28
GeneWatch THE MAGAZINE OF THE COUNCIL FOR RESPONSIBLE GENETICS | ADVANCING THE PUBLIC INTEREST IN BIOTECHNOLOGY SINCE 1983 VOLUME 25 NUMBER 3 | APRIL-MAY 2012 ISSN 0740-9737 Inside >> Trivializing Extinction by Rob DeSalle The Frozen Zoo by Oliver Ryder Interview: Mark Stoeckle on DNA barcoding

GeneWatch Vol. 25 No. 3

Embed Size (px)

DESCRIPTION

Genetics and Animal Biodiversity

Citation preview

Page 1: GeneWatch Vol. 25 No. 3

GeneWatchTHE MAGAZINE OF THE COUNCIL FOR RESPONSIBLE GENETICS | ADVANCING THE PUBLIC INTEREST IN BIOTECHNOLOGY SINCE 1983

Volume 25 Number 3 | April-mAy 2012

ISSN 0740-9737

Inside >> Trivializing Extinction by Rob DeSalle

The Frozen Zoo by Oliver Ryder

Interview: Mark Stoeckle on DNA barcoding

Page 2: GeneWatch Vol. 25 No. 3

2 GeneWatch april-May 2012

GeneWatch Vol. 25 No. 3

5 Trivializing Extinction Genetic technologies may make it possible for us to resurrect some species—but what happens when we let ourselves stop thinking of extinction as permanent? By Rob DeSalle

8 Smugglers, Poachers, and DNA Barcoding An emerging technology and an expanding reference library make it possible to identify species from samples of sushi, bushmeat, maybe even leather—and it’s accessible enough to turn high school students into bona fide conservation sleuths. Interview with Mark Stoeckle

12 The Genetic Jungle If releasing genetically modified organisms for conservation purposes proves effective, there’s still another question to ask: Isn’t there an easier way? By Guy Reeves

14 Hatch and Release In 2009, genetically modified mosquitoes were released in the Cayman Islands. It was another year before the world knew about it. By Camilo Rodríguez-Beltrán

17 GM Mosquitoes: Flying Through the Regulatory Gaps? As a company sells its genetically modified mosquitoes to developing countries, regulators try to figure out how to handle a novel technology released into the wild. By Lim Li Ching

Page 3: GeneWatch Vol. 25 No. 3

GeneWatch 3VoluMe 25 nuMber 3

19 The Frozen Zoo Wildlife gene banks provide a tool for studying species and monitoring conservation efforts. By Oliver Ryder

21 A Primer on GMOs and International Law Two different international frameworks attempt to govern genetically modified organisms ... and they don’t always agree with one another. By Phil Bereano

***

Topic Updates

25 Forensic DNA: Database Expansion in New York

25 Forensic DNA: Protections for the Innocent in Massachusetts

26 Gene Patents: Supreme Court Orders New Review of Myriad Gene Patents

26 Animal Biodiversity: Rhino DNA Database Leads to Poacher Arrests

Page 4: GeneWatch Vol. 25 No. 3

4 GeneWatch april-May 2012

GeneWatch is published by the Council for Responsible Genetics (CRG), a national, nonprofit, tax-exempt organization. Founded in 1983, CRG’s mission

is to foster public debate on the social, ethical, and environmental implications of new genetic technologies. The views expressed herein do not necessarily represent

the views of the staff or the CRG Board of Directors.

address 5 Upland Road, Suite 3 Cambridge, MA 02140 phone 617.868.0870 Fax 617.491.5344

www.councilforresponsiblegenetics.org

boArd of directors

sheldoN KrimsKy, phd, boArd chAir Tufts University

peter shorett, mpp treAsurer

The Chartis Group

eVAN bAlAbAN, phdMcGill University

pAul billiNgs, md, phdLife Technologies Corporation

sujAthA byrAVAN, phd

Centre for Development Finance, India

robert desAlle, phd

American Museum of Natural History

robert greeN, md, mphHarvard University

jeremy gruber, jdCouncil for Responsible Genetics

rAyNA rApp, phdNew York University

pAtriciA WilliAms, jdColumbia University

stAff

Jeremy Gruber, President and Executive DirectorSheila Sinclair, Manager of Operations

Samuel Anderson, Editor of GeneWatchAndrew Thibedeau, Senior Fellow

Magdalina Gugucheva, Fellow

editoriAl & creAtiVe coNsultANt

Grace Twesigye

GeneWatchapril-May 2012

VoluMe 25 nuMber 3

editor ANd desigNer: Samuel W. AndersoneditoriAl committee: Jeremy Gruber, Sheldon Krimsky,

Ruth Hubbard

Unless otherwise noted, all material in this publication is protected by copyright by the Council for Responsible Genetics. All rights reserved. GeneWatch 25,3

0740-973

This issue of GeneWatch explores technologies varying widely in their method, but connected by their impacts on the biodiversity of our planet’s fauna. Certainly there are technologies which have the very real potential to harm biodiversity—or, in the case of genetically modified Atlantic salmon, possibly entire ecosystems. But some DNA technologies can be helpful conservation tools, particularly when they help us better understand and monitor the species we are trying to preserve. Wildlife gene banks provide researchers with samples to conduct genetic studies on endangered and even extinct species, and genome sequencing provides new insights into what makes a species tick. Building reference libraries of DNA sequences—and important-ly, making them publicly available—allow scientists and conserva-tionists to identify a species based on a tiny biological sample. Mark Stoeckle talks in this issue of GeneWatch about DNA barcoding, an approach that simplifies species identification from DNA samples to the point that high school students can uncover mislabeled seafood.

These advances can provide real benefits to global biodiversity, but many of those who have worked to create and improve these tech-nologies will also be the first to tell you that no technology is a silver bullet when it comes to conservation. DNA barcoding can be an es-sential tool in prosecuting wildlife smugglers or monitoring animal populations, but it won’t preserve habitat. Even if mosquitoes can be genetically modified to reduce disease pressure on endangered ani-mals, they won’t stop the advance of invasive species. And even if we can resurrect an extinct species from a cryogenically frozen DNA sample, once that preserved species steps off the Ark, what kind of home will be left for it?

Complex problems caused by human actions—and with current extinction rates estimated at anywhere from 1,000 to over 10,000 times faster than just a few centuries ago, loss of biodiversity is very much a man-made problem—cannot be solved without corrections in the human behaviors that got us there. Wind and solar energy play an important role in reducing our use of fossil fuels, but they are no substitute for changing our behaviors to use less energy. Genetic technologies are increasingly useful for understanding and protect-ing biodiversity, but they are no substitute for protecting habitats. As in so many other cases, it’s not just about whether we have the technologies—it’s about how we use them. nnn

Editor’s NoteSamuel W. anderSon

Write to (or for) GeneWatchGeneWatch welcomes article submissions, comments and letters to the editor. Please email [email protected] if you would like to submit a letter or with any other comments or queries, including proposals for article submissions.

Cover photograph: Barney Mosswww.flickr.com/photos/barneymosswww.wheresbarney.com

Page 5: GeneWatch Vol. 25 No. 3

GeneWatch 5VoluMe 25 nuMber 3

It has been suggested that 99.9% of all the species that have ever exist-ed on this planet have gone extinct. Given that there are about 1.8 mil-lion named species and an estimated order of magnitude more unnamed, this means that there have been 2 bil-lion or so species that have existed on our planet since life began here 3.5 billion years ago. The numbers could be even higher when the full extent of microbial life on our planet is real-ized. Extinction prior to humans has been an ongoing process, considered part of the natural way of existence on this planet. Scientists have been able to characterize past extinctions and have concluded that there have been five major extinction periods

since organisms began to diverge on the planet. Each of these five previ-ous extinctions is presumed to have occurred as a result of natural con-sequences, like extreme geological change such as volcanism or asteroid impact.

In 1993, the Harvard biologist E.O. Wilson estimated that each year about 30,000 species go extinct. If you do the math, Wilson’s estimate in the 1990s meant that three species went extinct every hour. Since this estimate, things have not gotten bet-ter but rather worse, and the impact of extinction on biodiversity on our planet can be described as extremely grim. The current rate of extinction is so high that some biologists call it

the Sixth Extinction. My colleague Niles Eldredge has written exten-sively on the subject and points out that the current massive number of extinctions is different than the pre-vious five. The current Sixth Extinc-tion is different in that the source of the extinctions are almost entirely bi-otic—that is, caused by humans as a result of our changing the landscape, overexploiting wildlife, polluting the environment and challenging pris-tine environments with introduced species.

The problem begs our attention, and over the past several decades the discipline of conservation biology was birthed and matured. The imme-diacy of the problem has prompted

Trivializing ExtinctionGenetic technologies may make it possible for us to resurrect some species—but what happens when we let ourselves stop thinking of extinction as permanent? By roB deSalle

Page 6: GeneWatch Vol. 25 No. 3

The Uses and Abuses of Biology Programme is inviting students and recent graduates aged 30 or younger to enter its 2012 essay competition.

The UAB Programme investigates contemporary non-scientific uses and abuses of biological thought in the domains of philosophy, the social sciences, the media, religion and politics.

1st prize = £1000

2nd prize = £500 3rd prize = £250

AUBAUBAUBAUB

Essay Competition

“Explore the ways in which contemporary genetics both challenges and underpins notions of human freedom, value and

identity”

of BiologyUses and Abuses

[email protected]

Page 7: GeneWatch Vol. 25 No. 3

GeneWatch 7VoluMe 25 nuMber 3

scientists to call conservation biol-ogy a “crisis discipline” like cancer bi-ology or infectious disease research. Crisis disciplines often work under the “desperate times call for desper-ate measures” principle. The problem is that most people don’t understand just how desperate the times have become, nor what a “desperate mea-sure” really is.

Take for example the recent sug-gestion that cloning (writ large) can be used as a conservation tool. This suggestion fits exactly with our cur-rent Western “throwaway” society. The idea garnered a lot of media at-tention after being suggested a few years ago, and it seemed like every month some critically endangered species was being cloned or a proj-ect was being announced that was targeted at an endangered species. Many saw cloning as a drastic mea-sure whose time had come. While I do not want to disparage the intent of scientists who made this suggestion, I think it trivializes what extinction really is and gets us back to the im-portance of understanding just how intense our desperate times really are.

In addition to cloning, a large pro-portion of conservation biologists would rather be called conservation geneticists. This moniker points to the use of modern genetic technol-ogy to assay and screen populations that are threatened and endangered. The use of genetics to understand populations and to characterize vari-ability has also been viewed as a dras-tic measure by many conservation biologists. It has been challenged by some conservation biologists as un-necessary, even to the point where it is called “conversation genetics” by those who are critical of genetics as a tool in conservation. Nevertheless, these genetic approaches are viewed by the public as drastic measures that, I guess, are soothing with respect to

the crisis. It is easy for the public to say, as a result of these technologies being applied to conservation: “Hey, look, we are throwing everything we have at the problem!” But here is where the widespread lack of under-standing of the current mass extinc-tion comes back into the story.

My colleague Mike Novacek has suggested that much of the problem is on the shoulders of educators who have failed to make clear the role of biodiversity in a healthy planet. The general public’s lack of knowledge about what extinction is can be dem-onstrated by an interesting survey that City College of New York re-searcher and educator Yael Wyner has conducted. In the survey she asked over 1,500 New York area undergraduate students if the cur-rent demise of the panda in China is natural selection. The obvious answer to anyone who has stud-ied evolutionary biology and the process of natural selection is a resounding “No stupid, it’s us.” How-ever, nearly 50% of the students have answered that it IS natural selection, and a large proportion of those who answer “no” cannot properly explain why. This is yet another way we are trivializing extinction: through the inadequate education of our students and the public in evolutionary biol-ogy. The fact that a majority of Re-publican candidates for President of the United States this year take on a biblical interpretation of diversity on our planet trivializes extinction even more.

Another colleague of mine, George Amato, has strongly argued that it all ultimately comes down to fund-ing. When it comes to funding, the study of biodiversity is a weak sister discipline to the more reductionist approaches such as molecular biol-ogy and genomics. The piggybacking of biodiversity funding with business (the European Union has initiated a

Business@Biodiversity program) or with basic science such as genomics or infectious disease research are ex-amples of funding trends currently in place by government agencies. Two examples of the latter trend are the piggybacking of biodiversity fund-ing with research on emerging zoo-notic diseases and the so called “one health” initiatives; and The Barcode of Life program, which piggybacks biodiversity studies with genetic technology. By providing only mar-ginal funding for research that could help to slow the rate of extinction, it

is as if governments are trying to put a band-aid on a bullet wound.

What can we do? First, we have to resist thinking that technology will solve the problem of large-scale ex-tinction. Technology will help us un-derstand the problem better, but it is not the silver bullet. Second, we need to educate the general public better so that we all understand the im-mediacy and breadth of extinction. Finally, we need to get governments to focus on the problem. The longer biodiversity is underfunded and pig-gybacked with other disciplines, the more species appear in the rearview mirror. nnn

Rob DeSalle, PhD, is a curator in the American Museum of Natural History’s Division of Invertebrate Zoology and co-director of its molecular laboratories and a member of CRG’s Board of Directors.

The problem is that most people don’t understand

just how desperate the times have become,

nor what a ‘desperate measure’ really is.

Page 8: GeneWatch Vol. 25 No. 3

8 GeneWatch april-May 2012

Mark Stoeckle, MD, is an Adjunct Faculty Member in the Program for the Human Environment at The Rockefeller University. He has been involved in the DNA Barcoding Initiative since its beginnings in 2003. DNA barcoding is a technique for identifying species from DNA samples using a short genetic marker at a standard and agreed-upon position in the genome.

GeneWatch: What do you need in order to scan an organism’s DNA barcode? How easy is it to do?

Mark Stoeckle: DNA barcoding is

just a simple, standardized way of identifying species by DNA. With animals, for instance, you analyze one specific gene region and you try to match that sequence to your ref-erence library. That makes it impor-tant to have a really good reference library—and that’s where most of the effort is, in building that library, so that when you get a DNA sequence from a sample there is something in the library to match it to. It’s like a database of fingerprints: you need fingerprints in the database in order to identify your sample.

The sequencing technology is

pretty simple, and it’s getting simpler. It still requires a laboratory, but I can imagine it will only get easier. That part is a straightforward technol-ogy. You can get a result in about a day, and it could be faster in the fu-ture. Then you match that result to the library just like you would use a Google search to look something up. Since these are public databases, you’re just entering the sequence and seeing what in the database it’s matched to.

With people and labs all over the world doing this, how do you make

Smugglers, Poachers, and DNA BarcodingAn emerging technology and an expanding reference library make it possible to identify species from samples of sushi, bushmeat, maybe even leather—and it’s accessible enough to turn high school students into bona fide conservation sleuths. IntervIeW WIth mark Stoeckle

Page 9: GeneWatch Vol. 25 No. 3

GeneWatch 9VoluMe 25 nuMber 3

sure everyone is looking at the right location on the genome?

It’s a standardized approach, so the idea is that everybody is going to use the same gene region. There’s no rule exactly; it’s kind of a social agree-ment among scientists that it makes sense, if you’re analyzing a new spe-cies or going through your museum specimens, to analyze this specific gene or portion of the gene and to put that into the public databases. There’s a social agreement that for DNA barcoding, we all use the same gene region.

What is the margin of error? Is it very easy to confuse two very simi-lar species, or would an unexpected mutation make it harder to identify species?

For animals, I would say something like 95% of the time it’s very straight-forward and there is no ambiguity about the boundaries between spe-cies. The groups that you would get from just doing the DNA sequence alone fall under sets of similar se-quences, and those turn out to match very closely one-to-one with the same groups that biologists have identified as being the species. So it’s amazingly close.

There are maybe 5% of cases where two species are genetically very similar, and there it might be harder but not impossible to tell them apart; and there are some cases of organisms that biologists call dif-ferent species, but in this particular gene region they are identical. Biolo-gists call them two different species, maybe they don’t interbreed, but by analyzing this gene region there’s just not enough information to tell them apart. You know it’s one of the two species, but you don’t know which one it is.

Mutations haven’t turned out to be

much of an issue. For instance, if you were to analyze this same region of the genome among people, any two individuals might differ in one or two positions in the barcode; but people differ from chimpanzees by about 50 positions. So you wouldn’t be confus-ing one with the other.

So, in practical terms, it’s very good. The limitation right now is the library. There are lots of species—there are two million named species of plants and animals, and the most recent estimate is there are another eight million that we haven’t named yet.

So DNA barcoding can be used to identify species, but it doesn’t go beyond that.

Right, it’s not good for identifying in-dividuals. Obviously crime labs use DNA to identify individuals, but you have to analyze more gene regions in order to do that. The goal here is to make it as small and simple as possi-ble, accepting the fact that in the few percent of cases you’re not going to be able to distinguish closely related species from this gene region alone.

DNA barcoding has been used to identify bushmeat species and sea-food and even tea. How far can it be taken? How about, say, alligator boots?

You know, DNA is an amazingly har-dy molecule. Scientists have recov-ered DNA from very ancient speci-mens that are tens of thousands of years old. More recently, people have tried things that are very processed—leather is certainly one of those. I think we’re just beginning to look at that. I know that George Amato’s group at the American Museum of Natural History has retrieved DNA, and specifically DNA barcodes, out of leather products—information

that the U.S. Fish and Wildlife Ser-vice has then used in prosecution of people importing products made from endangered species.

So is DNA barcoding generally be-ing accepted as evidence in court?

The FDA, as recently as last fall, pub-lished DNA barcoding as their of-ficial method for seafood identifica-tion, and the FDA does investigate seafood fraud. That’s the first govern-ment agency that I know of that has said, “This is our legal standard,” but I think that’s going to increase. FDA is a model for agencies in other coun-

tries, and I know that other govern-ment agencies like USDA are looking at this. INTERPOL is also looking at it for detecting commercial fraud. In individual cases, the U.S. Fish and Wildlife Service has certainly used it in court. So yes, it has been used in court cases, and I think it will get ad-opted by more agencies.

What’s the alternative? What’s it re-placing? What is, say, USDA using instead of DNA barcoding?

They were using something called isoelectric focusing and protein elec-trophoresis for identifying seafood. It’s really not a very robust method. And I think for a lot of seafood, there just isn’t any other method [besides

There’s nothing magical about DNA barcoding. It aims to be a toaster: a

technology that you don’t have to read the instructions to

use.

Page 10: GeneWatch Vol. 25 No. 3

10 GeneWatch april-May 2012

supervising them, they can do it on the spot. It doesn’t require extensive training. Then the samples are sent to a lab that does sequencing, sort of like how we used to send film for processing. So the students send the DNA to lab to do the actual sequenc-ing. It’s really pretty simple. It’s not a toaster yet, but it’s getting there.

So you send out the sample with the DNA isolated and the barcode section amplified, the lab sends you back the sequence, and you go and match it against the reference library.

Right, and that’s done on the Inter-net. They send you sequence, and the mapping is something you do just like you’d do a Google search, us-ing GenBank or the Barcode of Life database.

There’s nothing magical about DNA barcoding. It aims to be a toaster: a technology that you don’t have to read the instructions to use. We’re not quite there yet, but the basic principles of it are there. The work for the scientific community is to build up the reference library, be-cause the method is only as good as the library. High school students like my daughter were able to see that su-shi labeled as white tuna was actually tilapia, and they could only do that because researchers had deposited sequences of tuna and tilapia in the public database.

The education potential is very big. High school students—anybody, but say high school students—can discover things that no one else knows. Most science projects, the teacher knows the answer; but with this technology, students can think of an investigation, collect the samples, and until you do the DNA investiga-tion, you really don’t know what you have. And that’s just such a fun thing … it’s discovery. nnn

certain species in it. That’s hard to do, but it might be easier to test it by just taking a water sample—don’t even look at what’s in there—just sort of spin it down and get some DNA out of it.

Another way that’s just starting to be used with water, along the same lines, is to not try to collect organ-isms, just collect water. For instance, in Europe, the American bullfrog is an invasive species. If you want to know if there are bullfrogs in the pond, you can just collect a water sample and see if there is American bullfrog DNA in the water sample. Again, that’s capitalizing on this ability to use very small amounts of DNA.

You have worked with high school students to use this tool for some conservation sleuthing—first “Su-shigate,” when you helped your daughter and a friend uncover mis-labeled sushi, and now with the Ur-ban Barcode Project. That raised a question for me: How easy is it for someone to do their own DNA de-tective work?

The Urban Barcode Project is a re-ally fun project, run by Cold Spring Harbor Laboratories. It’s a compe-tition among high schools in New York City, mainly public schools, to use DNA barcoding to do an investi-gation that they think is interesting.

It’s pretty simple. In this project, the students are thinking of what they want to know and they’re col-lecting the samples. They then bring them to a laboratory in a classroom that’s set up with the right equip-ment, where they go through the steps to isolate the DNA and am-plify the barcode gene. The equip-ment costs a few thousand dollars—it’s not ten thousand or a hundred thousand—and students can sort of walk in and, if they have someone

DNA barcoding]. Once you cut up a fish, you don’t know what it is. Once you cut a fin off of a shark, no one can identify it. In that area, barcoding is a completely new technology.

How is DNA barcoding prov-ing most useful for conservation purposes?

I think we’re at the beginning of the practical uses of it. Major uses are trade in products of regulated or en-dangered species, such as fish and bushmeat, where you need to be able to identify which species the product comes from. Most of those samples are from a product in the form that people use—they’re hard to identify because they’re cut up into pieces or processed in some way.

I think another area where barcod-ing could be useful in conservation is for conducting biosurveys—trying to figure out what lives in a certain area. Say you get a thousand samples of in-vertebrates and you send the crickets to a cricket specialist and the moths to a moth specialist to try to figure out what they are. Instead, you could run the DNA on all of them and you wouldn’t need an expert. It would potentially be an easier, cheaper and faster way to do biosurveys. Barcod-ing is already being used that way for freshwater quality surveys. The best indicator of the health of a wa-tershed, what’s most sensitive, is the life forms in the pond or the stream. And those are hard even for experts to identify, so that’s where people are using DNA barcoding.

It has also been mentioned as a potential tool for invasive species. How would that work?

One of the ways we get invasive spe-cies is in ballast water in ships. You’re supposed to be checking the ballast water to make sure it doesn’t have

Page 11: GeneWatch Vol. 25 No. 3

CHANGE EVERYTHING YOU KNOW ABOUT AUTISM.

“A new approach to understanding how autism affects many

different body systems . . . Essential reading.”

—Temple Grandin, author of Thinking in Pictures and The Way I See It “Gives parents two life-changing gifts.

First is the gift of realistic hope; the second is the gift of knowledge. . . . Dr. Herbert helps parents understand their

child in a new, holistic way.”—John L. Martin, director, Ohio Department of

Developmental Disabilities

“Dr. Martha Herbert is a brilliant researcher, clinician, and thinker.”

—Robert L. Hendren, Director, Child and Adolescent Psychiatry, University of California, San Francisco

In this paradigm-changing book, prominent Harvard researcher and clinician Dr. Martha Herbert offers a revolutionary and transformative strategy for living with autism.

Autism is not hardwired into a child’s genes and destined to remain fi xed forever, as parents are often told. Instead, Dr. Herbert approaches autism as a collection of problems that can be tackled—with talents that can be developed. Her specifi c recommendations aim to provide optimal nutrition, reduce toxic exposures, shore up the immune system, reduce stress, and open the door to learning and creativity—all by understanding and truly meeting your child’s needs.

www.AutismRevolution.orgT A Ballantine Books Hardcover and eBook | THE RANDOM HOUSE PUBLISHING GROUP

On sale now!

Page 12: GeneWatch Vol. 25 No. 3

12 GeneWatch april-May 2012

people engender a fair degree of dis-quiet, including myself, I will never-theless argue in favor of developing (though not necessarily using) an ap-proach that could reasonably be seen as a first step down this road.

The Hawaiian Islands are the bird extinction capital of the world and approximately 70% of native Hawai-ian bird species are already extinct or endangered. Bird malaria (avian malaria) is a significant factor in the continued loss of bird species. The mosquito that spreads avian malaria was accidently transported to the Hawaiian Islands in the 1820s, with the avian malaria parasite arriving about a hundred years later. Since then, many bird species have been in a race to evolve resistance to the

disease. Some species have already won this evolutionary race with the malaria parasite and continue to thrive, while some have lost and gone extinct, while others teeter on the edge of extinction.

The last known Po’ouli individual died in 2004, its extinction was at-tributed to a combination of malar-ia, non-native predators and habitat loss.

Part of the reason why avian ma-laria in Hawaii cannot be controlled is because both of our most effective tools against insect-spread diseases cannot be used. The application of chemical insecticides in the forest is not only extremely difficult, but any killing of non-target insects has the potential to disrupt the ecosystem.

I have an image of walking through a tropical jungle sometime in the fu-ture. It looks and sounds just like an idealized jungle should: birds singing, luxuriously green, with the percepti-ble sound of insects doing the myriad of things that insects do. Yet despite the idyllic vista, I experience a sense of disquiet knowing how some of the organisms got into the picture. There are butterflies genetically modified to be resistant to a viral disease; an iconic orchid which was genetically modified to protect it from hybrid-izing its way out of existence with an accidentally introduced relative; and there are frogs with a genetically modified bacteria growing on their skins to help protect them from a fun-gal plague that had previously devas-tated their populations. The jungle certainly looks and feels more like a jungle in having its native butterflies, orchids and frogs. However, the pres-ence of these GM organisms can also be seen as evidence of past failures to protect the environment, rather than as technological triumphs.

A case for genetically modified mosquitoes in saving endangered species

With the full knowledge that the above view of the future will in most

The Genetic JungleIf releasing genetically modified organisms for conservation purposes proves effective, there’s still another question to ask: Isn’t there an easier way? By Guy reeveS

Page 13: GeneWatch Vol. 25 No. 3

GeneWatch 13VoluMe 25 nuMber 3

A synthetic gene that is substantially effective in preventing mosquitoes from transmitting avian malaria has already been developed.

While it is difficult to predict if techniques to drive genes into popu-lations of mosquitoes will ultimately prove effective, it is clear that con-siderable effort and progress is being made towards this goal. However, if this approach is ever to be used in conservation, it should at some point be possible to argue to local resi-dents that the introduction of about 5,000 bases of foreign DNA into the 600,000,000 base genome of the Cu-lex quinquefasciatus mosquito (every base of which is foreign to Hawaii) is a safe and attractive alternative to other available solutions.

Are there alternative solutions?

Avian malaria is only one factor in the continued loss of bird species and there are intensive (but rather under resourced) efforts to address the im-pact of habitat loss and non-native predators. However, efforts to devel-op ways to directly reduce the impact of malaria have made little progress. This is despite a range of realistic proposals, which include:

(1) Eradicating the mosquito through the coordinated large-scale release of sterile males, which effectively prevent wild females from having any offspring (males never blood feed). This technique, first devel-oped by American scientists in the 1950s, was used to eradicate screwworms (which were non-na-tive flies) from most of continen-tal USA, in an area more than 100 times the total size of the Hawaiian Islands.

(2) Eliminating feral pigs from some reserve areas may create malaria free refuges for birds because the action of feral pigs feeding creates pools of water where mosquitoes

breed. (3) Providing animals with small

amounts of food that includes a drug that is harmless to birds and mammals but is lethal to mosqui-toes when ingested during blood feeding could result in areas with fewer mosquitoes. A number of drugs routinely used for treating worm infections in people and ani-mals have this property.

Surprisingly, none of these have been vigorously evaluated in the context of Hawaiian conservation, despite the fact that this could be achieved for much less than the cost of building one mile of freeway. In contrast, the development of GM mosquitoes has over the last decade attracted more than $30 million to address pressing human health prob-lems like human malaria and dengue fever. The possibility that the result-ing technological advances could also be harnessed to save species is an appealing one. However, if GM mosquitoes are ever used for con-servation purposes, failure to vigor-ously pursue alternative approaches is likely to prove critical in retrospec-tively resolving the question posed at the start of this article. If simple and available solutions appear to have been ignored in favor of complex ge-netic techniques, the imagined ‘ge-netic jungle’ may in fact represent a realistic outcome.

Guy Reeves, PhD, is a researcher at the Max Planck Institute for Evolutionary Biology in Germany and is currently developing genetic systems that could drive disease resistance genes into in-sect populations in a reversible manner. However, he would be content if the ap-plication of alternative, more attractive approaches meant that these techniques were not used outside of the laboratory. Email: [email protected]

Despite decades of sustained effort to develop a vaccine for human malaria, there has been little success and, as such, there is no reason to expect that a vaccine for bird malaria will become available. Given these reali-ties, is it inevitable that bird extinc-tions will continue to the point that almost an entire level of the Hawai-ian ecosystem is permanently lost?

What are genetically modified mosquitoes and how could they be useful?

A number of diverse genetic ap-proaches to control insect-spread diseases are at various stages of de-velopment. For example, more than 13 million Aedes aegypti mosquitoes, which are genetically modified to be partially sterile, have already been re-leased into the wild in Malaysia, Bra-zil and the Cayman Islands (releases have also recently been proposed in Key West, Florida). Another interest-ing approach is the development of a fungus which infects mosquitoes, that has been genetically modified to block malaria transmission.

However, the use of GM mosqui-toes in Hawaiian bird conservation is based on a different approach, in which synthetic disease-resistant genes are introduced into popula-tions of the mosquito species that spreads avian malaria. These genes would sustainably stay in the chro-mosomes of the wild mosquito pop-ulation and stop them spreading ma-laria. The driving of genes into wild populations is not easy, but already two systems have been developed in the laboratory, and more are in devel-opment. Some systems allow these introduced genes to be completely removed from the wild if desired. This work is largely motivated by the goal of controlling human diseases spread by insects, like human malar-ia and lymphatic filariasis, and not by the conservation role discussed here.

Page 14: GeneWatch Vol. 25 No. 3

14 GeneWatch april-May 2012

British biotech company Oxitec’s venture in developing GE mosqui-toes was known, the astonishment came from the sudden jump to field release. I soon realized that I was not the only one missing a year of surveillance on this exercise: the re-lease remained a de facto confiden-tial test for a year.

It was difficult to understand the

silence, intentional or not, on the experimental release of these mos-quitoes, in particular because there were not hidden military or obscure purposes underlying the technol-ogy. In fact, the intended use was described as a tool to tackle dengue fever, one of the major public health issues in many developing countries. With over 50 million infections every

In 2010 I read for the first time about the initial field experiments of genetically engineered mosquitoes that had taken place a year earlier in the Cayman Islands. This news came as a surprise to me, as I consid-ered myself part of the independent scientific community continuously monitoring modern biotechnology advances and applications. Although

Hatch and ReleaseIn 2009, genetically modified mosquitoes were released in the Cayman Islands. It was another year before the world knew about it. By camIlo rodríGuez-Beltrán

Page 15: GeneWatch Vol. 25 No. 3

GeneWatch 15VoluMe 25 nuMber 3

year, the fight against this disease is one of the most important priorities for societies not only in the develop-ing world but also in some regions of the developed world. Strategies range from vector management to early and accurate diagnosis, and while the research on vaccines and viral drugs is under development, no commercial vaccine is available for the moment.

Aedes aegypti is the principal, but not only, species of mosquito capa-ble of transmitting the virus through bites from the female to humans. For this specific case, the techno-logical strategy revolves around the release of mainly male engineered A. aegypti mosquitoes. This tech-nology is called RIDL—Release with a Dominant Lethal—where the in-sects carry a specific genetic switch that under certain conditions causes death at the larval stage of their offspring. This application aims to reduce the incidence of dengue fe-ver by suppressing the mosquito population.

At the molecular level, these GE mosquitoes have been designed with two transgenes. The first one (DsRed2) produces a red fluores-cency in the organism under white light. This is a useful marker for selection and also for monitoring. But the most interesting, and also complex, piece of the system comes from the second transgene, the RIDL regulation system.

Imagine your office door slightly open on a windy day: little by little the door opens more and more as the wind pushes through. You can stand up to try to close it but the wind is so strong that it will reopen it again, and at one stage the door will be so wide open that the wind will be strong enough to create a chaos

(flying pages, knocking over the cof-fee cup etc.). But suddenly you find the key to that door, and by closing the door you have reduced the flow necessary to create the chaos. Well, that is the RIDL system, a positive genetic feedback loop that produces a protein (tTAV) that is able to guide more production of itself (by acting positively on its own genetic pro-moter). This results in an over ex-pression of tTAV, at a concentration that becomes lethal to mosquitoes’ larvae. However there is one anti-dote, a chemical called tetracycline, which if present will bind the tTAV protein, reducing its presence in a free form to activate its promoter. tTAV will still be produced, but at a lower concentration with no toxic effect for the larvae. Just like the ab-sence of a key allowed wind to knock over the coffee cup inside the office, absence of tetracycline will produce a lethal effect at the larval stage of the mosquitoes.

From a biosafety standpoint, risks related to these organisms follow some general issues:

(1) On modified mosquitoes: What will be the consequences in the ecological network of mosquitoes? What will be the effect on preys and predators? What will be the influence in other species of dis-ease carrying mosquitoes? Could they benefit from a reduction in competition? Can the virus adapt better to other vectors because of this selection pressure?

(2) On GE organisms: What is the effect of the exposure to the DSRed2 and the tTAV proteins? What is the likelihood of instabil-ity of the genetically added trait? Could it evolve resistance to the lethal mechanism?

There are other specific issues re-lated to the ability of flying and the difficulty of monitoring the distribu-tion of the mosquitoes (in particular during transboundary movements), as well as issues related to the as-sociated technology (for example, the need to act under absence of tetracycline).

Some of these uncertainties re-garding the implications on eco-systems and health have apparently been accepted by some risk asses-sors, who have given approvals for the field release of the GM mosqui-toes not only in the Cayman Islands

but also in Malaysia and Brazil, with further approvals pending in the United States. Some have highly criticized the scientific approach used on these regulatory processes, and another article in this issue of GeneWatch addresses the regula-tory gaps in these experiences.

I believe that the issues related to the associated technology are of particular interest. It has been

The fact is that the potential of having survivors is a reality, and

some of these will be females—and

female mosquitoes, genetically

engineered or not, bite humans.

Page 16: GeneWatch Vol. 25 No. 3

16 GeneWatch april-May 2012

acknowledged by Oxitec that in the absence of tetracycline, the sur-vival rate of the GM mosquito lar-vae is about 3% under laboratory conditions (the specific reasons for this percentage of survival are un-known). It is interesting that some of the strongest discussions with the promoter of these technologies are about the numbers of surviving mosquitoes: Does it matter? Is it sig-nificant? Is it negligible? Debates are currently ongoing and will continue, but the fact is that the potential of having survivors is a reality, and some of these will be females—and female mosquitoes, genetically en-gineered or not, bite humans.

Another interesting factor is that the survival rate of GE mosquitoes can be underestimated in real con-ditions—not only because of the possibility of building a genetic re-sistance, but in particular because the antidote, tetracycline, is one of the major antibiotics used both for human health and agricultural practices. The major concentra-tion of tetracycline in urban areas is likely to be in sewage systems, and recent literature has shown that A. aegypti does breed in dirty water; therefore the scenario of breed-ing and development in potentially tetracycline-contaminated aquatic environments, with the risk of sup-pressing the lethal system, should now be considered. One could ar-gue that the concentrations in these environments will not be enough to trigger survival, but in order to know this a meticulous surveillance system of tetracycline concentration over time will be needed in the re-gions intended for release. For the moment I am not aware of any such initiatives, and I believe these will be very expensive and hard to put in

place.Aside from these questions, what

has not been covered is a thorough analysis of the appropriateness of this strategy. It seems that the con-text is not ready for the technology. The RIDL system was not developed to tackle dengue; before mosquitoes, the technology was designed for cotton bollworms, and it seems that other agricultural pests will be tar-geted in the future. In other words, rather than developing a technology for the purpose of reducing the inci-dence of dengue fever, Oxitec devel-oped the technology first and then looked for situations where it could be put into use. In this particular case, the use of tetracycline as an antidote makes things out in the en-vironment a little bit more compli-cated. If the technological solution had started from the real challenge or opportunity then it seems very unlikely that it would rely on an an-tidote that is currently available ex-actly where you don’t want it: in the waters where mosquito larvae grow.

I advocate for challenging solu-tions that rely solely on technology and forget to start from a context-centered approach. I put the weight on the challenge not really to the private companies, but on the gov-ernments and public research initia-tives that should be deciding the best for all. Before asking “Does it work?” we need to ask: “Is it appropriate?”

Camilo Rodríguez-Beltrán, MSc, is co-founder of the Taleo Initiative and was awarded the TEDGlobal2010 Fellowship.

“New techniques and new approach-es can and will tell us an enormous amount about the biological history of our species; but they also teach us that this history was a very complex one that is very inaccurately – indeed, distortingly – summed up by any at-tempt to classify human variety on the basis of discrete races. While we can acknowledge that our ideas of race do in some sense reflect a historical real-ity, and that human variety does indeed have biological underpinnings, it is im-portant to realize that those biologi-cal foundations are both transitory and epiphenomenal. Despite cultural barri-ers that uniquely help slow the process down in our species, the reintegration of Homo sapiens is proceeding apace. And this places the notion of “races” as anything other than sociocultural con-structs ever more at odds with real-ity. Increasingly, it seems, we are simply who we think we are.”

- from Race? Debunking a

Scientific MythBy Ian Tattersall and CRG Board

member Rob DeSalle

Available from Texas A&M University Press. Order by calling 800-826-8911, or visit www.tamupress.com.

Race? Debunking a Scientific Myth

Page 17: GeneWatch Vol. 25 No. 3

GeneWatch 17VoluMe 25 nuMber 3

In December 2010, 6,000 geneti-cally modified mosquitoes were re-leased in my country, Malaysia. This followed releases of large numbers of mosquitoes engineered with the same modification—a dominant le-thal gene—in the Cayman Islands, where over 3.3 million GM mosqui-toes were released in 2009 and 2010. Since February 2011, more than 3 million of these mosquitoes were re-leased in the city of Juaziero in north-eastern Brazil. The release of these same mosquitoes is currently being considered in the Florida Keys in the United States. Many other countries are reportedly evaluating the GM mosquitoes for laboratory research and possible future field releases.

The genetic modification in ques-tion targets Aedes aegypti, commonly known as the yellow fever mosquito, which is a vector of dengue fever and other diseases. The so-called RIDL technology involves a genetic regula-tion that, in the absence of the antibi-otic tetracycline, causes death at the larval stage of the offspring. The re-lease of mainly male GM mosquitoes carrying this lethal gene is intended to result in mosquito population sup-pression, with the consequent aim of reducing the incidence of dengue fever.

The GM mosquitoes were de-veloped and the associated tech-nology patented by the UK-based company Oxitec, which appears to be approaching many countries and offering the mosquitoes as a poten-tial solution to the dengue problem.

Dengue fever is a serious problem in many countries, and authorities are increasingly looking for alternatives, as tools such as pesticides are ren-dered ineffective due to resistance development.

However, the release of these GM mosquitoes into the environment raises many scientific, social, ethical and regulatory concerns. Even while these issues are still being debated, it seems that there is a headlong rush to release the GM mosquitoes.

The situation is compounded by the fact that the international regula-tory and risk assessment frameworks governing GM insects in general, and GM mosquitoes in particular, are still immature. So much so that in the US, discussion is on-going as to which

agency should regulate the proposed release of GM mosquitoes in Florida, since this is a completely new area which the regulatory world is unfa-miliar with.

Moreover, under the Cartagena Protocol on Biosafety—the only in-ternational law dealing exclusively with genetic engineering and geneti-cally modified organisms—a techni-cal expert group revised its guidance last year for GM mosquito risk as-sessment. This guidance, part of a larger package of guidance on risk assessment, will be forwarded to the Parties of the Cartagena Protocol for consideration in October 2012. To my knowledge, a correspond-ing group which convened under the World Health Organization to

GM Mosquitoes: Flying Through the Regulatory Gaps?As a company sells its genetically modified mosquitoes to developing countries, regulators try to figure out how to handle a novel technology released into the wild.By lIm lI chInG

Page 18: GeneWatch Vol. 25 No. 3

18 GeneWatch april-May 2012

not appear to have been done. Public information, consultation

and participation have been also lacking. In the case of the Cayman Islands, while Oxitec and the local Mosquito Research and Control Unit claim that adequate information was provided to the public prior to the re-lease of the GM mosquitoes, the vid-eo information provided by MRCU for outreach does not once mention that the mosquitoes in question are genetically modified. Moreover, given the significance of the first re-lease of GM mosquitoes in the world, it is puzzling as to why Oxitec only announced the fact of the release more than a year after they occurred, catching even scientists in the field of transgenic insects off guard.

It is clear that the regulatory pro-cesses that have governed the release of GM mosquitoes into the envi-ronment so far have been lacking. While international guidance may have recently been completed, the implementation at national level still suffers from a lack of adequate expe-rience in dealing with this novel ap-plication of genetic engineering, a lack of rigorous risk assessment and robust investigation of unanswered questions and a lack of effective and meaningful public consultation and participation. In light of this, the push to release the GM mosquitoes in var-ious countries is grossly premature.

Lim Li Ching, M.Phil., works in the biosafety program at Third World Net-work and is Deputy Editor of Science in Society.

develop guidance principles for GM mosquito evaluation has yet to finish this task.

At the national level, the first release of GM mosquitoes in the world, which occurred in the Cay-man Islands, was conducted in the absence of a biosafety law. While the release was approved by the authori-ties concerned, the Cayman Islands only had a draft biosafety bill at the time. Moreover, the provisions of the Cartagena Protocol did not apply to the Cayman Islands, even though the UK, under which the Caymans are a British Overseas Territory, is a Par-ty to the Protocol. This meant that specific biosafety questions may not have been fully considered nor evalu-ated, because of the absence of a de-tailed and comprehensive biosafety regulatory framework.

Indeed, the risk assessment that was used to support the approval of the releases in the Cayman Islands has been roundly criticized. Scien-tists at the Max Planck Institute for Evolutionary Biology in Germany conducted a thorough examination of the regulatory procedures and documents. They concluded that the risk assessment was incomplete, with no provision of experimental data on the releases; that there was poor referencing (unlikely to meet peer review standards); and worst of all, that there was a marked absence of discussion of the potential health or environmental hazards specific to the GM mosquito in question.

This trend of substandard regu-latory oversight is regrettably not a one-off. The Max Planck scientists assessed the regulatory process in the first three countries (US, Cay-man Islands, Malaysia) permitting releases of GM insects (including GM mosquitoes in the latter two countries) in terms of pre-release transparency and scientific quality, and found the process wanting. They

suggest deficits in the scientific qual-ity of the regulatory documents and a general absence of accurate experi-mental descriptions available to the public prior to the releases.

Worryingly, they judged the world’s first environmental impact statement on GM insects, produced by US authorities in 2008, to be sci-entifically deficient. This assertion is made on the basis that (1) by and large, the consideration of environ-mental risk was too generic to be sci-entifically meaningful; (2) it relied on unpublished data to establish central scientific points; and (3) despite the approximately 170 scientific pub-lications cited, the endorsement of the majority of novel transgenic ap-proaches was based on just two labo-ratory studies of only one of the four species covered by the document. However, the environmental impact statement appears to be used as the basis for regulatory approvals around the world, including that of the GM mosquitoes.

One of the most obvious questions to ask is whether humans can be bit-ten by the GM mosquitoes. In public information available on the Cayman Islands and Malaysian trials, how-ever, this question is either conspicu-ously ignored or it is implied that there is no biting risk, ‘as only male mosquitoes are released and they cannot bite.’

However, as detailed by the Max Planck scientists, it is probable that transgenic daughters of the released males will bite humans. This is be-cause the males are only partially sterile as the technology is not 100 percent effective. Furthermore, if the mosquitoes encounter tetracycline contamination in the wild, the num-bers of survivors could increase. The likely presence of transgenic females in the environment requires the con-sideration of a more complex series of potential hazards, but this does

Page 19: GeneWatch Vol. 25 No. 3

GeneWatch 19VoluMe 25 nuMber 3

In the course of their work, field biologists, veterinarians, and zoo scientists often collect biological specimens in order to assist ongoing studies on the biology and health of species. If, in doing so, they make ad-ditional efforts to bank specimens for future studies, they provide future scientists—who may have access to technologies undreamed of by their forbears—with opportunities to gain insights that may contribute to con-servation efforts for declining spe-cies. With the declines in biological diversity that have been well known for the better part of a century, these biobanking efforts have rather qui-etly been underway for over thirty-five years at the institution where I work. The Frozen Zoo at the San Di-ego Zoo’s Institute for Conservation Research now encompasses gam-etes, embryos and cell cultures from over 9,000 animals, comprising more than 1,000 species.

The frozen cultures of viable cells may be thawed, grown and divided into more cells, which can be frozen again. Although not an infinitely ex-pendable resource, it provides the opportunity to conduct studies now, while still keeping supplies for the future.

Hundreds of scientific studies have used samples from the Frozen Zoo. New species have been identi-fied after their distinctiveness was re-vealed by genetic studies of biobank

Wildlife gene banks provide a tool for studying species and

monitoring conservation efforts.By olIver ryder

The Frozen Zoo

Page 20: GeneWatch Vol. 25 No. 3

20 GeneWatch april-May 2012

samples. Studies of species and indi-vidual identity, for wildlife manage-ment and forensic applications, have been undertaken and have expanded the database of DNA profiles and barcodes. Infertile animals have been identified from genetic testing, par-entage relationships identified and incorporated into species manage-ment programs; and now, whole ge-nome sequencing and studies of the repertoire of expressed genes—the “transcriptome”—are being studied using samples banked over the last four decades. From most of the small population management programs of zoos, questions have arisen that are answerable by genetic testing, if appropriate samples are available. The Frozen Zoo has played a crucial role in all these activities.

Opportunities for the future

Before Dolly, the sheep cloned by Dr. Ian Wilmut’s team in 1996, most scientists—myself included—considered that the differentiated adult cells of the body could not be reprogrammed and proceed again through or guide the mammali-an development. It was a surprise again when Dr. Shinya Yamanaka’s team demonstrated that cultured fibroblast (skin) cells could be re-programmed by transiently activat-ing as few as four genes introduced into these cells. If the techniques for producing induced pluripotent stem cells (iPS, cells capable of becoming any cell type in the body) could be adapted to provide similar results with fibroblasts from many other species, the Frozen Zoo potentially represents the source of the largest and most diverse collection of stem cells anywhere.

After more than a year of dedi-cated work, Dr. Inbar Ben-Nun, in Professor Jeanne Loring’s group at The Scripps Research Institute,

and a team from U.C. San Diego and the San Diego Zoo Institute for Conservation Research announced the production of iPS from two en-dangered species, the drill and the northern white rhinoceros. The drill is a large African monkey with a de-clining population in U.S. zoos and endangered in its habitats in Cam-eroon because of habitat loss and il-legal poaching. The northern white rhinoceros is the most endangered form of rhinoceros. Studies of skull characteristics and genetics analy-ses resulted recently in this African rhinoceros being named a separate species, distinct from its southern relatives.

These studies demonstrate the potential for stem cells to be used in veterinary medicine and for treating illnesses. The question also arises of the potential for applying new ap-proaches in assisted reproductive technologies. These might include producing sperm and eggs in tissue culture flasks, the production of em-bryos, and down the road, northern white rhino babies. But it is a long road to travel. Time is running out for the northern white rhino, and al-though it may be one of the last tools remaining, technology may not be sufficient to prevent its extinction.

Impacts of induced pluripotency

It is altogether reasonable that we pause and consider what we might do, for what reason, and for whose benefit.

The use of DNA banks to produce living animals, restore to life extinct species, and provide novel life forms is entrenched in popular musings. The broadly reported effort to clone a mammoth seems to be known by people of all ages.

Given the limitations of our times and our global society in addressing the declines in biological diversity

and loss of species, should we strive to produce a living animal that went extinct ten thousand years ago? Could it play a role today as it did in its native ecosystem?

If the motivation for developing advances in assisted reproductive technologies were to be based on preventing ongoing losses of biologi-cal diversity and reducing the risk of extinction of species that have undergone dramatic recent declines because of human activities, the in-vestments and benefits would reflect a different set of values.

The way to sustain and conserve species is in natural habitats. How-ever, without invoking additional and alternative strategies, losses of biological diversity will surely be large, for many species continue to disappear from their habitats.

The controversies that arise around the discussions of conserva-tion strategies and methods stand in a different realm from the efforts of museums, zoos, and research insti-tutes to prospectively bank biological samples that can assist conservation assessments, monitoring and man-agement. Efforts devoted to banking cells desperately need to be expand-ed by and for the global community. nnn

Oliver A. Ryder, PhD, is Director of Ge-netics at the San Diego Zoo Institute for Conservation Research and an Adjunct Professor of Biology at The University of California, San Diego.

Page 21: GeneWatch Vol. 25 No. 3

GeneWatch 21VoluMe 25 nuMber 3

Introduction

Two international instruments changed the playing field in the past decade regarding the international regulation of genetically engineered organisms. One is the Cartagena Protocol on Biosafety, which is in-tended to regulate the international transfer of “living modified organ-isms” (LMOs). The second is a set of guidelines, the Risk Analysis Princi-ples for Foods Derived from Biotech-nology, established by a little-known United Nations body called the Co-dex Alimentarius Commission.

These two instruments signal at-tempts by the world community to establish rules governing the pro-duction, trade and use of genetically modified foodstuffs. Both agree-ments emphasize the rights of con-sumers and farmers, and the protec-tion of ecosystems. However, it is still not completely clear how their provisions will work alongside the free-trade rules of the World Trade Organization (WTO).

The Cartagena Protocol: a greener way

By joining the WTO, countries agree to limit their freedom to im-pose restrictions on foreign trade. The Cartagena Protocol, however, stresses that trade considerations need not always be given prece-dence over other national objectives. It recognizes that the need to pro-tect biodiversity, the environment and human health are valid priori-ties in decision-making. As of today, some 163 countries (minus several of the most important agricultural

exporters, including the United States, Canada, Argentina and Aus-tralia) have ratified the Protocol, which came into force on 11 Sep-tember 2003.

The Protocol establishes a pro-cedure called Advanced Informed Agreement. Under an AIA, those planning to export LMOs for intro-duction into the environment must notify the country to which they are being sent. That country is then en-titled to authorize or refuse permis-sion for the shipment, based on a risk assessment. Furthermore, the Proto-col allows the recipient nation to in-voke precautionary regulation if, in its judgment, there is not enough sci-entific information to make a proper assessment:

“Lack of scientific certainty due to insufficient relevant scientific in-formation and knowledge regarding the extent of the potential adverse effects of a living modified organ-ism on the conservation and sus-tainable use of biological diversity in the Party of import, taking also into account risks to human health, shall not prevent that Party from taking a decision, as appropriate, with regard to the import of that living modified organism...”

The Protocol does not specify how to resolve any conflict between its own rules allowing an importing country to control trade in LMOs and that country’s obligations not to impede trade if it is also a member of the WTO.

The state of international law regarding LMOs is intentionally fuzzy in some respects; diplomatic concerns for the WTO resulted in

having a Protocol Preamble contain-ing three intentionally conflicting provisions: that trade and the envi-ronment should be “mutually sup-portive”; that the agreement does not change any Party’s international rights and obligations; and that the Protocol should not be interpreted as being “subordinate” to any other treaty. In particular, the Protocol’s adoption of the precautionary prin-ciple—the idea that an action should not be carried out if the consequenc-es of it are unknown but highly likely to be negative—is claimed by trade interests to run counter to the WTO mandate.

Those involved in drafting the Protocol, along with other observ-ers, also acknowledge that there are a number of outstanding issues re-lating to the oversight of genetic ma-nipulation technologies even after adoption of the Protocol text. These include:

• “Living modified organisms” (LMOs) is a more restricted cat-egory than “genetically modified organisms” (GMOs), since it ex-cludes those no longer alive, and their products.

• “Intentional introduction into the environment” may not address sit-uations where the exporter knows that some shipped modified grain, for instance, will be planted within the importing country, but does not necessarily “intend” this to happen.

• Many important countries are not members of the Protocol, including the largest growers and exporters of LMOs: the United

A Primer on GMOs and International LawTwo different international frameworks attempt to govern genetically modified organisms ... and they don’t always agree with one another. By PhIl Bereano

Page 22: GeneWatch Vol. 25 No. 3

22 GeneWatch april-May 2012

States, Canada, Argentina and Australia.

• The Protocol’s provisions on trade in LMOs between a party and a non-party state does not require that its procedures be followed.

• The Protocol says nothing about any regulatory oversight within a country.

In the fall of 2010, a Supplemental Protocol on issues of liability and re-dress for damages caused by LMOs was adopted after 7 years of intense negotiations, and is in the process of being ratified by the requisite 40 countries.

The Codex Alimentarius: focus on food safety

Two months before the Protocol entered into force, a separate break-through took place. In July 2003, with the backing of all its 168 mem-ber nations, the Codex Alimentarius Commission produced the first set of international guidelines for assessing and managing any health risks posed by GM foods.

A relatively obscure United Na-tions agency, the Commission is charged with the key global task of setting international guidelines for food quality and safety. It was estab-lished in 1963 by the Food and Agri-culture Organization (FAO) and the World Health Organization (WHO), and given the mandate of “protect-ing the health of the consumers and ensuring fair practices in the food trade”. The Commission draws up voluntary international food guide-lines through negotiations in ap-proximately 30 committees and task forces.

The most significant element of the 2003 guidelines is that they call for safety assessments of all GM foods prior to their approval for commercial sale. This has important

implications for WTO members. In 1995, the WTO had agreed that Co-dex norms should be the reference point for evaluating the legitimacy of food regulatory measures that are challenged as restrictions on trade. Thus, although the Codex guidelines are strictly voluntary, they have legal significance for WTO members as a defense to charges of “unfair trade.” Also significant is that all of the ma-jor countries growing GMOs—the US, Canada, Argentina, and Austra-lia—are Codex members and agreed to these risk assessment guidelines.

The Codex risk assessment guide-lines contain much language about the need for a “scientific” evaluation of the actual hazards presented by the new foods. But they also recom-mend that “risk managers should take into account the uncertainties identified in the risk assessment and implement appropriate measures to manage these uncertainties”. This wording appears to acknowledge the validity of a precautionary regulato-ry regime, similar to that allowed for international shipments under the Cartagena Protocol.

The Codex also recognizes that “Other Legitimate Factors”—non-scientific in nature—can form a valid basis for regulations, such as using halal or kosher standards. Other

provisions within the guidelines call for a “transparent” safety assess-ment, that should be communicated to “all interested parties” that have opportunities to participate in “in-teractive” and “responsive consulta-tive processes” where their views are “sought” by the regulators.

These non-scientific aspects are consistent with the second prong of the Codex mandate, namely its role in deterring deceptive practices. Such practices might, for example, include selling or distributing GM foods to consumers without labeling them as such. As a top world food exporter, the United States has vig-orously advocated that only “objec-tive” and “scientific” health claims be used as the basis for regulating GM foods, but consumer groups have vigorously contested this position. In the summer of 2011, after 18 years of struggle, Codex finally adopted a guidance document recognizing that countries can adopt laws and regu-lations covering the labeling of GE foods, including mandatory labeling.

Too rich a mix?

It is not obvious how the Protocol, the Codex guidelines and WTO rules mesh together. Seeking a simple an-swer to this question assumes that the negotiation of these agreements

Page 23: GeneWatch Vol. 25 No. 3

GeneWatch 23VoluMe 25 nuMber 3

however, several nations which are Parties to the Protocol seem to be acting to protect the interests of these exporters.

As a result, the Protocol is likely to lead to rules that focus on protecting biodiversity and health more than any rules devised by the WTO. On that basis, there are grounds for be-lieving that the future will see better environmental and health protection than exists at present.

A different situation, however, is likely to unfold behind the scenes as GM food exporters—particularly the United States—put pressure on countries, one by one, to waive their rights under international law. This already happened before the Proto-col was enacted, where weak nations such as Croatia and Thailand had been subjected to pressure by the United States. And last year, Kenya—under enormous pressures from the US, Monsanto, the Gates Founda-tion and GE interests in South Afri-ca—adopted a very weak “biosafety” law that will likely lead to the large-scale introduction of GE crops being grown in that country. Thus the re-sponses of civil society will be crucial to ensure democratic and transpar-ent oversight of this technology.

Phil Bereano, JD, is Professor of Tech-nology and Public Policy at the Uni-versity of Washington, Seattle. He is on the roster of experts for the Cartagena Protocol, co-founder of the Council for Responsible Genetics, and currently rep-resents the Washington Biotechnology Action Council and the 49th Parallel Biotechnology Consortium at interna-tional meetings.

An earlier version of this article ap-peared in the April 2004 issue of Seed-ling magazine, published by Genetic Re-sources Action International (GRAIN).

because these call for the safety of a GM food to be analyzed before it is produced and sold.

The governments blocking the inclusion of the precautionary prin-ciple into the Codex guidelines have argued that if it were to be applied to regulating GM foods, it could be used to justify regulations intended primarily to protect domestic indus-tries from foreign competitors — in violation of the WTO agreements. Others point out, however, that it is not the purpose of the Codex guide-lines to stimulate trade, but rather, to protect consumers. The WTO is supposed to follow Codex norms, not vice versa.

Whither GMO politics?

The political storm raging round GM foods continues to grow in in-tensity, largely because the economic stakes rise steadily while scientific debate remains unresolved. Given the frameworks described above, what conclusion can one draw about the prospects for adequate regula-tory supervision of the technology, and for proper protection of human health and the environment?

The four countries keen to export GM crops—the United States, Cana-da, Argentina and Australia—are all Codex members, but none of them are a party to the Cartagena Proto-col. Therefore, one could argue that it would be inappropriate for such countries to object about others that choose to use the Codex risk assess-ments, since they all voted in Codex to adopt them.

On the other hand, as the coun-tries that signed the Protocol meet to work out the details for carrying out risk assessments under its aegis, and to set rules on traceability and liability, none of these four nations will be legally able to block action taken under the Protocol. In reality,

was guided by a logical process. In fact, they were produced at different times, by delegations from differ-ent national ministries with various missions (trade, environment, food, agriculture, health, etc), and without any reference to the bigger picture. These agreements also reflect the different configurations of indus-try and public interest groups that helped shape them.

Environmentalists argue that the new Codex guidelines on GM foods simply underscore how easy it has been for industry to bring GM foods to market without regulatory super-vision, for example in the US. This practice has been criticized by many activist organizations and a growing number of scientists, as well as sev-eral international authorities on food safety matters.

Many of these critics point out that there is virtually no peer-re-viewed, published scientific research on the risks or benefits of GM food that would allow for safety claims to be tested. They argue that the lack of evidence of risk is not the same as ev-idence of no risk. Many civil society organizations have insisted that pre-cautionary steps should be taken to avert potential risks. Even the WTO Appellate Body, which settles its dis-putes, has recognized that divergent scientific views may be considered in making assessments, such as those evaluating food risks.

Using the precautionary prin-ciple to manage risks also puts the burden of proof on those seeking to introduce the new technology. The United States and other exporters of GM foods have blocked efforts to incorporate the principle explicitly into the Codex guidelines. But some commentators and activists believe that, despite no actual mention of it in those guidelines, the precaution-ary principle is implicit in the docu-ment’s suggestions for risk analysis

Page 24: GeneWatch Vol. 25 No. 3

FGPI is a collaboration of the following organizations:

Page 25: GeneWatch Vol. 25 No. 3

GeneWatch 25VoluMe 25 nuMber 3

Database Expansion in New York

This past March, New York State enacted S. 6733, “DNA testing of certain offenders convicted of a crime,” becoming the first state in the country to adopt an “all crimes” forensic DNA database. Expected to double the size of the current data-base, this dramatic expansion now allows police in the state of New York to collect DNA samples from individuals convicted of even petty crimes, including loitering, reckless speeding or writing a bad check. The state already collects DNA samples from people convicted of felonies and class A misdemeanors, but Governor Andrew Cuomo success-fully claimed that including people convicted of such minor crimes—for which DNA evidence isn’t even rele-vant—is vital because the bigger the database the better chance of catch-ing a criminal in the future. Despite offering only a handful of anecdotal examples, the Governor received the support of all 62 district attorneys in the state as well as a litany of law en-forcement notables and legislative leaders.

Despite active and vocal attempts by several civil society organizations (including the Council for Respon-sible Genetics) raising the obvi-ous civil liberties issues as well as concerns about an overburdened system and risk of error and fraud, public discussion and criticism were largely ignored. Seemingly embold-ened, many other states are moving forward with similar proposals.

Protections for the Innocent in Massachusetts

The first exoneration through DNA testing occurred in 1989. Since then, over 280 people in the United States alone have been ex-onerated by DNA testing after they were convicted of a crime, including a number of individuals who origi-nally pled guilty and almost twenty people on death row. The innocent individuals had served an average of 13 ½ years in prison before exonera-tion and release.

This past February, Governor Pat-rick of Massachusetts signed into law S.1987, “an Act providing ac-cess to forensic and scientific analy-sis.” Championed by Representative John Fernandes and Senator Cynthia Creem, this vital legislation allows Massachusetts to join the 48 other states that grant their citizens the statutory right to their own DNA to prove their innocence after they have been convicted of a crime, with Oklahoma the lone state remaining.

This new law was the result of dil-igent efforts over three years by the Massachusetts ACLU, the New Eng-land Innocence Project, the Mas-sachusetts Bar Association and the Council for Responsible Genetics, which twice testified before the Mas-sachusetts legislature in support.

Topic update: Forensic DNA

Genetic Justice:DNA Data Banks, Criminal

Investigations, and Civil Liberties

National DNA databanks were initially established to catalogue the identities of violent criminals and sex offenders. However, since the mid-1990s, forensic DNA databanks have in some cases expanded to include people merely arrested, regardless of whether they’ve been charged or convicted of a crime. The public is largely unaware of these changes and the advances that biotechnology and forensic DNA science have made possible. Yet many citizens are beginning to realize that the unfettered collection of DNA profiles might compromise our basic freedoms and rights.

Two leading authors on medical ethics, science policy, and civil liberties take a hard look at how the United States has balanced the use of DNA technology, particularly the use of DNA databanks in criminal justice, with the privacy rights of its citizenry.

Sheldon Krimsky is a founding member of the CRG Board of Directors, Professor of urban and environmental policy and planning at Tufts University, and author of eight books and over 175 published essays and reviews.

Tania Simoncelli is a former member of the CRG Board of Directors and Science Advisor at the American Civil Liberties Union. She currently works for the U.S. Food and Drug Administration.

Page 26: GeneWatch Vol. 25 No. 3

26 GeneWatch april-May 2012

The legal case against patents on hu-man genes was given new life this spring after the Supreme Court va-cated a federal appeals court’s rul-ing that upheld Myriad Genetics’ patents on human genes linked to inherited forms of breast and ovar-ian cancer.

With the backing of medical, re-search and patient advocacy groups, the American Civil Liberties Union and the Public Patent Foundation originally filed suit against Myriad’s BRCA1 and BRCA2 gene patents in 2009. A US District Court invalidat-ed the patents in 2010, but last fall the US Court of Appeals for the Fed-eral Circuit upheld Myriad’s patents in a 2-1 decision. Now the Supreme Court has ordered those same three judges to reconsider their ruling.

It was significant that the Su-preme Court even agreed to take up the matter, says ACLU attorney San-dra Park.

“Most petitions that are filed at the court are denied—the vast, vast

majority of them,” Park says. “It was for us a very good development that the court granted the petition and vacated the lower court’s decision.”

Gene patent opponents have an-other case to thank for the devel-opment. Shortly before vacating the Myriad decision, the Supreme Court ruled on Mayo v. Prometheus, unanimously finding invalid Pro-metheus Labs’ patents on methods of evaluating patients’ drug respons-es. Prometheus, having patented a method of adjusting dosages based on patients’ biological response, originally brought the case against Mayo Collaborative Services, which had developed a similar test. The Supreme Court’s ruling in favor of Mayo hinged on its finding that Pro-metheus’ claimed patent matter was a law of nature—reaffirming, Park points out, its well established prec-edent of finding laws of nature and natural phenomena not patentable.

A key element of the Mayo deci-sion is the Supreme Court’s assertion

Topic update: Gene Patents

Supreme Court Orders New Review of Myriad Gene Patents

A rhinoceros DNA database in South Africa has led to 380 arrests and 25 sentences for poachers. The project, called RhODIS (for “rhino DNA index system”) helps to link samples from recovered horns to samples taken from the carcasses of poached rhinos.

The operation, based on the FBI’s Combined DNA Index System for humans (CODIS), was developed through a collaboration between

Topic update: Animal Biodiversity

Rhino DNA Database Leads to Poacher Arrests

that patents of natural phenomena hamper future innovation. The Myr-iad plaintiffs—who also filed an am-icus brief siding with Mayo—have made the same argument against gene patents, and Park sees it as a good sign that the Supreme Court cited it before vacating the Myriad decision.

“We have argued all along that patents on DNA impede innovation because they prevent others from working or examining the DNA itself.”

However, the future of the case against gene patents is far from certain.

“No court has yet applied Pro-metheus to the issue of DNA pat-ents, so we don’t know yet how that will play out,” Park says. And either way, it’s doubtful the case will stop there: “Regardless of the outcome at the Federal Circuit, it’s likely that whatever side does not win will ap-peal the case further.”

wildlife officers and the South Af-rican Police Service’s forensics lab. Between samples collected from horns, the carcasses left behind by poachers, and rhinos living on game preserves, the database contains DNA profiles for over 4,000 individ-ual rhinoceroses.

The system was recently adopted in Kenya, with other countries such as India and Botswana showing an interest as well.

In South Africa, those working on RhODIS say funding is a primary problem, with expensive laboratory costs and over 17,000 rhinos still to be profiled—and 171 rhinos poached in the first 100 days of this year.

To learn more about WWF’s African Rhino Programme, visit www.wwf.org.za/what_we_do/species/arp/

Page 27: GeneWatch Vol. 25 No. 3

C O L U M B I A U N I V E R S I T YP R E S STel: 800-343-4499 Fax: 800-351-5073 cup.columbia.edu

ORDER ONLINE AND SAVE 30%

“I can hardly wait for this book to begin circulation. It should be read and taught as widely as possible.” —Adolph Reed, Jr., University

of Pennsylvania

Divided into six major categories, the collec-tion begins with the historical origins and current uses of the concept of “race” in sci-ence. It follows with an analysis of the role of race in DNA databanks and its reflection of racial disparities in the criminal justice system. Essays then consider the rise of rec-reational genetics in the form of for-profit testing of genetic ancestry and the intro-duction of racialized medicine, specifically through an FDA-approved heart drug called BiDil, marketed to African American men. Concluding sections discuss the contradic-tions between our scientific and cultural understandings of race and the continu-ing significance of race in educational and criminal justice policy, not to mention the ongoing project of a society that has no use for racial stereotypes.

SHELDON KRIMSKY is professor of urban and environmental policy and planning and adjunct professor of public health and community medi-cine at Tufts University. He is the author of Science in the Private Interest: Has the Lure of Profit Corrupted Biomedical Research?

KATHLEEN SLOAN is a human rights advocate specializing in global feminism. She has run non-profit organizations for more than twenty years and has directed communications and public relations functions for multinational corporations and nonprofits.

Race and the Genetic RevolutionScience, Myth, and Culture

Edited by Sheldon Krimsky and Kathleen Sloan

$35.00 / £24.00 paper 978-0-231-15697-4

$105.00 / £72.50 cloth 978-0-231-15696-7

304 pages, 1 line drawings, 4 tablesA PROJECT OF THE COUNCIL FOR RESPONSIBLE GENETICS

“Novel and forward thinking, this book will be a valu-able addition to a literature that needs to be brought up to speed.” —David Rosner, Columbia University and

Mailman School of Public Health

To order online: www.cup.columbia.eduEnter Code: RACKR for 30% discount

Race and the Genetic Revolution Edited by Krimsky Sloan (304 pages)

paper ISBN 978-0-231-15697-4 regular price $35.00, now $24.50

Regular shipping and handling costs apply.

Page 28: GeneWatch Vol. 25 No. 3

council For responsible Genetics

5 Upland Road, Suite 3Cambridge, MA 02140

GENEWATCH IS FREEON THE WEB - BUT YOU CAN STILL SUBSCRIBE TO RECEIVE PRINT ISSUES.

ANNUAL PRINT SUBSCRIPTION RATES

INDIVIDUAL: $35NON - PROFIT: $50 LIBRARY: $70 CORPORATION: $100

All figures in US dollars. Future issues of GeneWatch will be exclusively electronic and free at www.councilforresponsiblegenetics.org. If you wish to obtain a special print copy of a single issue, please send a request to the address below or call 617.868.0870.

PAYMENT INFORMATION

Please send a subscription request with your full name, address and email address, along with a check or credit card (MasterCard or Visa) information, to:

Council for Responsible Genetics5 Upland Road, Suite 3 Cambridge, MA 02140

Orders are also accepted online(www.councilforresponsiblegenetics.org)and by phone (617.868.0870). Additional donations are welcome. Thank you!

IMPORTANT TAX INFORMATION: THE FAIR MARKET VALUE OF A ONEYEAR SUBSCRIPTION TO GENEWATCH IS $35.00

(PROOF OF NON-PROFIT STATUS MAY BE REQUIRED)

Support from people like you makes CRG’s work possible.

Much of our income comes from individuals. Your support helps keep our programs free of the restrictions that come with funding from pharmaceutical and health care companies or government sources. We are the watchdogs for accurate and unbiased information about biotechnology, even when the truth doesn’t suit current political or commercial agendas. And we depend on you to be able to do what we do.

There are many ways you can help CRG. You can become a donor: an annual gift in quarterly installments of $25, $50 or $100 gives us a wonderful and predictable support with a minimal shock to your budget. You may also be able to designate CRG through your workplace giving program, including the Combined Federal Campaign. Many companies will actually match or even double-match your donation. Check with your employer about its matching gift program. You might also consider making an investment in a future where biotechnology is properly used by remembering CRG in your will with a bequest or charitable trust gift.

To learn more about helping CRG, please call us at 617.868.0870. You may also write the Council for Responsible Genetics at 5 Upland Road, Suite 3, Cambridge MA 02140, or visit www.councilforresponsiblegenetics.org.