I from E. coli and T4 DNA ligase, Smith and his team<br />

demonstrated that it was possible to introduce a permanent mutation in the circular DNA<br />

of the phage, resulting in a phenotypic change. The prescient quote above shows the authors’<br />

understanding of the manifold applications for the new methodology.<br />

Since its development, a number of modifications and improvements have been made<br />

to the SDM technique. The greatest advance came with the invention of the polymerase<br />

chain reaction (PCR) in 1983. Since then, SDM and PCR have been inextricably linked,<br />

a circumstance reflected in the 1993 Nobel in Chemistry being shared by Kary Mullis,<br />

the pioneer of PCR, and Michael Smith. This advance made SDM both quicker and more<br />

flexible and has enabled it to become a standard and necessary technique employed in<br />

labs worldwide.<br />

In this booklet, we bring together a collection of influential papers that all depended<br />

on SDM for their success. As one of the most used techniques in molecular biology, it<br />

plays an often unacknowledged–but essential–role in many research projects. There is little<br />

doubt that future refinements and improvements will be made to the technique, allowing<br />

for an even broader range of applications. As these modifications are commercialized<br />

and made available to all scientists, new and innovative uses for them will be discovered,<br />

facilitating enrichment of our knowledge in many areas of research, from basic cellular<br />

processes to complex diseases such as neurodegenerative disorders and cancer.<br />

Sean Sanders, Ph.D.<br />

Commercial Editor, <strong>Science</strong><br />

1 C. A. Hutchison III, et al. <strong>Mutagenesis</strong> at a specific position in a DNA sequence. J. Biol. Chem. 253:6551-6560 (1978).<br />

2


The Power of <strong>Mutagenesis</strong><br />

In vitro mutagenesis is a very powerful tool for studying protein structure-function<br />

relationships, altering protein activity, and for modifying vector sequences to incorporate<br />

affinity tags and correct frame shift errors. A variety of mutagenesis techniques<br />

have been developed over the past decade that allow researchers to generate point mutations,<br />

modify (insert, delete or replace) one or more codons, swap domains between<br />

related gene sequences, and create diverse collections of mutant clones.<br />

<strong>Mutagenesis</strong> strategies can be divided into two main types, random or site-directed.<br />

With random mutagenesis, point mutations are introduced at random positions in a geneof-interest,<br />

typically through PCR employing an error-prone DNA polymerase (errorprone<br />

PCR). Randomized sequences are then cloned into a suitable expression vector,<br />

and the resulting mutant libraries can be screened to identify mutants with altered or improved<br />

properties. By correlating changes in function to alterations in protein sequence,<br />

researchers can create a preliminary map of protein functional domains.<br />

Site-directed mutagenesis is the method of choice for altering a gene or vector sequence<br />

at a selected location. Point mutations, insertions, or deletions are introduced by<br />

incorporating primers containing the desired modification(s) with a DNA polymerase in<br />

an amplification reaction. In more complex protein engineering experiments, researchers<br />

can design mutagenic primers to incorporate degenerate codons (site-saturation mutagenesis).<br />

Stratagene, now an Agilent Technologies company, is a worldwide leader in developing<br />

innovative products and technologies for life science research. Our expertise in enzyme<br />

engineering has translated into the most efficient and easy-to-use mutagenesis kits<br />

available today. Our QuikChange ® Site-Directed <strong>Mutagenesis</strong> products provide the most<br />

reliable and accurate method for performing targeted mutagenesis, saving time and valuable<br />

resources. These products have been cited in thousands of publications, including<br />

the papers featured in this booklet. In addition, our GeneMorph ® Random <strong>Mutagenesis</strong><br />

products offer an easy, robust method for creating diverse mutant collections.<br />

Our mutagenesis product offering has been substantially enhanced by our newest<br />

product introduction, the QuikChange ® Lightning Site-Directed <strong>Mutagenesis</strong> Kit. This<br />

kit offers significant time-savings by incorporating new, exclusive fast enzymes which<br />

allow researchers to shorten the duration of the thermal cycling and digestion steps while<br />

maintaining ultra-high fidelity and mutation efficiency.<br />

The efficacy of mutagenesis as a tool for studying protein function and structure may<br />

lead to new experimental therapeutic approaches to diseases such as cancer. Precise mapping<br />

of protein-protein interactions may ultimately lead to a greater understanding of<br />

disease mechanics. This <strong>Science</strong> <strong>Mutagenesis</strong> <strong>Techniques</strong> Collection booklet illustrates<br />

the utility of mutagenesis in a wide variety of research studies. We are pleased to have the<br />

opportunity to sponsor this publication.<br />

Patricia Sardina<br />

Product Manager<br />

Stratagene, An Agilent Technologies Company<br />

3


Secretory and transmembrane pro- functional domains (4). The most N-terteins<br />

traversing the endoplasmic minal domain, which resides in the ER<br />

reticulum (ER) during their biogen- lumen, senses elevated levels of unfolded<br />

esis fold to their native states in this com- ER proteins. Dissociation of ER chaperpartment<br />

(1). This process is facilitated by ones from Ire1 as they become engaged<br />

a plethora of ER-resident activities (the with unfolded proteins is thought to trig-<br />

protein folding machinery) (2). Insuffiger Ire1 activation (3).<br />

cient protein folding capacity causes an Ire1’s most C-terminal domain is a reg-<br />

accumulation of unfolded proteins in the ulated endoribonuclease (RNase), which<br />

ER lumen (a condition referred to as ER has a single known substrate in yeast: the<br />

stress) and triggers a transcriptional pro- HAC1<br />

gram called the UPR. In yeast, UPR targets<br />

include genes encoding chaperones,<br />

oxido-reductases, phospholipid biosynthetic<br />

enzymes, ER-associated degradation<br />

components, and proteins functioning<br />

downstream in the secretory pathway (3).<br />

Together, UPR target activities afford proteins<br />

passing through the ER an extended<br />

opportunity to fold and assemble properly,<br />

dispose of unsalvageable unfolded polypeptides,<br />

and increase the capacity for ER<br />

export.<br />

The yeast UPR is signaled through the<br />

ER stress sensor Ire1, a single-spanning<br />

ER transmembrane protein with three<br />

u mRNA (u stands for uninduced),<br />

which encodes the Hac1 transcriptional<br />

activator necessary for activation of UPR<br />

targets (5). HAC1u mRNA is constitutively<br />

transcribed, but not translated, because it<br />

contains a nonconventional translation-inhibitory<br />

intron (6). Upon activation, Ire1’s<br />

RNase cleaves HAC1u mRNA at two specific<br />

sites, excising the intron (7). The 5′<br />

and 3′ exons are rejoined by tRNA ligase<br />

(8), resulting in spliced HAC1i Unfolded proteins in the endoplasmic reticulum cause trans-autophosphorylation<br />

of the bifunctional transmembrane kinase Ire1, which induces its endoribonuclease<br />

activity. The endoribonuclease initiates nonconventional splicing<br />

of HAC1 messenger RNA to trigger the unfolded-protein response (UPR). We<br />

explored the role of Ire1’s kinase domain by sensitizing it through site-directed<br />

mutagenesis to the ATP-competitive inhibitor 1NM-PP1. Paradoxically, rather<br />

than being inhibited by 1NM-PP1, drug-sensitized Ire1 mutants required 1NM-<br />

PP1 as a cofactor for activation. In the presence of 1NM-PP1, drug-sensitized<br />

Ire1 bypassed mutations that inactivate its kinase activity and induced a full<br />

UPR. Thus, rather than through phosphorylation per se, a conformational<br />

change in the kinase domain triggered by occupancy of the active site with a<br />

ligand leads to activation of all known downstream functions.<br />

Secretory and transmembrane proteins traversing<br />

the endoplasmic reticulum (ER) during<br />

their biogenesis fold to their native states<br />

in this compartment (1). This process is facilitated<br />

by a plethora of ER-resident activities<br />

(the protein folding machinery) (2). Insufficient<br />

protein folding capacity causes an<br />

accumulation of unfolded proteins in the ER<br />

lumen (a condition referred to as ER stress)<br />

and triggers a transcriptional program called<br />

the UPR. In yeast, UPR targets include genes<br />

encoding chaperones, oxido-reductases,<br />

phospholipid biosynthetic enzymes, ERassociated<br />

degradation components, and proteins<br />

functioning downstream in the secretory<br />

pathway (3). Together, UPR target activities<br />

afford proteins passing through the ER an<br />

extended opportunity to fold and assemble<br />

properly, dispose of unsalvageable unfolded<br />

polypeptides, and increase the capacity for<br />

ER export.<br />

The yeast UPR is signaled through the ER<br />

stress sensor Ire1, a single-spanning ER<br />

transmembrane protein with three functional<br />

domains (4). The most N-terminal domain,<br />

which resides in the ER lumen, senses elevated<br />

levels of unfolded ER proteins. Dissociation<br />

of ER chaperones from Ire1 as they<br />

mRNA<br />

become engaged with unfolded proteins is (i stands for induced), which is actively<br />

thought to trigger Ire1 activation (3). translated to produce the Hac1 transcrip-<br />

Ire1’s most C-terminal domain is a regutional<br />

activator, in turn upregulating UPR<br />

lated endoribonuclease (RNase), which has a<br />

single known substrate in yeast: the HAC1 target genes (5).<br />

A functional kinase domain precedes<br />

the RNase domain on the cytosolic side<br />

of the ER membrane (9). Activation of<br />

Ire1 leads to its oligomerization in the<br />

ER membrane, followed by trans-autophosphorylation<br />

(9, 10). Mutations of<br />

catalytically essential kinase active-site<br />

u<br />

mRNA (u stands for uninduced), which encodes<br />

the Hac1 transcriptional activator necessary<br />

for activation of UPR targets (5).<br />

HAC1u mRNA is constitutively transcribed,<br />

but not translated, because it contains a nonconventional<br />

translation-inhibitory intron<br />

(6). Upon activation, Ire1’s RNase cleaves<br />

HAC1u mRNA at two specific sites, excising<br />

the intron (7). The 5� and 3� exons are rejoined<br />

by tRNA ligase (8), resulting in<br />

spliced HAC1i mRNA (i stands for induced),<br />

which is actively translated to produce the<br />

Hac1 transcriptional activator, in turn upregulating<br />

UPR target genes (5).<br />

A functional kinase domain precedes the<br />

RNase domain on the cytosolic side of the ER<br />

membrane (9). Activation of Ire1 leads to its<br />

oligomerization in the ER membrane, followed<br />

by trans-autophosphorylation (9, 10).<br />

Mutations of catalytically essential kinase<br />

active-site residues or mutations of residues<br />

known to become phosphorylated prevent<br />

HAC1u mRNA splicing and abrogate UPR<br />

signaling, demonstrating that Ire1’s kinase<br />

phosphotransfer function is essential for<br />

RNase activation (4, 9, 11). Thus, Ire1 communicates<br />

an unfolded-protein signal from<br />

the ER to the cytosol using the lumenal domain<br />

as the sensor and the RNase domain as<br />

the effector. Although clearly important for<br />

the circuitry of this machine, it is unknown<br />

why and how Ire1’s kinase is required for<br />

activation of its RNase. To dissect the role of<br />

Ire1’s kinase function, we used a recently<br />

developed strategy that allows us to sensitize<br />

Ire1 to specific kinase inhibitors (12).<br />

Unexpected behavior of Ire1 mutants<br />

sensitized to an ATP-competitive drug.<br />

The mutation of Leu745 1 2 Department of Medicine, Department of Cellular<br />

and Molecular Pharmacology,<br />

—situated at a conserved<br />

position in the adenosine 5�-triphosphate (ATP)–<br />

binding site—to Ala or Gly is predicted to sen-<br />

3Department of Biochemistry<br />

and Biophysics, 4 Feroz R. Papa,<br />

Howard Hughes Medical<br />

Institute, University of California, San Francisco, CA<br />

94143–2200, USA.<br />

*To whom correspondence should be addressed. Email:<br />

frpapa@medicine.ucsf.edu<br />

1,3* Chao Zhang, 2 Kevan Shokat, 2 Peter Walter3,4 Unfolded proteins in the endoplasmic reticulum cause trans-autophosphorylation of the<br />

bifunctional transmembrane kinase Ire1, which induces its endoribonuclease activity. The<br />

endoribonuclease initiates nonconventional splicing of HAC1 messenger RNA to trigger<br />

the unfolded protein response (UPR). We explored the role of Ire1’s kinase domain by<br />

sensitizing it through site-directed mutagenesis to the ATP-competitive inhibitor 1NM-<br />

PP1. Paradoxically, rather than being inhibited by 1NM-PP1, drug-sensitized Ire1 mutants<br />

required 1NM-PP1 as a cofactor for activation. In the presence of 1NM-PP1, drugsensitized<br />

Ire1 bypassed mutations that inactivate its kinase activity and induced a full<br />

UPR. Thus, rather than through phosphorylation per se, a conformational change in the<br />

kinase domain triggered by occupancy of the active site with a ligand leads to activation<br />

of all known downstream functions.<br />

4<br />

Bypassing a Kinase Activity with an<br />

ATP-Competitive Drug<br />

Feroz R. Papa, 1,3 * Chao Zhang, 2 Kevan Shokat, 2 Peter Walter3,4 Bypassing a Kinase Activity with<br />

an ATP-Competitive Drug<br />

sitize Ire1<br />

butyl-3-nap<br />

d]pyrimidin<br />

an enlarged<br />

wild-type k<br />

fected by<br />

partially o<br />

Ire1, these<br />

UPR sign<br />

vivo repo<br />

creased k<br />

[Ire1(L745<br />

relative to<br />

Ire1(L745<br />

approachin<br />

(Fig. 1A).<br />

Unexpe<br />

to cells<br />

Ire1(L745A<br />

concentrat<br />

1NM-PP1–<br />

stead, 1NM<br />

slightly (F<br />

prise, 1NM<br />

stored sig<br />

Ire1(L745G<br />

presents a<br />

compound<br />

an inhibito<br />

tant enzym<br />

UPR ac<br />

by dithioth<br />

cin (18); 1<br />

indicating<br />

directly af<br />

over, the e<br />

other stru<br />

naphthylm<br />

and 1-naph<br />

ene group<br />

the naphth<br />

hibited wil<br />

Activat<br />

ATP-comp<br />

tion of wh<br />

sary in the<br />

this end, w<br />

mutation<br />

D828A, w<br />

due needed<br />

as the Ire<br />

was detect<br />

wild-type c<br />

detectable,<br />

and could<br />

Ire1(L745A<br />

the UPR r<br />

addition o<br />

levels appr<br />

tivation lev<br />

binding of<br />

Ire1 rende<br />

www.sciencemag.org SCIENCE VOL 302 28 NOVEMBER 2003


esidues or mutations of residues known<br />

to become phosphorylated prevent HAC1 u<br />

mRNA splicing and abrogate UPR signaling,<br />

demonstrating that Ire1’s kinase<br />

phosphotransfer function is essential for<br />

RNase activation (4, 9, 11). Thus, Ire1<br />

communicates an unfolded-protein signal<br />

from the ER to the cytosol using the lumenal<br />

domain as the sensor and the RNase<br />

domain as the effector. Although clearly<br />

important for the circuitry of this machine,<br />

it is unknown why and how Ire1’s kinase<br />

is required for activation of its RNase. To<br />

dissect the role of Ire1’s kinase function,<br />

we used a recently developed strategy that<br />

allows us to sensitize Ire1 to specific kinase<br />

inhibitors (12).<br />

Unexpected behavior of Ire1 mutants<br />

sensitized to an ATP-competitive<br />

drug. The mutation of Leu 745 —situated<br />

at a conserved position in the adenosine<br />

5′-triphosphate (ATP)–binding site—to<br />

Ala or Gly is predicted to sensitize Ire1 to<br />

the ATP-competitive drug 1-tert-butyl-3naphthalen-1-ylmethyl-1H-pyrazolo[3,4d]pyrimidin-4-ylemine<br />

(1NM-PP1) by<br />

creating an enlarged active-site pocket not<br />

found in any wild-type kinase (13). Most<br />

kinases are not affected by these mutations,<br />

but others are partially or severely<br />

impaired (14, 15). For Ire1, these substitutions<br />

markedly decreased UPR signaling<br />

(assayed through an in vivo reporter),<br />

which is indicative of decreased kinase activity.<br />

Ire1(L 745 →A 745 ) [Ire1(L745A) (16)]<br />

showed a 40% decrease relative to the<br />

wild-type kinase, whereas Ire1(L745G)<br />

showed a >90% decrease, approaching<br />

that of the Δire1 control (Fig. 1A)<br />

Unexpectedly, the addition of 1NM-<br />

PP1 to cells expressing partially active<br />

Ire1(L745A) caused no inhibition, even<br />

at concentrations that completely inhibit<br />

other 1NM-PP1–sensitized kinases (15,<br />

17). Instead, 1NM-PP1 increased reporter<br />

activity slightly (Fig. 1A, bar 4). To our<br />

further surprise, 1NM-PP1 provided significantly<br />

restored signaling to the severely<br />

crippled Ire1(L745G) (Fig. 1A, bar 6).<br />

This result presents a paradox: An ATPcompetitive<br />

compound that was expected<br />

to function as an inhibitor of the rationally<br />

engineered mutant enzyme instead permits<br />

activation.<br />

UPR activation strictly required induction<br />

by dithiothreitol (DTT) (Fig. 1) or<br />

tunicamycin (18); 1NM-PP1 by itself had<br />

no effect, indicating that it does not induce<br />

the UPR by directly affecting ER protein<br />

folding. Moreover, the effects of 1NM-<br />

PP1 were specific: other structurally similar<br />

compounds, 2-naphthylmethyl PP1,<br />

which is a regio-isomer, and 1-naphthyl<br />

PP1, which lacks the methylene group between<br />

the heterocyclic ring and the naphthyl<br />

group, neither activated nor inhibited<br />

wild-type or mutant Ire1 (18).<br />

Activation rather than inhibition by<br />

an ATP-competitive inhibitor raised the<br />

question of whether the kinase activity<br />

is necessary in the 1NM-PP1–sensitized<br />

mutants. To this end, we combined<br />

the L745A or L745G mutation with a<br />

“kinase-dead” variant D828A, which<br />

lacks an active-site Asp residue needed<br />

to chelate Mg 2+ (11, 19). Whereas the<br />

Ire1(L745A,D828A) double mutant was<br />

detectable in cells at levels near those of<br />

wild-type cells, Ire1(L745G,D828A) was<br />

undetectable, probably because of instability,<br />

and could not be assayed. As expected,<br />

Ire1(L745A,D828A) was unable to induce<br />

the UPR reporter with DTT alone, but<br />

the addition of 1NM-PP1 allowed activation<br />

to levels approaching 80% of the<br />

wild-type activation level (Fig. 1B). This<br />

suggests that the binding of 1NM-PP1 to<br />

1NM-PP1–sensitized Ire1 renders the kinase<br />

activity dispensable.<br />

Trans-autophosphorylation by Ire1 of<br />

two activation-segment Ser residues (S 840<br />

and S 841 ) is necessary for signaling (9),<br />

prompting the question of whether this<br />

requirement is also bypassed in 1NM-<br />

PP1–sensitized mutants. As expected, the<br />

unphosphorylatable Ire1 (S840A,S841A)<br />

mutant was inactive in the absence and<br />

presence of 1NM-PP1 (Fig. 1C). In<br />

5


of two<br />

nd S 841 )<br />

ting the<br />

is also<br />

mutants.<br />

le Ire1<br />

the abig.<br />

1C).<br />

d Ire1<br />

(L745G,<br />

tly actisphoryld.<br />

Two<br />

lative to<br />

ignaling<br />

745 was<br />

to Gly<br />

ent with<br />

creasing<br />

n reducresence<br />

nsitized<br />

in the<br />

perhaps<br />

ding by<br />

ain was<br />

equires<br />

mRNA<br />

ation by<br />

wed the<br />

mRNA<br />

expectith<br />

DTT<br />

ealed by<br />

HAC1 i<br />

M-PP1<br />

Fig. 2A,<br />

h HAC1<br />

ted sig-<br />

P1–senmRNA<br />

a signifprovidlone<br />

did<br />

xpected,<br />

ng (e.g.,<br />

M-PP1<br />

pressing<br />

t, which<br />

roduced<br />

xhibited<br />

duction<br />

(Fig. 2,<br />

re1 pronts,exbserved<br />

nsitized<br />

Fig. 1. (A to D) In vivo UPR assays using a UPR element–driven lacZ reporter. The relevant<br />

genotypes and the presence or absence of 1NM-PP1 and the UPR inducer DTT are indicated. For<br />

(A) to (C), DTT was used in all assays. wt, wild type; a.u., arbitrary units.<br />

contrast, the 1NM-PP1–sensitized Ire1<br />

(L745A,S840A,S841A) and Ire1(L745G,<br />

S840A,S841A) mutants were significantly<br />

activated by 1NM-PP1, indicating that<br />

phosphorylation of S<br />

6<br />

840 and S841 stress. Thus, in the genetic background of<br />

1NM-PP1–sensitized IRE1 alleles, 1NM-PP1<br />

is, in effect, a cofactor for signaling in the UPR.<br />

1NM-PP1 uncouples the kinase and<br />

RNase activities of 1NM-PP1–sensitized<br />

can be<br />

Ire1. To rule out indirect effects, we assayed<br />

bypassed. Two trends are apparent: With<br />

the activities of the Ire1 mutants described<br />

above DTT inalone, vitro. relative The cytosolic to the portion parent mutant of Ire1<br />

consisting Ire1(S840A,S841A), of the kinasesignaling domain and decreased RNase<br />

domain as the side (Ire1*) chain wasof expressed residue 745 andwas purified pro-<br />

from gressively Escherichia reduced coli from (7). Leu Using to Ala [�- to Gly<br />

(Fig. 1C, bars 3, 5, and 7). This is consistent<br />

with the results of Fig. 1A, which<br />

32P]-la beled ATP, we assayed wild-type Ire1* and<br />

Ire1*(L745G) for autophosphorylation in the<br />

presence or absence of 1NM-PP1. Wild-type<br />

Ire1* exhibited strong autophosphorylation,<br />

which was not inhibited by 1NM-PP1 (Fig. 3A,<br />

lanes 1 and 2). In contrast, Ire1*(L745G) ex-<br />

Fig. 2. In vivo HAC1<br />

mRNA splicing and<br />

Hac1 protein production.<br />

Northern blots<br />

showing (A) in vivo<br />

HAC1 mRNA splicing<br />

and (B) SCR1 ribosomal<br />

RNA (rRNA) as<br />

loading control. Western<br />

blots showing (C)<br />

Hac1 and (D) Ire1 protein<br />

levels. The arrow in<br />

(C), lane 6, calls attention<br />

to the small<br />

amount of Hac1 protein<br />

present in these cells.<br />

Ire1*(L745G) show an increasing is diminished sensitivity in its ability of the toacuse ATP tive-site as to substrate side-chain and reduction that 1NM-PP1 at residue is, as<br />

predicted, 745. Reciprocally, an ATP competitor. in the presence of 1NM-<br />

PP1, In contrast, activation when of provided 1NM-PP1–sensitized<br />

with [�-<br />

Ire1(S840A,S841A) mutants increased in<br />

the same order (Fig. 1C, bars 4, 6, and 8),<br />

perhaps because of progressively enhanced<br />

binding by 1NM-PP1 as the residue 745<br />

side chain was trimmed back<br />

1NM-PP1–sensitized Ire1 requires<br />

1NM-PP1 as a cofactor for HAC1 mRNA<br />

splicing. We next confirmed that activa-<br />

32P]-la beled N-6 benzyl ATP, an ATP analog with a<br />

bulky substituent (Fig. 3E), Ire1*(L745G) displayed<br />

enhanced autophosphorylation activity<br />

relative to wild-type Ire1* (Fig. 3B, lane 1 and<br />

2). Therefore, the L745G mutation does not destroy<br />

Ire1’s catalytic function, but rather alters its<br />

substrate specificity from ATP to ATP analogs<br />

having complementary features that permit binding<br />

in the expanded active site.<br />

To ask how the RNase activity of Ire1 is<br />

affected by manipulation of its kinase domain,<br />

we incubated wild-type Ire1* and Ire1*(L745G)<br />

with a truncated in vitro transcribed HAC1 RNA


tion by 1NM-PP1–sensitized Ire1 mutants<br />

followed the expected pathway of activating<br />

HAC1 mRNA splicing and Hac1 protein<br />

production. As expected, UPR induction<br />

of wild-type cells with DTT caused<br />

splicing of HAC1 mRNA, as revealed by<br />

near-complete conversion of HAC1 u to<br />

HAC1 i mRNA (Fig. 2A, lanes 1 and 2) (5);<br />

1NM-PP1 had no effect by itself or with<br />

DTT (Fig. 2A, lanes 3 and 4). In strict correlation<br />

with HAC1 mRNA splicing, Hac1<br />

protein accumulated significantly (Fig.<br />

2C, lanes 2 and 4)<br />

In contrast, cells expressing 1NM-<br />

PP1–sensitized Ire1(L745G) spliced HAC1<br />

mRNA weakly with DTT alone, but displayed<br />

a significant increase when 1NM-<br />

PP1 was also provided (Fig. 2A, lanes 5 to<br />

8). 1NM-PP1 alone did not activate HAC1<br />

mRNA splicing. As expected, Hac1 protein<br />

levels correlated with splicing (e.g., Fig.<br />

2C, lanes 6 and 8). Activation by 1NM-<br />

PP1 was even more pronounced in cells<br />

expressing the Ire1(L745A,D828A) double<br />

mutant, which neither spliced HAC1<br />

mRNA nor produced Hac1 protein with<br />

DTT alone, but exhibited robust splicing<br />

and Hac1 protein production when provided<br />

both 1NM-PP1and DTT (Fig. 2, A<br />

and C, compare lanes 10 and 12).<br />

Importantly, steady-state levels of Ire1<br />

proteins were comparable in all experiments,<br />

excluding as a trivial explanation<br />

for the observed effects the possibility<br />

that 1NM-PP1–sensitized Ire1 mutants<br />

are only stably expressed in the presence<br />

of 1NM-PP1 (Fig. 2D). Taken together,<br />

our data show that 1NM-PP1 permits but<br />

does not instruct 1NM-PP1–sensitized<br />

Ire1 mutants to splice HAC1 mRNA in response<br />

to ER stress. Thus, in the genetic<br />

background of 1NM-PP1–sensitized IRE1<br />

alleles, 1NM-PP1 is, in effect, a cofactor<br />

for signaling in the UPR<br />

1NM-PP1 uncouples the kinase and<br />

RNase activities of 1NM-PP1–sensitized<br />

Ire1. To rule out indirect effects, we<br />

assayed the activities of the Ire1 mutants<br />

described above in vitro. The cytosolic<br />

portion of Ire1 consisting of the kinase<br />

domain and RNase domain (Ire1*) was<br />

expressed and purified from Escherichia<br />

coli (7). Using γ- 32 P]-labeled ATP, we assayed<br />

wild-type Ire1* and Ire1*(L745G)<br />

for autophosphorylation in the presence or<br />

absence of 1NM-PP1. Wild-type Ire1* exhibited<br />

strong autophosphorylation, which<br />

was not inhibited by 1NM-PP1 (Fig. 3A,<br />

lanes 1 and 2). In contrast, Ire1*(L745G)<br />

exhibited markedly diminished autophosphorylation,<br />

consistent with poor signaling<br />

by Ire1(L745G) in vivo (Figs. 1 and<br />

2), which was completely extinguished by<br />

1NM-PP1 (Fig. 3A, lane 3 and 4). This<br />

suggested that Ire1*(L745G) is diminished<br />

in its ability to use ATP as substrate<br />

and that 1NM-PP1 is, as predicted, an ATP<br />

competitor.<br />

In contrast, when provided with [γ-<br />

32 P]-labeled N-6 benzyl ATP, an ATP<br />

analog with a bulky substituent (Fig. 3E),<br />

Ire1*(L745G) displayed enhanced autophosphorylation<br />

activity relative to wildtype<br />

Ire1* (Fig. 3B, lane 1 and 2). Therefore,<br />

the L745G mutation does not destroy<br />

Ire1’s catalytic function, but rather alters its<br />

substrate specificity from ATP to ATP analogs<br />

having complementary features that<br />

permit binding in the expanded active site.<br />

To ask how the RNase activity of Ire1<br />

is affected by manipulation of its kinase<br />

domain, we incubated wild-type Ire1*<br />

and Ire1*(L745G) with a truncated in<br />

vitro transcribed HAC1 RNA substrate<br />

(HAC1 600 ) and monitored the production<br />

of cleavage products (7). Wild-type Ire1*<br />

cleaved HAC1 600 RNA at both splice junctions,<br />

producing the expected fragments<br />

corresponding to the singly and doubly cut<br />

species (Fig. 3C, lane 2). The cleavage reaction<br />

was not affected by 1NM-PP1 (Fig.<br />

3C, lane 3). In contrast, Ire1*(L745G) was<br />

completely inactive, whereas the addition<br />

of 1NM-PP1, at the same concentration<br />

that was shown in Fig. 3A (lane 4) to completely<br />

inhibit the kinase activity, restored<br />

RNase activity significantly (Fig. 3C,<br />

lanes 4 and 5), strongly supporting the no-<br />

7


expressions for every mRNA betwe<br />

ified genotypes.<br />

The comparison of wild-ty<br />

expressing cells with �ire1 cells<br />

distinct set of up-regulated UPR tar<br />

wild-type IRE1-expressing cells<br />

genes that were up-regulated more t<br />

are shown in pink). These IRE1<br />

genes include previously described<br />

scriptional targets (21, 22). Notabl<br />

set of (pink-colored) UPR target<br />

conformational change up-regulated occurs to a similar ma<br />

IRE1(L745A,D828A)-expressing c<br />

in response to cofactor bind-<br />

to �ire1 cells (Fig. 5B), suggesting<br />

lane 2). The cleavage reaction was not affected ent velocity sedimentation. ing Ire1* by sedimented analyzing as nonical Ire1* UPRand was induced in the<br />

by 1NM-PP1 (Fig. 3C, lane 3). In contrast, a broad peak in the gradient Ire1*(L745G) (Fig. 4). Upon the through sensitized, glycerol- kinase-dead mutant. This<br />

Ire1*(L745G) was completely inactive, whereas addition of ADP (Fig. 4B), but not 1NM-PP1 was confirmed by the tight superim<br />

the addition of 1NM-PP1, at the same concen- (Fig. 4C), the peak of Ire1* gradient shifted as avelocity result of sedimentation.<br />

the scatter diagonals of both UPR<br />

tration that was shown in Fig. 3A (lane 4) to faster sedimentation. Conversely, Ire1* sedimented the peak of as the a broad rest ofpeak the transcriptome evid<br />

completely inhibit the kinase activity, restored Ire1*(L745G) shifted by a corresponding expression profiles of wild-type I<br />

tion that 1NM-PP1 uncouples the kinase in the gradient (Fig. 4). Upon the addition<br />

RNase activity significantly (Fig. 3C, lanes 4 and amount upon the addition of 1NM-PP1 (Fig. 4F) pared with IRE1(L745A,D828A)<br />

5), and strongly RNase supporting activities the notionof that1NM-PP1–sensi 1NM-PP1 but not upon the of addition ADP (Fig. of ADP 4B), (Fig. but 4E). This not 1NM-PP1 cells (Fig. 5C). (Fig.<br />

uncouples tized Ire1. the kinase and RNase activities of shift may be4C), indicative the peak of aof conformational Ire1* shifted as Building a result anof instructive UPR<br />

1NM-PP1–sensitized Ire1.<br />

change and/or change in the oligomeric state of 1NM-PP1–sensitized IRE1 mutants d<br />

We We have have previously previously shown thatshown cleavage that of the cleav- enzymes faster in response sedimentation. to the binding ofConversely, their far act the permissively, peak because activat<br />

HAC1 age mRNA of HAC1 by wild-type mRNA Ire1* by is stimulated wild-type respective Ire1* cofactors. of Ire1*(L745G) shifted by a both correspond- 1NM-PP1 and protein-misfold<br />

by adenosine 5�-diphosphate (ADP) (7). This is 1NM-PP1–sensitized, kinase-dead Ire1 This indicates that an unfolded-pr<br />

is stimulated by adenosine 5′-diphosphate ing amount upon the addition of 1NM-PP1<br />

consistent with the notion derived from in vivo signals a canonical UPR. To ask whether needs to be received by the ER lume<br />

studies (ADP) that Ire1 (7). requires This is anconsistent active kinasewith do- the bypassing no- the(Fig. Ire1 4F) kinasebut activity not upon with 1NM- the addition for activation. of ADP We asked if we could<br />

main,tion because derived we know from that in recombinant vivo studies wild- that PP1 Ire1 allows the (Fig. induction 4E). This of a full shift UPR, may we be requirement indicative using of a constitutively o<br />

type Ire1* is already phosphorylated, thereby profiled mRNA expression on yeast genomic IRE1 (called IRE1<br />

alleviating requires a necessity an active for de kinase novo phosphoryl- domain, because microarrays. a Toconformational this end, mRNA from change wild- and/or change<br />

ation we inknow vitro. ADP that recombinant efficiently stimulated wild-type the type Ire1* IRE1, �ire1, in the and oligomeric IRE1(L745A,D828A) state of the enzymes in<br />

RNase activity of Ire1* (Fig. 3D, lane 1) but not cells was isolated at different points in time<br />

is already phosphorylated, thereby allevi- response to the binding of their respective<br />

that of Ire1*(L745G) (Fig. 3D, lane 2). after the addition of 1NM-PP1 and DTT. cDating<br />

For wild-type a necessity Ire1 and 1NM-PP1–sensitized<br />

for de novo phosphory- NAs derivedcofactors. from these mRNA populations<br />

Ire1(L745G), lation in ADP vitro. andADP 1NM-PP1, efficiently respectively, stimulated were labeled with 1NM-PP1–sensitized, different fluorescent dyes, kinase-dead<br />

function as stimulatory cofactors. Therefore, we mixed pairwise, and hybridized to microarrays<br />

asked the whether RNase a conformational activity of Ire1* change (Fig. occurs 3D, (20). lane Scatter Ire1 plots of signals pairwise comparisons a canonical of UPR. To ask<br />

in 1) response but not to cofactor that of binding Ire1*(L745G) by analyzing(Fig.<br />

mRNA 3D, expression whether levels bypassing are shown the Ire1 in kinase activity<br />

Ire1* and Ire1*(L745G) through glycerol-gradi- Fig. 5; the x-y scatter plots represent relative<br />

lane 2).<br />

with 1NM-PP1 allows the induction of a<br />

For wild-type Ire1 and 1NM-PP1–sen- full UPR, we profiled mRNA expression<br />

sitized Ire1(L745G), ADP and 1NM-PP1, on yeast genomic microarrays. To this end,<br />

respectively, function as stimulatory co- mRNA from wildtype IRE1, Δire1, and<br />

factors. Therefore, we asked whether a IRE1(L745A,D828A) cells was isolated<br />

C ) isolated previou<br />

random mutagenesis of the ER lume<br />

(18, 21). IRE1C significantly induce<br />

without DTT, resulting in about 7<br />

maximal activity of wild-type IRE1<br />

(Fig. 1D, bars 2 and 5). We pred<br />

chimera of the IRE1C Fig. 3. In vitro assays of Ire1 mutant proteins. Ire1* autophos<br />

assays with [�and<br />

1NM-PP1<br />

kinase-dead IRE1(L745A,D828A<br />

would activate the UPR with 1NM<br />

even in the absence of DTT. The dat<br />

(bars 13 to 16) show that this pred<br />

32P] ATP (A) and [�-32P]-labeled N-6 benzyl AT<br />

phorylated Ire1* proteins (wild-type and L745G) are indicated. I<br />

assays (against in vitro transcribed HAC1600 RNA) with stimula<br />

tors 1NM-PP1 (C) and ADP (D). Icons in (C) and (D) indicate HA<br />

cleavage products, denoting all combinations of singly and doub<br />

(E) The chemical structures of 1NM-PP1 and N-6 benzyl ATP a<br />

expressions for every mRNA between the specified<br />

genotypes.<br />

Fig. 4. Glycerol-density The comparison of wild-type IRE1-<br />

gradient velocity expressingsedicells with �ire1 cells revealed a<br />

mentation of distinct wild-type set of up-regulated UPR target genes in<br />

and mutant wild-type Ire1*. Ire1* IRE1-expressing cells (Fig. 5A;<br />

proteins (A genes to C) that and were up-regulated more than twofold<br />

Ire1*(L745G) are proteins shown in pink). These IRE1-dependent<br />

(D to F) were genes incubated include previously described UPR tran-<br />

in the presence scriptional of ADP targets (21, 22). Notably, the same<br />

[(B) and (E)] or set1NM-PP1 of (pink-colored) UPR target genes was<br />

[(C) and (F)] and up-regulated subject- to a similar magnitude in<br />

ed to velocity IRE1(L745A,D828A)-expressing sedimen-<br />

cells relative<br />

tation analysis. to �ire1Sedi cells (Fig. 5B), suggesting that a ca-<br />

lane 2). The cleavage reaction was not affected ent velocity sedimentation. Ire1* sedimented mentation as isnonical from UPR left was induced in the 1NM-PP1–<br />

by 1NM-PP1 (Fig. 3C, lane 3). In contrast, a broad peak in the gradient (Fig. 4). to Upon right. the sensitized, kinase-dead mutant. This conclusion<br />

Ire1*(L745G) was completely inactive, whereas addition of ADP (Fig. 4B), but not 1NM-PP1 was confirmed by the tight superimposition of<br />

the addition of 1NM-PP1, at the same concen- (Fig. 4C), the peak of Ire1* shifted as a result of the scatter diagonals of both UPR targets and<br />

tration that was shown in Fig. 3A (lane 4) to faster sedimentation. Conversely, the peak of the rest of the transcriptome evident in the<br />

completely inhibit the kinase activity, restored Ire1*(L745G) shifted by a corresponding expression profiles of wild-type IRE1- com-<br />

RNase activity significantly (Fig. 3C, lanes 4 and amount upon the addition of 1NM-PP1 (Fig. 4F) pared with IRE1(L745A,D828A)-expressing<br />

5), strongly supporting the notion that 1NM-PP1 but not upon the addition of ADP (Fig. 4E). This cells (Fig. 5C).<br />

uncouples the kinase and RNase activities of shift may be indicative of a conformational Building an instructive UPR switch.The<br />

1NM-PP1–sensitized Ire1.<br />

change and/or change in the oligomeric state of 1NM-PP1–sensitized IRE1 mutants described so<br />

We have previously shown that cleavage of the enzymes in response to the binding of their far act permissively, because activation requires<br />

HAC1 mRNA by wild-type Ire1* is stimulated respective cofactors.<br />

both 1NM-PP1 and protein-misfolding agents.<br />

by adenosine 5�-diphosphate (ADP) (7). This is 1NM-PP1–sensitized, kinase-dead Ire1 This indicates that an unfolded-protein signal<br />

consistent with the notion derived from in vivo signals a canonical UPR. To ask whether needs to be received by the ER lumenal domain<br />

studies that Ire1 requires an active kinase do- bypassing the Ire1 kinase activity with 1NM- for activation. We asked if we could bypass this<br />

main, because we know that recombinant wild- PP1 allows the induction of a full UPR, we requirement using a constitutively on version of<br />

type Ire1* is already phosphorylated, thereby profiled mRNA expression on yeast genomic IRE1 (called IRE1<br />

alleviating a necessity for de novo phosphoryl- microarrays. To this end, mRNA from wildation<br />

in vitro. ADP efficiently stimulated the type IRE1, �ire1, and IRE1(L745A,D828A)<br />

RNase activity of Ire1* (Fig. 3D, lane 1) but not cells was isolated at different points in time<br />

that of Ire1*(L745G) (Fig. 3D, lane 2). after the addition of 1NM-PP1 and DTT. cD-<br />

For wild-type Ire1 and 1NM-PP1–sensitized NAs derived from these mRNA populations<br />

Ire1(L745G), ADP and 1NM-PP1, respectively, were labeled with different fluorescent dyes,<br />

function as stimulatory cofactors. Therefore, we mixed pairwise, and hybridized to microarrays<br />

asked whether a conformational change occurs (20). Scatter plots of pairwise comparisons of<br />

in response to cofactor binding by analyzing mRNA expression levels are shown in<br />

Ire1* and Ire1*(L745G) through glycerol-gradi- Fig. 5; the x-y scatter plots represent relative<br />

www.sciencemag.org SCIENCE VOL 302 28 NOVEMBER 2003<br />

C ) isolated previously through<br />

random mutagenesis of the ER lumenal domain<br />

(18, 21). IRE1C significantly induced the UPR<br />

without DTT, resulting in about 75% of the<br />

maximal activity of wild-type IRE1 with DTT<br />

(Fig. 1D, bars 2 and 5). We predicted that a<br />

chimera of the IRE1C assays with [�and<br />

1NM-PP1–sensitized,<br />

kinase-dead IRE1(L745A,D828A) mutants<br />

would activate the UPR with 1NM-PP1 alone,<br />

even in the absence of DTT. The data in Fig. 1D<br />

(bars 13 to 16) show that this prediction holds<br />

32P] ATP (A) and [�-32P]-labeled N-6 benzyl ATP (B). Phosphorylated<br />

Ire1* proteins (wild-type and L745G) are indicated. Ire1* RNase<br />

assays (against in vitro transcribed HAC1600 RNA) with stimulatory cofactors<br />

1NM-PP1 (C) and ADP (D). Icons in (C) and (D) indicate HAC1600 RNA<br />

cleavage products, denoting all combinations of singly and doubly cut RNA.<br />

(E) The chemical structures of 1NM-PP1 and N-6 benzyl ATP are shown.<br />

Fig. 4. Glycerol-density expressions for every mRNA between the spec-<br />

gradient velocity sediified genotypes.<br />

mentation of wild-type The comparison of wild-type IRE1-<br />

and mutant Ire1*. Ire1*<br />

expressing cells with �ire1 cells revealed a<br />

proteins (A to C) and<br />

Ire1*(L745G) proteins<br />

distinct set of up-regulated UPR target genes in<br />

(D to F) were incubated wild-type IRE1-expressing cells (Fig. 5A;<br />

in the presence of ADP genes that were up-regulated more than twofold<br />

[(B) and (E)] or 1NM-PP1 are shown in pink). These IRE1-dependent<br />

[(C) and (F)] and subject- genes include previously described UPR traned<br />

to velocity sedimenscriptional<br />

targets (21, 22). Notably, the same<br />

tation analysis. Sedimentation<br />

is from left<br />

set of (pink-colored) UPR target genes was<br />

to right.<br />

up-regulated to a similar magnitude in<br />

IRE1(L745A,D828A)-expressing cells relative<br />

to �ire1 cells (Fig. 5B), suggesting that a ca-<br />

lane 2). The cleavage reaction was not affected ent velocity sedimentation. Ire1* sedimented as nonical UPR was induced in the 1NM-PP1–<br />

by 1NM-PP1 (Fig. 3C, lane 3). In contrast, a broad peak in the gradient (Fig. 4). Upon the sensitized, kinase-dead mutant. This conclusion<br />

Ire1*(L745G) was completely inactive, whereas addition of ADP (Fig. 4B), but not 1NM-PP1 was confirmed by the tight superimposition of<br />

the addition of 1NM-PP1, at the same concen- (Fig. 4C), the peak of Ire1* shifted as a result of the scatter diagonals of both UPR targets and<br />

tration that was shown in Fig. 3A (lane 4) to faster sedimentation. Conversely, the peak of the rest of the transcriptome evident in the<br />

completely inhibit the kinase activity, restored Ire1*(L745G) shifted by a corresponding expression profiles of wild-type IRE1- com-<br />

RNase activity significantly (Fig. 3C, lanes 4 and amount upon the addition of 1NM-PP1 (Fig. 4F) pared with IRE1(L745A,D828A)-expressing<br />

5), strongly supporting the notion that 1NM-PP1 but not upon the addition of ADP (Fig. 4E). This cells (Fig. 5C).<br />

uncouples the kinase and RNase activities of shift may be indicative of a conformational Building an instructive UPR switch.The<br />

1NM-PP1–sensitized Ire1.<br />

change and/or change in the oligomeric state of 1NM-PP1–sensitized IRE1 mutants described so<br />

We have previously shown that cleavage of the enzymes in response to the binding of their far act permissively, because activation requires<br />

HAC1 mRNA by wild-type Ire1* is stimulated respective cofactors.<br />

both 1NM-PP1 and protein-misfolding agents.<br />

by adenosine 5�-diphosphate (ADP) (7). This is 1NM-PP1–sensitized, kinase-dead Ire1 This indicates that an unfolded-protein signal<br />

consistent with the notion derived from in vivo signals a canonical UPR. To ask whether needs to be received by the ER lumenal domain<br />

studies that Ire1 requires an active kinase do- bypassing the Ire1 kinase activity with 1NM- for activation. We asked if we could bypass this<br />

main, because we know that recombinant wild- PP1 allows the induction of a full UPR, we requirement using a constitutively on version of<br />

type Ire1* is already phosphorylated, thereby profiled mRNA expression on yeast genomic IRE1 (called IRE1<br />

alleviating a necessity for de novo phosphoryl- microarrays. To this end, mRNA from wildation<br />

in vitro. ADP efficiently stimulated the type IRE1, �ire1, and IRE1(L745A,D828A)<br />

RNase activity of Ire1* (Fig. 3D, lane 1) but not cells was isolated at different points in time<br />

that of Ire1*(L745G) (Fig. 3D, lane 2). after the addition of 1NM-PP1 and DTT. cD-<br />

For wild-type Ire1 and 1NM-PP1–sensitized NAs derived from these mRNA populations<br />

Ire1(L745G), ADP and 1NM-PP1, respectively, were labeled with different fluorescent dyes,<br />

function as stimulatory cofactors. Therefore, we mixed pairwise, and hybridized to microarrays<br />

asked whether a conformational change occurs (20). Scatter plots of pairwise comparisons of<br />

in response to cofactor binding by analyzing mRNA expression levels are shown in<br />

Ire1* and Ire1*(L745G) through glycerol-gradi- Fig. 5; the x-y scatter plots represent relative<br />

www.sciencemag.org SCIENCE VOL 302 28 NOVEMBER 2003 1535<br />

C ) isolated previously through<br />

random mutagenesis of the ER lumenal domain<br />

(18, 21). IRE1C significantly induced the UPR<br />

without DTT, resulting in about 75% of the<br />

maximal activity of wild-type IRE1 with DTT<br />

(Fig. 1D, bars 2 and 5). We predicted that a<br />

chimera of the IRE1C assays with [�and<br />

1NM-PP1–sensitized,<br />

kinase-dead IRE1(L745A,D828A) mutants<br />

would activate the UPR with 1NM-PP1 alone,<br />

even in the absence of DTT. The data in Fig. 1D<br />

(bars 13 to 16) show that this prediction holds<br />

32P] ATP (A) and [�-32 Fig. 3. In vitro assays P]-labeledof N-6Ire1 benzyl mutant ATP (B). Phosphorylated<br />

Ire1* proteins (wild-type and L745G) are indicated. Ire1* RNase<br />

proteins. Ire1* autophosphorylation<br />

assays (against in vitro transcribed HAC1600 RNA) with stimulatory cofactors<br />

1NM-PP1 (C) assays and ADP with (D). Icons [γ-in (C) and (D) indicate HAC1600 RNA<br />

cleavage products, denoting all combinations of singly and doubly cut RNA.<br />

(E) The chemical structures of 1NM-PP1 and N-6 benzyl ATP are shown.<br />

Fig. 4. Glycerol-density<br />

gradient velocity sedimentation<br />

of wild-type<br />

and mutant Ire1*. Ire1*<br />

proteins (A to C) and<br />

Ire1*(L745G) proteins<br />

(D to F) were incubated<br />

in the presence of ADP<br />

[(B) and (E)] or 1NM-PP1<br />

[(C) and (F)] and subjected<br />

to velocity sedimentation<br />

analysis. Sedimentation<br />

is from left<br />

to right.<br />

Fig. 4. Glycerol-density gradient velocity sedimentation<br />

of wild-type and mutant Ire1*. Ire1* proteins (A to C)<br />

and Ire1*(L745G) proteins (D to F) were incubated in the<br />

presence of ADP [(B) and(E)] or 1NM-PP1 [(C) and(F)] and<br />

subjected to velocity sedimentation analysis. Sedimentation<br />

is from left to right.<br />

www.sciencemag.org SCIENCE VOL 302 28 NOVEMBER 2003 1535<br />

32P] ATP (A) and[ γ-<br />

32P]-labeled N-6 benzyl ATP (B).<br />

Phosphorylated Ire1* proteins (wildtype<br />

and L745G) are indicated.<br />

Ire1* RNase assays (against in vitro<br />

transcribed HAC1 RNA) with<br />

600<br />

stimulatory cofactors 1NM-PP1<br />

(C) and ADP (D). Icons in (C) and<br />

(D) indicate HAC1 RNA cleavage<br />

600<br />

products, denoting all combinations<br />

of singly and doubly cut RNA. (E) The<br />

chemical structures of 1NM-PP1 and<br />

N-6 benzyl ATP are shown.<br />

8<br />

R E S E A R C H A<br />

R E S E A R C H A R T I C L E<br />

Fig. 3. In vitro assays of Ire1 mutant proteins. Ire1* autophosphorylation


1NM-PP1 alone (i.e., even in the absence of<br />

ER stress).<br />

identified. Here, we report that both the<br />

kinase activity and activation-segment phos-<br />

Fig. 5. (A to C) x-y scatter plot analysis of mRNA abundance. The plots (log 10 ) compare yeast<br />

genomic microarray analyses between the genotypes indicated. In each case, mRNA was purified<br />

from cells that were provided both DTT and 1NM-PP1. Pink dots represent genes with expression<br />

more than two times as high in wild-type cells as in �ire1 �ire1 cells.<br />

Fig. 6. Model of activation of 1NM-PP1–sensitized and wild-type Ire1. For both proteins, chaperone<br />

dissociation resulting from unfolded-protein accumulation causes oligomerization in the plane of<br />

the ER membrane, which is a precondition for further activation. For mutant Ire1 (A), (A), 1NM-PP1<br />

binding to the inactive kinase domain active site causes a conformational change, which activates<br />

the RNase. For wild-type Ire1 (B), (B), trans-autophosphorylation, with ATP, of the activation segment<br />

fully opens the kinase domain for binding of ADP or ATP, which activates the RNase.<br />

1536<br />

at different points in time after the 28 28addi NOVEMBER labeled 2003 with VOL different 302 SCIENCE fluorescent www.sciencemag.org<br />

dyes,<br />

tion of 1NM-PP1 and DTT. cDNAs de- mixed pairwise, and hybridized to mirived<br />

from these mRNA populations were croarrays (20). Scatter plots of pairwise<br />

9<br />

diated by<br />

RNase a<br />

Ire1, th<br />

mechani<br />

One m<br />

proposed<br />

model,<br />

Ire1 occ<br />

after the<br />

chaperon<br />

by latera<br />

membra<br />

ation of<br />

activatio<br />

(24, 25)<br />

allowing<br />

leading<br />

that acti<br />

In co<br />

the activ<br />

occurs e<br />

ation of<br />

sible tha<br />

PP1 can<br />

ment, w<br />

to the a<br />

ATP mo<br />

1NM-PP<br />

the unp<br />

opens tr<br />

frequenc<br />

rate of<br />

sensitize<br />

ADP and<br />

could b<br />

Based o<br />

drug bin<br />

rearrang<br />

the RNa<br />

phospho<br />

mimics<br />

phospho<br />

The m<br />

why 1NM<br />

1NM-PP1<br />

vate them<br />

by activa<br />

nal doma<br />

the obser<br />

may be e<br />

free muta<br />

well to c<br />

(or cons<br />

may be<br />

neighbori<br />

formation


comparisons of mRNA expression levels<br />

are shown in Fig. 5; the x-y scatter plots<br />

represent relative expressions for every<br />

mRNA between the specified genotypes.<br />

The comparison of wild-type IRE1-<br />

expressing cells with Δire1 cells revealed<br />

a distinct set of up-regulated UPR target<br />

genes in wild-type IRE1-expressing cells<br />

(Fig. 5A; genes that were up-regulated<br />

more than twofold are shown in pink).<br />

These IRE1-dependent genes include<br />

previously described UPR transcriptional<br />

targets (21, 22). Notably, the same set<br />

of (pink-colored) UPR target genes was<br />

up-regulated to a similar magnitude in<br />

IRE1(L745A,D828A)-expressing cells<br />

relative to Δire1 cells (Fig. 5B), suggesting<br />

that a canonical UPR was induced in<br />

the 1NM-PP1–sensitized, kinase-dead<br />

mutant. This conclusion was confirmed<br />

by the tight superimposition of the scatter<br />

diagonals of both UPR targets and the rest<br />

of the transcriptome evident in the expression<br />

profiles of wild-type IRE1- compared<br />

with IRE1(L745A,D828A)-expressing<br />

cells (Fig. 5C).<br />

Building an instructive UPR switch.<br />

The 1NM-PP1–sensitized IRE1 mutants<br />

described so far act permissively, because<br />

activation requires both 1NM-PP1<br />

and protein-misfolding agents. This indicates<br />

that an unfolded-protein signal<br />

needs to be received by the ER lumenal<br />

domain for activation. We asked if we<br />

could bypass this requirement using a<br />

constitutively on version of IRE1 (called<br />

IRE1 C ) isolated previously through random<br />

mutagenesis of the ER lumenal domain<br />

(18, 21). IRE1 C significantly induced<br />

the UPR without DTT, resulting in about<br />

75% of the maximal activity of wild-type<br />

IRE1 with DTT (Fig. 1D, bars 2 and 5).<br />

We predicted that a chimera of the IRE1 C<br />

and 1NM-PP1–sensitized, kinase-dead<br />

IRE1(L745A,D828A) mutants would activate<br />

the UPR with 1NM-PP1 alone, even<br />

in the absence of DTT. The data in Fig. 1D<br />

(bars 13 to 16) show that this prediction<br />

holds true. Whereas wild-type IRE1 re-<br />

10<br />

quires DTT for induction and is immune<br />

to 1NM-PP1 (Fig. 1D, bars 1 to 4), and<br />

IRE1(L745A,D828A) requires both DTT<br />

and 1NM-PP1 (Fig. 1D, bars 9 to 12), the<br />

chimeric IRE1 C (L745A,D828A) requires<br />

only 1NM-PP1 for activation, irrespective<br />

of the addition of DTT (Fig. 1D, bars 13<br />

to 16). Thus, IRE1 C (L745A,D828A) is an<br />

instructive switch that turns on the UPR<br />

upon the addition of 1NM-PP1 alone (i.e.,<br />

even in the absence of ER stress).<br />

Summary and implications. Since<br />

our discovery that Ire1’s RNase initiates<br />

splicing of HAC1 mRNA, the function of<br />

its kinase domain has remained an enigma<br />

(7). Although mutational analyses have<br />

indicated a requirement for an enzymatically<br />

active kinase and have defined activation-segment<br />

phosphorylation sites (9),<br />

to date no targets besides Ire1 itself have<br />

been identified. Here, we report that both<br />

the kinase activity and activation-segment<br />

phosphorylation can be bypassed if the<br />

small ATP-mimic 1NM-PP1 is provided to<br />

a mutant enzyme to which it can bind. Instead<br />

of inhibiting the function served by<br />

ATP, 1NM-PP1 rectifies the signaling defect<br />

of 1NM-PP1–sensitized Ire1 mutants,<br />

leading to the induction of a canonical<br />

UPR (21). Thus, ligand binding to the kinase<br />

domain, rather than a phosphotransfer<br />

function mediated by the kinase activity,<br />

is required for RNase activation. The<br />

kinase module of Ire1, therefore, uses an<br />

unprecedented mechanism to propagate<br />

the UPR signal.<br />

One model that could explain our data<br />

is proposed in Fig. 6B. According to this<br />

model, autophosphorylation of wild-type<br />

Ire1 occurs in trans between Ire1 molecules<br />

after they have been released from<br />

their chaperone anchors and have oligomerized<br />

by lateral diffusion in the plane of<br />

the ER membrane (9, 23). Trans-autophosphorylation<br />

of the activation segment locks<br />

the activation segment in the “swung-out”<br />

state (24, 25), and fully opens the active<br />

site, allowing ADP or ATP to bind efficiently,<br />

leading in turn to conformational


changes that activate the RNase domain.<br />

In contrast, the binding of 1NM-PP1<br />

to the active site of 1NM-PP1–sensitized<br />

Ire1 occurs even in the absence of phosphorylation<br />

of the activation segment. It<br />

is plausible that because of its small size,<br />

1NMPP1 can bypass the “closed” activation<br />

segment, which is proposed to occlude<br />

access to the active site for the larger ADP<br />

and ATP molecules (Fig. 6A). Alternatively,<br />

1NM-PP1 may enter the active site<br />

when the unphosphorylated activation segment<br />

opens transiently, which occurs with<br />

low frequency in other kinases (26). If the<br />

offrate of 1NM-PP1 binding to 1NM-PP1–<br />

sensitized Ire1 is slow compared to that of<br />

ADP and ATP, most mutant Ire1 molecules<br />

could be trapped in a drug-bound state.<br />

Based on either scenario, we propose that<br />

drug binding alone causes conformational<br />

rearrangements in the kinase that activate<br />

the RNase. The nucleotide-bound state of<br />

phosphorylated wild-type Ire1, therefore,<br />

mimics the 1NM-PP1–bound state of unphosphorylated<br />

1NM-PP1–sensitized Ire1.<br />

The model discussed so far does not<br />

explain why 1NM-PP1 should not bind<br />

to individual 1NM-PP1–sensitized Ire1<br />

molecules and activate them, even without<br />

oligomerization induced by activation of<br />

function.<br />

the UPR<br />

Currently,<br />

through<br />

our<br />

the<br />

data<br />

ER<br />

do<br />

lumenal<br />

not allow<br />

domain.<br />

us to<br />

distinguish between these possibilities.<br />

Two<br />

Our<br />

possibilities<br />

findings provoke<br />

could<br />

the<br />

account<br />

question<br />

for<br />

of<br />

the<br />

what<br />

ob-<br />

the servations: “natural” stimulatory (i) The binding ligandof of 1NM-PP1 Ire1’s kinase<br />

may domain be enhanced may be. in One oligomerized, likely candidate chap- is<br />

ADP, erone-free which mutants, is generated or (ii) by Ire1 1NM-PP1 from ATP may<br />

during bind equally the autophosphorylation well to chaperone-anchored<br />

reaction. In<br />

vitro,<br />

and<br />

ADP<br />

chaperone-free<br />

is a better activator<br />

(or constitutive-on)<br />

of Ire1 than<br />

ATP (7). Ire1 would therefore be “self priming,”<br />

Ire1,<br />

generating<br />

but oligomerization<br />

an efficient activator<br />

may be required<br />

already<br />

bound because to itsinteraction active site. between In many professional neighboring<br />

secretory molecules cells is required such as the for �-cells the conforma- of the<br />

endocrine tional change pancreas that (which activates contain the RNase high funcconcentrationstion. Currently, of Ire1), our data ADPdo levels not allow rise (and us to<br />

ATP levels decline) temporarily in proportion<br />

distinguish between these possibilities.<br />

to nutritional stress (27). ADP therefore is<br />

physiologically<br />

Our findings<br />

poised<br />

provoke<br />

to serve<br />

the<br />

as<br />

question<br />

a cofactor<br />

of<br />

that what could the “natural” signal a starvation stimulatory state. ligand It is of<br />

known Ire1’s that kinase protein domain folding may becomes be. One ineffi- likely<br />

cient candidate as the nutritional is ADP, which status of is cells generated declines, by<br />

triggering Ire1 from the ATP UPR during (28). the Thus, autophosphory-<br />

in the face of<br />

ATP depletion, ADP-mediated conformalation<br />

reaction. In vitro, ADP is a better<br />

tional changes might increase the dwell time<br />

of activated Ire1. As such, Ire1 could have<br />

evolved this regulatory mechanism to monitor<br />

the energy balance of the cell and couple<br />

activator of Ire1 than ATP (7). Ire1 would<br />

therefore be “self priming,” generating an<br />

efficient activator already bound to its active<br />

site. In many professional secretory<br />

cells such as the β-cells of the endocrine<br />

pancreas (which contain high concentrations<br />

of Ire1), ADP levels rise (and ATP<br />

levels decline) temporarily in proportion<br />

to nutritional stress (27). ADP therefore<br />

is physiologically poised to serve as a cofactor<br />

that could signal a starvation state.<br />

It is known that protein folding becomes<br />

inefficient as the nutritional status of cells<br />

declines, triggering the UPR (28). Thus, in<br />

the face of ATP depletion, ADP-mediated<br />

conformational changes might increase<br />

the dwell time of activated Ire1. As such,<br />

Ire1 could have evolved this regulatory<br />

mechanism to monitor the energy balance<br />

of the cell and couple this information to<br />

activation of the UPR.<br />

Although unprecedented in proteins<br />

containing active kinase domains, our<br />

findings are consistent with previous observations<br />

of RNase L, which is closely<br />

related to Ire1, bearing a kinase-like domain<br />

followed C-terminally by an RNase<br />

domain (29, 30). Similar to Ire1, RNase<br />

L is activated by dimerization (31). However,<br />

RNase L’s kinase domain is naturally<br />

Similar<br />

inactive<br />

to<br />

(29),<br />

Ire1,<br />

whereas<br />

RNase<br />

its<br />

L<br />

RNase<br />

is activated<br />

activity<br />

by<br />

is<br />

dimerization (31). However, RNase L’s kinase<br />

stimulated<br />

domain is<br />

by<br />

naturally<br />

adenosine<br />

inactive<br />

nucleotide<br />

(29), wherebindasing<br />

itsto RNase the kinase activitydomain. is stimulated Therefore, by adeno- like<br />

sine Ire1, nucleotide the ligand-occupied binding to the kinase domain. domain<br />

Therefore, of RNase like L serves Ire1, the as ligand-occupied a module that parkinaseticipates domain in activation of RNase Land serves regulation as a module of the<br />

that<br />

RNase<br />

participates<br />

function.<br />

in<br />

Insights<br />

activation<br />

gained<br />

and regulation<br />

from Ire1<br />

of the RNase function. Insights gained from<br />

Ire1<br />

and<br />

and<br />

RNase<br />

RNase<br />

L may<br />

L may<br />

extend<br />

extend<br />

to other<br />

to other<br />

proteins<br />

proteins<br />

containing containing kinase kinase or or enzymatically enzymaticallyinac inactivetive<br />

pseudokinase domains (32).<br />

References and Notes<br />

1. M. J. Gething, J. Sambrook, Nature 355, 33 (1992).<br />

2. F. J. Stevens, Y. Argon, Semin. Cell Dev. Biol. 10, 443<br />

(1999).<br />

3. C. Patil, P. Walter, Curr. Opin. Cell Biol. 13, 349 (2001).<br />

4. J. S. Cox, C. E. Shamu, P. Walter, Cell 73, 1197 (1993).<br />

5. J. S. Cox, P. Walter, Cell 87, 391 (1996).<br />

6. U. Ruegsegger, J. H. Leber, P. Walter, Cell 107, 103<br />

(2001).<br />

7. C. Sidrauski, P. Walter, Cell 90, 1031 (1997).<br />

8. C. Sidrauski, J. S. Cox, P. Walter, Cell 87, 405 (1996).<br />

9. C. E. Shamu, P. Walter, EMBO J. 15, 3028 (1996).<br />

10. A. Weiss, J. Schlessinger, Cell 94, 277 (1998).<br />

11. K. Mori, W. Ma, M. J. Gething, J. Sambrook, Cell 74,<br />

743 (1993).<br />

11<br />

12. K. Shah, Y. Liu, C. Deirmengian, K. M. Shokat, Proc.<br />

Natl. Acad. Sci. U.S.A. 94, 3565 (1997).<br />

18. F. R. Papa,<br />

19. M. Huse,<br />

20. J. L. DeRisi,<br />

21. K. J. Trave<br />

22. Materials<br />

material o<br />

23. F. Urano, A<br />

24. S. R. Hubb<br />

Chem. 27<br />

25. T. Schindl<br />

26. M. Porter<br />

Chem. 27<br />

27. F. Schuit,<br />

Chem. 27<br />

28. R. J. Kaufm<br />

(2002).<br />

29. B. Dong, M<br />

361 (200<br />

30. S. Naik, J.<br />

Res. 26, 1<br />

31. B. Dong, R.<br />

32. M. Kroihe<br />

(2001).<br />

33. We thank<br />

Ullrich fo<br />

tants, and<br />

Shokat la<br />

grants fro<br />

(AI44009<br />

Molecular<br />

support a<br />

Medical In<br />

Supporting O


12<br />

statistically and experimentally (1). How-<br />

ever, characterizing polymorphisms locat-<br />

esineffi- declines,<br />

e face of<br />

onforma-<br />

well time<br />

uld have<br />

to moni-<br />

d couple<br />

UPR.<br />

eins con-<br />

findings<br />

ations of<br />

to Ire1,<br />

owed C-<br />

(29, 30).<br />

6. U. Ruegsegger, J. H. Leber, P. Walter, Cell 107, 103<br />

(2001).<br />

7. C. Sidrauski, P. Walter, Cell 90, 1031 (1997).<br />

8. C. Sidrauski, J. S. Cox, P. Walter, Cell 87, 405 (1996).<br />

9. C. E. Shamu, P. Walter, EMBO J. 15, 3028 (1996).<br />

10. A. Weiss, J. Schlessinger, Cell 94, 277 (1998).<br />

11. K. Mori, W. Ma, M. J. Gething, J. Sambrook, Cell 74,<br />

743 (1993).<br />

12. K. Shah, Y. Liu, C. Deirmengian, K. M. Shokat, Proc.<br />

Natl. Acad. Sci. U.S.A. 94, 3565 (1997).<br />

13. A. C. Bishop et al., Curr. Biol. 8, 257 (1998).<br />

14. E. L. Weiss, A. C. Bishop, K. M. Shokat, D. G. Drubin,<br />

Nature Cell Biol. 2, 677 (2000).<br />

15. A. C. Bishop et al., Nature 407, 395 (2000).<br />

16. Single-letter abbreviations for the amino acid resi-<br />

dues are as follows: A, Ala; D, Asp; G, Gly; L, Leu; and<br />

S, Ser.<br />

17. A. S. Carroll, A. C. Bishop, J. L. DeRisi, K. M. Shokat, E. K.<br />

O’Shea, Proc. Natl. Acad. Sci. U.S.A. 98, 12578 (2001).<br />

33. We thank M. Stark for constructing the IRE1 allele, S.<br />

Ullrich for his initial effort in constructing Ire1 mu-<br />

tants, and J. Kuriyan and members of the Walter and<br />

Shokat labs for valuable discussions. Supported by<br />

grants from the NIH to P.W. (GM32384) and K.S.<br />

(AI44009), and postdoctoral support from the UCSF<br />

Molecular Medicine Program (to F.P.). P.W. receives<br />

support as an investigator of the Howard Hughes<br />

Medical Institute.<br />

Supporting Online Material<br />

www.sciencemag.org/cgi/content/full/1090031/DC1<br />

Materials and Methods<br />

Microarray Excel Spreadsheets S1 to S3<br />

References and Notes<br />

4 August 2003; accepted 1 October 2003<br />

Published online 16 October 2003;<br />

10.1126/science.1090031<br />

Include this information when citing this paper.<br />

REPORTS<br />

ion of the Inverse<br />

ppler Effect<br />

ddon* and T. Bearpark<br />

ervation of an inverse Doppler shift, in which the<br />

ased on reflection from a receding boundary. This<br />

een produced by reflecting a wave from a moving<br />

transmission line. Doppler shifts produced by this<br />

eproducible manner by electronic control of the<br />

ically five orders of magnitude greater than those<br />

with kinematic velocities. Potential applications<br />

tunable and multifrequency radiation sources.<br />

wn phe-<br />

a wave is<br />

locity of<br />

the source and the observer (1, 2). Our con-<br />

ventional understanding of the Doppler ef-<br />

fect, from the schoolroom to everyday expe-<br />

rience of passing vehicles, is that increased<br />

frequencies are measured when a source and<br />

observer approach each other. Applications<br />

of the effect are widely established and in-<br />

clude radar, laser vibrometry, blood flow<br />

measurement, and the search for new astro-<br />

nomical objects. The inverse Doppler effect<br />

refers to frequency shifts that are in the op-<br />

posite sense to those described above; for<br />

example, increased frequencies would be<br />

measured on reflection of waves from a re-<br />

ceding boundary. Demonstration of this<br />

counterintuitive phenomenon requires a fun-<br />

damental change in the way that radiation is<br />

reflected from a moving boundary, and, de-<br />

spite a wide range of theoretical work that<br />

spans the past 60 years, the effect has not<br />

previously been verified experimentally.<br />

Although inverse Doppler shifts have<br />

been predicted to occur in particular disper-<br />

sive media (3–8) and in the near zone of<br />

three-dimensional dipoles (9–11), these<br />

schemes are difficult to implement experi-<br />

mentally and have not yet been realized.<br />

However, recent advances in the design of<br />

composite condensed media (metamaterials)<br />

offers new and exciting possibilities for the<br />

control of radiation. In particular, materials<br />

with a negative refractive index (NRI) (12–<br />

14) and the use of shock discontinuities in<br />

photonic crystals (15) and transmission lines<br />

, Advanced<br />

ffice Box 5,<br />

dressed. E-<br />

www.sciencemag.org SCIENCE VOL 302 28 NOVEMBER 2003 1537<br />

ivated by<br />

e L’s ki-<br />

), where-<br />

yadeno- e domain.<br />

upied ki-<br />

a module<br />

egulation<br />

ined from<br />

ther pro-<br />

tically in-<br />

33 (1992).<br />

iol. 10, 443<br />

349 (2001).<br />

197 (1993).<br />

.<br />

ell 107, 103<br />

97).<br />

405 (1996).<br />

8 (1996).<br />

998).<br />

ok, Cell 74,<br />

hokat, Proc.<br />

98).<br />

. G. Drubin,<br />

00).<br />

o acid resi-<br />

; L, Leu; and<br />

Shokat, E. K.<br />

78 (2001).<br />

18. F. R. Papa, C. Zhang, K. Shokat, P. Walter, data not shown.<br />

19. M. Huse, J. Kuriyan, Cell 109, 275 (2002).<br />

20. J. L. DeRisi, V. R. Iyer, P. O. Brown, <strong>Science</strong> 278, 680 (1997).<br />

21. K. J. Travers et al., Cell 101, 249 (2000).<br />

22. Materials and methods are available as supporting<br />

material on <strong>Science</strong> Online.<br />

23. F. Urano, A. Bertolotti, D. Ron, J. Cell Sci. 113, 3697 (2000).<br />

24. S. R. Hubbard, M. Mohammadi, J. Schlessinger, J. Biol.<br />

Chem. 273, 11987 (1998).<br />

25. T. Schindler et al., Mol. Cell 3, 639 (1999).<br />

26. M. Porter, T. Schindler, J. Kuriyan, W. T. Miller, J. Biol.<br />

Chem. 275, 2721 (2000).<br />

27. F. Schuit, K. Moens, H. Heimberg, D. Pipeleers, J. Biol.<br />

Chem. 274, 32803 (1999).<br />

28. R. J. Kaufman et al., Nature Rev. Mol. Cell Biol. 3, 411<br />

(2002).<br />

29. B. Dong, M. Niwa, P. Walter, R. H. Silverman, RNA 7,<br />

361 (2001).<br />

30. S. Naik, J. M. Paranjape, R. H. Silverman, Nucleic Acids<br />

Res. 26, 1522 (1998).<br />

31. B. Dong, R. H. Silverman, Nucleic Acids Res. 27, 439 (1999).<br />

32. M. Kroiher, M. A. Miller, R. E. Steele, Bioessays 23, 69<br />

(2001).<br />

33. We thank M. Stark for constructing the IRE1 C allele, S.<br />

Ullrich for his initial effort in constructing Ire1 mu-<br />

tants, and J. Kuriyan and members of the Walter and<br />

Shokat labs for valuable discussions. Supported by<br />

grants from the NIH to P.W. (GM32384) and K.S.<br />

(AI44009), and postdoctoral support from the UCSF<br />

Molecular Medicine Program (to F.P.). P.W. receives<br />

support as an investigator of the Howard Hughes<br />

Medical Institute.<br />

Supporting Online Material<br />

www.sciencemag.org/cgi/content/full/1090031/DC1<br />

Materials and Methods<br />

Microarray Excel Spreadsheets S1 to S3<br />

References and Notes<br />

4 August 2003; accepted 1 October 2003<br />

Published online 16 October 2003;<br />

10.1126/science.1090031<br />

Include this information when citing this paper.<br />

REPORTS<br />

he<br />

his<br />

ing<br />

nomical objects. The inverse Doppler effect<br />

refers to frequency shifts that are in the op-<br />

posite sense to those described above; for<br />

example, increased frequencies would be<br />

measured on reflection of waves from a re-<br />

ceding boundary. Demonstration of this<br />

counterintuitive phenomenon requires a fun-<br />

damental change in the way that radiation is<br />

reflected from a moving boundary, and, de-<br />

spite a wide range of theoretical work that<br />

R E S E A R C H A R T I C L E<br />

Ire1 than<br />

elf prim-<br />

r already<br />

fessional<br />

s of the<br />

high con-<br />

rise (and<br />

roportion<br />

refore is<br />

cofactor<br />

te. It is<br />

es ineffi-<br />

declines,<br />

e face of<br />

onforma-<br />

well time<br />

uld have<br />

to moni-<br />

d couple<br />

UPR.<br />

eins con-<br />

findings<br />

ations of<br />

to Ire1,<br />

owed C-<br />

(29, 30).<br />

that participates in activation and regulation<br />

of the RNase function. Insights gained from<br />

Ire1 and RNase L may extend to other pro-<br />

teins containing kinase or enzymatically in-<br />

active pseudokinase domains (32).<br />

References and Notes<br />

1. M. J. Gething, J. Sambrook, Nature 355, 33 (1992).<br />

2. F. J. Stevens, Y. Argon, Semin. Cell Dev. Biol. 10, 443<br />

(1999).<br />

3. C. Patil, P. Walter, Curr. Opin. Cell Biol. 13, 349 (2001).<br />

4. J. S. Cox, C. E. Shamu, P. Walter, Cell 73, 1197 (1993).<br />

5. J. S. Cox, P. Walter, Cell 87, 391 (1996).<br />

6. U. Ruegsegger, J. H. Leber, P. Walter, Cell 107, 103<br />

(2001).<br />

7. C. Sidrauski, P. Walter, Cell 90, 1031 (1997).<br />

8. C. Sidrauski, J. S. Cox, P. Walter, Cell 87, 405 (1996).<br />

9. C. E. Shamu, P. Walter, EMBO J. 15, 3028 (1996).<br />

10. A. Weiss, J. Schlessinger, Cell 94, 277 (1998).<br />

11. K. Mori, W. Ma, M. J. Gething, J. Sambrook, Cell 74,<br />

743 (1993).<br />

12. K. Shah, Y. Liu, C. Deirmengian, K. M. Shokat, Proc.<br />

Natl. Acad. Sci. U.S.A. 94, 3565 (1997).<br />

13. A. C. Bishop et al., Curr. Biol. 8, 257 (1998).<br />

14. E. L. Weiss, A. C. Bishop, K. M. Shokat, D. G. Drubin,<br />

Nature Cell Biol. 2, 677 (2000).<br />

15. A. C. Bishop et al., Nature 407, 395 (2000).<br />

16. Single-letter abbreviations for the amino acid resi-<br />

dues are as follows: A, Ala; D, Asp; G, Gly; L, Leu; and<br />

S, Ser.<br />

17. A. S. Carroll, A. C. Bishop, J. L. DeRisi, K. M. Shokat, E. K.<br />

O’Shea, Proc. Natl. Acad. Sci. U.S.A. 98, 12578 (2001).<br />

25. T. Schindler et al., Mol. Cell 3, 639 (1999).<br />

26. M. Porter, T. Schindler, J. Kuriyan, W. T. Miller, J. Biol.<br />

Chem. 275, 2721 (2000).<br />

27. F. Schuit, K. Moens, H. Heimberg, D. Pipeleers, J. Biol.<br />

Chem. 274, 32803 (1999).<br />

28. R. J. Kaufman et al., Nature Rev. Mol. Cell Biol. 3, 411<br />

(2002).<br />

29. B. Dong, M. Niwa, P. Walter, R. H. Silverman, RNA 7,<br />

361 (2001).<br />

30. S. Naik, J. M. Paranjape, R. H. Silverman, Nucleic Acids<br />

Res. 26, 1522 (1998).<br />

31. B. Dong, R. H. Silverman, Nucleic Acids Res. 27, 439 (1999).<br />

32. M. Kroiher, M. A. Miller, R. E. Steele, Bioessays 23, 69<br />

(2001).<br />

33. We thank M. Stark for constructing the IRE1 C allele, S.<br />

Ullrich for his initial effort in constructing Ire1 mu-<br />

tants, and J. Kuriyan and members of the Walter and<br />

Shokat labs for valuable discussions. Supported by<br />

grants from the NIH to P.W. (GM32384) and K.S.<br />

(AI44009), and postdoctoral support from the UCSF<br />

Molecular Medicine Program (to F.P.). P.W. receives<br />

support as an investigator of the Howard Hughes<br />

Medical Institute.<br />

Supporting Online Material<br />

www.sciencemag.org/cgi/content/full/1090031/DC1<br />

Materials and Methods<br />

Microarray Excel Spreadsheets S1 to S3<br />

References and Notes<br />

4 August 2003; accepted 1 October 2003<br />

Published online 16 October 2003;<br />

10.1126/science.1090031<br />

Include this information when citing this paper.<br />

REPORTS<br />

ion of the Inverse<br />

ppler Effect<br />

ddon* and T. Bearpark<br />

ervation of an inverse Doppler shift, in which the<br />

ased on reflection from a receding boundary. This<br />

een produced by reflecting a wave from a moving<br />

transmission line. Doppler shifts produced by this<br />

eproducible manner by electronic control of the<br />

ically five orders of magnitude greater than those<br />

with kinematic velocities. Potential applications<br />

tunable and multifrequency radiation sources.<br />

wn phe-<br />

a wave is<br />

locity of<br />

the source and the observer (1, 2). Our con-<br />

ventional understanding of the Doppler ef-<br />

fect, from the schoolroom to everyday expe-<br />

rience of passing vehicles, is that increased<br />

frequencies are measured when a source and<br />

observer approach each other. Applications<br />

of the effect are widely established and in-<br />

clude radar, laser vibrometry, blood flow<br />

measurement, and the search for new astro-<br />

nomical objects. The inverse Doppler effect<br />

refers to frequency shifts that are in the op-<br />

posite sense to those described above; for<br />

example, increased frequencies would be<br />

measured on reflection of waves from a re-<br />

ceding boundary. Demonstration of this<br />

counterintuitive phenomenon requires a fun-<br />

damental change in the way that radiation is<br />

reflected from a moving boundary, and, de-<br />

spite a wide range of theoretical work that<br />

spans the past 60 years, the effect has not<br />

previously been verified experimentally.<br />

Although inverse Doppler shifts have<br />

been predicted to occur in particular disper-<br />

sive media (3–8) and in the near zone of<br />

three-dimensional dipoles (9–11), these<br />

schemes are difficult to implement experi-<br />

mentally and have not yet been realized.<br />

However, recent advances in the design of<br />

composite condensed media (metamaterials)<br />

offers new and exciting possibilities for the<br />

control of radiation. In particular, materials<br />

with a negative refractive index (NRI) (12–<br />

14) and the use of shock discontinuities in<br />

photonic crystals (15) and transmission lines<br />

, Advanced<br />

fice Box 5,<br />

dressed. E-<br />

www.sciencemag.org SCIENCE VOL 302 28 NOVEMBER 2003 1537<br />

n squirrels<br />

pared with<br />

od supple-<br />

reds have<br />

size or a<br />

ce that the<br />

bserved in<br />

increased<br />

). We hy-<br />

resource-<br />

investment<br />

ounts, re-<br />

ure fitness<br />

roduction,<br />

ailable re-<br />

augmenta-<br />

show that<br />

rger litters<br />

cause off-<br />

ver, when<br />

juveniles is<br />

offspring,<br />

population<br />

n of young<br />

come with<br />

ause over-<br />

r years of<br />

for Amer-<br />

e previous<br />

s surviving<br />

t 15 = –0.7,<br />

portion of<br />

ersus adult<br />

.37 ± 0.27,<br />

amp-and-<br />

tion, then<br />

ion growth<br />

y the pred-<br />

tive output<br />

current but<br />

ggest that<br />

ms is more<br />

an present<br />

seed mast-<br />

vivalpros- e resource<br />

allel in re-<br />

predators<br />

28, 863<br />

Univ. of<br />

tal Care<br />

).<br />

May, Ed.<br />

don Ser. B<br />

M. Schauber,<br />

, 232 (2000).<br />

9. L. M. Curran, M. Leighton, Ecol. Monogr. 70, 101 (2000).<br />

10. W. D. Koenig, J. M. H. Knops, Am. Sci. 93, 340 (2005).<br />

11. D. H. Janzen, Annu. Rev. Ecol. Syst. 2, 465 (1971).<br />

12. A. F. Hedlin, For. Sci. 10, 124 (1964).<br />

13. M. J. McKone, D. Kelly, A. L. Harrison, J. J. Sullivan,<br />

A. J. Cone, N.Z. J. Zool. 28, 89 (2001).<br />

14. P. R. Wilson, B. J. Karl, R. J. Toft, J. R. Beggs, Biol. Conserv.<br />

83, 175 (1998).<br />

15. T. Ruf, J. Fietz, W. Schlund, C. Bieber, Ecology 87, 372<br />

(2006).<br />

16. F. H. Bronson, Reproductive Biology (Univ. of Chicago<br />

Press, Chicago, 1989).<br />

17. Materials and methods are available on <strong>Science</strong> Online.<br />

18. J. O. Wolff, J. Mammal. 77, 850 (1996).<br />

19. A. S. McNeilly, Reprod. Fertil. Dev. 13, 583 (2001).<br />

20. J. D. Ligon, Nature 250, 80 (1974).<br />

21. P. J. Berger, N. C. Negus, E. H. Sanders, P. D. Gardner,<br />

<strong>Science</strong> 214, 69 (1981).<br />

22. S. Eis, J. Inkster, Can. J. For. Res. 2, 460 (1972).<br />

23. B. M. Fitzgerald, M. J. Efford, B. J. Karl, N.Z. J. Zool. 31,<br />

167 (2004).<br />

24. M. C. Smith, J. Wildl. Manag. 32, 305 (1968).<br />

25. D. A. Rusch, W. G. Reeder, Ecology 59, 400 (1978).<br />

26. L. A. Wauters, C. Swinnen, A. A. Dhondt, J. Zool. 227, 71<br />

(1992).<br />

27. L. A. Wauters, A. A. Dhondt, J. Zool. 217, 93 (1989).<br />

28. C. D. Becker, Can. J. Zool. 71, 1326 (1993).<br />

29. K. W. Larsen, C. D. Becker, S. Boutin, M. Blower,<br />

J. Mammal. 78, 192 (1997).<br />

30. M. M. Humphries, S. Boutin, Ecology 81, 2867<br />

(2000).<br />

31. A. G. McAdam, S. Boutin, Evolution 57, 1689 (2003).<br />

32. L. A. Wauters, E. Matthysen, F. Adriaensen, G. Tosi,<br />

J. Anim. Ecol. 73, 11 (2004).<br />

33. We thank J. Herbers, C. Aumann, J. LaMontagne,<br />

M. Andruskiw, J. Lane, R. Boonstra, K. McCann,<br />

B. Thomas, Kluane Red Squirrel Project 2005 annual<br />

meeting participants, and three anonymous reviewers for<br />

very useful comments. We are indebted to the many field<br />

workers who helped collect the data and to A. Sykes and<br />

E. Anderson for management of the long-term data.<br />

Financed by Natural <strong>Science</strong>s and Engineering Research<br />

Council of Canada, NSF (American reds), a Concerted<br />

Action of the Belgian Ministry of Education, and a<br />

Training and Mobility of Researchers grant from the<br />

Commission of the European Union (Eurasian reds). This<br />

is publication 30 of the Kluane Red Squirrel Project.<br />

Supporting Online Material<br />

www.sciencemag.org/cgi/content/full/314/5807/1928/DC1<br />

Materials and Methods<br />

SOM Text<br />

Fig. S1<br />

Table S1<br />

References<br />

25 September 2006; accepted 15 November 2006<br />

10.1126/science.1135520<br />

Human Catechol-O-Methyltransferase<br />

Haplotypes Modulate Protein Expression<br />

by Altering mRNA Secondary Structure<br />

A. G. Nackley, 1 S. A. Shabalina, 2 I. E. Tchivileva, 1 K. Satterfield, 1 O. Korchynskyi, 3<br />

S. S. Makarov, 4 W. Maixner, 1 L. Diatchenko 1 *<br />

Catechol-O-methyltransferase (COMT) is a key regulator of pain perception, cognitive function, and<br />

affective mood. Three common haplotypes of the human COMT gene, divergent in two synonymous<br />

and one nonsynonymous position, code for differences in COMT enzymatic activity and are<br />

associated with pain sensitivity. Haplotypes divergent in synonymous changes exhibited the largest<br />

difference in COMT enzymatic activity, due to a reduced amount of translated protein. The major<br />

COMT haplotypes varied with respect to messenger RNA local stem-loop structures, such that the<br />

most stable structure was associated with the lowest protein levels and enzymatic activity. Site-<br />

directed mutagenesis that eliminated the stable structure restored the amount of translated<br />

protein. These data highlight the functional significance of synonymous variations and suggest the<br />

importance of haplotypes over single-nucleotide polymorphisms for analysis of genetic variations.<br />

T<br />

he ability to predict the downstream<br />

effects of genetic variation is critically<br />

important for understanding both the<br />

evolution of the genome and the molecular<br />

basis of human disease. The effects of non-<br />

synonymous polymorphisms have been wide-<br />

ly characterized; because these variations<br />

directly influence protein function, they are<br />

relatively easy to study statistically and<br />

experimentally (1). However, characterizing<br />

polymorphisms located in regulatory regions,<br />

which are much more common, has proved to<br />

be problematic (2). Here, we focus on the<br />

mechanism whereby polymorphisms of the<br />

cathechol-O-methyltransferase (COMT ) gene<br />

regulate gene expression.<br />

COMT is an enzyme responsible for de-<br />

grading catecholamines and thus represents a<br />

critical component of homeostasis maintenance<br />

(3). The human COMT gene encodes two<br />

distinct proteins: soluble COMT (S-COMT)<br />

and membrane-bound COMT (MB-COMT)<br />

1 Center for Neurosensory Disorders, University of North<br />

Carolina, Chapel Hill, NC 27599, USA. 2 National Center for<br />

Biotechnology Information, National Institutes of Health,<br />

Bethesda, MD 20894, USA. 3 Thurston Arthritis Center,<br />

University of North Carolina, Chapel Hill, NC 27599, USA.<br />

4 Attagene, Inc., 7030 Kit Creek Road, Research Triangle Park,<br />

NC 27560, USA.<br />

The ability to predict the downstream<br />

effects of genetic variation is criti-<br />

cally important for understanding<br />

both the evolution of the genome and the<br />

molecular basis of human disease. The ef-<br />

fects of nonsynonymous polymorphisms<br />

have been widely characterized; because<br />

these variations directly influence protein<br />

function, they are relatively easy to study<br />

feeding on seed (26, 27). Lastly, food supple-<br />

mentation experiments of American reds have<br />

failed to produce increases in litter size or a<br />

second litter, providing strong evidence that the<br />

increased reproductive investment observed in<br />

our study cannot be triggered by increased<br />

energy availability alone (17, 28, 29). We hy-<br />

pothesize that, rather than following a resource-<br />

tracking strategy where reproductive investment<br />

is determined by current resource amounts, re-<br />

productive rates are driven by future fitness<br />

payoffs. During years of low seed production,<br />

competition among juveniles for available re-<br />

sources is intense, and, although litter augmenta-<br />

tion experiments in American reds show that<br />

females are capable of supporting larger litters<br />

(30), they refrain from doing so because off-<br />

spring recruitment is low (30). However, when<br />

mast years occur, competition among juveniles is<br />

reduced, and females produce more offspring,<br />

which successfully recruit into the population<br />

(31, 32). Further, the increased production of young<br />

by females in mast years does not come with<br />

any obvious cost to the female, because over-<br />

winter survival is not reduced after years of<br />

increased reproductive investment (for Amer-<br />

ican reds, offspring production in the previous<br />

year versus proportion of adult females surviving<br />

to spring has slope = –0.023 ± 0.034, t 15 = –0.7,<br />

and P = 0.51; for Eurasian reds, proportion of<br />

estrous females in the previous year versus adult<br />

female survival to spring has slope = 0.37 ± 0.27,<br />

t 15 = 1.4, and P = 0.19).<br />

If masting has evolved as a swamp-and-<br />

starve adaptation against seed predation, then<br />

anticipatory reproduction and population growth<br />

represent a potent counteradaptation by the pred-<br />

ators. Given that increased reproductive output<br />

in these systems coincides with low current but<br />

high future resources, our results suggest that<br />

reproductive investment in these systems is more<br />

responsive to future fitness prospects than present<br />

energetic constraints. The evolution of seed mast-<br />

ing in trees is also driven by the survival pros-<br />

pects for progeny rather than simple resource<br />

tracking, suggesting an intriguing parallel in re-<br />

productive strategies of trees and the predators<br />

that consume their seed.<br />

References and Notes<br />

1. J. L. Gittleman, S. D. Thompson, Am. Zool. 28, 863<br />

(1988).<br />

2. T. H. Clutton-Brock, Reproductive Success (Univ. of<br />

Chicago Press, Chicago, 1988).<br />

3. T. H. Clutton-Brock, The Evolution of Parental Care<br />

(Princeton Univ. Press, Princeton, NJ, 1999).<br />

4. G. Caughley, in Theoretical Ecology, R. M. May, Ed.<br />

(Blackwell, Oxford, 1981), ch. 6.<br />

5. P. Bayliss, D. Choquenot, Proc. R. Soc. London Ser. B<br />

357, 1233 (2002).<br />

6. C. G. Jones, R. S. Ostfeld, M. P. Richard, E. M. Schauber,<br />

J. O. Wolff, <strong>Science</strong> 279, 1023 (1998).<br />

7. R. S. Ostfeld, F. Keesing, Trends Ecol. Evol. 15, 232 (2000).<br />

8. D. Kelly, V. L. Sork, Annu. Rev. Ecol. Syst. 33, 427<br />

(2002).<br />

12. A. F. Hedlin, For. Sci. 10, 124 (1964).<br />

13. M. J. McKone, D. Kelly, A. L. Harrison, J. J. Sullivan,<br />

A. J. Cone, N.Z. J. Zool. 28, 89 (2001).<br />

14. P. R. Wilson, B. J. Karl, R. J. Toft, J. R. Beggs, Biol. Conserv.<br />

83, 175 (1998).<br />

15. T. Ruf, J. Fietz, W. Schlund, C. Bieber, Ecology 87, 372<br />

(2006).<br />

16. F. H. Bronson, Reproductive Biology (Univ. of Chicago<br />

Press, Chicago, 1989).<br />

17. Materials and methods are available on <strong>Science</strong> Online.<br />

18. J. O. Wolff, J. Mammal. 77, 850 (1996).<br />

19. A. S. McNeilly, Reprod. Fertil. Dev. 13, 583 (2001).<br />

20. J. D. Ligon, Nature 250, 80 (1974).<br />

21. P. J. Berger, N. C. Negus, E. H. Sanders, P. D. Gardner,<br />

<strong>Science</strong> 214, 69 (1981).<br />

22. S. Eis, J. Inkster, Can. J. For. Res. 2, 460 (1972).<br />

23. B. M. Fitzgerald, M. J. Efford, B. J. Karl, N.Z. J. Zool. 31,<br />

167 (2004).<br />

24. M. C. Smith, J. Wildl. Manag. 32, 305 (1968).<br />

25. D. A. Rusch, W. G. Reeder, Ecology 59, 400 (1978).<br />

26. L. A. Wauters, C. Swinnen, A. A. Dhondt, J. Zool. 227, 71<br />

(1992).<br />

27. L. A. Wauters, A. A. Dhondt, J. Zool. 217, 93 (1989).<br />

28. C. D. Becker, Can. J. Zool. 71, 1326 (1993).<br />

29. K. W. Larsen, C. D. Becker, S. Boutin, M. Blower,<br />

J. Mammal. 78, 192 (1997).<br />

32. L. A. Wa<br />

J. Anim.<br />

33. We than<br />

M. Andru<br />

B. Thom<br />

meeting<br />

very use<br />

workers<br />

E. Ander<br />

Financed<br />

Council<br />

Action o<br />

Training<br />

Commiss<br />

is public<br />

Supporting<br />

www.sciencem<br />

Materials and<br />

SOM Text<br />

Fig. S1<br />

Table S1<br />

References<br />

25 September<br />

10.1126/scien<br />

Human Catechol-O-Methy<br />

Haplotypes Modulate Prot<br />

by Altering mRNA Secon<br />

A. G. Nackley, 1 S. A. Shabalina, 2 I. E. Tchivileva, 1 K. Satte<br />

S. S. Makarov, 4 W. Maixner, 1 L. Diatchenko 1 *<br />

Catechol-O-methyltransferase (COMT) is a key regulator of pai<br />

affective mood. Three common haplotypes of the human COM<br />

and one nonsynonymous position, code for differences in CO<br />

associated with pain sensitivity. Haplotypes divergent in synon<br />

difference in COMT enzymatic activity, due to a reduced amo<br />

COMT haplotypes varied with respect to messenger RNA loca<br />

most stable structure was associated with the lowest protein<br />

directed mutagenesis that eliminated the stable structure res<br />

protein. These data highlight the functional significance of sy<br />

importance of haplotypes over single-nucleotide polymorphis<br />

T<br />

he ability to predict the downstream<br />

effects of genetic variation is critically<br />

important for understanding both the<br />

evolution of the genome and the molecular<br />

basis of human disease. The effects of non-<br />

synonymous polymorphisms have been wide-<br />

ly characterized; because these variations<br />

directly in<br />

relatively<br />

experimen<br />

polymorph<br />

which are<br />

be proble<br />

mechanism<br />

cathechol-<br />

regulate g<br />

COMT<br />

grading ca<br />

critical com<br />

(3). The<br />

distinct pr<br />

and mem<br />

through th<br />

1 Center for Neurosensory Disorders, University of North<br />

Carolina, Chapel Hill, NC 27599, USA. 2 National Center for<br />

Biotechnology Information, National Institutes of Health,<br />

Bethesda, MD 20894, USA. 3 Thurston Arthritis Center,<br />

University of North Carolina, Chapel Hill, NC 27599, USA.<br />

4 Attagene, Inc., 7030 Kit Creek Road, Research Triangle Park,<br />

NC 27560, USA.<br />

*To whom correspondence should be addressed. E-mail:<br />

lbdiatch@email.unc.edu<br />

22 DECEMBER 2006 VOL 314 SCIENCE www.sciencemag.org<br />

1930<br />

Ire1 than<br />

elf prim-<br />

r already<br />

fessional<br />

s of the<br />

high con-<br />

rise (and<br />

roportion<br />

refore is<br />

cofactor<br />

te. It is<br />

es ineffi-<br />

declines,<br />

e face of<br />

onforma-<br />

well time<br />

uld have<br />

to moni-<br />

d couple<br />

UPR.<br />

eins con-<br />

findings<br />

ations of<br />

to Ire1,<br />

owed C-<br />

(29, 30).<br />

that participates in activation and regulation<br />

of the RNase function. Insights gained from<br />

Ire1 and RNase L may extend to other pro-<br />

teins containing kinase or enzymatically in-<br />

active pseudokinase domains (32).<br />

References and Notes<br />

1. M. J. Gething, J. Sambrook, Nature 355, 33 (1992).<br />

2. F. J. Stevens, Y. Argon, Semin. Cell Dev. Biol. 10, 443<br />

(1999).<br />

3. C. Patil, P. Walter, Curr. Opin. Cell Biol. 13, 349 (2001).<br />

4. J. S. Cox, C. E. Shamu, P. Walter, Cell 73, 1197 (1993).<br />

5. J. S. Cox, P. Walter, Cell 87, 391 (1996).<br />

6. U. Ruegsegger, J. H. Leber, P. Walter, Cell 107, 103<br />

(2001).<br />

7. C. Sidrauski, P. Walter, Cell 90, 1031 (1997).<br />

8. C. Sidrauski, J. S. Cox, P. Walter, Cell 87, 405 (1996).<br />

9. C. E. Shamu, P. Walter, EMBO J. 15, 3028 (1996).<br />

10. A. Weiss, J. Schlessinger, Cell 94, 277 (1998).<br />

11. K. Mori, W. Ma, M. J. Gething, J. Sambrook, Cell 74,<br />

743 (1993).<br />

12. K. Shah, Y. Liu, C. Deirmengian, K. M. Shokat, Proc.<br />

Natl. Acad. Sci. U.S.A. 94, 3565 (1997).<br />

13. A. C. Bishop et al., Curr. Biol. 8, 257 (1998).<br />

14. E. L. Weiss, A. C. Bishop, K. M. Shokat, D. G. Drubin,<br />

Nature Cell Biol. 2, 677 (2000).<br />

15. A. C. Bishop et al., Nature 407, 395 (2000).<br />

16. Single-letter abbreviations for the amino acid resi-<br />

dues are as follows: A, Ala; D, Asp; G, Gly; L, Leu; and<br />

S, Ser.<br />

17. A. S. Carroll, A. C. Bishop, J. L. DeRisi, K. M. Shokat, E. K.<br />

O’Shea, Proc. Natl. Acad. Sci. U.S.A. 98, 12578 (2001).<br />

25. T. Schindler et al., Mol. Cell 3, 639 (1999).<br />

26. M. Porter, T. Schindler, J. Kuriyan, W. T. Miller, J. Biol.<br />

Chem. 275, 2721 (2000).<br />

27. F. Schuit, K. Moens, H. Heimberg, D. Pipeleers, J. Biol.<br />

Chem. 274, 32803 (1999).<br />

28. R. J. Kaufman et al., Nature Rev. Mol. Cell Biol. 3, 411<br />

(2002).<br />

29. B. Dong, M. Niwa, P. Walter, R. H. Silverman, RNA 7,<br />

361 (2001).<br />

30. S. Naik, J. M. Paranjape, R. H. Silverman, Nucleic Acids<br />

Res. 26, 1522 (1998).<br />

31. B. Dong, R. H. Silverman, Nucleic Acids Res. 27, 439 (1999).<br />

32. M. Kroiher, M. A. Miller, R. E. Steele, Bioessays 23, 69<br />

(2001).<br />

33. We thank M. Stark for constructing the IRE1 C allele, S.<br />

Ullrich for his initial effort in constructing Ire1 mu-<br />

tants, and J. Kuriyan and members of the Walter and<br />

Shokat labs for valuable discussions. Supported by<br />

grants from the NIH to P.W. (GM32384) and K.S.<br />

(AI44009), and postdoctoral support from the UCSF<br />

Molecular Medicine Program (to F.P.). P.W. receives<br />

support as an investigator of the Howard Hughes<br />

Medical Institute.<br />

Supporting Online Material<br />

www.sciencemag.org/cgi/content/full/1090031/DC1<br />

Materials and Methods<br />

Microarray Excel Spreadsheets S1 to S3<br />

References and Notes<br />

4 August 2003; accepted 1 October 2003<br />

Published online 16 October 2003;<br />

10.1126/science.1090031<br />

Include this information when citing this paper.<br />

REPORTS<br />

ion of the Inverse<br />

ppler Effect<br />

ddon* and T. Bearpark<br />

ervation of an inverse Doppler shift, in which the<br />

ased on reflection from a receding boundary. This<br />

een produced by reflecting a wave from a moving<br />

transmission line. Doppler shifts produced by this<br />

eproducible manner by electronic control of the<br />

ically five orders of magnitude greater than those<br />

with kinematic velocities. Potential applications<br />

tunable and multifrequency radiation sources.<br />

wn phe-<br />

a wave is<br />

locity of<br />

the source and the observer (1, 2). Our con-<br />

ventional understanding of the Doppler ef-<br />

fect, from the schoolroom to everyday expe-<br />

rience of passing vehicles, is that increased<br />

frequencies are measured when a source and<br />

observer approach each other. Applications<br />

of the effect are widely established and in-<br />

clude radar, laser vibrometry, blood flow<br />

measurement, and the search for new astro-<br />

nomical objects. The inverse Doppler effect<br />

refers to frequency shifts that are in the op-<br />

posite sense to those described above; for<br />

example, increased frequencies would be<br />

measured on reflection of waves from a re-<br />

ceding boundary. Demonstration of this<br />

counterintuitive phenomenon requires a fun-<br />

damental change in the way that radiation is<br />

reflected from a moving boundary, and, de-<br />

spite a wide range of theoretical work that<br />

spans the past 60 years, the effect has not<br />

previously been verified experimentally.<br />

Although inverse Doppler shifts have<br />

been predicted to occur in particular disper-<br />

sive media (3–8) and in the near zone of<br />

three-dimensional dipoles (9–11), these<br />

schemes are difficult to implement experi-<br />

mentally and have not yet been realized.<br />

However, recent advances in the design of<br />

composite condensed media (metamaterials)<br />

offers new and exciting possibilities for the<br />

control of radiation. In particular, materials<br />

with a negative refractive index (NRI) (12–<br />

14) and the use of shock discontinuities in<br />

photonic crystals (15) and transmission lines<br />

, Advanced<br />

fice Box 5,<br />

dressed. E-<br />

www.sciencemag.org SCIENCE VOL 302 28 NOVEMBER 2003 1537<br />

Ire1 than<br />

elf prim-<br />

r already<br />

fessional<br />

s of the<br />

high con-<br />

rise (and<br />

roportion<br />

refore is<br />

cofactor<br />

te. It is<br />

es ineffi-<br />

declines,<br />

e face of<br />

onforma-<br />

well time<br />

uld have<br />

to moni-<br />

d couple<br />

UPR.<br />

eins con-<br />

findings<br />

ations of<br />

to Ire1,<br />

owed C-<br />

(29, 30).<br />

that participates in activation and regulation<br />

of the RNase function. Insights gained from<br />

Ire1 and RNase L may extend to other pro-<br />

teins containing kinase or enzymatically in-<br />

active pseudokinase domains (32).<br />

References and Notes<br />

1. M. J. Gething, J. Sambrook, Nature 355, 33 (1992).<br />

2. F. J. Stevens, Y. Argon, Semin. Cell Dev. Biol. 10, 443<br />

(1999).<br />

3. C. Patil, P. Walter, Curr. Opin. Cell Biol. 13, 349 (2001).<br />

4. J. S. Cox, C. E. Shamu, P. Walter, Cell 73, 1197 (1993).<br />

5. J. S. Cox, P. Walter, Cell 87, 391 (1996).<br />

6. U. Ruegsegger, J. H. Leber, P. Walter, Cell 107, 103<br />

(2001).<br />

7. C. Sidrauski, P. Walter, Cell 90, 1031 (1997).<br />

8. C. Sidrauski, J. S. Cox, P. Walter, Cell 87, 405 (1996).<br />

9. C. E. Shamu, P. Walter, EMBO J. 15, 3028 (1996).<br />

10. A. Weiss, J. Schlessinger, Cell 94, 277 (1998).<br />

11. K. Mori, W. Ma, M. J. Gething, J. Sambrook, Cell 74,<br />

743 (1993).<br />

12. K. Shah, Y. Liu, C. Deirmengian, K. M. Shokat, Proc.<br />

Natl. Acad. Sci. U.S.A. 94, 3565 (1997).<br />

13. A. C. Bishop et al., Curr. Biol. 8, 257 (1998).<br />

14. E. L. Weiss, A. C. Bishop, K. M. Shokat, D. G. Drubin,<br />

Nature Cell Biol. 2, 677 (2000).<br />

15. A. C. Bishop et al., Nature 407, 395 (2000).<br />

16. Single-letter abbreviations for the amino acid resi-<br />

dues are as follows: A, Ala; D, Asp; G, Gly; L, Leu; and<br />

S, Ser.<br />

17. A. S. Carroll, A. C. Bishop, J. L. DeRisi, K. M. Shokat, E. K.<br />

O’Shea, Proc. Natl. Acad. Sci. U.S.A. 98, 12578 (2001).<br />

25. T. Schindler et al., Mol. Cell 3, 639 (1999).<br />

26. M. Porter, T. Schindler, J. Kuriyan, W. T. Miller, J. Biol.<br />

Chem. 275, 2721 (2000).<br />

27. F. Schuit, K. Moens, H. Heimberg, D. Pipeleers, J. Biol.<br />

Chem. 274, 32803 (1999).<br />

28. R. J. Kaufman et al., Nature Rev. Mol. Cell Biol. 3, 411<br />

(2002).<br />

29. B. Dong, M. Niwa, P. Walter, R. H. Silverman, RNA 7,<br />

361 (2001).<br />

30. S. Naik, J. M. Paranjape, R. H. Silverman, Nucleic Acids<br />

Res. 26, 1522 (1998).<br />

31. B. Dong, R. H. Silverman, Nucleic Acids Res. 27, 439 (1999).<br />

32. M. Kroiher, M. A. Miller, R. E. Steele, Bioessays 23, 69<br />

(2001).<br />

33. We thank M. Stark for constructing the IRE1 C allele, S.<br />

Ullrich for his initial effort in constructing Ire1 mu-<br />

tants, and J. Kuriyan and members of the Walter and<br />

Shokat labs for valuable discussions. Supported by<br />

grants from the NIH to P.W. (GM32384) and K.S.<br />

(AI44009), and postdoctoral support from the UCSF<br />

Molecular Medicine Program (to F.P.). P.W. receives<br />

support as an investigator of the Howard Hughes<br />

Medical Institute.<br />

Supporting Online Material<br />

www.sciencemag.org/cgi/content/full/1090031/DC1<br />

Materials and Methods<br />

Microarray Excel Spreadsheets S1 to S3<br />

References and Notes<br />

4 August 2003; accepted 1 October 2003<br />

Published online 16 October 2003;<br />

10.1126/science.1090031<br />

Include this information when citing this paper.<br />

REPORTS<br />

ion of the Inverse<br />

ppler Effect<br />

ddon* and T. Bearpark<br />

ervation of an inverse Doppler shift, in which the<br />

ased on reflection from a receding boundary. This<br />

een produced by reflecting a wave from a moving<br />

transmission line. Doppler shifts produced by this<br />

eproducible manner by electronic control of the<br />

ically five orders of magnitude greater than those<br />

with kinematic velocities. Potential applications<br />

tunable and multifrequency radiation sources.<br />

wn phe-<br />

a wave is<br />

locity of<br />

the source and the observer (1, 2). Our con-<br />

ventional understanding of the Doppler ef-<br />

fect, from the schoolroom to everyday expe-<br />

rience of passing vehicles, is that increased<br />

frequencies are measured when a source and<br />

observer approach each other. Applications<br />

of the effect are widely established and in-<br />

clude radar, laser vibrometry, blood flow<br />

measurement, and the search for new astro-<br />

nomical objects. The inverse Doppler effect<br />

refers to frequency shifts that are in the op-<br />

posite sense to those described above; for<br />

example, increased frequencies would be<br />

measured on reflection of waves from a re-<br />

ceding boundary. Demonstration of this<br />

counterintuitive phenomenon requires a fun-<br />

damental change in the way that radiation is<br />

reflected from a moving boundary, and, de-<br />

spite a wide range of theoretical work that<br />

spans the past 60 years, the effect has not<br />

previously been verified experimentally.<br />

Although inverse Doppler shifts have<br />

been predicted to occur in particular disper-<br />

sive media (3–8) and in the near zone of<br />

three-dimensional dipoles (9–11), these<br />

schemes are difficult to implement experi-<br />

mentally and have not yet been realized.<br />

However, recent advances in the design of<br />

composite condensed media (metamaterials)<br />

offers new and exciting possibilities for the<br />

control of radiation. In particular, materials<br />

with a negative refractive index (NRI) (12–<br />

14) and the use of shock discontinuities in<br />

photonic crystals (15) and transmission lines<br />

, Advanced<br />

fice Box 5,<br />

dressed. E-<br />

www.sciencemag.org SCIENCE VOL 302 28 NOVEMBER 2003 1537


A common SNP in codon 158 (val met) notype. Because variation in the S-COMT for the s<br />

has been associated with pain ratings and promoter region does not contribute to pain structure,<br />

longest, m<br />

in the va<br />

Fig. 1. Common haplotypes<br />

of the human COMT<br />

gene differ with respect<br />

to mRNA secondary structure<br />

and enzymatic activity.<br />

(A) A schematic<br />

diagram illustrates COMT<br />

MB-COM<br />

loop struc<br />

~17 kcal/m<br />

LPS haplo<br />

(Fig. 1B a<br />

porting p<br />

obtained<br />

genomic organization and<br />

ondary str<br />

SNP composition for the<br />

of COMT<br />

three haplotypes. Per-<br />

species (f<br />

cent frequency of each<br />

structures<br />

haplotype in a cohort of<br />

healthy Caucasian females,<br />

and percent independent<br />

SNP contribution<br />

to pain sensitivity, are<br />

indicated. (B) The local<br />

stem-loop structures associated<br />

with each of<br />

the three haplotypes are<br />

shown. Relative to the<br />

LPS and APS haplotypes,<br />

the HPS local stem-loop<br />

structure had a higher<br />

folding potential. (C and<br />

D) The LPS haplotype exhibited<br />

the highest, while<br />

the HPS haplotype exhib-<br />

highly sta<br />

gous to t<br />

stantial d<br />

observed<br />

notable f<br />

studies w<br />

modeling<br />

We co<br />

cDNA cl<br />

tors that<br />

correspon<br />

lotypes (<br />

were tran<br />

six constr<br />

tein expr<br />

measured<br />

ited the lowest enzymatic<br />

HPS hap<br />

activity and protein lev-<br />

reduction<br />

els in cells expressing<br />

MB-COM<br />

COMT. ***P < 0.001, ≠<br />

LPS.<br />

and fig. S<br />

ited mark<br />

protein ex<br />

APS hapl<br />

fold redu<br />

MB-COM<br />

+++ A common SNP in codon 158 (val met) notype. Because variation in the S-COMT for the<br />

has been associated with pain ratings and promoter region does not contribute to pain structur<br />

longest<br />

in the<br />

Fig. 1. Common haplotypes<br />

of the human COMT<br />

gene differ with respect<br />

to mRNA secondary structure<br />

and enzymatic activity.<br />

(A) A schematic<br />

diagram illustrates COMT<br />

MB-CO<br />

loop stru<br />

~17 kca<br />

LPS hap<br />

(Fig. 1B<br />

porting<br />

obtained<br />

genomic organization and<br />

ondary<br />

SNP composition for the<br />

of COM<br />

three haplotypes. Per-<br />

species<br />

cent frequency of each<br />

structur<br />

haplotype in a cohort of<br />

healthy Caucasian females,<br />

and percent independent<br />

SNP contribution<br />

to pain sensitivity, are<br />

indicated. (B) The local<br />

stem-loop structures associated<br />

with each of<br />

the three haplotypes are<br />

shown. Relative to the<br />

LPS and APS haplotypes,<br />

the HPS local stem-loop<br />

structure had a higher<br />

folding potential. (C and<br />

D) The LPS haplotype exhibited<br />

the highest, while<br />

the HPS haplotype exhib-<br />

highly<br />

gous to<br />

stantial<br />

observe<br />

notable<br />

studies<br />

modelin<br />

We c<br />

cDNA<br />

tors th<br />

corresp<br />

lotypes<br />

were tr<br />

six con<br />

tein ex<br />

measure<br />

ited the lowest enzymatic<br />

HPS h<br />

activity and protein lev-<br />

reductio<br />

els in cells expressing<br />

MB-CO<br />

COMT. ***P < 0.001, ≠<br />

LPS. P < 0.001, ≠ APS.<br />

and fig<br />

ited ma<br />

protein<br />

APS ha<br />

fold red<br />

MB-CO<br />

+++ P < 0.001, ≠ APS.<br />

group demonstrated that three common<br />

haplotypes of the human COMT gene are<br />

associated with pain sensitivity and the<br />

likelihood of developing temporomandibular<br />

joint disorder (TMJD), a common<br />

chronic musculoskeletal pain condition<br />

(4). Three major haplotypes are formed<br />

by four single-nucleotide polymorphisms<br />

(SNPs): one located in the S-COMT promoter<br />

region (A/G; rs6269) and three in<br />

the S- and MB-COMT coding region at<br />

codons his62his (C/T; rs4633), leu136leu (C/<br />

G; rs4818), and val158 ed in regulatory regions, which www.sciencemag.org are www.sciencemag.org much<br />

SCIENCE VOL VOL314 314 22 22 DECEMBER DECEMBER200 20<br />

more common, has proved to be problematic<br />

(2). Here, we focus on the mechanism<br />

whereby polymorphisms of the cathechol-<br />

O-methyltransferase (COMT) gene regulate<br />

gene expression.<br />

COMT is an enzyme responsible<br />

for degrading catecholamines and thus<br />

represents a critical component of homeostasis<br />

maintenance (3). The human<br />

COMT gene encodes two distinct proteins:<br />

soluble COMT (S-COMT) and<br />

membrane-bound COMT (MB-COMT)<br />

met (A/G; rs4680)<br />

through the use of alternative translation (Fig. 1A). On the basis of subjects’ pain<br />

initiation sites and promoters (3). Re- responsiveness, haplotypes were designatcently,<br />

COMT has been implicated in the ed as low (LPS; GCGG), average (APS;<br />

modulation of persistent pain (4–7). Our ATCA), or high (HPS; ACCG) pain sen-<br />

13


1932<br />

ture and converts convertsit it into a LPS haplotype–like<br />

structure (HPS Lsm). Double mutation of<br />

Fig. 2. Site-directed mutagenesis<br />

that destroys<br />

the stable stem-loop structure<br />

corresponding to to the<br />

HPS haplotype restores<br />

COMT enzymatic activity<br />

and protein expression.<br />

(A) The mRNA structure<br />

corresponding to to the HPS<br />

haplotype was converted<br />

to to an LPS haplotype-like<br />

structure (HPS Lsm) by<br />

single mutation of of 403C<br />

to to G. The original HPS<br />

haplotype structure (HPS<br />

dm) was restored by double<br />

mutation of of interacting<br />

nucleotides 403C 403Cto to G<br />

and 479G 479Gto toC. C. (B and C)<br />

The HPS Lsm exhibited<br />

COMT enzymatic activity<br />

and protein levels equivalentalentto<br />

to those thoseof of the LPS<br />

haplotype, whereas the<br />

HPS dm exhibited reduced<br />

enzymatic activity.<br />

***P < 0.001, ≠ LPS.<br />

sitive. Individuals<br />

carrying HPS/APS<br />

or APS/APS diplotypes<br />

were nearly<br />

2.5 times as likely<br />

to develop TMJD.<br />

Previous data further<br />

suggest that<br />

COMT haplotypes code for differences in<br />

COMT enzymatic activity (4); however,<br />

the molecular mechanisms involved have<br />

remained unknown.<br />

A common SNP in codon 158 (val158met) has been associated with pain ratings and<br />

μ-opioid system responses (7) as well as<br />

addiction, cognition, and common affective<br />

disorders (3, 8–10). The substitution<br />

of valine (Val) by methionine (Met) results<br />

in reduced thermostability and activity of<br />

the enzyme (3). It is generally accepted<br />

that val158met is the main source of individual<br />

variation in human COMT activity;<br />

numerous studies have identified associations<br />

between the low-activity met allele<br />

and several complex phenotypes (3, 8, 10).<br />

14<br />

verified by an alternate approach of modifying<br />

mRNA secondary structure (fig. S5).<br />

However, observed associations between<br />

these conditions and the met allele are often<br />

modest and occasionally inconsistent<br />

(3). This suggests that additional SNPs in<br />

the COMT gene modulate COMT activity.<br />

Indeed, we found that haplotype rather<br />

than an individual SNP better accounts for<br />

variability in pain sensitivity (4). The HPS<br />

and LPS haplotypes that both code for the<br />

stable val 158 variant were associated with<br />

the two extreme-pain phenotypes; thus,<br />

the effect of haplotype on pain sensitivity<br />

in our study cannot be explained by<br />

the sum of the effects of functional SNPs.<br />

Instead, the val 158 met SNP interacts with<br />

other SNPs to determine phenotype. Because<br />

variation in the S-COMT promoter<br />

Refere Refer<br />

1. 1. L. L. Y. Y. Ya Y<br />

Hum. M<br />

2. 2. J. J. C. C. Kn K<br />

3. 3. P. P. T. T. M<br />

(1999) (1999<br />

4. 4. L. L. Diat Dia<br />

(2005) (2005<br />

5. 5. J. J. J. J. Ma M<br />

(1976) (1976<br />

6. 6. T. T. T. T. Ra R<br />

7. 7. J. J. K. K. Zu Z<br />

8. 8. G. G. Win<br />

134 (2 (2<br />

9. 9. B. B. Funk Fun<br />

10. G. G. Oros Oro<br />

(2004) (2004<br />

11. J. J. Duan Dua<br />

(2003) (2003<br />

12. L. L. X. X. Sh S<br />

Acad. S<br />

13. I. I. Puga Pug<br />

14. K. K. Mita Mit<br />

203, 9<br />

15. T. T. D. D. S<br />

H. H. J. J. Le L<br />

(1994) (1994<br />

16. A. A. Shal Sha<br />

(2002) (2002<br />

17. Materia Materi<br />

materia materi<br />

18. D. D. H. H. M<br />

Biol. 2<br />

19. M. M. Zuk<br />

20. A. A. Y. Y. O<br />

M. M. A. A. R<br />

21. Sequen Seque<br />

CA4894 CA489<br />

haploty haplot<br />

three threeh<br />

site sitewe we<br />

enzyme enzym<br />

three threeS<br />

through throug<br />

(BI835<br />

ends endsas a<br />

22. S. S. A. A. Sh S<br />

Acids AcidsR<br />

22 DECEMBER 2006 VOL 314 SCIENCE www.sciencemag.or<br />

www.sciencemag.o


egion does not contribute to pain phenotype<br />

(Fig. 1A), we suggest that the rate of<br />

mRNA degradation or protein synthesis<br />

is affected by the structural properties of<br />

the haplotypes, such as haplotype-specific<br />

mRNA secondary structure.<br />

Previous reports have shown that<br />

polymorphic alleles can markedly affect<br />

mRNA secondary structure (11, 12), which<br />

can then have functional consequences on<br />

the rate of mRNA degradation (11, 13).<br />

It is also plausible that polymorphic alleles<br />

directly modulate protein translation<br />

through alterations in mRNA secondary<br />

structure, because protein translation efficiency<br />

is affected by mRNA secondary<br />

structure (14–16). To test these possibilities,<br />

we evaluated the effect of LPS, APS,<br />

and HPS haplotypes on the stability of the<br />

corresponding mRNA secondary structures<br />

(17).<br />

Secondary structures of the full-length<br />

LPS, APS, and HPS mRNA transcripts<br />

were predicted by means of the RNA<br />

Mfold (18, 19) and Afold (20) programs.<br />

The mRNA folding analyses demonstrated<br />

that the major COMT haplotypes<br />

differ with respect to mRNA secondary<br />

structure. The LPS haplotype codes for<br />

the shortest, least stable local stem-loop<br />

structure, and the HPS haplotype codes<br />

for the longest, most stable local stemloop<br />

structure in the val 158 region for both<br />

S-COMT and MB-COMT. Gibbs free energy<br />

(ΔG) for the stem-loop structure associated<br />

with the HPS haplotype is ~17<br />

kcal/mol less than that associated with<br />

the LPS haplotype for both S-COMT and<br />

MB-COMT (Fig. 1B and fig. S1A). Additional<br />

evidence supporting predicted<br />

RNA folding structures was obtained by<br />

generating consensus RNA secondary<br />

structures based on comparative analysis<br />

of COMT sequences from eight mammalian<br />

species (fig. S2). The consensus RNA<br />

folding structures were LPS-like and did<br />

not contain highly stable local stem-loop<br />

structures analogous to the human HPSlike<br />

form. Thus, substantial deviation<br />

from consensus structure, as observed for<br />

the HPS haplotype, should have notable<br />

functional consequences. Additional studies<br />

were conducted to test this molecular<br />

modeling.<br />

We constructed full-length S- and MB-<br />

COMT cDNA clones in mammalian expression<br />

vectors that differed only in three<br />

nucleotides corresponding to the LPS, APS,<br />

and HPS haplotypes (17, 21). Rat adrenal<br />

(PC-12) cells were transiently transfected<br />

with each of these six constructs. COMT<br />

enzymatic activity, protein expression, and<br />

mRNA abundance were measured. Relative<br />

to the LPS haplotype, the HPS haplotype<br />

showed a 25- and 18-fold reduction in<br />

enzymatic activity for S- and MB-COMT<br />

constructs, respectively (Fig. 1C and fig.<br />

S1B). The HPS haplotype also exhibited<br />

marked reductions in S- and MB-COMT<br />

protein expression (Fig. 1D and fig. S1C).<br />

The APS haplotype displayed a moderate<br />

2.5- and 3-fold reduction in enzymatic<br />

activity for S- and MB-COMT constructs,<br />

respectively, while protein expression levels<br />

did not differ. The moderate reduction<br />

in enzymatic activity produced by the APS<br />

haplotype is most likely due to the previously<br />

reported decrease in protein thermostability<br />

coded by the met 158 allele (3).<br />

These data illustrate that the reduced enzymatic<br />

activity corresponding to the HPS<br />

haplotype is paralleled by reduced protein<br />

levels, an effect that could be mediated by<br />

local mRNA secondary structure at the level<br />

of protein synthesis and/or mRNA degradation.<br />

Because total RNA abundance<br />

and RNA degradation rates did not parallel<br />

COMT protein levels (fig. S2), differences<br />

in protein translation efficiency likely results<br />

from differences in the local secondary<br />

structure of corresponding mRNAs<br />

To directly assess this hypothesis, we<br />

performed site-directed mutagenesis (17).<br />

The stable stem-loop structure of S- and<br />

MB-COMT mRNA corresponding to the<br />

HPS haplotype is supported by base pairs<br />

between several critical nucleotides, including<br />

403C and 479G in S-COMT and<br />

15


n thermo-<br />

3). These<br />

ic activity<br />

paralleled<br />

t could be<br />

tructure at<br />

r mRNA<br />

dance and<br />

el COMT<br />

n protein<br />

m differe<br />

of cor-<br />

esis, we<br />

sis (17).<br />

f S- and<br />

the HPS<br />

between<br />

ng 403C<br />

701G in<br />

Mutation<br />

to G in<br />

op structype–like<br />

tation of<br />

625C and 701G in MB-COMT (Fig. 2A<br />

and fig. S3A). Mutation of 403C to G<br />

in S-COMT or 625C to G in MB-COMT<br />

destroys the stable stem-loop structure<br />

and converts it into a LPS haplotype–like<br />

structure (HPS Lsm). Double mutation of<br />

mRNA in position 403C to G and 479G to<br />

C in S-COMT or 625C to G and 701G to<br />

C in MB-COMT reconstructs the original<br />

long stem-loop structure (HPS dm). The<br />

single- and double-nucleotide HPS mutants<br />

(HPS Lsm and HPS dm, respectively)<br />

were transiently transfected to PC-12<br />

cells. As predicted by the mRNA secondary-structure<br />

folding analyses, the HPS<br />

Lsm exhibited increased COMT enzymatic<br />

activity and protein levels equivalent to<br />

those of the LPS haplotype, whereas the<br />

HPS dm exhibited reduced enzymatic<br />

activity and protein levels equivalent to<br />

those of the original HPS haplotype (Fig.<br />

2, B and C; fig. S3, B and C). These data<br />

rule out the involvement of RNA sequence<br />

recognition motifs or codon usage in the<br />

regulation of translation. In contrast to<br />

the HPS haplotype, protein levels did not<br />

parallel COMT enzymatic activity for the<br />

APS haplotype and site-directed mutagenesis<br />

confirmed that the met158 structure (HPS dm). The single- and doublenucleotide<br />

HPS mutants (HPS Lsm and HPS<br />

dm, respectively) were transiently transfected to<br />

PC-12 cells. As predicted by the mRNA<br />

secondary-structure folding analyses, the HPS<br />

Lsm exhibited increased COMT enzymatic<br />

activity and protein levels equivalent to those<br />

of the LPS haplotype, whereas the HPS dm<br />

exhibited reduced enzymatic activity and protein<br />

levels equivalent to those of the original<br />

HPS haplotype (Fig. 2, B and C; fig. S3, B and<br />

C). These data rule out the involvement of RNA<br />

sequence recognition motifs or codon usage in<br />

the regulation of translation. In contrast to the<br />

HPS haplotype, protein levels did not parallel<br />

COMT enzymatic activity for the APS haplotype<br />

and site-directed mutagenesis confirmed that the<br />

met<br />

allele,<br />

not a more stable mRNA secondary structure,<br />

drives the reduced enzymatic activity<br />

observed for the APS haplotype (fig.<br />

S4). This difference is moderate relative<br />

to the mRNA structure–dependent difference<br />

coded by LPS and HPS haplotypes.<br />

These results were verified by an alternate<br />

approach of modifying mRNA secondary<br />

structure (fig. S5).<br />

Our data have very broad evolutionary<br />

and medical implications for the analysis of<br />

variants common in the human population.<br />

The fact that alterations in mRNA secondary<br />

structure resulting from synonymous<br />

changes have such a pronounced effect on<br />

the level of protein expression emphasizes<br />

the critical role of synonymous nucleotide<br />

positions in maintaining mRNA secondary<br />

structure and suggests that the mRNA<br />

secondary structure, rather than indepen-<br />

158 allele, not a more stable mRNA secondary<br />

structure, drives the reduced enzymatic<br />

activity observed for the APS haplotype (fig.<br />

S4). This difference is moderate relative to the<br />

mRNA structure–dependent difference coded by<br />

LPS and HPS haplotypes. These results were<br />

verified by an alternate approach of modifying<br />

mRNA secondary structure (fig. S5).<br />

2 DECEMBER 162006<br />

VOL 314 SCIENCE www.sciencemag.org<br />

The fact that alterations in mRNA secondary<br />

structure resulting from synonymous changes<br />

have such a pronounced effect on the level of<br />

protein expression emphasizes the critical role<br />

of synonymous nucleotide positions in maintaining<br />

dent nucleotides mRNA secondary in the synonymous structure andposi suggeststions,<br />

that should theundergo mRNA substantial secondary selective structure,<br />

rather pressure than (22). independent Furthermore, nucleotides our data in stress the<br />

synonymous the importance positions, of synonymous should undergo SNPs sub- as<br />

stantial<br />

potential<br />

selective<br />

functional<br />

pressure<br />

variants<br />

(22).<br />

in<br />

Furthermore,<br />

the area of<br />

our data stress the importance of synonymous<br />

SNPs<br />

human<br />

as<br />

medical<br />

potential<br />

genetics.<br />

functional<br />

Although<br />

variants in<br />

non-<br />

the<br />

area synonymous of humanSNPs medical are genetics. believed Although to have<br />

nonsynonymous the strongest impact SNPs on arevariation believed in to gene have<br />

the function, strongest our data impact clearly on variation demonstrate in gene that<br />

function, haplotypic ourvariants data clearly of common demonstrate synony- that<br />

haplotypic variants of common synonymous<br />

mous SNPs can have stronger effects on<br />

SNPs can have stronger effects on gene function<br />

gene than function nonsynonymous than nonsynonymous variations varia- and<br />

play tions an and important play an important role in disease role in onset disease and<br />

progression. onset and progression.<br />

References and Notes<br />

1. L. Y. Yampolsky, F. A. Kondrashov, A. S. Kondrashov,<br />

Hum. Mol. Genet. 14, 3191 (2005).<br />

2. J. C. Knight, J. Mol. Med. 83, 97 (2005).<br />

3. P. T. Mannisto, S. Kaakkola, Pharmacol. Rev. 51, 593<br />

(1999).<br />

4. L. Diatchenko et al., Hum. Mol. Genet. 14, 135<br />

(2005).<br />

5. J. J. Marbach, M. Levitt, J. Dent. Res. 55, 711<br />

(1976).<br />

6. T. T. Rakvag et al., Pain 116, 73 (2005).<br />

7. J. K. Zubieta et al., <strong>Science</strong> 299, 1240 (2003).<br />

8. G. Winterer, D. Goldman, Brain Res. Brain Res. Rev. 43,<br />

134 (2003).<br />

9. B. Funke et al., Behav. Brain Funct. 1, 19 (2005).<br />

10. G. Oroszi, D. Goldman, Pharmacogenomics 5, 1037<br />

(2004).<br />

11. J. Duan et al., Hum. Mol. Genet. 12, 205<br />

(2003).<br />

12. L. X. Shen, J. P. Basilion, V. P. Stanton Jr., Proc. Natl.<br />

Acad. Sci. U.S.A. 96, 7871 (1999).<br />

13. I. Puga et al., Endocrinology 146, 2210 (2005).<br />

14. K. Mita, S. Ichimura, M. Zama, T. C. James, J. Mol. Biol.<br />

203, 917 (1988).<br />

15. T. D. Schmittgen, K. D. Danenberg, T. Horikoshi,<br />

H. J. Lenz, P. V. Danenberg, J. Biol. Chem. 269, 16269<br />

(1994).<br />

16. A. Shalev et al., Endocrinology 143, 2541<br />

(2002).<br />

17. Materials and methods are available as supporting<br />

material on <strong>Science</strong> Online.<br />

18. D. H. Mathews, J. Sabina, M. Zuker, D. H. Turner, J. Mol.<br />

Biol. 288, 911 (1999).<br />

19. M. Zuker, Nucleic Acids Res. 31, 3406 (2003).<br />

20. A. Y. Ogurtsov, S. A. Shabalina, A. S. Kondrashov,<br />

M. A. Roytberg, Bioinformatics 22, 1317 (2006).<br />

21. Sequence accession numbers: S-COMT clones BG290167,<br />

CA489448, and BF037202 represent LPS, APS, and HPS<br />

haplotypes, respectively. Clones corresponding to all<br />

three haplotypes and including the transcriptional start<br />

site were constructed by use of the unique restriction<br />

enzyme Bsp MI. The gene coding region containing all<br />

three SNPs was cut and inserted into the plasmids<br />

through use of the S-COMT (BG290167) and MB-COMT<br />

(BI835796) clones containing the entire COMT 5′ and 3′<br />

ends as a backbone vector.<br />

22. S. A. Shabalina, A. Y. Ogurtsov, N. A. Spiridonov, Nucleic<br />

Acids Res. 34, 2428 (2006).


e of Child<br />

the NIH, National Center for Biotechnology<br />

SOM Text<br />

nal Institute of Information.<br />

Figs. S1 to S6<br />

National<br />

References<br />

arch at the NIH Supporting Online Material<br />

esearch was www.sciencemag.org/cgi/content/full/314/5807/1930/DC1 13 June 2006; accepted 16 November 2006<br />

rch Program of Materials and Methods<br />

10.1126/science.1131262<br />

cidophilic<br />

aled by Community<br />

ysis<br />

ichard I. Webb,<br />

17<br />

3 Judith Flanagan, 2 *<br />

, 2 ‡ Jillian F. Banfield 1,2 motes further pyrite dissolution, leading to AMD<br />

generation. AMD solutions forming underground<br />

in the Richmond Mine at Iron Mountain,<br />

California, are warm (30° to 59°C), acidic (pH<br />

~0.5 to 1.5), metal-rich [submolar Fe<br />

§<br />

cies can be easily overlooked in standard polymerase chain<br />

sed community genomic data obtained without PCR or<br />

ents bearing unusual 16S ribosomal RNA (rRNA) and proteinging<br />

to novel archaeal lineages. The organisms are minor<br />

in pH 0.5 to 1.5 solutions within the Richmond Mine,<br />

RNA showed that the fraction less than 0.45 micrometers in<br />

anisms. Transmission electron microscope images revealed that<br />

ual folded membrane protrusions and have apparent volumes<br />

variety of derived from multispecies consortia (4–6) and<br />

late natural whole environments (7–9) have provided new<br />

ed by the information about diversity and metabolic<br />

in reaction potential. However, PCR-based methods have<br />

nt methods limited ability to detect organisms whose genes<br />

enes (1–3). are significantly divergent relative to gene<br />

e fragments sequences in databases, and most cultivationindependent<br />

genomic sequencing approaches<br />

are relatively insensitive to organisms that<br />

occur at low abundance. Consequently, it is<br />

likely that low-abundance microorganisms distantly<br />

related to known species will be undetected<br />

members of natural consortia, even in low<br />

complexity systems such as acid mine drainage<br />

(AMD) (10).<br />

An important way in which microorganisms<br />

affect geochemical cycles is by accelerating the<br />

dissolution of minerals. For example, microorganisms<br />

can derive metabolic energy by oxidizing<br />

iron released by the dissolution of<br />

pyrite (FeS2). The ferric iron by-product pro-<br />

2+ REPORTS<br />

f Child<br />

the NIH, National Center for Biotechnology<br />

SOM Text<br />

l Institute of Information.<br />

Figs. S1 to S6<br />

tional<br />

References<br />

h at the NIH Supporting Online Material<br />

earch was www.sciencemag.org/cgi/content/full/314/5807/1930/DC1 13 June 2006; accepted 16 November 2006<br />

Program of Materials and Methods<br />

10.1126/science.1131262<br />

idophilic<br />

and micromolar<br />

As and Cu (11)] and host active<br />

led by Community<br />

microbial communities. Extensive cultivationindependent<br />

sequence analysis of functional and<br />

sis<br />

rRNA genes (11, 12) revealed that biofilms contain<br />

a significant number of Archaea, but the<br />

hard I. Webb, diversity reported to date has been limited to the<br />

order Thermoplasmatales (10). Current models<br />

for AMD generation thus include only these<br />

species.<br />

The genomes of the five dominant members<br />

of one biofilm community from the “5-way”<br />

region of the Richmond Mine (fig. S1) were<br />

largely reconstructed through the assembly of<br />

76 Mb of shotgun genomic sequence (4). Previously<br />

unreported is a genome fragment that<br />

encodes part of the 16S rRNA gene of a novel<br />

archaeal lineage: Archaeal Richmond Mine<br />

Acidophilic Nanoorganism (ARMAN-1). Using<br />

an expanded data set that now comprises more<br />

than 100 Mb of genomic sequence, we reconstructed<br />

a contiguous 4.2-kb fragment adjacent<br />

to this gene. A second 13.2-kb genome fragment<br />

encoding a 16S rRNA gene from an organism<br />

that is related to ARMAN-1 (ARMAN-2) was<br />

reconstructed from 117 Mb of community ge-<br />

s, University of<br />

Environmental<br />

nomic sequence derived from a biofilm from the<br />

y of California,<br />

A drift (fig. S1). Within the data sets from each<br />

icroscopy and<br />

site, results to date indicate that each ARMAN<br />

ogy and Para-<br />

population is near-clonal.<br />

isbane 4072,<br />

Comparison of the ARMAN-1 and -2 DNA<br />

San Francisco,<br />

fragments revealed some gene rearrangements,<br />

insertions, and deletions (Fig. 1). Genes present<br />

Joint Genome<br />

in both organisms encode putative inorganic<br />

pyrophosphatases, a transcription regulator, and<br />

Diego, La Jolla,<br />

a gene shown to be an arsenate reductase (13).<br />

essed. E-mail:<br />

Comparative analysis of these genes with sequences<br />

in the public databases consistently<br />

ARMAN-1<br />

pyrophosphatase insert<br />

1<br />

4,226<br />

31% 69% 85% 90% 97%<br />

13,236<br />

3 Judith Flanagan, 2 *<br />

‡ Jillian F. Banfield 1,2 motes further pyrite dissolution, leading to AMD<br />

generation. AMD solutions forming underground<br />

in the Richmond Mine at Iron Mountain,<br />

California, are warm (30° to 59°C), acidic (pH<br />

~0.5 to 1.5), metal-rich [submolar Fe<br />

§<br />

ies can be easily overlooked in standard polymerase chain<br />

d community genomic data obtained without PCR or<br />

nts bearing unusual 16S ribosomal RNA (rRNA) and proteining<br />

to novel archaeal lineages. The organisms are minor<br />

pH 0.5 to 1.5 solutions within the Richmond Mine,<br />

NA showed that the fraction less than 0.45 micrometers in<br />

isms. Transmission electron microscope images revealed that<br />

al folded membrane protrusions and have apparent volumes<br />

ariety of derived from multispecies consortia (4–6) and<br />

te natural whole environments (7–9) have provided new<br />

d by the information about diversity and metabolic<br />

reaction potential. However, PCR-based methods have<br />

t methods limited ability to detect organisms whose genes<br />

nes (1–3). are significantly divergent relative to gene<br />

fragments sequences in databases, and most cultivationindependent<br />

genomic sequencing approaches<br />

are relatively insensitive to organisms that<br />

occur at low abundance. Consequently, it is<br />

likely that low-abundance microorganisms distantly<br />

related to known species will be undetected<br />

members of natural consortia, even in low<br />

complexity systems such as acid mine drainage<br />

(AMD) (10).<br />

An important way in which microorganisms<br />

affect geochemical cycles is by accelerating the<br />

dissolution of minerals. For example, microorganisms<br />

can derive metabolic energy by oxidizing<br />

iron released by the dissolution of<br />

pyrite (FeS2). The ferric iron by-product pro-<br />

2+ 23. We are grateful to the National Institute of Child<br />

the NIH, National Center for Biotechnology<br />

SOM Text<br />

Health and Human Development, National Institute of Information.<br />

Figs. S1 to S6<br />

Neurological Disorders and Stroke, and National<br />

References<br />

Institute of Dental and Craniofacial Research at the NIH Supporting Online Material<br />

for financial support of this work. This research was www.sciencemag.org/cgi/content/full/314/5807/1930/DC1 REPORTS 13 June 2006<br />

also supported by the Intramural Research Program of Materials and Methods<br />

10.1126/scien<br />

Lineages of Acidophilic<br />

Archaea Revealed by Community<br />

Genomic Analysis<br />

Brett J. Baker,<br />

and micromolar<br />

As and Cu (11)] and host active<br />

microbial communities. Extensive cultivationindependent<br />

sequence analysis of functional and<br />

rRNA genes (11, 12) revealed that biofilms contain<br />

a significant number of Archaea, but the<br />

diversity reported to date has been limited to the<br />

order Thermoplasmatales (10). Current models<br />

for AMD generation thus include only these<br />

species.<br />

The genomes of the five dominant members<br />

of one biofilm community from the “5-way”<br />

region of the Richmond Mine (fig. S1) were<br />

largely reconstructed through the assembly of<br />

76 Mb of shotgun genomic sequence (4). Previously<br />

unreported is a genome fragment that<br />

encodes part of the 16S rRNA gene of a novel<br />

archaeal lineage: Archaeal Richmond Mine<br />

Acidophilic Nanoorganism (ARMAN-1). Using<br />

an expanded data set that now comprises more<br />

than 100 Mb of genomic sequence, we reconstructed<br />

a contiguous 4.2-kb fragment adjacent<br />

to this gene. A second 13.2-kb genome fragment<br />

encoding a 16S rRNA gene from an organism<br />

that is related to ARMAN-1 (ARMAN-2) was<br />

reconstructed from 117 Mb of community ge-<br />

University of<br />

vironmental<br />

nomic sequence derived from a biofilm from the<br />

f California,<br />

A drift (fig. S1). Within the data sets from each<br />

roscopy and<br />

site, results to date indicate that each ARMAN<br />

y and Para-<br />

population is near-clonal.<br />

ane 4072,<br />

Comparison of the ARMAN-1 and -2 DNA<br />

n Francisco,<br />

fragments revealed some gene rearrangements,<br />

insertions, and deletions (Fig. 1). Genes present<br />

int Genome<br />

in both organisms encode putative inorganic<br />

pyrophosphatases, a transcription regulator, and<br />

ego, La Jolla,<br />

a gene shown to be an arsenate reductase (13).<br />

sed. E-mail:<br />

Comparative analysis of these genes with sequences<br />

in the public databases consistently<br />

ARMAN-1<br />

pyrophosphatase insert<br />

1<br />

4,226<br />

31% 69% 85% 90% 97%<br />

1 Gene W. Tyson, 2 Richard I. Webb, 3 Judith Flanagan, 2 *<br />

Philip Hugenholtz, 2 † Eric E. Allen, 2 ‡ Jillian F. Banfield 1,2 motes furth<br />

generation<br />

ground in t<br />

California,<br />

~0.5 to 1.5<br />

cromolar<br />

microbial<br />

independen<br />

rRNA gene<br />

§<br />

tain a sign<br />

diversity re<br />

Novel, low-abundance microbial species can be easily overlooked in standard polymerase chain order Ther<br />

reaction (PCR)–based surveys. We used community genomic data obtained without PCR or for AMD<br />

cultivation to reconstruct DNA fragments bearing unusual 16S ribosomal RNA (rRNA) and protein- species.<br />

coding genes from organisms belonging to novel archaeal lineages. The organisms are minor The ge<br />

components of all biofilms growing in pH 0.5 to 1.5 solutions within the Richmond Mine, of one bio<br />

California. Probes specific for 16S rRNA showed that the fraction less than 0.45 micrometers in region of<br />

diameter is dominated by these organisms. Transmission electron microscope images revealed that largely rec<br />

the cells are pleomorphic with unusual folded membrane protrusions and have apparent volumes 76 Mb of<br />

of


of decay from NMD (Fig. 1).<br />

The turnover of normal mRNAs requires<br />

deadenylation followed by Dcp1pmediated<br />

decapping and degradation by<br />

the major 5′-to-3′ exonuclease Xrn1p (6,<br />

7). NMD is distinguished in that these<br />

events occur without prior deadenylation<br />

(8, 9). Nonstop-PGK1 transcripts showed<br />

rapid decay in strains lacking Xrn1p or<br />

Dcp1p activity (Fig. 1), providing further<br />

evidence that they are not subject to<br />

NMD. This result was surprising since<br />

both of these factors are also required<br />

for the turnover of normal mRNAs after<br />

deadenylation by the major deadenylase<br />

Ccr4p (10). Nonstop decay was also un-<br />

Fig. 1. Nonstop-PGK1 transcripts are rapidly<br />

degraded by a novel mechanism. Half-lives<br />

(t 1/2 ) of WT-PGK1, PTC(22)-PGK1, nonstop-<br />

PGK1, and Ter-poly(A)-PGK1 transcripts<br />

are shown (27). Half-lives were performed<br />

in a wild-type yeast strain ( WT ), strains<br />

deleted for Upf1p (upf1Δ), Xrn1p (xrn1Δ), or<br />

Ccr4p (ccr4Δ), a strain lacking activity of the<br />

decapping enzyme Dcp1p (dcp1-2), and a WT<br />

yeast strain treated for 1 hour with 100 μg/ml<br />

of cycloheximide (+CHX ).<br />

18<br />

altered in a strain lacking Ccr4p (Fig. 1).<br />

Therefore, degradation of nonstop-PGK1<br />

transcripts requires none of the factors involved<br />

in the pathway for degradation of<br />

wild-type or nonsense mRNAs.<br />

Additional experiments were performed<br />

to assess the role of translation in nonstop<br />

decay. Treatment with cycloheximide<br />

(CHX) or depletion of charged tRNAs in<br />

yeast harboring the conditional cca1-1 allele<br />

(11) grown at the nonpermissive temperature<br />

substantially increased the stability<br />

of nonstop-PGK1 transcripts (Fig. 1)<br />

(12). It has been shown that translation<br />

into the 3′ UTR of selected transcripts can<br />

displace bound trans-factors that are posi-<br />

tive determinants of message stability (13,<br />

14). To determine whether this mechanism<br />

is relevant to nonstop decay, we examined<br />

into the 3′ UTR of selected transcripts can


major substantial (threefold) stabilization (Fig. 1).<br />

decay Unlike nonstop-PGK1 transcripts, the half-<br />

Ccr4p life of Ter-poly(A)-PGK1 was increased in<br />

onstop- the absence of Xrn1p (Fig. 1), indicating<br />

hefac- that tive this determinants transcript is of degraded message stability by the path- (13,<br />

egrada-way 14). for To determine normal mRNA whether and this not mechanism by the<br />

As. nonstop pathway. These data suggest that<br />

is relevant to nonstop decay, we examined<br />

formed the instability of nonstop-PGK1 transcripts<br />

onstop cannot<br />

the performance<br />

fully be explained<br />

of transcript<br />

by ribosomal<br />

Ter-poly(A)dis<br />

ximide placement PGK1 which of factors contains bound a termination to the 3� codon UTR.<br />

NAs in inserted There is one ancodon emerging upstream viewof that the the site 5� of<br />

1-1 al- and polyadenylate 3� ends of mRNAs [poly(A)] interact addition to form (15) in a<br />

etem- closed transcript loopnonstop-PGK1. and that this conformation The addition is of<br />

stabili- required a termination for efficient codon translation at the 3′ end initiation of the<br />

Fig. 1) and normal mRNA stability. Participants in<br />

3′ UTR of a nonstop transcript resulted in<br />

on into the interaction include components of the<br />

an dis- translation<br />

substantial<br />

initiation<br />

(threefold)<br />

complex<br />

stabilization<br />

eIF4F as well<br />

(Fig.<br />

ositive as1). Pab1p. Unlike The nonstop-PGK1 absence of transcripts, a termination the<br />

3, 14). codon half-life in nonstop-PGK1 of Ter-poly(A)-PGK1 is predicted to was allow in-<br />

ism is the creased ribosome in the to continue absence translating of Xrn1p (Fig. through 1),<br />

ned the the indicating 3� UTR and that poly(A) this transcript tail, potentially is degraded reoly(A)sulting<br />

by the inpathway displacement for normal of Pab1p, mRNA disruption and not<br />

codon of normal mRNP structure, and consequently<br />

by the nonstop pathway. These data suggest<br />

that the instability of nonstop-PGK1<br />

transcripts cannot fully be explained by<br />

ribosomal displacement of factors bound<br />

to the 3′ UTR.<br />

There is an emerging view that the<br />

5′and 3′ ends of mRNAs interact to form<br />

a closed loop and that this conformation is<br />

required for efficient translation initiation<br />

and normal mRNA stability. Participants<br />

in the interaction include components of<br />

the translation initiation complex eIF4F as<br />

well as Pab1p. The absence of a termination<br />

codon in nonstop-PGK1 is predicted<br />

to allow the ribosome to continue translating<br />

through the 3′ UTR and poly(A)<br />

tail, potentially resulting in displacement<br />

of Pab1p, disruption of normal mRNP<br />

structure, and consequently accelerated<br />

degradation of the transcript. However,<br />

while mRNAs in pab1Δ strains do undergo<br />

accelerated decay, the rapid turnover is<br />

a result of premature decapping and can<br />

be suppressed by mutations in XRN1 (16),<br />

allowing distinction from nonstop decay.<br />

rapidly Lability of nonstop transcripts might also<br />

alf-lives be a consequence of ribosomal stalling at<br />

onstop- the 3′ end of the transcript. It is tempting<br />

pts are<br />

d in a to speculate that 3′-to-5′ exonucleolytic<br />

eted for activity underlies the decapping- and 5′-<br />

Ccr4p to-3′ exonuclease–independent, transla-<br />

decap-<br />

T yeast tion-dependent accelerated decay of non-<br />

ml of cycloheximide stop transcripts. (�CHX ). An appealing candidate<br />

lowing distinction from nonstop decay. Lability<br />

of nonstop transcripts might also be a<br />

consequence of ribosomal stalling at the 3�<br />

end of the transcript. It is tempting to speculate<br />

is the that exosome, 3�-to-5�a collection exonucleolytic of proteins activity<br />

underlies with 3′-to-5′ the exoribonuclease decapping- and 5�-to-3� activity that exonuclease–independent,<br />

translation-dependent<br />

functions in the processing of 5.8S RNA,<br />

accelerated decay of nonstop transcripts. An<br />

appealing<br />

rRNA, small<br />

candidate<br />

nucleolar<br />

is the<br />

RNA<br />

exosome,<br />

(snoRNA),<br />

a collection<br />

small of proteins nuclear with RNA 3�-to-5� (snRNA), exoribonuclease and other<br />

activity transcripts. thatData functions presented in thein processing van Hoof et of<br />

5.8S al. (17) RNA, validate rRNA, this prediction. small nucleolar RNA<br />

Fig. 2. Nonstop decay is conserved in mammalian<br />

cells. (A) Steady-state abundance of �-glucuronidase<br />

transcripts derived from wild type–<br />

( WT ), nonsense-, nonstop-, and Ter-poly(A)-<br />

�gluc minigene constructs after transient<br />

transfection into HeLa cells (28). Zeocin (zeo)<br />

transcripts served as a control for loading and<br />

transfection efficiency. For each construct, the<br />

�gluc/Zeo transcript ratio is expressed relative<br />

to that observed for WT-�gluc. (B) Steadystate<br />

abundance of WT-�gluc, nonstop-�gluc,<br />

and nonsense-�gluc transcripts were determined<br />

in the nuclear (N) and cytoplasmic (C)<br />

fractions of transfected cells (28). The ratio of<br />

normalized nonstop- or nonsense-�gluc transcript<br />

levels to WT-�gluc transcript levels are<br />

shown for each cellular compartment. Reported<br />

results are the average of three independent<br />

experiments.<br />

www.sciencemag.org SCIENCE VOL 295 22 MARCH 2002 19 2259<br />

Downloaded from<br />

www.sciencemag.org on June 29, 2007


In order to determine whether nonstop<br />

decay functions in mammals, we assessed<br />

the performance of transcripts derived<br />

from β-glucuronidase (βgluc) minigene<br />

constructs containing exons 1, 10, 11, and<br />

12 separated by introns derived from the<br />

endogenous gene (18). The abundance of<br />

transcripts derived from a nonstop-βgluc<br />

version of this minigene was significantly<br />

reduced relative to WT-βgluc, suggesting<br />

that a termination codon is also essential<br />

for normal mRNA stability in mammalian<br />

cells (Fig. 2A). Addition of a stop<br />

codon one codon upstream from the site<br />

of poly(A) addition in nonstop-βgluc<br />

[Ter-poly(A)-βgluc (15)] increased the<br />

abundance of this transcript to near–wildtype<br />

levels (Fig. 2A). Thus, all functional<br />

characteristics of nonstop transcripts in<br />

yeast appear to be relevant to mammalian<br />

systems.<br />

Both the abundance and stability of<br />

most nonsense transcripts, including nonsense-βgluc<br />

(Fig. 2B), are reduced in the<br />

nuclear fraction of mammalian cells (19,<br />

20). This has been interpreted to suggest<br />

that translation and NMD initiate while<br />

the mRNA is still associated with (if<br />

not within) the nuclear compartment. If<br />

translation is initiated on nucleus-associated<br />

transcripts, so might nonstop decay.<br />

Subcellular fractionation studies localized<br />

mammalian nonstop decay to the cytoplasmic<br />

compartment, providing further<br />

distinction from NMD (Fig. 2B).<br />

The conservation of nonstop decay<br />

in yeast and mammals suggests that the<br />

pathway serves an important biologic<br />

role. There are many potential physiologic<br />

sources of nonstop transcripts that warrant<br />

consideration. Mutations in bona fide termination<br />

codons would not routinely initiate<br />

nonstop decay due to the frequent occurrence<br />

of in-frame termination codons in<br />

the 3′ UTR and could not plausibly provide<br />

the evolutionary pressure for maintenance<br />

of nonstop decay. In contrast, alternative<br />

use of 3′-end processing signals embedded<br />

in coding sequence has been documented<br />

20<br />

in many genes including CBP1 in yeast<br />

and the growth hormone receptor (GHR)<br />

gene in fowl (21–23). Moreover, a computer<br />

search of the human mRNA and S.<br />

cerevisiae open reading frame (ORF) databases<br />

revealed many additional genes that<br />

contain a strict consensus sequence for 3′<br />

end cleavage and polyadenylation within<br />

their coding region (Fig. 3A). Utilization<br />

of these premature signals would direct<br />

formation of truncated transcripts that<br />

might be substrates for the nonstop decay<br />

pathway. Indeed, analysis of 3425 random<br />

yeast cDNA clones sequenced from the 3′<br />

end (i.e., 3′ ESTs) revealed that 40 showed<br />

apparent premature polyadenylation within<br />

the coding region (24).<br />

The truncated GHR transcript in fowl<br />

is apparently translated, as evidenced by<br />

its association with polysomes (22). As<br />

predicted for a substrate for nonstop decay,<br />

the ratio of truncated–to–full-length<br />

GHR transcripts was dramatically increased<br />

after treatment of cultured chicken<br />

hepatocellular carcinoma cells with the<br />

translational inhibitor emetine (Fig. 3B).<br />

Other truncated transcripts, both bigger<br />

and smaller than the predicted 0.7-kb nonstop<br />

transcript, did not show a dramatic<br />

increase in steady-state abundance upon<br />

translational arrest suggesting that they<br />

are derived from other mRNA processing<br />

events, perhaps alternative splicing (Fig.<br />

3B) (12). To directly test whether the nonstop<br />

mRNA pathway degrades prematurely<br />

polyadenylated mRNAs, we analyzed<br />

CBP1 transcripts in yeast. The CBP1 gene<br />

produces a 2.2-kb full-length mRNA and<br />

a 1.2-kb species with premature 3′ end<br />

processing and polyadenylation within the<br />

coding region (25). As expected from the<br />

observation that nonstop decay requires<br />

the exosome in yeast (17), deletion of the<br />

gene encoding the exosome component<br />

Ski7p stabilized the prematurely polyadenylated<br />

mRNA but had no effect on the<br />

stability of the full-length mRNA (Fig.<br />

3C). These data indicate that physiologic<br />

transcripts arising from premature poly-


Fig. 3. Many physiologic transcripts are<br />

substrates for nonstop decay. (A) 239<br />

human mRNAs contain a polyadenylation<br />

signal consisting of either of the two most<br />

common words for the positioning, cleavage,<br />

and downstream elements (29), separated by<br />

optimal distances (30), where the cleavage<br />

site occurs between the start and stop<br />

codons as determined from the cDNA coding<br />

sequence (CDS). Similarly 52 S. cerevisiae ORFs<br />

contained the yeast polyadenylation signal<br />

consisting of optimal upstream, positioning,<br />

and cleavage signals followed by any of nine<br />

common “U-rich” signals (24) with optimal<br />

spacing of the elements. Computer search<br />

was performed using the human mRNA<br />

database and the S. cerevisiae ORF database<br />

and an analysis program written in PERL that<br />

is available upon request. (B) Abundance of<br />

nonstop and full-length cGHR transcripts in<br />

chicken hepatocellular carcinoma cells (CRL-<br />

2117 purchased from the American Type<br />

Culture Collection, Manassas, VA) treated<br />

with 100 μg/ml of emetine for 6 hours. Northern analysis was performed as described (22). The<br />

relative ratio of each transcript in untreated and treated cells is reported. (C) Half-lives (t 1/2 ) of<br />

the full-length and nonstop CBP1 mRNAs measured in wild-type–(WT) and SKI7-deleted (ski7Δ)<br />

yeast strains carrying plasmid pG::-26 (25, 31). (D) Half-lives of WT-PGK1 and Ter-poly(A)-PGK1<br />

transcripts in wild-type–(WT ) and SKI7-deleted (ski7Δ) yeast strains treated with the indicated<br />

dose of paromomycin (mg/ml) (PM) for 20 hours. For cycloheximide (CHX ) experiments, yeast<br />

were treated identically except 100 μg/ml of CHX was added during the last hour before half-life<br />

determination. Half-lives were determined as described (27).<br />

21


adenylation are subject to nonstop mRNA<br />

decay. The cytoplasmic localization of<br />

ski7p is consistent with our observation<br />

that nonstop decay occurs within the cytoplasm<br />

(Fig. 2B).<br />

Any event that diminishes translational<br />

fidelity and promotes readthrough of termination<br />

codons could plausibly result in<br />

the generation of substrates for nonstop<br />

decay. In view of recent attempts to treat<br />

genetic disorders resulting from PTCs<br />

with long-term and high-dose aminoglycoside<br />

regimens, this may achieve medical<br />

significance. As a proof-of-concept experiment,<br />

we examined the performance of<br />

transcripts with one [Ter-poly(A)-PGK1]<br />

or multiple (WT-PGK1) in-frame termination<br />

codons (including those in the 3′<br />

prematurely polyadenylated mRNA but had<br />

no UTR) effect in yeast on thestrains stability after of treatment the full-length with<br />

mRNA paromomycin, (Fig. 3C). which Theseinduces data indicate ribosomal that<br />

physiologic readthrough. transcripts Remarkably, arising both from transcripts premature<br />

showed polyadenylation a dose-dependent are subject decrease to nonstop in sta-<br />

mRNA bility that decay. could Thebe cytoplasmic reversed by localization inhibiting<br />

of ski7p is consistent with our observation<br />

translational elongation with CHX or pre-<br />

that nonstop decay occurs within the cytoplasmvented<br />

(Fig. by deletion 2B). of the gene encoding<br />

Ski7p Any(Fig. event3D). that These diminishes data suggest translational that<br />

fidelity nonstop and decay promotes can limit readthrough the efficiency of termi- of<br />

nation therapeutic codons strategies could plausibly aimed at result enhancing in the<br />

generation nonsense suppression of substratesand for nonstop might contrib- decay.<br />

In view of recent attempts to treat genetic<br />

ute to the toxicity associated with amino-<br />

disorders resulting from PTCs with longtermglycoside<br />

and high-dose therapy. aminoglycoside regimens,<br />

The this degradation may achieve of nonstop medical transcripts significance.<br />

may be As regulated. a proof-of-concept The relative experiment, expression<br />

we level examined of truncated the performance GHR transcripts of transcripts com-<br />

with pared one to full-length [Ter-poly(A)-PGK1] transcripts or varies multiple in a<br />

(WT-PGK1) in-frame termination codons<br />

tissue-, gender-, and developmental stage-<br />

(including those in the 3� UTR) in yeast<br />

strains specific after manner treatment (22) and with the paromomycin,<br />

relative abun-<br />

which dance of induces truncated ribosomal CBP1 readthrough. transcripts varies Remarkably,<br />

with growth both condition transcripts (21). showed Data present- a dosedependented<br />

here warrant decrease the hypothesis in stability that thatregula could<br />

be tion reversed may occur by inhibiting at the level translational of nonstop elontrangation with CHX or prevented by deletion<br />

script stability rather than production. The<br />

of the gene encoding Ski7p (Fig. 3D).<br />

These conservation data suggest of the that GHR nonstop coding decay sequence can<br />

limit 3′- end theprocessing efficiency signal of therapeutic throughout strategies avian<br />

aimed phylogeny at enhancing and conservation nonsenseof suppression premature<br />

and polyadenylation might contribute of the to yeast the toxicity RNA14 associtranatedscript with and aminoglycoside its fruitfly homolog therapy. su(f ) sup-<br />

The degradation of nonstop transcripts may<br />

port speculation that nonstop transcripts<br />

be regulated. The relative expression level of<br />

truncated or derived GHR protein transcripts products compared serve essential to fulllength<br />

transcripts varies in a tissue-, gender-,<br />

and 22 developmental stage-specific manner (22)<br />

and the relative abundance of truncated CBP1<br />

developmental and/or homeostatic functions<br />

that are regulated by nonstop decay<br />

(21, 26).<br />

Many processes contribute to the precise<br />

control of gene expression including<br />

transcriptional and translational control<br />

mechanisms. In recent years, mRNA stability<br />

has emerged as a major determinant<br />

of both the magnitude and fidelity of gene<br />

expression. Perhaps the most striking and<br />

comprehensively studied example is the<br />

accelerated decay of transcripts harboring<br />

PTCs by the NMD pathway. Nonstop decay<br />

now serves as an additional example<br />

of the critical role that translation plays<br />

in monitoring the fidelity of gene expression,<br />

the stability R E Pof Oaberrant R T S<br />

or atypical<br />

transcripts, and hence the abundance of<br />

gene expression, the stability of aberrant or<br />

atypical<br />

truncated<br />

transcripts,<br />

proteins.<br />

and hence the abundance<br />

of truncated proteins.<br />

References and Notes<br />

1. A. W. Karzai, E. D. Roche, R. T. Sauer, Nature Struct.<br />

Biol. 7, 449 (2000).<br />

2. K. C. Keiler, P. R. Waller, R. T. Sauer, <strong>Science</strong> 271, 990<br />

(1996).<br />

3. The nonstop-PGK1 construct was generated from<br />

WT-PGK1 [pRP469 (8)] by creating three point mutations,<br />

using site-directed mutagenesis (Quik-<br />

Change Site-Directed <strong>Mutagenesis</strong> Kit, Stratagene)<br />

which eliminated the bona fide termination codon<br />

and all in-frame termination codons in the 3� UTR.<br />

The Ter-poly(A)-PGK1 construct was created from<br />

the nonstop-PGK1 construct, using site-directed mutagenesis<br />

(Quik-Change Site-Directed <strong>Mutagenesis</strong><br />

Kit, Stratagene). All changes were confirmed by sequencing.<br />

Primer sequences are available upon request.<br />

The PTC(22)-PGK1 construct is described in (8)<br />

(pRP609).<br />

4. P. Leeds, J. M. Wood, B. S. Lee, M. R. Culbertson, Mol.<br />

Cell. Biol. 12, 2165 (1992).<br />

5. Y. Cui, K. W. Hagan, S. Zhang, S. W. Peltz, Genes Dev.<br />

9, 423 (1995).<br />

6. D. Muhlrad, C. J. Decker, R. Parker, Genes Dev. 8, 855<br />

(1994).<br />

7. C. A. Beelman et al., Nature 382, 642 (1996).<br />

8. D. Muhlrad, R. Parker, Nature 370, 578 (1994).<br />

9. G. Caponigro, R. Parker, Microbiol. Rev. 60, 233<br />

(1996).<br />

10. M. Tucker et al., Cell 104, 377 (2001).<br />

11. C. L. Wolfe, A. K. Hopper, N. C. Martin, J. Biol. Chem.<br />

271, 4679 (1996).<br />

12. P. A. Frischmeyer, H. C. Dietz, data not shown.<br />

13. I. M. Weiss, S. A. Liebhaber, Mol. Cell. Biol. 14, 8123<br />

(1994).<br />

14. X. Wang, M. Kiledjian, I. M. Weiss, S. A. Liebhaber,<br />

Mol. Cell. Biol. 15, 1769 (1995).<br />

15. We performed 3� rapid amplification of cDNA ends<br />

(RACE) using the Marathon cDNA Amplification Kit<br />

(Clontech) and the Advantage 2 Polymerase Mix<br />

(Clontech) according to the manufacturer’s instructions.<br />

Poly(A) RNA isolated from cells (Oligotex<br />

mRNA midi kit, Qiagen) transiently transfected with<br />

WT-�gluc or from a wild-type yeast strain carrying a<br />

plasmid expressing WT-PGK1 was used as template.<br />

A single RACE product was generated, cloned ( TOPO<br />

TA 2.1 kit, Invitrogen), and sequenced.<br />

16. G. Caponigro, R. Parker, Genes Dev. 9, 2421 (1995).<br />

21. K. A. Spa<br />

4676 (19<br />

22. E. R. Ol<br />

Endocrin<br />

23. C. Hilge<br />

J. Virol.<br />

24. J. H. Gr<br />

Nucleic<br />

25. K. A. Spa<br />

Biol. 17,<br />

26. A. Audib<br />

U.S.A. 9<br />

27. All PGK1<br />

the GAL1<br />

were gro<br />

media–u<br />

scription<br />

SC-ura m<br />

off) to a<br />

removed<br />

10 �g of<br />

od (4) w<br />

dehyde g<br />

Screen P<br />

nucleotid<br />

cally det<br />

transcrip<br />

the perce<br />

log plot.<br />

for the c<br />

formed a<br />

1 hour] e<br />

28. HeLa ce<br />

bovine s<br />

confluen<br />

RPMI 16<br />

dithiothr<br />

and cells<br />

Electrop<br />

a capaci<br />

kit, Qiag<br />

tion, wa<br />

gel, tran<br />

with a po<br />

prising e<br />

had bee<br />

(Random<br />

heim). T<br />

with boi<br />

labeled<br />

quantita<br />

subcellu<br />

HeLa cel


ty associ- 15. We performed 3� rapid amplification of cDNA ends<br />

.<br />

(RACE) using the Marathon cDNA Amplification Kit<br />

(Clontech) and the Advantage 2 Polymerase Mix<br />

ripts may (Clontech) according to the manufacturer’s instruc-<br />

n level of tions. Poly(A) RNA isolated from cells (Oligotex<br />

d to full- mRNA midi kit, Qiagen) transiently transfected with<br />

WT-�gluc or from a wild-type yeast strain carrying a<br />

, gender-, plasmid expressing WT-PGK1 was used as template.<br />

nner (22) A single RACE product was generated, cloned ( TOPO<br />

ted CBP1 TA 2.1 kit, Invitrogen), and sequenced.<br />

16. G. Caponigro, R. Parker, Genes Dev. 9, 2421 (1995).<br />

ition (21). 17. A. van Hoof, P. A. Frischmeyer, H. C. Dietz, R. Parker,<br />

thesis that <strong>Science</strong> 295, 2262 (2002).<br />

f nonstop 18. To create the WT-�gluc minigene construct, mouse<br />

genomic DNA was used as template for amplifying<br />

ction. The<br />

three different portions of the �-glucuronidase gene:<br />

uence 3�- from nucleotide 2313 to 3030 encompassing exon 1<br />

nphylog- and part of intron 1 (including the splice donor selyadenylquence),<br />

from nucleotide 11285 to 13555 encompassing<br />

part of intron 9 (including the splice acceptor sept<br />

and its quence ), exon 10, and part of intron 10 (including the<br />

lation that splice donor sequence), and from 13556 to 15813<br />

products which included the remainder of intron 10, exon 11,<br />

intron 11, and exon 12. All three fragments were TA<br />

r homeo- cloned ( Topo TA 2.1 kit, Invitrogen), sequenced, and<br />

y nonstop then cloned into pZeoSV2 (Invitrogen). A single basepair<br />

deletion corresponding to the gus<br />

e precise<br />

ing tranlmechability<br />

has<br />

f both the<br />

pression.<br />

mprehenerateddes<br />

by the<br />

serves as<br />

l role that<br />

fidelity of<br />

23<br />

mps tagenesis (Quik-Change Site-Directed <strong>Mutagenesis</strong><br />

gel, transferred to a nylon membrane, and hybridized<br />

ide regi- Kit, Stratagene). All changes were confirmed by se-<br />

with a polymerase chain reaction (PCR) product comsignifiquencing.<br />

Primer sequences are available upon reprising<br />

exons 10 through 12 of the �-gluc gene that<br />

quest. The PTC(22)-PGK1 construct is described in (8)<br />

eriment,<br />

had been radioactively labeled by nick translation<br />

(pRP609).<br />

(Random Primed DNA Labeling Kit, Boehringer Mann-<br />

anscripts 4. P. Leeds, J. M. Wood, B. S. Lee, M. R. Culbertson, Mol.<br />

heim). The membrane was subsequently stripped<br />

multiple Cell. Biol. 12, 2165 (1992).<br />

with boiling 0.5% SDS and probed with radioactively<br />

5. Y. Cui, K. W. Hagan, S. Zhang, S. W. Peltz, Genes Dev.<br />

codons<br />

labeled zeocin cDNA. Northern blot results were<br />

9, 423 (1995).<br />

quantitated using an Instant Imager (Packard). For<br />

in yeast 6. D. Muhlrad, C. J. Decker, R. Parker, Genes Dev. 8, 855 subcellular fractionation, transiently transfected<br />

omycin, (1994).<br />

HeLa cells were trypsinized, washed in cold 1� phos-<br />

7. C. A. Beelman et al., Nature 382, 642 (1996).<br />

ugh. Rephate<br />

buffered saline, and pelleted by centrifuging at<br />

8. D. Muhlrad, R. Parker, Nature 370, 578 (1994).<br />

2040g for 5 min at 4°C. Cells were then resuspended<br />

adose- 9. G. Caponigro, R. Parker, Microbiol. Rev. 60, 233 in 200 �l of 140 mM NaCl, 1.5 mM MgCl2 , 10 mM<br />

at could (1996).<br />

Tris-HCl (pH 8.6), 0.5% NP-40, and 1 mM DTT,<br />

nal elon- 10. M. Tucker et al., Cell 104, 377 (2001).<br />

vortexed, and incubated on ice for 5 min. Nuclei were<br />

11. C. L. Wolfe, A. K. Hopper, N. C. Martin, J. Biol. Chem.<br />

deletion<br />

pelleted by centrifuging at 12,000g for 45 s at 4°C<br />

271, 4679 (1996).<br />

and subsequently separated from the aqueous (cyto-<br />

ig. 3D). 12. P. A. Frischmeyer, H. C. Dietz, data not shown.<br />

plasmic) phase.<br />

ecay can 13. I. M. Weiss, S. A. Liebhaber, Mol. Cell. Biol. 14, 8123 29. J. H. Graber, C. R. Cantor, S. C. Mohr, T. F. Smith, Proc.<br />

(1994).<br />

trategies<br />

Natl. Acad. Sci. U.S.A. 96, 14055 (1999).<br />

14. X. Wang, M. Kiledjian, I. M. Weiss, S. A. Liebhaber, 30. F. Chen, C. C. MacDonald, J. Wilusz, Nucleic Acids Res.<br />

pression Mol. Cell. Biol. 15, 1769 (1995).<br />

23, 2614 (1995).<br />

y associ- 15. We performed 3� rapid amplification of cDNA ends 31. Plasmid pG-26 (25) containing the CBP1 gene un-<br />

.<br />

(RACE) using the Marathon cDNA Amplification Kit der the control of the GAL promoter and the LEU2<br />

(Clontech) and the Advantage 2 Polymerase Mix selectable marker was introduced into wild-type<br />

ripts may (Clontech) according to the manufacturer’sallele instruc- was and SKI7-deleted strains. Yeast were grown in me-<br />

level of tions. generated Poly(A) by site-directed RNA isolated mutagenesis from cells (Quik-Change (Oligotex dia lacking leucine and containing 2% galactose at<br />

d to full- mRNA Site-Directed midi kit, <strong>Mutagenesis</strong> Qiagen) transiently Kit, Stratagene) transfected of thewith WT- 30°C to an OD600 of 0.5. Cells were pelleted and<br />

WT-�gluc �gluc construct or fromto a create wild-type nonsense-�gluc. yeast strain carrying Nonstop- a resuspended in media lacking a carbon source.<br />

, gender-, plasmid �gluc was expressing generatedWT-PGK1 from WT-�gluc was used by two as template. rounds of Glucose was then added to a final concentration of<br />

nner (22) Asite-directed single RACEmutagenesis product was(Quik-Change generated, cloned Site-Directed ( TOPO 2% at time 0 and time points were taken. RNA was<br />

ted CBP1 TA <strong>Mutagenesis</strong> 2.1 kit, Invitrogen), Kit, Stratagene), and sequenced. which removed the bona extracted and analyzed by northern blotting with a<br />

16. G. fide Caponigro, termination R. Parker, codon Genes and allDev. in-frame 9, 2421 termination (1995).<br />

tion (21).<br />

labeled oligo probe oRP1067 (5�-CTCGGTCCTG-<br />

17. A. codons van Hoof, in theP. 3�A. UTR. Frischmeyer, Ter-poly(A)-�gluc H. C. Dietz, was R. generated Parker, TACCGAACGAGACGAGG-3�).<br />

thesis that <strong>Science</strong> from nonstop-�gluc 295, 2262 (2002). via a single point mutation, which 32. We thank N. Martin, A. Atkin, C. Dieckmann, and M.<br />

f nonstop 18. To created create a stop the one WT-�gluc codon minigene upstream of construct, the poly(A) mouse tail. Sands for the provision of valuable reagents. This<br />

genomic All changes DNA were was confirmed used as by template sequencing. for amplifying<br />

tion. The<br />

Primer se- work was supported by a grant from NIH (GM55239)<br />

three quences different are available portions upon of the request. �-glucuronidase gene: (H.C.D.), the Howard Hughes Medical Institute<br />

uence 3�- 19. from P. Belgrader, nucleotide J. Cheng, 2313 to X. 3030 Zhou, encompassing L. S. Stephenson, exonL. 1E.<br />

(H.C.D, R.P.), and the Medical Scientist Training Pronphylog-<br />

and Maquat, part Mol. of intron Cell. 1 Biol. (including 14, 8219 the(1994). splice donor segram (P.A.F.).<br />

lyadenyl 20. quence), O. Kessler, fromL. nucleotide A. Chasin, 11285 Mol. toCell. 13555 Biol. encompass- 16, 4426<br />

ing (1996). part of intron 9 (including the splice acceptor se- 22 October 2001; accepted 31 January 2002<br />

t and its<br />

berrant or 21. quence K. A. Sparks, ), exonC. 10, L. and Dieckmann, part of intron Nucleic 10 Acids (including Res. the 26,<br />

ation that splice 4676 (1998). donor sequence), and from 13556 to 15813<br />

bundance<br />

products www.sciencemag.org 22. which E. R. Oldham, includedSCIENCE the B. Bingham, remainder VOL W. of intron R. 295 Baumbach, 10, 22 exon MARCH Mol. 11, 2002 2261<br />

intron Endocrinol. 11, and 7, exon 1379 12. (1993). All three fragments were TA<br />

r homeo-<br />

23. cloned C. Hilger, ( Topo I. Velhagen, TA 2.1 kit, H. Invitrogen), Zentgraf, sequenced, C. H. Schroder, and<br />

nonstop then J. Virol. cloned 65, into 4284 pZeoSV2 (1991). (Invitrogen). A single base-<br />

24. pair J. H. deletion Graber, corresponding C. R. Cantor, to S. the C. Mohr, gus T. F. Smith,<br />

ature Struct.<br />

e precise Nucleic Acids Res. 27, 888 (1999).<br />

25. K. A. Sparks, S. A. Mayer, C. L. Dieckmann, Mol. Cell.<br />

ing ce 271, tran- 990 Biol. 17, 4199 (1997).<br />

lmecha- 26. A. Audibert, M. Simonelig, Proc. Natl. Acad. Sci.<br />

erated from<br />

bility has U.S.A. 95, 14302 (1998).<br />

e point mu- 27. All PGK1 minigene constructs are under the control of<br />

esis both(Quik- the<br />

the GAL1 upstream activation sequence (UAS). Cultures<br />

pression. Stratagene) were grown to mid-log phase in synthetic complete<br />

ation codon mprehen- media–uracil (SC-ura) containing 2% galactose (tran-<br />

the 3� UTR. scription on). Cells were pelleted and resuspended in<br />

reated rated from de-<br />

SC-ura media. Glucose was then added (transcription<br />

sirected by the mu- off) to a final concentration of 2% and aliquots were<br />

<strong>Mutagenesis</strong> serves as removed at the indicated time points. Approximately<br />

rmed by se- 10 �g of total RNA isolated with the hot phenol methlerole<br />

upon that reod (4) was electrophoresed on a 1.2% agarose formalscribed<br />

idelityin of (8) dehyde gel, transferred to a nylon membrane (Gene-<br />

Screen Plus, NEN), and hybridized with a 23 bp oligo-<br />

ertson, Mol. nucleotide end-labeled radioactive probe that specifically<br />

detects the polyG tract in the 3� UTR of the PGK1<br />

, Genes Dev. transcripts (8). Half-lives were determined by plotting<br />

the percent of mRNA remaining versus time on a semi-<br />

Dev. 8, 855 log plot. All half-lives were determined at 25°C except<br />

for the ccr4� (performed at 30°C) and dcp1-2 [per-<br />

996).<br />

(1994).<br />

v. 60, 233<br />

formed after shift to the nonpermissive temp (37°C) for<br />

1 hour] experiments.<br />

28. HeLa cells [maintained in DMEM with 10% fetal<br />

bovine serum (FBS)] were grown to �80 to 90%<br />

confluency, trypsinized, and resuspended in 300 �l of<br />

. Biol. Chem. RPMI 1640 without FBS, 10 mM glucose and 0.1 mM<br />

dithiothreitol (DTT). Plasmid DNA (10 �g) was added<br />

shown.<br />

ol. 14, 8123<br />

and cells were electroporated (Bio-Rad, Gene Pulser II<br />

Electroporation System) at a voltage of 0.300 kV and<br />

. Liebhaber,<br />

a capacitance of 500 �F. Poly(A) RNA (Oligotex midi<br />

kit, Qiagen, 2 �g), isolated 36 hours after transfection,<br />

was separated on a 1.6% agarose formaldehyde<br />

mps t genetic The Ter-poly(A)-PGK1 construct was created from removed at the indicated time points. Approximately<br />

the nonstop-PGK1 construct, using site-directed mu- 10 �g of total RNA isolated with the hot phenol meth-<br />

ith longtagenesis (Quik-Change Site-Directed <strong>Mutagenesis</strong> od (4) was electrophoresed on a 1.2% agarose formalideregi-<br />

Kit, Stratagene). All changes were confirmed by sedehyde gel, transferred to a nylon membrane (Genelsignifiquencing.<br />

Primer sequences are available upon re- Screen Plus, NEN), and hybridized with a 23 bp oligoquest.<br />

The PTC(22)-PGK1 construct is described in (8) nucleotide end-labeled radioactive probe that specifieriment,<br />

(pRP609).<br />

cally detects the polyG tract in the 3� UTR of the PGK1<br />

anscripts 4. P. Leeds, J. M. Wood, B. S. Lee, M. R. Culbertson, Mol. transcripts (8). Half-lives were determined by plotting<br />

multiple Cell. Biol. 12, 2165 (1992).<br />

the percent of mRNA remaining versus time on a semi-<br />

5. Y. Cui, K. W. Hagan, S. Zhang, S. W. Peltz, Genes Dev. log plot. All half-lives were determined at 25°C except<br />

codons 9, 423 (1995).<br />

for the ccr4� (performed at 30°C) and dcp1-2 [per-<br />

in yeast 6. D. Muhlrad, C. J. Decker, R. Parker, Genes Dev. 8, 855 formed after shift to the nonpermissive temp (37°C) for<br />

omycin, (1994).<br />

1 hour] experiments.<br />

7. C. A. Beelman et al., Nature 382, 642 (1996).<br />

ugh. Re-<br />

28. HeLa cells [maintained in DMEM with 10% fetal<br />

8. D. Muhlrad, R. Parker, Nature 370, 578 (1994).<br />

bovine serum (FBS)] were grown to �80 to 90%<br />

adose- 9. G. Caponigro, R. Parker, Microbiol. Rev. 60, 233 confluency, trypsinized, and resuspended in 300 �l of<br />

at could (1996).<br />

RPMI 1640 without FBS, 10 mM glucose and 0.1 mM<br />

nal elon- 10. M. Tucker et al., Cell 104, 377 (2001).<br />

dithiothreitol (DTT). Plasmid DNA (10 �g) was added<br />

11. C. L. Wolfe, A. K. Hopper, N. C. Martin, J. Biol. Chem. and cells were electroporated (Bio-Rad, Gene Pulser II<br />

deletion 271, 4679 (1996).<br />

Electroporation System) at a voltage of 0.300 kV and<br />

ig. 3D). 12. P. A. Frischmeyer, H. C. Dietz, data not shown.<br />

a capacitance of 500 �F. Poly(A) RNA (Oligotex midi<br />

ecay can 13. I. M. Weiss, S. A. Liebhaber, Mol. Cell. Biol. 14, 8123 kit, Qiagen, 2 �g), isolated 36 hours after transfec-<br />

(1994).<br />

tion, was separated on a 1.6% agarose formaldehyde<br />

trategies<br />

14. X. Wang, M. Kiledjian, I. M. Weiss, S. A. Liebhaber, gel, transferred to a nylon membrane, and hybridized<br />

pression Mol. Cell. Biol. 15, 1769 (1995).<br />

with a polymerase chain reaction (PCR) product com-<br />

y associ- 15. We performed 3� rapid amplification of cDNA ends prising exons 10 through 12 of the �-gluc gene that<br />

had been radioactively labeled by nick translation<br />

.<br />

(RACE) using the Marathon cDNA Amplification Kit<br />

(Clontech) and the Advantage 2 Polymerase Mix (Random Primed DNA Labeling Kit, Boehringer Mann-<br />

ripts may (Clontech) according to the manufacturer’s instrucheim). The membrane was subsequently stripped<br />

level of tions. Poly(A) RNA isolated from cells (Oligotex with boiling 0.5% SDS and probed with radioactively<br />

labeled zeocin cDNA. Northern blot results were<br />

d to full- mRNA midi kit, Qiagen) transiently transfected with<br />

WT-�gluc or from a wild-type yeast strain carrying a quantitated using an Instant Imager (Packard). For<br />

, gender-, plasmid expressing WT-PGK1 was used as template. subcellular fractionation, transiently transfected<br />

nner (22) A single RACE product was generated, cloned ( TOPO HeLa cells were trypsinized, washed in cold 1� phos-<br />

ted CBP1 TA 2.1 kit, Invitrogen), and sequenced.<br />

phate buffered saline, and pelleted by centrifuging at<br />

16. G. Caponigro, R. Parker, Genes Dev. 9, 2421 (1995). 2040g for 5 min at 4°C. Cells were then resuspended<br />

tion (21). 17. A. van Hoof, P. A. Frischmeyer, H. C. Dietz, R. Parker, in 200 �l of 140 mM NaCl, 1.5 mM MgCl2 , 10 mM<br />

thesis that <strong>Science</strong> 295, 2262 (2002).<br />

Tris-HCl (pH 8.6), 0.5% NP-40, and 1 mM DTT,<br />

f nonstop 18. To create the WT-�gluc minigene construct, mouse vortexed, and incubated on ice for 5 min. Nuclei were<br />

genomic DNA was used as template for amplifying pelleted by centrifuging at 12,000g for 45 s at 4°C<br />

tion. The<br />

three different portions of the �-glucuronidase gene: and subsequently separated from the aqueous (cyto-<br />

uence 3�- from nucleotide 2313 to 3030 encompassing exon 1 plasmic) phase.<br />

nphylog- and part of intron 1 (including the splice donor se- 29. J. H. Graber, C. R. Cantor, S. C. Mohr, T. F. Smith, Proc.<br />

Natl. Acad. Sci. U.S.A. 96, 14055 (1999).<br />

lyadenylquence), from nucleotide 11285 to 13555 encompassing<br />

part of intron 9 (including the splice acceptor se- 30. F. Chen, C. C. MacDonald, J. Wilusz, Nucleic Acids Res.<br />

t and its quence ), exon 10, and part of intron 10 (including the 23, 2614 (1995).<br />

lation that splice donor sequence), and from 13556 to 15813 31. Plasmid pG-26 (25) containing the CBP1 gene under<br />

the control of the GAL promoter and the LEU2<br />

products which included the remainder of intron 10, exon 11,<br />

intron 11, and exon 12. All three fragments were TA selectable marker was introduced into wild-type<br />

r homeo- cloned ( Topo TA 2.1 kit, Invitrogen), sequenced, allele was and and SKI7-deleted strains. Yeast were grown in menonstop<br />

generated then cloned byinto site-directed pZeoSV2 mutagenesis (Invitrogen). A(Quik-Change single basedia lacking leucine and containing 2% galactose at<br />

Site-Directed pair deletion <strong>Mutagenesis</strong> corresponding Kit, toStratagene) the gus of the WT- 30°C to an OD600 of 0.5. Cells were pelleted and<br />

�gluc construct to create nonsense-�gluc. Nonstop- resuspended in media lacking a carbon source.<br />

e precise �gluc was generated from WT-�gluc by two rounds of Glucose was then added to a final concentration of<br />

ing tran- site-directed mutagenesis (Quik-Change Site-Directed 2% at time 0 and time points were taken. RNA was<br />

<strong>Mutagenesis</strong> Kit, Stratagene), which removed the bona<br />

lmecha- extracted and analyzed by northern blotting with a<br />

fide termination codon and all in-frame termination labeled oligo probe oRP1067 (5�-CTCGGTCCTGbility<br />

has codons in the 3� UTR. Ter-poly(A)-�gluc was generated TACCGAACGAGACGAGG-3�).<br />

both the from nonstop-�gluc via a single point mutation, which 32. We thank N. Martin, A. Atkin, C. Dieckmann, and M.<br />

pression. created a stop one codon upstream of the poly(A) tail. Sands for the provision of valuable reagents. This<br />

All changes were confirmed by sequencing. Primer se- work was supported by a grant from NIH (GM55239)<br />

mprehenquences are available upon request.<br />

(H.C.D.), the Howard Hughes Medical Institute<br />

ratedde- 19. P. Belgrader, J. Cheng, X. Zhou, L. S. Stephenson, L. E. (H.C.D, R.P.), and the Medical Scientist Training Pros<br />

by the Maquat, Mol. Cell. Biol. 14, 8219 (1994).<br />

gram (P.A.F.).<br />

20. O. Kessler, L. A. Chasin, Mol. Cell. Biol. 16, 4426<br />

serves as (1996).<br />

22 October 2001; accepted 31 January 2002<br />

l role that<br />

idelity of<br />

www.sciencemag.org SCIENCE VOL 295 22 MARCH 2002 2261<br />

mps SC-ura media. Glucose was then added (transcription<br />

off) to a final concentration of 2% and aliquots were<br />

removed at the indicated time points. Approximately<br />

10 �g of total RNA isolated with the hot phenol method<br />

(4) was electrophoresed on a 1.2% agarose formaldehyde<br />

gel, transferred to a nylon membrane (Gene-<br />

Screen Plus, NEN), and hybridized with a 23 bp oligonucleotide<br />

end-labeled radioactive probe that specifically<br />

detects the polyG tract in the 3� UTR of the PGK1<br />

transcripts (8). Half-lives were determined by plotting<br />

the percent of mRNA remaining versus time on a semilog<br />

plot. All half-lives were determined at 25°C except<br />

for the ccr4� (performed at 30°C) and dcp1-2 [performed<br />

after shift to the nonpermissive temp (37°C) for<br />

1 hour] experiments.<br />

28. HeLa cells [maintained in DMEM with 10% fetal<br />

bovine serum (FBS)] were grown to �80 to 90%<br />

confluency, trypsinized, and resuspended in 300 �l of<br />

RPMI 1640 without FBS, 10 mM glucose and 0.1 mM<br />

dithiothreitol (DTT). Plasmid DNA (10 �g) was added<br />

and cells were electroporated (Bio-Rad, Gene Pulser II<br />

Electroporation System) at a voltage of 0.300 kV and<br />

a capacitance of 500 �F. Poly(A) RNA (Oligotex midi<br />

kit, Qiagen, 2 �g), isolated 36 hours after transfection,<br />

was separated on a 1.6% agarose formaldehyde<br />

gel, transferred to a nylon membrane, and hybridized<br />

with a polymerase chain reaction (PCR) product comprising<br />

exons 10 through 12 of the �-gluc gene that<br />

had been radioactively labeled by nick translation<br />

(Random Primed DNA Labeling Kit, Boehringer Mannheim).<br />

The membrane was subsequently stripped<br />

with boiling 0.5% SDS and probed with radioactively<br />

labeled zeocin cDNA. Northern blot results were<br />

quantitated using an Instant Imager (Packard). For<br />

subcellular fractionation, transiently transfected<br />

HeLa cells were trypsinized, washed in cold 1� phosphate<br />

buffered saline, and pelleted by centrifuging at<br />

2040g for 5 min at 4°C. Cells were then resuspended<br />

in 200 �l of 140 mM NaCl, 1.5 mM MgCl2 , 10 mM<br />

Tris-HCl (pH 8.6), 0.5% NP-40, and 1 mM DTT,<br />

vortexed, and incubated on ice for 5 min. Nuclei were<br />

pelleted by centrifuging at 12,000g for 45 s at 4°C<br />

and subsequently separated from the aqueous (cytoplasmic)<br />

phase.<br />

29. J. H. Graber, C. R. Cantor, S. C. Mohr, T. F. Smith, Proc.<br />

Natl. Acad. Sci. U.S.A. 96, 14055 (1999).<br />

30. F. Chen, C. C. MacDonald, J. Wilusz, Nucleic Acids Res.<br />

23, 2614 (1995).<br />

31. Plasmid pG-26 (25) containing the CBP1 gene under<br />

the control of the GAL promoter and the LEU2<br />

selectable marker was introduced into wild-type<br />

allele was and SKI7-deleted strains. Yeast were grown in me-<br />

generated by site-directed mutagenesis (Quik-Change dia lacking leucine and containing 2% galactose at<br />

Site-Directed <strong>Mutagenesis</strong> Kit, Stratagene) of the WT- 30°C to an OD600 of 0.5. Cells were pelleted and<br />

�gluc construct to create nonsense-�gluc. Nonstop- resuspended in media lacking a carbon source.<br />

�gluc was generated from WT-�gluc by two rounds of Glucose was then added to a final concentration of<br />

site-directed mutagenesis (Quik-Change Site-Directed 2% at time 0 and time points were taken. RNA was<br />

<strong>Mutagenesis</strong> Kit, Stratagene), which removed the bona extracted and analyzed by northern blotting with a<br />

fide termination codon and all in-frame termination labeled oligo probe oRP1067 (5�-CTCGGTCCTGcodons<br />

in the 3� UTR. Ter-poly(A)-�gluc was generated TACCGAACGAGACGAGG-3�).<br />

from nonstop-�gluc via a single point mutation, which 32. We thank N. Martin, A. Atkin, C. Dieckmann, and M.<br />

created a stop one codon upstream of the poly(A) tail. Sands for the provision of valuable reagents. This<br />

All changes were confirmed by sequencing. Primer se- work was supported by a grant from NIH (GM55239)<br />

quences are available upon request.<br />

(H.C.D.), the Howard Hughes Medical Institute<br />

19. P. Belgrader, J. Cheng, X. Zhou, L. S. Stephenson, L. E. (H.C.D, R.P.), and the Medical Scientist Training Pro-<br />

Maquat, Mol. Cell. Biol. 14, 8219 (1994).<br />

gram (P.A.F.).<br />

20. O. Kessler, L. A. Chasin, Mol. Cell. Biol. 16, 4426<br />

(1996).<br />

22 October 2001; accepted 31 January 2002<br />

www.sciencemag.org SCIENCE VOL 295 22 MARCH 2002 2261<br />

iencemag.org on June 29, 2007<br />

Downloaded from<br />

from www.sciencemag.org<br />

on June on Downloaded June 29, 2007 29, 2007 from www.sc


The old notion of natural selection are usually opaque, and a lack of response<br />

as an omnipotent force in biologi- to selection may reflect nothing more than<br />

cal evolution has given way to one a lack of heritable variation (4, 7). If the<br />

where adaptive processes are constrained cause of a constraint is to be elucidated,<br />

by physical, chemical, and biological exi- it must be for a simple phenotype whose<br />

gencies (1–4). Whether constraint and/or relationship to fitness is understood.<br />

stabilizing selection explain phenotypic Coenzyme use by β-isopropylmalate<br />

stasis, in the fossil record and in phylog- dehydrogenase (IMDH) is a simple pheenies,<br />

remains an open question (5). Direct notype whose etiology and relationship<br />

experimental tests of constraint are scarce to fitness are understood (10, 11). IMDHs<br />

(6–9). Even tight correlations among catalyze the oxidative decarboxylation of<br />

traits, at once suggestive of constraint, can β-isopropylmalate to α-ketoisocaproate<br />

be broken by artificial selection to produce during the biosynthesis of leucine, an es-<br />

new phenotypic combinations (8, 9). Desential amino acid. All IMDHs use nicospite<br />

all circumstantial evidence, results tinamide adenine dinucleotide (NAD) as a<br />

from direct experimental tests imply that coenzyme (cosubstrate). This invariance of<br />

selection is largely unconstrained. function among IMDHs hints at the pres-<br />

The direct experimental test for conence of ancient constraints, even though<br />

straint is conceptually simple. A pheno- some related isocitrate dehydrogenases<br />

type is subjected to selection (natural or (IDHs) use NADP instead (12, 13).<br />

artificial) in an attempt to break the pos- Structural comparisons with related<br />

tulated constraint (6–9). A response to se- NADP-using IDHs identify amino aclection<br />

indicates a lack of constraint. No ids controlling coenzyme use (14–16)<br />

response to selection indicates the presence<br />

of a constraint. However, the cause of<br />

a constraint is rarely specified because the<br />

etiologies of most phenotypes are not well<br />

(Fig. 1A). Introducing five replacements<br />

(Asp<br />

understood, their relationships to fitness<br />

236 → Arg, Asp289 → Lys, Ile290→ Tyr,<br />

Ala296→ Val, and Gly337→ Tyr) into the<br />

coenzyme-binding pocket of Escherichia<br />

coli leuB–encoded IMDH by site-directed<br />

mutagenesis causes a complete reversal<br />

in specificity (10, 11): NAD performance<br />

(k NAD<br />

cat /K NAD<br />

m ,where K is the Michae-<br />

m<br />

lis constant) is reduced by a factor of<br />

340, from 68 × 103 M –1 s –1 to 0.2 × 103 (16), and controls the activation and differentiation<br />

of T cells (17). Our finding that a mycobacterial<br />

3. Y. Wang et al., J. Immunol. 169, 2422 (2002).<br />

4. M. J. Gething, J. Sambrook, Nature 355, 33 (1992).<br />

5. P. Srivastava, Annu. Rev. Immunol. 20, 395 (2002).<br />

26 May 2006;<br />

10.1126/scien<br />

Direct Demonstration of an<br />

Adaptive Constraint<br />

Stephen P. Miller, 1 Mark Lunzer, 1 Antony M. Dean 1,2 than a lack<br />

cause of a<br />

be for a sim<br />

fitness is u<br />

Coenzym<br />

*<br />

genase (IM<br />

etiology and<br />

The role of constraint in adaptive evolution is an open question. Directed evolution of an engineered (10, 11). IM<br />

b-isopropylmalate dehydrogenase (IMDH), with coenzyme specificity switched from nicotinamide ation of b<br />

adenine dinucleotide (NAD) to nicotinamide adenine dinucleotide phosphate (NADP), always produces during the<br />

mutants with lower affinities for NADP. This result is the correlated response to selection for relief amino acid.<br />

from inhibition by NADPH (the reduced form of NADP) expected of an adaptive landscape subject to dinucleotide<br />

three enzymatic constraints: an upper limit to the rate of maximum turnover (kcat ), a correlation in This invaria<br />

NADP and NADPH affinities, and a trade-off between NAD and NADP usage. Two additional constraints, at the presen<br />

high intracellular NADPH abundance and the cost of compensatory protein synthesis, have ensured some relate<br />

the conserved use of NAD by IMDH throughout evolution. Our results show that selective mechanisms use NADP<br />

and evolutionary constraints are to be understood in terms of underlying adaptive landscapes.<br />

Structur<br />

using IDHs<br />

he old notion of natural selection as an constraint, can be broken by artificial selection to enzyme use<br />

omnipotent force in biological evolution produce new phenotypic combinations (8, 9). De- replacemen<br />

Thas<br />

given way to one where adaptive prospite all circumstantial evidence, results from Ile<br />

cesses are constrained by physical, chemical, and direct experimental tests imply that selection is<br />

biological exigencies (1–4). Whether constraint largely unconstrained.<br />

and/or stabilizing selection explain phenotypic The direct experimental test for constraint is<br />

stasis, in the fossil record and in phylogenies, re- conceptually simple. A phenotype is subjected to<br />

mains an open question (5). Direct experimental selection (natural or artificial) in an attempt to<br />

tests of constraint are scarce (6–9). Even tight break the postulated constraint (6–9). A response<br />

correlations among traits, at once suggestive of to selection indicates a lack of constraint. No response<br />

to selection indicates the presence of a<br />

constraint. However, the cause of a constraint is<br />

rarely specified because the etiologies of most<br />

phenotypes are not well understood, their relationships<br />

to fitness are usually opaque, and a lack of<br />

response to selection may reflect nothing more<br />

290 Y Ty<br />

into the coe<br />

coli leuB–e<br />

genesis cau<br />

(10, 11): N<br />

Km is the<br />

factor of 34<br />

103 M –1<br />

iDCs (Fig. 4C). Furthermore, pretreatment of<br />

iDCs with pertussis toxin or TAK-779 inhibited<br />

the ability of peptide-loaded myHsp70 to stimu-<br />

further connection between the innate and adaptive<br />

immune responses during mycobacterial infection.<br />

The cellular aggregation induced by myHsp70<br />

K. T. Jean<br />

9. J. E. Hildr<br />

Eur. J. Im<br />

late DCs to generate influenza peptide–specific signaling through CCR5 may play an important 10. P. Revy, M<br />

Nat. Imm<br />

CTLs (Fig. 4D). To examine the effect of CCR5- role in the formation of granulomas, the hallmark<br />

11. M. L. Dus<br />

mediated signaling in mycobacterial infection, we of mycobacterial infection.<br />

Direct Demonstration of an<br />

363 (200<br />

used the model pathogen M. bovis BCG-lux (12). Microbial-induced DC responses need to be 12. B. Kampm<br />

As reported for murine DCs (13), human DCs highly regulated, reflecting a balance between a 13. K. A. Bod<br />

69, 800<br />

from Adaptive immune people wereConstraint unable to kill inter- rapid and appropriate response to invading mi-<br />

14. J. G. Cyst<br />

nalized mycobacteria, even in the presence of crobes and the inducement of immunopathology 15. M. Nieto<br />

autologous T cells. TAK-779 inhibition of CCR5 (2)—a particular problem in mycobacterial in- 16. F. Castelli<br />

led to a dose-dependent enhancement of intrafection. An increasing number of human patho- 17. S. A. Luth<br />

18. T. Dragic<br />

cellular mycobacterial replication (Fig. 4E and fig. gens, including HIV (18), toxoplasma (19), and<br />

19. J. Aliberti<br />

S6A) at all concentrations of mycobacteria tested M. tuberculosis (as described here), target the 20. E. de Silva<br />

(fig. S6B), suggesting an important role for this CCR5 receptor. This role of CCR5 as a pattern- 21. We are gr<br />

receptor in controlling mycobacterial infection. recognition receptor for myHsp70 may, at least Spectroph<br />

The identification of CCR5 as the critical re- in part, be responsible for the maintenance of the the Wellco<br />

(R.A.F.), th<br />

ceptor for myHsp70-mediated DC stimulation has high CCR5D32 allele frequency (10 to 15%) in FEBS (E.H<br />

implications for both mycobacterial infection and Northern European populations (20) and may<br />

the therapeutic use of myHsp70. CCR5 is impor- alter the pattern of disease seen in people with Supporting O<br />

tant in immune cell cross talk. Interaction with its the CCR5D32 allele.<br />

www.sciencema<br />

naturally occurring ligand MIP-1b promotes the<br />

Materials and M<br />

recruitment of cells to sites of inflammation (14), References and Notes<br />

Figs. S1 to S7<br />

facilitates immune synapse formation (15), orches- 1. J. L. Flynn, J. Chan, Trends Microbiol. 13, 98 (2005).<br />

Table S1<br />

Videos S1 to S<br />

2. J. L. Flynn, J. Chan, Annu. Rev. Immunol. 19, 93 (2001).<br />

trates T cell interactions within lymph nodes<br />

3. Y. Wang et al., J. Immunol. 169, 2422 (2002).<br />

References<br />

(16), and controls the activation and differentiation 4. M. J. Gething, J. Sambrook, Nature 355, 33 (1992). 26 May 2006;<br />

of T cells (17). Our finding that a mycobacterial 5. P. Srivastava, Annu. Rev. Immunol. 20, 395 (2002). 10.1126/scienc<br />

Direct Demonstration of an<br />

Adaptive Constraint<br />

(kNADP cat /Km NA<br />

1 2 from 0.49 �<br />

Stephen BioTechnology P. Miller, Institute, Department of Ecology, Evolution<br />

and Behavior, University of Minnesota, St. Paul, MN<br />

The engine<br />

55108, USA.<br />

final R re<br />

*To whom correspondence should be addressed. E-mail:<br />

wild-type E<br />

deanx024@umn.edu<br />

cific toward<br />

458<br />

20 OCTOBER 2006 VOL 314 SCIENCE www.sciencemag.org<br />

1 Mark Lunzer, 1 Antony M. Dean 1,2 than a lack<br />

cause of a c<br />

be for a sim<br />

fitness is un<br />

Coenzym<br />

*<br />

genase (IM<br />

etiology and<br />

The role of constraint in adaptive evolution is an open question. Directed evolution of an engineered (10, 11). IM<br />

b-isopropylmalate dehydrogenase (IMDH), with coenzyme specificity switched from nicotinamide ation of b-<br />

adenine dinucleotide (NAD) to nicotinamide adenine dinucleotide phosphate (NADP), always produces during the<br />

mutants with lower affinities for NADP. This result is the correlated response to selection for relief amino acid.<br />

from inhibition by NADPH (the reduced form of NADP) expected of an adaptive landscape subject to dinucleotide<br />

three enzymatic constraints: an upper limit to the rate of maximum turnover (kcat ), a correlation in This invaria<br />

NADP and NADPH affinities, and a trade-off between NAD and NADP usage. Two additional constraints, at the presen<br />

high intracellular NADPH abundance and the cost of compensatory protein synthesis, have ensured some relate<br />

the conserved use of NAD by IMDH throughout evolution. Our results show that selective mechanisms use NADP<br />

and evolutionary constraints are to be understood in terms of underlying adaptive landscapes.<br />

Structura<br />

using IDHs<br />

he old notion of natural selection as an constraint, can be broken by artificial selection to enzyme use<br />

omnipotent force in biological evolution produce new phenotypic combinations (8, 9). De- replacemen<br />

Thas<br />

given way to one where adaptive prospite all circumstantial evidence, results from Ile<br />

cesses are constrained by physical, chemical, and direct experimental tests imply that selection is<br />

biological exigencies (1–4). Whether constraint<br />

and/or stabilizing selection explain phenotypic<br />

stasis, in the fossil record and in phylogenies, remains<br />

an open question (5). Direct experimental<br />

tests of constraint are scarce (6–9). Even tight<br />

largely unconstrained.<br />

The direct experimental test for constraint is<br />

conceptually simple. A phenotype is subjected to<br />

selection (natural or artificial) in an attempt to<br />

break the postulated constraint (6–9). A response<br />

correlations among traits, at once suggestive of to selection indicates a lack of constraint. No response<br />

to selection indicates the presence of a<br />

constraint. However, the cause of a constraint is<br />

rarely specified because the etiologies of most<br />

phenotypes are not well understood, their relationships<br />

to fitness are usually opaque, and a lack of<br />

response to selection may reflect nothing more<br />

290 Y Ty<br />

into the coe<br />

coli leuB–e<br />

genesis cau<br />

(10, 11): N<br />

Km is the<br />

factor of 34<br />

103 M –1 s<br />

1 2<br />

BioTechnology Institute, Department of Ecology, Evolution<br />

and Behavior, University of Minnesota, St. Paul, MN<br />

55108, USA.<br />

*To whom correspondence should be addressed. E-mail:<br />

deanx024@umn.edu<br />

(kNADP cat /Km NA<br />

from 0.49 �<br />

The engine<br />

final R rep<br />

wild-type E<br />

cific toward<br />

458<br />

24<br />

20 OCTOBER 2006 VOL 314 SCIENCE www.sciencemag.org


M –1 s –1 , whereas NADP performance<br />

(k NADP<br />

cat /K NADP<br />

m ) is increased by a factor of<br />

70, from 0.49 × 103 M –1 s –1 to 34 × 103 M –1<br />

s –1 . The engineered LeuB[RKYVYR] mutant<br />

(the final R represents Arg341 , already<br />

present in wild-type E. coli IMDH) is as<br />

active and as specific toward NADP as the<br />

wild-type enzyme is toward NAD. Evidently,<br />

protein architecture has not constrained<br />

IMDH to use NAD since the last<br />

common ancestor.<br />

Despite similar in vitro performances,<br />

the NADP-specific LeuB[RKYVYR] mu-<br />

tant is less fit than the NAD-specific wild<br />

type (11). IMDHs, wild-type and mutant<br />

alike, display a factor of 30 higher affinity<br />

for the reduced form of NADP (NADPH)<br />

(coproduct) than for NADP (Fig. 1B). We<br />

suggested the LeuB[RKYVYR] mutant is<br />

subject to intense inhibition by intracellular<br />

NADPH, which is far more abundant<br />

in vivo than is NADP (11, 17). The inhibition<br />

slows leucine biosynthesis, reduces<br />

growth rate, and lowers Darwinian fitness<br />

(Fig. 1C). The wild type retains high fitness<br />

because its affinity for NADPH is<br />

REPORTS<br />

Fig. 1. The molecular anatomy of an adaptive constraint. (A) Structural and NADPH (K NADPH<br />

phosphorus) with labels designating i the ) seen with engin<br />

alignment of the coenzyme binding pockets of E. coli amino IMDH acid (14) and (brown site number in the in screened IMDH followed mutants (dots). (C) Th<br />

main chain) with bound NAD modeled from Thermus thermophilus by the amino IMDH acid in by IDH. IMDH Coenzyme (11) showing use the phenotype<br />

(15) and E. coli IDH (16) (green main chain) with bound is determined NADP. Only by keyH<br />

bonds in chemostat to NAD competition (brown and descr<br />

residues are shown (gray, carbon; red, oxygen; blue, lines) nitrogen; and to yellow, the 2’-phosphate predicted (2’P) distribution of NADP of performanc<br />

phosphorus) with labels designating the amino acid and (green sitelines). number (B) Mutations in binding in the pocket coenzyme (gray dots), with wil<br />

IMDH followed by the amino acid in IDH. Coenzyme use binding is determined pocket by that H dot), stabilize andNADP single also amino acid repla<br />

bonds to NAD (brown lines) and to the 2-phosphate (2P) stabilize of NADP NADPH. (greenThe<br />

binding ensuing pocket correlation (pink dots). The fitne<br />

lines). (B) Mutations in the coenzyme binding pocket that in affinities stabilize for NADP NADP be (K lower than in the wild type bec<br />

also stabilize NADPH. The ensuing correlation in affinities for NADP (KNADP m ) intracellular NADPH (11).<br />

toward NAD. Evidently, protein architecture has<br />

not constrained IMDH to use NAD since the last<br />

common ancestor.<br />

Despite similar in vitro performances, the<br />

NADP-specific LeuBERKYVYR^ mutant is less<br />

fit than the NAD-specific wild type (11). IMDHs,<br />

PH (K NADPH<br />

i wild-type<br />

) seen Fig. 1. with The and<br />

engineered molecular mutant alike,<br />

mutants anatomy display<br />

(circles) of a factor an (11) adaptive of<br />

is<br />

30<br />

retained<br />

reened mutants<br />

higher constraint. (dots).<br />

affinity<br />

(C) (A) The<br />

for<br />

adaptive Structural the reduced<br />

landscape alignment form<br />

for<br />

of<br />

coenzyme<br />

NADP of the use<br />

(11) showing<br />

(NADPH) coenzyme the phenotype-fitness<br />

(coproduct) binding pockets map<br />

than for of (blue<br />

NADP E. coli surface) IMDH determined<br />

(Fig. 1B). (14)<br />

ostat competition<br />

We (brown and<br />

suggested main described<br />

thechain) by<br />

LeuBERKYVYR^ with equation bound S5 NAD (11, 18) and the<br />

mutant modeled is sub-<br />

d distribution of performances for 512 mutants in the coenzyme<br />

pocket (gray<br />

ject from<br />

dots),<br />

to intense Thermus<br />

with wild<br />

inhibition thermophilus<br />

type (red<br />

by intracellular IMDH (15)<br />

dot), LeuB[RKYVYR]<br />

NADPH, and E.<br />

(black<br />

d single amino<br />

which coli is IDH<br />

acid<br />

far(16) more(green replacements<br />

abundant main<br />

in<br />

in vivo chain)<br />

LeuB[RKYVYR]<br />

than with is NADP bound<br />

coenzyme<br />

pocket (pink(11, NADP.<br />

dots). 17). The The<br />

Only<br />

fitness inhibition<br />

key residues<br />

of LeuB[RKYVYR] slows leucine<br />

are shown<br />

biosynthesis,<br />

(gray,<br />

carbon; red, oxygen; blue, nitrogen; is hypothesized yellow, to<br />

r than in the reduces wild type growth because rate, and of lowers strong Darwinian inhibition fitness by abundant<br />

ular NADPH (Fig. (11). 1C). The wild type retains high fitness because Fig. 2. The phenotypic basis of the genetic screen. Lowered expr<br />

its affinity for NADPH is low, whereas inhibition by brings IMDH to saturation with isopropylmalate, 25 increasing co<br />

NADH, which is far less abundant than NAD in expression in the chemostat-derived adaptive landscape [lower con<br />

NADP<br />

m ) and NADPH<br />

(K NADPH<br />

i ) seen with engineered mutants<br />

(circles) (11) is retained in the screened<br />

mutants (dots). (C) The adaptive landscape<br />

for coenzyme use by IMDH (11) showing<br />

the phenotype-fitness map (blue surface)<br />

determined in chemostat competition and<br />

described by equation S5 (11, 18) and the<br />

predicted distribution of performances<br />

for 512 mutants in the coenzyme binding<br />

pocket (gray dots), with wild type (red dot),<br />

LeuB[RKYVYR] (black dot), and single amino<br />

acid replacements in LeuB[RKYVYR] coenzyme<br />

binding pocket (pink dots). The fitness of<br />

LeuB[RKYVYR] is hypothesized to be lower than<br />

in the wild type because of strong inhibition by<br />

abundant intracellular NADPH (11)


. Coenzyme use is determined by H<br />

2-phosphate (2P) of NADP (green<br />

inding pocket that stabilize NADP<br />

ation in affinities for NADP (K m NADP )<br />

cture has<br />

e the last<br />

ces, the<br />

nt is less<br />

IMDHs,<br />

tor of 30<br />

f NADP<br />

Fig. 1B).<br />

nt is sub-<br />

NADPH,<br />

is NADP<br />

ynthesis,<br />

an fitness<br />

s because<br />

ibition by<br />

NAD in<br />

uence of<br />

ure (18),<br />

DPH inuse.<br />

rained to<br />

ssociated<br />

, identifynhibition)<br />

causes of<br />

P (maxireak<br />

the<br />

ties (Fig.<br />

oenzyme<br />

nefit the<br />

romising<br />

therefore<br />

use NAD<br />

at ADP , an<br />

f NADP<br />

de-off in<br />

low, whereas inhibition by NADH, which<br />

is far less abundant than NAD in vivo, is<br />

ineffective. Perhaps as a consequence of<br />

differences in Michaelis complex structure<br />

(18), IDH is not subject to such intense<br />

NADPH inhibition and hence could<br />

evolve NADP use.<br />

We hypothesize that IMDH is constrained<br />

to use NAD because the strong<br />

inhibition associated with NADP use reduces<br />

fitness. However, identifying the<br />

mechanism of selection (NADPH inhibition)<br />

is not synonymous with identifying<br />

the causes of constraint. Mutations that<br />

increase k NADP<br />

cat (maximum rate of NADP<br />

turnover), that break the correlation in<br />

NADP and NADPH affinities (Fig. 1B),<br />

or that eliminate the trade-off in coenzyme<br />

performances (Fig. 1C) could each<br />

benefit the LeuB[RKYVYR] mutant without<br />

compromising its performance with<br />

NADP (18). We therefore hypothesize<br />

that IMDH is constrained to use NAD by<br />

three causes: an upper limit to kNADP ing) (19–21) to test whether NADP-specific<br />

IMDHs with higher fitness could<br />

be isolated. Random substitutions were<br />

introduced into leuB[RKYVYR] by means<br />

of error-prone polymerase chain reaction<br />

(18). Mutated alleles were ligated downstream<br />

of the T7 promoter in pETcoco (a<br />

stable single-copy vector) and transformed<br />

into strain RFS[DE3] (leuA<br />

, an<br />

cat<br />

+ BamC + We used directed evolution (targeted random<br />

mutagenesis and selective screening) (19–21) to<br />

test whether NADP-specific IMDHs with higher<br />

fitness could be isolated. Random substitutions<br />

were introduced into leuB[ RKYVYR] by means<br />

of error-prone polymerase chain reaction (18).<br />

Mutated alleles were ligated downstream of the<br />

T7 promoter in pETcoco (a stable single-copy<br />

vector) and transformed into strain RFSEDE3^<br />

(leuA , with T7<br />

RNA polymerase expressed from a chromosomal<br />

lacUV5 promoter). Sequencing<br />

unscreened plasmids revealed that, on average,<br />

each 1110–base pair leuB[RKYVYR]<br />

received two nucleotide substitutions.<br />

From the pattern of base substitutions in<br />

þBamCþ assuming Poisson statistics, we estimate that only<br />

10.5 (0.4%) of the 2431 possible amino acid replacements<br />

were missing in our mutant library (18).<br />

Our experimental design used decreased<br />

IMDH expression to provide a simple selective<br />

screen for mutations in leuB[ RKYVYR] that increase<br />

growth and/or fitness. As predicted from<br />

the phenotype-fitness map (Fig. 2A), selection<br />

against leuB[ RKYVYR] intensified as IMDH<br />

, with T7 RNA polymerase ex- expression was lowered in the presence of excess<br />

pressed from a chromosomal lacUV5 promoter). glucose (Fig. 2B). Using 40 mM isopropyl-b-D-<br />

Sequencing unscreened plasmids revealed that, thiogalactopyranoside (IPTG) to induce a low<br />

on average, each 1110–base pair leuB[ RKYVYR] level of IMDH expression, we found that cells<br />

received two nucleotide substitutions. From the harboring leuB[ WT] formed large colonies at 24<br />

pattern of base substitutions in these mutants and hours, whereas cells harboring leuB[ RKYVYR]<br />

ww.sciencemag.org SCIENCE VOL 314 20 OCTOBER 2006 459<br />

unbreakable correlation in the affinities of<br />

NADP and NADPH, and an inescapable<br />

trade-off in coenzyme performance.<br />

We used directed evolution (targeted<br />

random mutagenesis and selective screen-<br />

26<br />

dot), and single amino acid replacements in LeuB[RKYVYR] coenzyme<br />

binding pocket (pink dots). The fitness of LeuB[RKYVYR] is hypothesized to<br />

be lower than in the wild type because of strong inhibition by abundant<br />

intracellular NADPH (11).<br />

Fig. 2. The phenotypic basis of the genetic screen. Lowered expression in the presence of excess glucose<br />

brings IMDH to saturation with isopropylmalate, increasing coenzyme affinities. (A) Lowering IMDH<br />

expression in the chemostat-derived adaptive landscape [lower concentration of E in equation S5 (11, 18)] is<br />

predicted to reduce the fitness of LeuB[RKYVYR] (white sphere) far more than that of the wild type (red<br />

sphere). (B) IPTG-controlled expression of IMDHs ligated downstream of the T7 promoter in pETcoco in strain<br />

RFS[DE3] (leuA þ B am C þ , with T7 RNA polymerase expressed from a chromosomal lacUV5 promoter) confirms<br />

that lower expression affects growth in minimal glucose medium of the LeuB[RKYVYR] (white) more than in<br />

the wild type (red). Plasmids lacking LeuB (black) are incapable of growth except in the presence of leucine.<br />

these mutants and assuming Poisson statistics,<br />

we estimate that only 10.5 (0.4%) of<br />

the 2431 possible amino acid replacements<br />

were missing in our mutant library (18).<br />

Our experimental design used decreased<br />

IMDH expression to provide a<br />

simple selective screen for mutations<br />

in leuB[RKYVYR] that increase growth<br />

and/or fitness. As predicted from the<br />

phenotype-fitness map (Fig. 2A), selection<br />

against leuB[RKYVYR] intensified<br />

as IMDH expression was lowered in the<br />

presence of excess glucose (Fig. 2B).<br />

Using 40 μM isopropyl-β-D-thiogalacto


eplacement in the coenzyme binding pocket, or<br />

both. Upstream substitutions occurred at three<br />

nucleotide positions: –3, –9, and –14 relative to<br />

eliminate H bonds to the 2-phosphate of NADP<br />

to cause striking reductions in NADP performance<br />

(Table 1). Although some mutants have improved<br />

Table 1. Kinetic effects of amino acid replacements in LeuB[RKYVYR] isolated at 48 hours growth.<br />

Standard errors are G13% of estimates. Single-letter abbreviations for amino acid residues: A, Ala;<br />

C, Cys; D, Asp; E, Glu; F, Phe; G, Gly; H, His; I, Ile; K, Lys; L, Leu; M, Met; N, Asn; P, Pro; Q, Gln; R,<br />

Arg; S, Ser; T, Thr; V, Val; W, Trp; Y, Tyr.<br />

NADP NAD<br />

Enzyme<br />

Performance<br />

(A)<br />

Performance<br />

(B)<br />

Preference<br />

(A/B)<br />

K NADPH<br />

i<br />

(mM)<br />

Km (mM)<br />

kcat (s –1 )<br />

kcat /Km (mM –1 s –1 )<br />

Km (mM)<br />

kcat (s –1 )<br />

kcat /Km (mM –1 s –1 )<br />

Site<br />

DDIAGR (wild type) 8400 4.10 0.4900 101 6.90 68.3200 0.007 254.30<br />

RKYVYR 183 6.20 33.8800 4108 0.83 0.2000 169.400 9.90<br />

Beneficial replacements<br />

K235N 3919 5.95 1.5200 6356 0.50 0.0800 19.000 129.00<br />

K235N,–3(M),A60V 1744 5.06 2.9000 5395 0.44 0.0800 35.700 56.00<br />

K235R 554 6.39 11.5000 3292 0.36 0.1100 106.100 26.00<br />

K235R,E63V 658 6.75 10.3000 3748 0.46 0.1200 84.100 28.00<br />

K289E* 3449 2.71 0.7900 4897 1.98 0.4000 2.000 133.50<br />

K289E*,D317E 3554 3.62 1.0200 5580 2.09 0.3700 2.800 112.00<br />

K289E*,D87Y,S182F 2551 2.54 1.0000 5750 3.65 0.6300 1.600 105.20<br />

K289E*,F102L 3891 3.66 0.9400 4823 1.10 0.2300 4.100 131.90<br />

K289M 2216 4.13 1.8600 4034 0.48 0.1200 15.500 78.60<br />

K289M,F170L 1945 2.50 1.2900 2947 0.48 0.1600 8.100 87.20<br />

K289N* 1141 6.32 5.5400 4997 1.14 0.2300 24.300 44.00<br />

K289N*,R187H 1146 5.83 5.0900 3069 0.79 0.2600 19.600 44.70<br />

K289N*,H367Q 1248 5.27 4.2200 4555 1.05 0.2300 18.300 57.80<br />

K289T 1596 6.18 3.8700 4056 0.78 0.1900 20.400 71.60<br />

K289T,Q157H 1696 5.66 3.3400 5038 0.93 0.1800 18.600 72.60<br />

Y290C 988 5.21 5.2700 3792 1.02 0.2700 19.600 49.00<br />

Y290C,R152C 910 3.15 3.4600 3844 0.90 0.2300 14.800 43.00<br />

Y290D 10213 3.13 0.3100 9040 0.82 0.0900 3.400 344.00<br />

Y290F* 1658 5.59 3.3700 3110 0.65 0.2100 16.000 59.70<br />

Y290F,P97S,D314E 1097 4.64 4.2300 4158 0.71 0.1700 24.800 57.40<br />

Y290N,N52T 2381 3.44 1.4400 10660 0.26 0.0200 59.500 90.00<br />

Replacements with beneficial 5-leader mutations<br />

RKYVYR 183 6.20 33.8800 4108 0.83 0.2000 169.400 9.90<br />

–3(M)†,E66D 208 2.97 14.2800 5002 0.30 0.0600 238.000 11.90<br />

–3(M)†,E66D,E331D 267 7.99 29.9000 3403 0.34 0.1000 296.900 14.00<br />

E82D 221 5.44 24.6200 4735 0.35 0.0700 351.700 14.10<br />

K100R,A229T,L248M 137 5.58 40.3700 4079 0.45 0.1100 370.300 11.50<br />

G156R,K289R,Y311H 292 5.40 18.4900 4464 0.54 0.1200 144.800 17.40<br />

E173D 219 6.39 29.1800 5304 0.89 0.1700 171.600 6.60<br />

Y337N,G131 171 5.71 33.3900 3367 0.24 0.0700 468.500 8.00<br />

Y337H,P85S,E181K 318 5.21 16.3800 2409 0.22 0.0900 183.600 12.80<br />

*Switch from NADP to NAD requires two base substitutions in one codon. This is a possible transitional amino acid produced by<br />

a single base substitution (11). †The –3(M) designates a G to A substitution at base –3 that creates a new start codon.<br />

460<br />

pyranoside (IPTG) to induce a low level terion compatible with reliably identifying<br />

20 OCTOBER 2006 VOL 314 SCIENCE www.sciencemag.or<br />

of IMDH expression, we found that cells beneficial mutations in leuB[RKYVYR].<br />

harboring leuB[WT] formed large colo- Longer periods of growth (or higher connies<br />

at 24 hours, whereas cells harboring centrations of IPTG) allowed unmutated<br />

leuB[RKYVYR] barely formed pinprick leuB[RKYVYR] to form colonies, whereas<br />

colonies at 48 hours. We chose colony for- shorter periods (or lower concentrations of<br />

mation at 48 hours as the least stringent cri- IPTG) produced no colonies.<br />

27<br />

are benef<br />

formance<br />

if there ar<br />

upper lim<br />

for NAD<br />

enzyme p<br />

with redu<br />

phenotyp<br />

remains f<br />

performan<br />

of NADP<br />

cial as co<br />

NADPH f<br />

As predi<br />

mutants h<br />

and NAD<br />

increases<br />

eficial (18<br />

that NAD<br />

vivo beca<br />

NADPH.<br />

and increa<br />

the predic<br />

map cons<br />

correlation<br />

and a trad<br />

Break<br />

NADP-sp<br />

increases<br />

for NADP<br />

trade-off<br />

clusions a<br />

is too stri<br />

distinct m<br />

limit, 17<br />

leader seq<br />

LeuBERK<br />

ments in<br />

zyme pe<br />

inhibition<br />

cial muta<br />

teria dem<br />

not overly<br />

Nor ar<br />

sampling<br />

adaptive<br />

sequential<br />

lution de<br />

mutants<br />

coli are n<br />

force mu<br />

vantageo


Of the 100,000 mutated plasmids<br />

screened, 134 (representing 107 distinct<br />

isolates) formed colonies within 48 hours<br />

(table S1). Each had either a substitution<br />

in the 5′ leader sequence upstream of<br />

leuB[RKYVYR], an amino acid replacement<br />

in the coenzyme binding pocket, or<br />

both. Upstream substitutions occurred at<br />

three nucleotide positions: –3, –9, and –14<br />

relative to the leuB AUG start codon (fig.<br />

S1). Ten amino acid replacements were<br />

found at three codons in the coenzyme<br />

binding pocket.<br />

At first glance, the positive response to<br />

selection might suggest that coenzyme use<br />

by IMDH is unconstrained. Upstream substitutions<br />

in the Shine-Dalgarno sequence<br />

(positions –9 and –14), as well as a new<br />

AUG start codon (position –3) that replaces<br />

the less efficient GUG start codon,<br />

presumably derive their benefits through<br />

increases in expression because their kinetics<br />

are unchanged. Unexpectedly, however,<br />

beneficial amino acid replacements<br />

in the coenzyme binding pocket eliminate<br />

H bonds to the 2′-phosphate of NADP to<br />

cause striking reductions in NADP performance<br />

(Table 1). Although some mutants<br />

have improved NAD performance, others<br />

remain unchanged and several show reduced<br />

NAD performance. Isolated amino<br />

acid replacements outside the coenzyme<br />

binding pocket, which might have been<br />

expected to increase kNADP or to break the<br />

cat<br />

correlation in affinities between NADP<br />

and NADPH (K NADP<br />

m and K NADPH<br />

i ), have<br />

no detectable functional effects (table S2,<br />

those associated with beneficial 5′-leader<br />

mutations in Table 1). No doubt they,<br />

along with 126 silent substitutions (table<br />

S1), hitchhiked through the genetic screen<br />

with the beneficial mutations.<br />

Our results suggest that increases in expression<br />

are beneficial, whereas increases<br />

in NADP performance are not. This seeming<br />

paradox is resolved if there are no<br />

mutations capable of breaking the upper<br />

limit to kNADP the correlation in affinities<br />

cat<br />

for NADP and NADPH, or the trade-off in<br />

28<br />

coenzyme performance. With these constraints,<br />

and with reduced expression in<br />

the genetic screen, the phenotype-fitness<br />

map near LeuB[RKYVYR] remains flat<br />

with respect to increases in NADP performance<br />

(Fig. 2A). By contrast, severe losses<br />

of NADP performance are predicted to<br />

be beneficial as correlated reductions in<br />

the affinities for NADPH free up IMDH<br />

for use with abundant NAD. As predicted,<br />

all beneficial LeuB[RKYVYR] mutants<br />

have reduced affinities for both NADP and<br />

NADPH (Table 1). Unaffected by constraints,<br />

increases in expression are unconditionally<br />

beneficial (18). These results<br />

support the hypothesis that NADP-specific<br />

IMDHs function poorly in vivo because of<br />

strong inhibition by abundant NADPH.<br />

That reductions in NADP performance<br />

and increases in expression are both beneficial<br />

are the predicted consequences of a<br />

phenotype-fitness map constrained by an<br />

upper limit to k NADP<br />

cat , a correlation in affinities<br />

for NADP and NADPH, and a tradeoff<br />

in coenzyme performance.<br />

Breaking any one constraint would allow<br />

NADP-specific IMDHs to evolve.<br />

Yet no mutant increases k NADP<br />

cat , no mutant<br />

uncouples the affinities for NADP and<br />

NADPH, and no mutant breaks the tradeoff<br />

in coenzyme performance. These conclusions<br />

are not the result of a selective<br />

screen that is too stringent. Of 35 colonies,<br />

representing 26 distinct mutants, that appeared<br />

after the 48-hour limit, 17 had no<br />

nucleotide substitutions in the 5′ leader<br />

sequence or amino acid replacements in<br />

LeuB[RKYVYR] (table S3). Amino acid<br />

replacements in the other mutants neither<br />

improved enzyme performance nor<br />

decreased NADPH inhibition (table S4).<br />

That no additional beneficial mutations<br />

were recovered with relaxed criteria demonstrates<br />

that the selective screen was not<br />

overly stringent.<br />

Nor are the results a consequence of<br />

inadequate sampling of protein sequence<br />

space. Natural adaptive evolution fixes<br />

advantageous mutations sequentially (22–


mpling of<br />

olutionary<br />

04 advane<br />

missing<br />

(18). We<br />

eaking mplingthe of<br />

re olutionary unlikely<br />

04 e they advan- are<br />

e), missing because<br />

oth. (18). We<br />

an aking NADP- the<br />

er unlikely NADPH<br />

ese they expres- are<br />

elieves , because the<br />

sources oth. of<br />

the n NADP- rest of<br />

rnefit NADPH to be<br />

se (Fig. expres- 1C)<br />

lieves such that the<br />

bove sources wild- of<br />

the rcome restthe of<br />

nefit resources to be<br />

d(Fig. compen- 1C)<br />

such a protein that<br />

the oveevolu wildcome<br />

the<br />

otypes, resources by<br />

ring, compen- is not<br />

constraints a protein<br />

he constraints evolu-<br />

e identified<br />

phenotype, types, by<br />

ing, landscape. is not<br />

constraints sed to test<br />

constraints we have<br />

lationships ion phenotype<br />

24). Indeed, experimental evolution demonstrates<br />

that advantageous double mutants<br />

in the evolved ß-galactosidase of E.<br />

coli are not evolutionarily accessible and<br />

perforce must be accumulated as sequential<br />

advantageous mutations (25). Hence,<br />

screening all single substitutions is a<br />

sufficient sampling of protein sequence<br />

space for robust evolutionary conclusions.<br />

We estimate that only 0.04 advantageous<br />

amino acid replacements are missing from<br />

the mutant leuB[RKYVYR] library (18).<br />

We conclude that mutations capable of<br />

breaking the limit, the correlation, or the<br />

trade-off are unlikely to ever be fixed in<br />

populations because they are exceedingly<br />

rare (they may not exist), because they are<br />

minimally advantageous, or both.<br />

The two remaining ways to evolve an<br />

underlying NADP-specific a conserved IMDH phenotype are to can reduce be identified intra-<br />

from cellular the relationships NADPH pool amongand, genotype, as our phenotype, results<br />

and show, fitness to increase that define expression. an adaptive Reducing landscape. in-<br />

Experimental tracellular NADPH evolution relieves can then the be used inhibition to test<br />

their<br />

but,<br />

existence.<br />

as experiments<br />

Using this<br />

deleting<br />

approach,<br />

sources<br />

we have<br />

underlying of<br />

shown howa certain conserved structure-function phenotype can be relationships identified<br />

from NADPH show (13), the disruption to the<br />

in IMDH the relationships have constrained among itsgenotype, coenzyme phenotype, phenotype<br />

and since rest fitness of themetabolism lastthat common defineancestor. costs an adaptive far more landscape. than the<br />

Experimental evolution can then be used to test<br />

their existence. Using this approach, we have<br />

shown References how certain andstructure-function Notes relationships<br />

1. S. J. Gould, R. C. Lewontin, Proc. R. Soc. London Ser. B<br />

in IMDH have constrained its coenzyme phenotype<br />

205, 581 (1979).<br />

since 2. J. the Antonovics, last common P. H. vanancestor. Tienderen, Trends Ecol. Evol. 6,<br />

166 (1991).<br />

3. M. R. Rose, G. V. Lauder, Adaptation (Academic Press,<br />

References San Diego, CA, and 1996). Notes<br />

1. 4. S. M. J. Pigliucci, Gould, R. J. C. Kaplan, Lewontin, Trends Proc. Ecol. R. Evol. Soc. 15, London 66 (2000). Ser. B<br />

5. 205, N. Eldredge 581 (1979). et al., Paleobiology 31, 133 (2005).<br />

2. 6. J. M. Antonovics, Travisano, J. P. A. H. Mongold, van Tienderen, A. F. Bennett, Trends Ecol. R. E. Evol. Lenski, 6,<br />

166 <strong>Science</strong> (1991). 267, 87 (1995).<br />

3. 7. M. H. R. Teotónio, Rose, G. M. V. R. Lauder, Rose, Nature Adaptation 408, (Academic 463 (2000). Press,<br />

8. San P. Beldade, Diego, CA, K. Koops, 1996). P. M. Brakefield, Nature 416, 844<br />

4. M. (2002). Pigliucci, J. Kaplan, Trends Ecol. Evol. 15, 66 (2000).<br />

5. 9. N. W. Eldredge A. Frankino, et al., B. J. Paleobiology Zwaan, D. L. 31, Stern, 133 P. (2005). M. Brakefield,<br />

6. M. <strong>Science</strong> Travisano, 307, J. 718 A. Mongold, (2005). A. F. Bennett, R. E. Lenski,<br />

10. <strong>Science</strong> R. Chen, 267, A. Greer, 87 (1995). A. M. Dean, Proc. Natl. Acad. Sci.<br />

7. H. U.S.A. Teotónio, 93, 12171 M. R. Rose, (1996). Nature 408, 463 (2000).<br />

11. 8. P. M. Beldade, Lunzer, S. K. P. Koops, Miller, P. R. M. Felsheim, Brakefield, A. Nature M. Dean, 416, <strong>Science</strong> 844<br />

(2002). 310, 499 (2005).<br />

12. 9. W. A. A. M. Frankino, Dean, G. B. J. Golding, Zwaan, Proc. D. L. Stern, Natl. Acad. P. M. Sci. Brakefield, U.S.A.<br />

<strong>Science</strong> 94, 3104 307, (1997). 718 (2005).<br />

10. 13. R. G. Chen, Zhu, G. A. B. Greer, Golding, A. M. A. Dean, M. Dean, Proc. <strong>Science</strong> Natl. Acad. 307, Sci. 1279<br />

U.S.A. (2005); 93, published 12171 (1996). online 13 January 2005 (10.1126/<br />

11. M. science.1106974).<br />

Lunzer, S. P. Miller, R. Felsheim, A. M. Dean, <strong>Science</strong><br />

14. 310, G. Wallon 499 (2005). et al., J. Mol. Biol. 266, 1016 (1997).<br />

12. 15. A. J. H. M. Hurley, Dean, G. A. B. M. Golding, Dean, Structure Proc. Natl. 2, 1007 Acad. (1994). Sci. U.S.A.<br />

16. 94, J. H. 3104 Hurley, (1997). A. M. Dean, D. E. Koshland Jr., R. M. Stroud,<br />

13. G. Biochemistry Zhu, G. B. 30, Golding, 8671 A. (1991). M. Dean, <strong>Science</strong> 307, 1279<br />

ole for 17. (2005); T. Penfound, LEDGF/p75<br />

published J. W. Foster, onlinein13 Escherichia January 2005 coli and (10.1126/ Salmonella:<br />

science.1106974).<br />

Cellular and Molecular Biology, F. C. Neidhardt, Ed. (ASM<br />

Press, Washington, DC, ed. 2, 1996), pp. 721–730.<br />

18. See supporting material on <strong>Science</strong> Online.<br />

19. O. Kuchner, F. H. Arnold, Trends Biotechnol. 58, 1 (1997).<br />

20. F. H. Arnold, P. L. Wintrobe, K. Miyazaki, A. Gershenson,<br />

Meehan, Phonphimon Wongthida, Mary Peretz,<br />

Trends Biochem. Sci. 26, 100 (2001).<br />

ole for LEDGF/p75<br />

REPORTS<br />

benefit to be gained. The phenotype-fitness<br />

map (Fig. 1C) imposes a law of diminishing<br />

returns such that LeuB[RKYVYR]<br />

must be expressed above wild-type levels<br />

by a factor of 100 to overcome the inhibition<br />

by NADPH (18). Diverting resources<br />

away from other metabolic needs toward<br />

compensatory protein synthesis would impose<br />

a protein burden (26–29) sufficient<br />

to prevent the evolution of NADP-specific<br />

IMDHs.<br />

The production of unnatural phenotypes,<br />

by artificial selection or molecular<br />

engineering, is not sufficient to conclude<br />

that evolutionary constraints are absent<br />

entirely. Rather, potential constraints underlying<br />

a conserved phenotype can be<br />

identified from the relationships among<br />

genotype, phenotype, and fitness that define<br />

14. G. an Wallon adaptive et al., J. Mol. landscape. Biol. 266, 1016 Experimental<br />

(1997).<br />

evolution 15. J. H. Hurley, can A. M. then Dean, be Structure used 2, to 1007 test (1994). their<br />

16. J. H. Hurley, A. M. Dean, D. E. Koshland Jr., R. M. Stroud,<br />

existence. Biochemistry Using 30, 8671 this (1991). approach, we have<br />

shown 17. T. Penfound, how J. certain W. Foster, in structure-function Escherichia coli and Salmonella: re-<br />

Cellular and Molecular Biology, F. C. Neidhardt, Ed. (ASM<br />

lationships 14. G. Press, Wallon Washington, etin al., IMDH J. DC, Mol. ed. Biol. have 2, 1996), 266, constrained 1016 pp. 721–730. (1997). its<br />

coenzyme 15. 18. J. See H. supporting Hurley, phenotype A. M. material Dean, on Structure since <strong>Science</strong>2, the Online. 1007 last (1994). com-<br />

16. 19. J. O. H. Kuchner, Hurley, F. A. H. M. Arnold, Dean, D. Trends E. Koshland Biotechnol. Jr., 58, R. M. 1 Stroud, (1997).<br />

mon<br />

20. Biochemistry ancestor.<br />

F. H. Arnold, 30, P. L. 8671 Wintrobe, (1991). K. Miyazaki, A. Gershenson,<br />

17. T. Trends Penfound, Biochem. J. W. Foster, Sci. 26, in100 Escherichia (2001). coli and Salmonella:<br />

21. Cellular L. G. Otten, and Molecular W. J. Quax, Biology, Biomol. F. C. Eng. Neidhardt, 22, 1 (2005). Ed. (ASM<br />

22. Press, S. Wright, Washington, in Proceedings DC, ed. 2, of 1996), the Sixth pp. International 721–730.<br />

18. See Genetics supporting Congress, material D. F. on Jones, <strong>Science</strong> Ed. (Brooklyn Online. Botanic<br />

19. O. Garden, Kuchner, Menasha, F. H. Arnold, WI, 1932), Trendspp. Biotechnol. 356–366. 58, 1 (1997).<br />

20. 23. F. J. H. Maynard Arnold, Smith, P. L. Wintrobe, Nature 225, K. Miyazaki, 563 (1970). A. Gershenson,<br />

24. Trends D. M. Weinreich, Biochem. Sci. N. F. 26, Delaney, 100 (2001). M. A. Depristo, D. L. Hartl,<br />

21. L. <strong>Science</strong> G. Otten, 312, W. 111 J. Quax, (2005). Biomol. Eng. 22, 1 (2005).<br />

22. 25. S. B. Wright, G. Hall, inAntimicrob. Proceedings Agents of theChemother. Sixth International 46, 3035<br />

Genetics (2002). Congress, D. F. Jones, Ed. (Brooklyn Botanic<br />

26. Garden, S. Zamenhof, Menasha, H. H. WI, Eichhorn, 1932), pp. Nature 356–366. 216, 456 (1967).<br />

23. 27. J. K. Maynard J. Andrews, Smith, G. D. Nature Hegeman, 225, J. 563 Mol. (1970). Evol. 8, 317 (1976).<br />

24. 28. D. M. E. Weinreich, Dykhuizen, N. Evolution F. Delaney, 32, M. 125 A. Depristo, (1978). D. L. Hartl,<br />

29. <strong>Science</strong> A. L. Koch, 312, J. 111 Mol. (2005). Evol. 19, 455 (1983).<br />

25. 30. B. Supported G. Hall, Antimicrob. by NIH grant Agents GM060611 Chemother. (A.M.D.). 46, We 3035 thank<br />

(2002). three anonymous reviewers whose constructive criticisms<br />

26. S. weZamenhof, found invaluable. H. H. Eichhorn, Nature 216, 456 (1967).<br />

27. K. J. Andrews, G. D. Hegeman, J. Mol. Evol. 8, 317 (1976).<br />

Supporting 28. D. E. Dykhuizen, OnlineEvolution Material 32, 125 (1978).<br />

www.sciencemag.org/cgi/content/full/314/5798/458/DC1<br />

29. A. L. Koch, J. Mol. Evol. 19, 455 (1983).<br />

Materials 30. Supported and Methods by NIH grant GM060611 (A.M.D.). We thank<br />

SOM three Text anonymous reviewers whose constructive criticisms<br />

Figs. we S1 found to S3 invaluable.<br />

Tables S1 to S5<br />

Supporting References Online Material<br />

www.sciencemag.org/cgi/content/full/314/5798/458/DC1<br />

4 August 2006; accepted 20 September 2006<br />

Materials and Methods<br />

10.1126/science.1133479<br />

SOM Text<br />

Figs. S1 to S3<br />

Tables S1 to S5<br />

References the HIV-1 IN protein from rapid degradation in<br />

4the August 26S2006; proteasome accepted 20(8). September In the2006 bona fide viral<br />

10.1126/science.1133479<br />

context, drastic knockdown of p75 changed the<br />

genomic pattern of HIV-1 integration by reducing<br />

the viral bias for active genes, which suggests that 29<br />

the p75HIV-1 influences IN protein integration fromtargeting rapid degradation (9). However, in<br />

the 26S proteasome (8). In the bona fide viral<br />

REPORTS<br />

REPORTS


Tech Notes<br />

Large Insertions: Two Simple Steps Using<br />

QuikChange® II Site-Directed <strong>Mutagenesis</strong> Kits<br />

Abstract<br />

This Technical Note describes a modified<br />

method to introduce large insertions in<br />

plasmids in two simple steps using the<br />

QuikChange ® II Site-Directed <strong>Mutagenesis</strong><br />

Kit. Dr. Wenge Wang in Dr. Wafik S.<br />

El-Deiry’s lab at the University of Pennsylvania<br />

has used the QuikChange kit and<br />

a modification of the method described<br />

by Geiser et al. (2001) 1 to perform a large<br />

insertion of a PCR product into a target<br />

plasmid. He had inserted the enhanced<br />

green fluorescent protein (EGFP) open<br />

reading frame (760 bp) following the Cterminus<br />

of death receptor TRAIL receptor<br />

1/death receptor 4 (TRAIL-R1/DR4)<br />

right after the signal sequence, using the<br />

QuikChange II site-directed mutagenesis<br />

kit with some modifications to the basic<br />

protocol. Dr. Wang transfected cells with<br />

the recombined plasmid and it generated<br />

a fusion protein with the correct molecular<br />

size (Figure 1).<br />

Background<br />

Dr. Wafik S. El-Deiry’s lab is focused on<br />

studying the mechanism of action of the<br />

tumor suppressor p53 and the contribution<br />

of its downstream target genes to<br />

control cell growth. Their lab has identified<br />

a number of genes that are directly<br />

regulated by p53 and which can inhibit<br />

cell cycle progression (p21WAF1), induce<br />

apoptosis (KILLER/DR5, Bid, caspase 6,<br />

Traf4 and others) or activate DNA repair<br />

(DDB2). The research has provided knowledge<br />

into the tissue specificity of the DNA<br />

damage response in vivo and into the<br />

mechanism by which wild-type p53 sensitizes<br />

cells to killing by anti-cancer drugs.<br />

High-Efficiency <strong>Mutagenesis</strong> Method<br />

All QuikChange ® kits a offer a one-day<br />

method to introduce point mutations, amino<br />

acid substitutions, small insertions, and<br />

deletions in virtually any double-stranded<br />

30<br />

plasmid template at efficiencies greater<br />

than 80% (Figure 2). The QuikChange II<br />

and QuikChange II XL kits feature our<br />

highest fidelity DNA polymerase, our gold<br />

standard for accuracy, and a linear amplification<br />

strategy. Together, they reduce the<br />

mutation frequency due to incorporation<br />

errors typically caused by using an exponential<br />

PCR amplification approach. The<br />

result is high-efficiency mutagenesis without<br />

unwanted errors.<br />

Ideal for Large Insertions and Deletions<br />

We have used the QuikChange kit method<br />

for small insertions and deletions (~12 bp)<br />

and observed greater than 80% efficiency 2 .<br />

Many of our customers have used the<br />

QuikChange method or have modified it<br />

to introduce deletions of 31 bp 3 , 87 bp 4 ,<br />

and 3 kb 5 or insertions of 31 bp 3 and up<br />

to ~1 kb 5 . In order to achieve large insertions<br />

of up to 1 kb, megaprimers must be<br />

generated from an initial PCR reaction. To<br />

ensure the highest mutagenesis efficiency,<br />

we recommend gel purification with our<br />

StrataPrep ® PCR Purification Kit. To create<br />

the megaprimers, a minimum of 20 bp<br />

upstream and downstream of the insertion<br />

sequence should be complementary to<br />

your template vector that will be used in<br />

the QuikChange kit reaction. Therefore,<br />

each oligo used in the initial PCR reaction<br />

will require 20 bp overhangs on the<br />

5’ ends.<br />

Using the QuikChange kit for large deletions<br />

is even easier since oligos can be<br />

synthesized, rendering the initial PCR reaction<br />

and fragment purification unnecessary.<br />

Again, 20 to 30 bp may be required<br />

to bind to your template DNA both upstream<br />

and downstream of the region that<br />

you want to delete. However, the largest<br />

oligo required should not exceed 60 bp.<br />

This Technical Note describes performing


a large insertion that can be completed<br />

in two simple steps. First, PCR-amplify<br />

the megaprimer. Second, add your gel<br />

purified megaprimer to our QuikChange<br />

site-directed mutagenesis kit (Figure 2).<br />

This strategy avoids tedious and time-consuming<br />

sub-cloning. With our QuikChange<br />

kits, you will have confidence in generating<br />

an error free clone while saving valuable<br />

time.<br />

Materials & Methods<br />

Part I. PCR reaction to generate<br />

the megaprimer<br />

Reaction and cycling conditions.<br />

PCR reagents: PfuUltra ® II Fusion HS DNA<br />

Polymerase (Stratagene)<br />

Thermal Cycler used: TECHNE: Touchgene<br />

® Gradient<br />

• The forward primer: GACTCCGAATCC<br />

CGGGAGCGCAGCGGTGAGCAAGG<br />

GCGAGGAG, the first 25 bp overlaps<br />

the target vector, the remaining primer<br />

sequence is complementary to EGFP.<br />

The primers were resuspended in 50<br />

mM NaCl.<br />

• The reverse primer: CGCGGCTGCCTC<br />

TGTCCCACTCTTGTACAGCTCGTCCA<br />

TGCC, the first 22 bp targets the vector<br />

and the remaining primer sequence is<br />

complementary to EGFP. The primers<br />

were resuspended in 50 mM NaCl.<br />

• Template for PCR reaction: 200 ng of<br />

plasmid DNA (pEGFP-N1; Clontech)<br />

• PfuUltra ® II Fusion HS DNA Polymerase<br />

(Stratagene)<br />

• PCR product was purified with<br />

QIAGEN’S QIAEX II Gel Extraction Kit<br />

• The quantity of the PCR product was<br />

estimated on agarose gel with Ethidium<br />

Bromide<br />

Note: The size of the PCR product is 760<br />

bp, which includes the insert (EGFP) of<br />

714 bp plus overlapping sequence of the<br />

target vector.<br />

Fig. 1 Enhanced green fluorescent protein (EGFP)<br />

was inserted in the open reading frame (760 bp) following<br />

the C-terminus of death receptor TRAIL receptor<br />

1/death receptor 4 (TRAIL-R1/DR4) right after the<br />

signal sequence. Cell staining shows the member<br />

protein fused with EGFP.<br />

1. Mutant Strand Synthesis<br />

Perform thermal cycling to:<br />

• Denature DNA template<br />

• Anneal mutagenic primers<br />

containing desired mutation<br />

• Extend and incorporate primers<br />

with high-fidelity DNA polymerase<br />

2. Dpn I Digestion of Template<br />

Digest parental methylated and<br />

hemimethylated DNA with Dpn I<br />

3. Transformation<br />

Transform mutated molecule into<br />

competant cells for nick repair<br />

Fig. 2. The QuikChange ® II One-Day Site-Directed<br />

<strong>Mutagenesis</strong> Method. 1. Mutant strand synthesis. 2.<br />

Dpn I Digestion of parental DNA template. 3. Transformation<br />

of the resulting annealed double-stranded<br />

nicked DNA molecules. After transformation, the XL-1<br />

Blue E. coli cell repairs nicks in the plasmid.<br />

31


Tech Notes<br />

Part II. The QuikChange II<br />

mutagenesis reaction can<br />

be set up using the following<br />

guidelines (Tables 1<br />

and 2).<br />

• Primer for the insertion<br />

reaction 300-500<br />

ng of PCR product (760<br />

bp EGFP) to serve as a<br />

megaprimer.<br />

Note: Follow reaction TABLE 2. Cycling Method<br />

with Dpn I digestion and<br />

transformation per the<br />

QuikChange<br />

Segment<br />

1<br />

Cycles<br />

1<br />

Temperature<br />

95°C<br />

Time<br />

30 seconds<br />

2 5 95°C 30 seconds<br />

52°C 1 minute<br />

68°C<br />

1 min/kb of<br />

plasmid* (7 min)<br />

3 13 95°C 30 seconds<br />

55°C 1 minute<br />

68°C<br />

1 min/kb of<br />

plasmid* (7 min)<br />

* For example, a 7 kb plasmid would have a 7 minute extension time<br />

® II manual.<br />

For PfuUltra ® II Fusion HS<br />

DNA Polymerase information<br />

please refer to<br />

the manual.<br />

Results<br />

The results of this mutagenesis<br />

experiment<br />

demonstrate that large<br />

PCR fragments can be successfully<br />

inserted into the target plasmid DNA. The 760 bp PCR fragment had at<br />

each end only short regions of homology where the insertion occurred. There were<br />

approximately 100 colonies produced. About 20 colonies were selected for further<br />

analysis of mutagenesis efficiency, and they were 100% positive by sequencing.<br />

References<br />

1. Geiser, M., et al. (2001) Bio<strong>Techniques</strong> 31:<br />

88-92.<br />

2. Papworth, C., Braman, J., and Wright, D.A.<br />

(1996) Strategies 9: 3-4.<br />

3. Wang, W. and Malcolm, B. (1999) Bio<strong>Techniques</strong><br />

26: 680-682.<br />

4. Burke, T., et al. (1998) Oncogene 16: 1031-1040.<br />

5. Makarova, O., et al. (2000) Bio<strong>Techniques</strong> 29:<br />

970-972.<br />

Legal<br />

QuikChange ® and StrataPrep ® are registered<br />

32<br />

TABLE 1. Reaction Set-up<br />

Amount<br />

Component per reaction<br />

Distilled water (dH 2 O) X µl<br />

10x QuikChange reaction buffer 5.0 µl<br />

dNTP mix 1 µl<br />

DNA template (50 ng) X µl<br />

Megaprimer (300 - 500 ng) X µl<br />

High Fidelity DNA polymerase (2.5 U/µl) 1.0 µl (2.5U)<br />

Total reaction volume 50 µl<br />

Acknowledgments<br />

This Technical Note was written based on data provided by Wenge Wang, M.D.,<br />

Ph.D., in Dr. Wafik S. El-Deiry’s laboratory, Division of Hematology and Oncology in<br />

the Department of Medicine at the University of Pennsylvania.<br />

trademarks of Stratagene, an Agilent Technologies<br />

company.<br />

QIAGEN ® is a registered trademark of QIAGEN.<br />

Touchgene ® is the registered trademark of Techne<br />

Incorporated.<br />

Clontech © logo is the trademark of Clontech<br />

Laboratories, Inc.<br />

a. U.S. Patent Nos. 7,176,004; 7,132,265;<br />

7,045,328; 6,734,293; 6,489,150; 6,444,428;<br />

6,391,548; 6,183,997; 5,948,663; 5,932,419;<br />

5,866,395; 5,789,166; 5,545,552, and patents<br />

pending.


��� ������� ����������<br />

With GeneMorph s<br />

®<br />

II mutagenesis kits, a balanced spectrum of mutations is right at your fingertips<br />

Our GeneMorph ®<br />

II Kits include Mutazyme ®<br />

II<br />

Polymerase, which delivers a balanced mutational<br />

spectrum.<br />

Mutazyme ®<br />

II<br />

DNA Polymerase<br />

Mutazyme ®<br />

I<br />

DNA Polymerase<br />

Taq<br />

DNA Polymerase<br />

AT to N GC to N<br />

Value 10% 20% 30% 40% 50% 60% 70% 80%<br />

u.s. and canada 1-800-424-5444 x3<br />

europe 00800-7000-7000<br />

www.stratagene.com<br />

U.S. Patent No. 7,045,328; 6,803,216; 6,734,293; 6,489,150; 6,444,428; 6,183,997; 5,489,523 and patents pending<br />

GeneMorph ®<br />

and Mutazyme ®<br />

are registered trademarks of Stratagene, an Agilent Technologies Company,<br />

in the United States.<br />

��������� �<br />

�� ������ ����������� ����� �������<br />

��� �������� �������� �<br />

�� ��� �����������<br />

����� �������� � �������� ���������� ��������<br />

���� ���� ������ ������ ���� ��� ����������<br />

����� ����������� ��� ����������� ���� ������<br />

��� �� �������� ���� ��� �������� ����������� ���<br />

������� �������� ������ ��� ������ ���� �������<br />

���� ��������� ��� ��������� �� ���� ��������<br />

� ������ �������� �� ������� �������� ���������<br />

� ��������� ����������� ����� �� � �� �� �����<br />

��� ��<br />

� �������� ���� ��� ����� ��� ���������� ����<br />

�� ��� ����������<br />

��������� �<br />

�� ����������� ����<br />

��������� �<br />

�� ������ ����������� ���<br />

�� ���� ������<br />

��������� �<br />

�� ������� ������ ����������� ���<br />

�� ���� ������


���������� ��� �����<br />

2X Faster<br />

���������� �<br />

��������� ����� ��� ����<br />

�� ��<br />

� ��<br />

� ��<br />

���������� �<br />

���������<br />

u.s. and canada 1-800-424-5444 x3<br />

europe 00800-7000-7000<br />

www.stratagene.com������������<br />

���������� �<br />

��<br />

� � � �� �����<br />

���� ���������� �� ���������� �<br />

��������� �� ���������� �<br />

��<br />

QuikChange ®<br />

is a registered trademark of Stratagene, an Agilent Technologies Company, in the United States.<br />

*U.S. Patent Nos 6,734,293; 6,489,150; 6,444,428; 6,391,548; 6,183,997; 5,948,663;<br />

5,932,419; 5,866,395; 5,789,166; 5,545,552; 6,713,285 and patents pending.<br />

��� ��� ���������� �<br />

��������� ������������� �����������<br />

���� ���������� ����� ���������� ����������� �� ���������<br />

������������� ������ ���� ��� �������� ����� ����� ��� ����<br />

�������� ������������� ������ ����������� ��<br />

����������� ��� ��������� ����� ������������ ��� �������<br />

�������� ��� ���� ������������� ������������<br />

� ������� ���� ��� ����������� ����������<br />

� ������������� ������ ������������� ��������� ��������<br />

������<br />

� ������ ��������� ��������� �� ����� � �����<br />

���� ����� ��� ������ �� �� ������������������������������<br />

���������� �<br />

��������� ������������� ����������� ��� �� ��� ������<br />

�� ��� ������

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!