Science in Society Review Fall 2022 Edition

Page 7

SCIENCE IN SOCIETY REVIEW

Fall 2022 | The University of Chicago

A Production of The Triple Helix

Conflict
Science
in

Editorial Board

2022-2023

Editors-in-Chief

Nikhil Kumar

Marissa McCollum

Managing Editors

Aman Majumdar

Rachel Quan

Emily Shi

Associate Editors

Jacob Botaish

Maryam Contractor

Robert Gorman

Lynn Li

Eva McCord

Shiloh Miller

Production Director

Alex Yoon

The Triple Helix UChicago

Tanya Cukierman, President

Sophie Hu, Vice President

02

Dear Reader,

Conflict often seems like an inevitable part of everyday life, and as 2022 comes to a close, we are no stranger to this fact. Whether events with global ramifications, such as the US midterm elections and the war in Ukraine or our personal struggles with growing responsibilities and maintaining our health, conflict is everywhere in this day and age.

When it comes to the realm of science, conflict is often what drives us forward. The scientific method relies on proving and disproving clashing theories. Ethical disputes lead us to be more responsible with our innovation. Political pressures can influence the direction of scientific progress and control how it shapes our future. Rather than being an obstacle to effective scientific progress, conflict is a quintessential

aspect of what makes science valuable.

Our authors in this issue demonstrate this idea. Science is not always a smooth process with clear goals—it is often fraught with opposing interests and struggles between inquiry and ethics. From space exploration to climate change, from gene editing to indigenous interests, we present a range of perspectives on various conflicts, and the impact they have on the process of scientific discovery. While conflicts are sure to perpetually rise up, it is our responsibility to constantly illuminate and address them in the pursuit of a better world.

Sincerely,

Letter from the Editors-In-Chief
The Privitization of Space Exploration Shiloh Miller 04 The Overlooked Involvement of Indigenous Communities in Psychedelic Medicine Daniela Santillan-Enriquez 11 Human Gene Editing: When to Say Stop? Lizzie Huffaker 17 The Implications of Distorted Science: Case Study of Climate Change Hazal Kara 07 The Bioethical Considerations of Archaeogenomic Studies Sarah Kim 14 03
Table of Contents

The Privatization of Space Exploration

Shiloh Miller is a first year at the University of Chicago considering majoring in Astrophysics or HIPS. Her academic interests include history, astronomy, and writing. When not working at Hallowed Grounds, you can find her rock climbing, reading Neil Gaiman books, or playing Frisbee on the quad.

In the 1960s, head-to-head in the Space Race and facing the prospect of nuclear war, the United States and the Soviet Union drafted what would eventually become the backbone of international space law. Produced in collaboration with the United Nations, the Outer Space Treaty (OST) declares space “the province of all mankind,” saying that space exploration “shall be carried out for the benefit and in the interests of all countries”.1 The fundamental ideas enshrined in the OST—peace, collaboration, anti-nationalism— were reiterated in four subsequent treaties, collectively becoming the core five pillars of space law.

These core five treaties were all written more than 40 years ago.2 None of their authors foresaw a world in which space exploration was spearheaded by anything other than the government. But over the past fifty years, the competitive and patriotic spirit surrounding space exploration has given way to a more entrepreneurial and commercial one, characterized by the rise of private space companies like Blue Origin and SpaceX. Government actors like NASA have begun sharing space—literally—with private space exploration companies. As space exploration transitions from the realm of the government to that of the private sector, we are entering uncharted territory both physically and legally. Much of the existing legislation surrounding space, including the OST, was written to be applicable to nations and not corporations. Private space exploration has its benefits, but letting it expand without providing a legal framework is not safe or sustainable. The value that private exploration provides—greatly reduced cost for launches, innovation through competition, rapid progress—can best be harnessed when the nuances of legal ownership, property, and liability are clearly established and enforced.

The “core five” treaties, one of which the United States never signed, are treaties, not legislation. They require that countries pass their own laws to enforce them. Additionally, they all operate under the assumption of government-led exploration. Even the treaties that admit the possibility of commercial use of space delay directly addressing it; one agreement says that if lunar resource mining becomes possible, an “international

A SpaceX rocket carrying two NASA astronauts to the International Space Station launches from Florida’s Kennedy Space Center on May 30, 2020, marking the first crewed mission to the ISS developed by a private company. (Image Courtesy of Wikimedia Commons)
04

regime must be established to govern how those resources are obtained and used”. 2 With NASA already contracting four companies to begin lunar mining by 2024, as well as space enthusiasts Jeff Bezos and Elon Musk expressing interest in asteroid mining, this may need to be established sooner rather than later.3 Space tourism, orbital hotels, and private space stations are all additional ideas proposed by companies that would operate within no existing legal framework, and private companies are liable to take advantage of the fact that the field of commercial space law is in its infancy.

Why? Businesses are businesses. Government-run space exploration, funded by taxpayers, often concerns itself with missions that benefit “national security, basic science, and national pride,” i.e. ideas that the citizens of the spacefaring country can get behind.4 Companies, on the other hand, don’t exist to please every citizen: they exist to make a profit. Market pressures, the desire to please shareholders, and the need to make a good return on investment mean that speed, efficiency, and cost naturally become priorities for companies.5 Though these priorities may go hand in hand with risk, having a profit motive is not inherently a bad thing. However, the potential benefits of this system—namely its reward of speed, growth, and innovation—can easily turn into downfalls when dealing with high-risk, low-regulation activities like private citizens’ space travel.

Many commercial-minded space enthusiasts don’t believe that this is an issue. In fact, a Harvard Business Review article published last year suggested deliberately loosening safety regulations in order to spur on the increase of private sector activities in space: “policymakers should consider allowing private space tourists and settlers to voluntarily take on more risk than states would tolerate for government-employed astronauts. In the long run, ensuring high safety levels will be essential to convince larger numbers of people to travel or live in space, but in the early years of exploration, too great an aversion to risk will stop progress before it starts”.4 But what happens when brushing off risks leads to one or many deaths? Destruction of technology and loss of human life should not be taken

as the cost of entry for a spacefaring society. For example, murky liability issues arose when a SpaceX rocket exploded on the ground in 2016.6 Imagine just how more complicated it would have been if it was a crewed mission.

As space tourism leaves the realm of science fiction and becomes our reality, ensuring safety will be paramount. In yet another gap in space law, government astronauts and private space tourists do not fall under the same legal guidelines. Government astronauts are required to go through rigorous physical training, examinations, and regular health checks to ensure their mental and physical well-being is maintained during their stint in space. Private companies sending space tourists, on the other hand, aren’t required to physically screen or prepare their spacefarers.7 Additionally, privately manufactured spacecraft don’t need to be reviewed for safety by a third party. This is due to the private space industry lobbying for a 2004 U.S. Congress moratorium on safety regulations.8

These hosts of legal concerns aren’t just theoretical. There are ways in which private corporations have already exploited the unregulated nature of space travel. One requirement of space law is that private companies apply to the government for licenses to launch and maintain satellites. The satellites are then registered with the government and the United Nations. This practice has been decreasing in recent years, despite corporations continuing to launch satellites at similar rates.9 Even the very definition of where airspace ends and outer space begins is unclear. Business’s rapid try, fail, and try again method may work well when prototyping new technology or developing ideas, but it has much more serious consequences when human life is brought into the mix. A lack of incentive to make their projects safe or sustainable, coupled with existing legislation full of holes and gray areas, makes it clear that legally regulating the private sector is the only way to ensure that their goals are executed safely.

This is an urgent goal. Currently, a completely privately run mission to other planets is still unrealistic without government help. The past few decades are littered with the failed plans of wealthy space

05
Representatives from the United States and Soviet Union sign the Outer Space Treaty, a key moment in the history of space law Photo courtesy of The United Nations

enthusiasts—for example, the Space Island Project’s space station designed to host 20,000 people, entrepreneur Bas Lansdorp’s Martian colonization mission called Mars One, or Golden Spike’s roundtrip lunar expeditions.10 But while companies may be dependent on the government for now, this is not guaranteed to continue. Many companies are using the government as a mere stepping stone, “selling to NASA with the hopes of eventually creating and expanding into a larger private market”.4 To put it plainly, “[p]rivate space companies will likely continue to push forward independently of public demands”.11

There is no doubt that the public and private sectors can mutually benefit from collaboration on space activities. NASA gets to execute missions faster and at a much lower cost, and companies can get the money they need to sustain their business while gaining legitimacy in the public eye by associating with a dependably popular sect of

the government.12 But this era of coexistence is coming to an end as companies begin to extricate themselves from their entanglement with the government. As the commercial space industry expands in size, scope, and influence, it is critical that its actions are heavily regulated through law. Space is where we need our ethics to be at our most rigorous. Whatever standards we set for ourselves now will go on to become precedent. Some have said that the OST’s ideas of space being humankind’s “common heritage” are just “statements of general goals and should be seen as mere moral or philosophical obligations” rather than legal ones.9 But what are laws if not our highest moral and philosophical intentions cast into an enforceable format? This new age of public and private space exploration is an opportunity to bring the very best of our world—the scientific merit of government missions, the innovation of private enterprise, the just and guiding goals of legislation—up to the stars and into the future.

References

1. Outer Space Treaty. United Nations General Assembly, 1967, https://outerspacetreaty.org/.

2. International Space Law. Space Foundation. https://www.spacefoundation.org/space_brief/international-space-law/

3. Gilbert, Alex. “Mining in Space Is Coming.” Milken Institute Review, 26 April 2021, https://www.milkenreview.org/articles/mining-inspace-is-coming.

4. Weinzierl, Matthew, and Mehak Sarang. “The Commercial Space Age Is Here.” Harvard Business Review, 12 Feb. 2021. hbr.org, https:// hbr.org/2021/02/the-commercial-space-age-is-here.

5. Sharma, Maanas. “The Privatized Frontier: The Ethical Implications and Role of Private Companies in Space Exploration.” The Space Reviews, SpaceNews, 7 Sep. 2021, https://www.thespacereview.com/article/4238/1.

6. Pasztor, Andy. “SpaceX Leads Probe Into Falcon 9 Rocket Explosion.” WSJ, The Wall Street Journal, 8 Sept. 2016, http://www.wsj.com/ articles/spacex-leads-probe-into-falcon-9-rocket-explosion-1473376404.

7. Dillon, Brandon. “Profitable Risk: The Dangers of Consumer Spaceflight and Space Tourism.” Viterbi Conversations in Ethics, USC, 12 Dec. 2020, https://vce.usc.edu/volume-4-issue-2/profitable-risk-the-dangers-of-consumer-spaceflight-and-space-tourism/.

8. Pultarova, Tereza. “Independent body proposed to ensure commercial spaceflight safety,” Space News, 26 April 2019, https://spacenews. com/independent-body-proposed-to-ensure-commercial-spaceflight-safety/.

9. Oduntan, Gbenga. “Aspects of the International Legal Regime Concerning Privatization and Commercialization of Space Activities.” Georgetown Journal of International Affairs, vol. 17, no. 1, 2016, pp. 79–90. JSTOR, http://www.jstor.org/stable/26396156.

10. Genta, Giancarlo. “Private Space Exploration: A New Way for Starting a Spacefaring Society?” Acta Astronautica, vol. 104, no. 2, Nov. 2014, pp. 480–86. ScienceDirect, https://doi.org/10.1016/j.actaastro.2014.04.008.

11. Pethokoukis, James, and John Konicki. “Do Americans Care About Space?” The New Atlantis, Spring 2022, https://www.thenewatlantis. com/publications/do-americans-care-about-space.

12. Howell, Elizabeth. “SpaceX: Facts about Elon Musk’s Private Spaceflight Company.” Space.Com, Future US Inc, 27 Apr. 2022, https:// www.space.com/18853-spacex.html.

06

THE IMPLICATIONS OF DISTORTED SCIENCE: CASE STUDY OF CLIMATE CHANGE

Hazal Kara is a first year at the University of Chicago and originally from Istanbul. She is considering double majoring in Physics and Environmental Studies. Her interests include astrophysics, climate science, environmental history, ethics, and honestly any niche topic she happens to come across. In her free time, you can find her writing short fiction, reading, agonizing over a problem set, and looking at pictures of her cats (whom she misses dearly).

Does science lead us towards objective truth? Historian and philosopher of science Michael Ruse does not believe so. He claims that, while many people view science as “just the facts,” it is in fact more than that.1 Explicitly, science offers us descriptions of reality—of what we sense, perceive, and experience. Furthermore, science has the potential to make predictions about the future. For instance, science can help us determine the possibility of rain in the future based on weather patterns. Science is interpretable and highly contingent on the social systems it serves. The science behind climate change carries these attributes as well; climate science has a long history of denialism and distortion, with ramifications that impact climate and environmental policy today.

For more than a century, scientists have known that certain gasses can trap heat in the atmosphere, ultimately altering our climate. Indeed, there are reports of climate change dating back to the mid-20th century. Throughout the 20th century, even the oil industry’s own scientists came to this conclusion.2 Yet why did it take so long to act—and why are current policies still not sufficient? While there are numerous factors influencing the answers to these questions, much of the delay has to do with climate change denial, which has recently taken on a new form: climate misinformation.

First of all, it is crucial to understand the impact of conflicts of interest (particularly financial ones) in scientific research, as well as the power dynamics in scientific research processes. Conflicts of interest can impact a scientist’s research conclusions, from how they phrase their results to the nature of the results themselves. Government institutions and corporations often have substantial authority over research, particularly in the environmental fields.3

One famous case is the work of Dr. Wei-Hock Soon, who was a researcher affiliated with the Harvard-Smithsonian Center for Astrophysics. Following outcry from the scientific community after Dr. Soon released a paper claiming climate change was a result of ‘solar variability’ (changes in the Sun’s activity), the Smithsonian Institute conducted an internal investigation. “As the contracts, proposals, reports, letters, and other documents reveal, Soon relied exclusively on grants from the fossil fuel industry for his entire salary and research budget”.2 This is a clear example of how cash can distort scientific research and upend scientists’ intentions.

In order to evaluate scientific research we must ask ourselves the following questions:

• What is being researched, and what isn’t?

• What information is being included, and what is left out?

• Who is funding the research?

• How is this research being communicated to the public?

An instance of external manipulation in the climate research sphere can be observed in a 2008 study by the Union of Concerned Scientists (UCS) on federal climate scientists. The study found that 73 percent of respondents “perceived” inappropriate interference with climate science research in the five years within the span of the survey. Forty-six percent of the respondents either “perceived” or “personally experienced” pressure to remove terms such as ‘climate change’ and ‘global warming’ from any content pertaining to climate communication. In a more recent UCS survey from 2015, 56 percent of National Oceanic and Atmospheric Administration (NOAA) scientists claimed too much weight was being given to political interests. On a more positive note, the survey shows improvement when it comes to the freedom to express ‘controversial’ scientific opinions in media and journals.4

Does it matter which phrases and words we use to describe the phenomena? A study by Geoffrey Supran and Naomi Oreskes from the Department of the History of Science at Harvard University certainly seems to indicate that our word choice is important. The study found that oil and gas corporation ExxonMobil’s “rhetoric of climate ‘risk’ downplays the reality and seriousness of climate change”.5 Posing climate change as a risk—rather than a reality—undermined the credibility of climate scientists and created uncertainty surrounding the idea of anthropogenic global warming. Exxon has been spreading climate misinformation and denialism since at least the 1970s. Furthermore, Exxon’s emphasis on meeting the demands of customers placed blame onto the general individuals, rather than the companies and countries that are causing the majority of greenhouse gas emissions. Supran and Oreskes also noted that ExxonMobil’s rhetorical choices mirror tactics used by the tobacco industry in response to the scientific consensus regarding the health consequences of tobacco use.

In addition to the linguistic aspects of climate misinformation

08
1958 Article on Climate Change. From Sunday Telegraph. Photo courtesy of The Sunday Telegraph
Science is not esoteric. Rather, it is influenced by a myriad of factors, including financial and political motivations.

First Global Land and Ocean Temperature Graph.

Second Global Land and Ocean Temperature Graph.

Third Global Land and Ocean Temperature Graph.

09
Photo courtesy of NOAA Photo courtesy of NOAA Photo courtesy of NOAA

and misleading narratives, statistics, data visualization, and graphs also play a role in undermining scientific consensus or distorting scientific results. The 1954 book How to Lie With Statistics by Darrell Huff (who, ironically, worked for tobacco lobbyists in the 1950s and 60s) highlights some key ways in which statistics (which are commonly considered to be objective—“the hard cold facts”) can be distorted. He delves into how a sample can be biased, how which “average” being discussed matters, how important it is to look at what figures are and aren’t shown, and more. 6 The first graph to the right shows a simple example of a misleading climate data graph.

This graph was created using data from NOAA, showing the annual global average temperature anomalies (land and ocean combined). A first look at this graph and one would think that rather than global warming, we have global cooling. After all, the graph is showing a downward trend. Upon further inspection, it’s clear that some cherry picking is involved. The graph does reflect actual data, but only data from 2016 to 2021. However, a climate is determined based on long-term weather patterns, on the scale of multiple decades, not merely a few years. This is why scientists study the climate trend over a range of 30 years or more, rather than looking at temperature and climatic changes within the past half a decade.

Throughout this article, we have looked at climate change denialism, the impact of rhetoric on climate change perception, and data manipulation or distortion in communication. Without the relevant background, it can be difficult to differentiate between what is true, and what isn’t. So how can one identify and evaluate misinformation, even without scientific expertise? With the amount of information that exists online, which can oftentimes be overwhelming, making sense of scientific claims

becomes even more difficult. Moreover, many individuals are stuck in digital echo chambers, meaning they only come across content that reinforces their preexisting beliefs. While there isn’t a magical fix to determining whether a piece of information is false or misleading, there are certain steps any individual can take. In the view of psychologists Gale Sinatra and Barbara Hofer, a push for increased digital and media literacy can mitigate confirmation bias and susceptibility to misinformation and other fallacies. Furthermore, individuals can (1) learn to prioritize “scientific consensus and empirical research over anecdotal accounts,” (2) expand their information literacy (e.g. by looking at multiple websites about one topic rather than a single one), (3) investigate the motives of a source, and (4) inform themselves about the role of algorithms in social media feed.7

Science is not esoteric. Rather, it is influenced by a myriad of factors, including financial and political motivations. Such motivations can also impact how science is presented to the public, and subsequently how it is perceived. The public perception of scientific matters has a role in determining policy and societal changes. Climate denialism, for instance, fuels political inaction. It prevents the necessary steps from being taken towards climate change solutions. This is evident in the aftermath of the 27th Conference of the Parties (COP27), the 27th in a series of yearly conferences held by the United Nations, which have been widely regarded as unsuccessful. The same goes for other similar social issues, such as the distrust of vaccines, which has led to many preventable outbreaks and has serious public health repercussions. Thus, scientists, science communicators, educators, and policymakers have a responsibility to present scientific research in an accurate (at least, as accurately as possible) and transparent manner. Without intending to sound too dramatic, perhaps the fate of the planet depends on their ability to do so.

References

1. Boudry, Maarten, and Massimo Pigliucci. Science Unlimited? The Challenges of Scientism. University of Chicago Press, 2018.

2. Mulvey, Kathy, and Seth Shulman. “The Climate Deception Dossiers.” Union of Concerned Scientists, 2015, https://www.ucsusa.org/ resources/climate-deception-dossiers.

3. Morgenthau, Hans. “Modern Science and Political Power .” JSTOR, Columbia Law Review, 1964, https://www.jstor.org/stable/1120764.

4. “Climate at a Glance: Global Time Series.” National Oceanic and Atmospheric Administration, NOAA National Centers for Environmental Information (NCEI), 2022, https://www.ncei.noaa.gov/access/monitoring/climate-at-a-glance/global/time-series.

5. Supran, Geoffrey, and Naomi Oreskes. “Rhetoric and Frame Analysis of ExxonMobil’s Climate Change Communications.” One Earth, Cell Press, 13 May 2021, https://www.sciencedirect.com/science/article/pii/S2590332221002335#.

6. Huff, Darrell, and Irving Geis. How to Lie with Statistics. W.W. Norton & Co., 1954.

7. Sinatra, Gale M., and Barbara K. Hofer. “How Do We Make Sense of Science Claims Online?” Science Denial, 2021, pp. 23–49., https:// doi.org/10.1093/oso/9780190944681.003.0002.

8. “Progress and Problems.” Union of Concerned Scientists, 28 Sept. 2015, https://www.ucsusa.org/resources/progress-and-problems.

9. “Voices of Federal Climate Scientists.” Union of Concerned Scientists, 2008, https://www.ucsusa.org/resources/voices-federal-climatescientists.

10

The Overlooked Involvement of Indigenous Communities in Psychedelic Medicine

Daniela Santillan-Enriquez is a first year majoring in chemistry. In addition to writing for SISR, Daniela is a cohort member of the Computation and Engineering in the Life Sciences (CELLS) program, and the Marketing Chair for the Indigenous Student Association. When not worrying about her chemistry assignments, she loves to bead, code on calculators, and listen to Mexican mom music.

Indigenous peoples have used psychedelic substances in medical and religious contexts long before doctors and scientists in the European-centric world discovered their benefits. Although research on the medical potential of psychedelics stagnated in the United States after they became illegal in 1968, recently, the Psychedelic Renaissance brought a resurgence of interest. This movement has allowed for the slow decriminalization of psychedelic substances; for example: as explained by Kat Eschner in the New York Times, “Oregon is legalizing mushrooms” for therapeutic uses. 1 Yet despite the potential medical benefits of psychedelics, one thing is clear: the white-dominated fields of science and medicine have erased the history that links psychedelics and Indigenous peoples together, whilst silencing their voices on the matter.

Psychedelics were an integral part of many indigenous communities in America long before they gained wider popularity in the 1960s. According to the Drug Policy Alliance, “Psychoactive mushrooms have been used for thousands of years and have a long history of both medicinal and ceremonial uses”. 2 Amongst

11

the many native groups that have historically used psychedelics, the Mazatecs commonly used psilocybin in religious rituals overseen by curanderos, “healers”, to “come closer to God, and experience the Divine”. 3

Another example is the Shipido People, who used Ayahuasca in rituals to “reveal the true patterns of reality”. The psychedelic movement was popularized in the US by Timothy Leary after he visited Zihuatanejo, Mexico Native guides “fed [him] a powerful hallucinogen every 72 hours or so, to expand [his] consciousness and strip [him] of [his] ego ‘game’”. 4 This was an idea that Leary keenly adopted and imported to the United States.

The lack of indigenous voices brings cultural appropriation and double standards.

in 1994 the Congress amended the American Indian Religious Freedom Act of 1978 (AIRFA), allowing for “the use, possession, or transportation of peyote [cacti native to Mexico and the southern U.S.A with hallucinogenic properties] by an Indian for bona fide traditional ceremonial purposes in connection with the practice of a traditional Indian religion is lawful, and shall not be prohibited by the United States”. 5

Politically, indigenous peoples have also been instrumental in ensuring the decriminalization and legalization of psychedelic substances in America. When various psychedelic substances were made illegal in 1968, indigenous communities who used them for religious rituals had two options: abstain or face legal consequences.

Fortunately, thanks to indigenous leaders and activists,

Whereas AIRFA allowed for indigenous peoples to continue their practices freely, the newfound interest in decriminalizing certain psychedelic substances has worried indigenous communities on the potential for non-natives to abuse and diminish these cultural practices. Ever since The Psychedelic Renaissance gained traction in the United States, many indigenous communities have expressed their concerns regarding the erasure of indigenous cultures. As explained by Colleen Roan, a Navajo Nation member, and peyote practitioner, “Indigenous communities have traditional peyote ceremony and ritual passed down through

12
“Hombres Voladores de Papantla,” natives from Zihuatanejo. Photo courtesy of zihuatanejo.net

generations. Our people survived the campaign of genocide, the extinction of distinct groups of people Indigenous to the land, and peyote is embraced as medicine for healing and not to be used recreationally, experimentally, or as an economic resource. Decriminalizing peyote would open the doors for peyote to be subject to recreational, experimental, and economic use”. 6 However, white-dominated scientific, political, and medical communities have continued to neglect indigenous voices in favor of research, a “misrecognition or absence of the voices of indigenous people and people of color which can be a form of imprisonment and oppression”. 7

The lack of indigenous voices brings cultural appropriation and double standards: where indigenous peoples have persistently been disparaged for their drug use, white folks have used psychedelics recreationally. According to Psychedelic Support, “One review of 18 trials found that over 82% of participants were non-hispanic whites.” 8 This is ridiculous. Indigenous peoples are the pioneers of psychedelic science, yet they and not non-natives have suffered oppression and legal consequences for doing the same things non-natives have. And cultural appropriation of sacred practices diminishes the historical and cultural importance that they possess to native folks. It also downplays the matter of resilience and oppression in native communities, with white and non-native individuals engaging in these practices without the same stigmas and legal

consequences that tribes would have faced over 30 years ago.

Decolonizing scientific and political perspectives is essential to respecting indigenous peoples and giving back to the communities that have allowed for the medical field to be where it is today. Indigenous voices in research are necessary to have an objective and respectful scientific environment. The scientific, political and medical community must address the long standing appropriation of practices that have been taken from indigenous peoples, as is also the case for acknowledging the damage that they have done in the past whilst upholding white supremacy in science.

Despite being pioneers of science, indigenous peoples to this day remain one of the most oppressed and silenced groups within scientific and medical communities. If we strive to have healthy and influential processes in science that truly benefit the entirety of the population, then we must give back to those who have been wronged in the process of learning. Whilst the damage has already been done, it is never too late to act ethically––especially for a field that perpetually kindles the idea of benefiting all.

References

1. Eschner, Kat. “The Promises and Perils of Psychedelic Health Care.” The New York Times, The New York Times, 5 Jan. 2022, https://www. nytimes.com/2022/01/05/well/psychedelic-drugs-mental-health-therapy.html.

2. What Is the History of Psychoactive Mushrooms?, Drug Policy Alliance, https://drugpolicy.org/drug-facts/history-psychoactivemushrooms.

3. Lewis-Healey, byEvan. “How 4 Indigenous Cultures Currently Use Psychedelics.” Psychedelic Spotlight, 15 Sep. 2021, https:// psychedelicspotlight.com/how-indigenous-cultures-currently-use-psychedelics/.

4. Burleigh, Nina. “A Return Trip to Timothy Leary’s Psychedelic, Day-Glo Mexico.” The New York Times, The New York Times, 6 May 2022, https://www.nytimes.com/2022/05/06/travel/mexico-timothy-leary-psychedelics.html.

5. The American Indian Religious Freedom Act (1978). National Parks Service, U.S. Department of the Interior, https://mylearning.nps. gov/library-resources/american-indian-religious-freedom-act-1978/.

6. Weiss, Suzannah. “‘a Real Slap in the Face’: The Push to Legalize Psychedelics Has Ignored Indigenous Communities.” Mic, Mic, 1 May 2021, https://www.mic.com/life/the-push-to-legalize-Psychedelics-has-ignored-indigenous-communities-75816090.

7. George, Jamilah R., et al. “The Psychedelic Renaissance and the Limitations of a White-Dominant Medical Framework: A Call for Indigenous and Ethnic Minority Inclusion.” Journal of Psychedelic Studies, vol. 4, no. 1, 2019, pp. 4–15., https://doi.org/10.1556/2054.2019.015.

8. How Can We Pay Indigenous People the Respect They Deserve?, Psychedelic Support, 21 Sept. 2022, https://psychedelic.support/ resources/psychedelic-medicine-how-can-we-pay-indigenous-people-the-respect-they-deserve/.

13

The Bioethical Considerations of Archaeogenomic Studies

Sarah Kim is a fourth-year at the University of Chicago majoring in Biological Sciences. Outside of classes, she works at UChicago Medicine as an Oncology Clinical Trials Assistant and Pharmacokinetics Coordinator. Sarah also plays for UChicago’s womxn’s rugby team, and is an executive board member of Korean Students Organization. She hopes to work as a clinical research coordinator after graduation. In her spare time, she enjoys writing poems, watching movies, and solving chess puzzles.

In February of 2017, a 14-member team of anthropologists, geneticists, and archaeologists from the American Museum of Natural History (AMNH) excavated numerous ancestral remains from a burial site in Chaco Canyon, New Mexico. This excavation was part of a larger endeavor to research the ancient Puebloans, a Native American civilization whose cultural influence spread across the Four Corners of the American Southwest during its peak from 850 to 1250 AD.1 The presence of thousands of shell beads and turquoise jewelry within the burial site suggested that these remains belonged to elite members of society.2 The research team successfully extracted and sequenced ancient DNA from each entombed individual, and found that everyone buried at Chaco Canyon shared a common mitochondrial DNA (mtDNA) lineage. Since mtDNA is only inherited maternally, the geneticists confirmed that the elite status of the Puebloans was also matrilineally inherited. Six of the individuals had sufficient DNA preservation to have their genomes mapped, leading to identification of motherdaughter and grandmother-grandson relationships amongst them.2

After publishing their work in scientific journals and discussing their research within academic circles, the AMNH team anticipated support from the surrounding Native American communities. Believing their research would be welcomed as evidence for modern Pueblo and Diné (Navajo) claims of descent from the ancient Chacoans, the AMNH team was unprepared for the dismay of the Native communities, who felt that the dignity of their ancestors and the sacredness of Chaco Canyon had been severely violated.

Other established researchers also condemned the project. “In recent decades, archaeologists have been working hard to build trusting, collaborative relationships with our Indigenous colleagues,” stated Ruth Van Dyke, a professor of anthropology at Binghamton University. “Any research that fails to respect Native rights and sensibilities can only undermine this progress”.2

In the United States, the Native American Graves Protection and Repatriation Act (NAGPRA) regulates outside interference with Native American ancestral remains and objects. NAGPRA mandates all federal agencies and museums to engage with “any tribal group that may have cultural affiliations” with any Native American historical artifacts of interest, and to “repatriate ancestral remains and objects” to any Native American or Indigenous community once a direct cultural relationship can be established.3 However, NAGPRA does not explicitly specify the manner in which this engagement or consultation must occur: a simple notification to the Native leaders of any excavation or transfer of their ancestral artifacts may technically fulfill the consultation requirement, even if no mutual agreement or exchange has occurred. Due to the ambiguity of this policy, the researchers escaped legal punishment. Nevertheless, many prominent tribal leaders, as well as other archaeologists and scientists, felt that the concealment of the Chaco Canyon excavation project, and the subjection of these individual remains to highly destructive chemical processes for sequencing, was highly unethical

14
The AMNH team was unprepared for the dismay of the Native communities.

and disrespectful to the Pueblo and Diné tribes near Chaco Canyon.

To extract and analyze DNA from ancient human remains, each sample must undergo highly destructive processes. First, samples of DNA can be taken by sawing cross-sectional pieces out of the skeletal remains sample. Then the process involves crushing the pieces into powder, immersing the powder in a series of organic solvents, and centrifuging the mixtures to isolate the DNA samples.4 Ancient oral microbiomes can also be extracted from ceramic artifacts, or even dental calculus, which is calcified dental plaque and mineral buildup. In fact, bacteria typically found in the upper respiratory tract and the digestive system can be extracted from dental calculus dating back several millennia.5

Initial attempts to sequence ancient genetic material in the 1980s were successful, but highly inefficient. The first prominent method was Sanger sequencing. Sanger sequencing involves attaching free-floating DNA bases to different colored fluorescent tags, then sequencing a sample of DNA by observing the color of light emitted. Before the first complete human genome was compiled and sequenced in 2003, scientists had no reference DNA to compare these ancient samples to.6 This made it exceptionally difficult to detect and separate genuine ancient DNA from modern DNA (possibly from cross-contamination during collection), and only one base pair could be sequenced at a time. However, the emergence of a more advanced sequencing method called nextgeneration sequencing (NGS), as well as the compilation of the world’s first entire human genome, revolutionized the field of genetic sequencing.

NGS can sequence multiple (up to one million) DNA fragments simultaneously per run. Compared to Sanger sequencing, NGS is way more efficient and cost-effective.7 It is capable of whole-genome sequencing as well as targeted sequencing that focuses only on a specific region of the DNA. The two main targeted NGS methods are hybridization capture and polymerase chain-reaction (PCR) amplification.4

In hybridization capture, DNA is heated in solution until it denatures, or, unravels. It is then randomly sheared at various points in the sample. The DNA fragment solution is then washed over a hybridization capture sensor, which has “baits” made of complementary DNA that are specific to the region of interest. The appropriate DNA fragments then bind onto the sensor, and residual DNA is flushed out to unveil any overlapping pieces. Multiple cycles of hybridization are carried out to ensure accuracy and adequate amounts of data.

In PCR amplification, fragments of genomic double-stranded DNA

15
NGS can sequence up to one million DNA fragments simultaneously per run.

are unraveled into single strands and treated with primers, before being bathed in a solution of free-floating DNA segments that are complementary to the region of interest. The free bases then bind onto the single-stranded DNA in their corresponding order, doubling the number of DNA fragments in the solution. The background DNA fragments are flushed out so that only the DNA of interest remains. Multiple cycles of PCR amplification can lead to a pool of available fragments to sequence. Compared to hybridization, although PCR amplification has quicker turnaround time and costs less per run, it is less sensitive and can only accommodate a limited number of targets.4 Depending on the size of the gene panel, different methods are suited for different studies. If an ancient DNA research project involves many different genes that need to be tagged in a sample, hybridization would be preferred. However, if only one gene of interest is involved in a study, then PCR would be much more efficient.

In 2021, a committee of archaeological and bioanthropological scientists from over 31 countries convened to draft a set of universal guidelines that ensured ancient DNA research remains sustainable and ethical.8 These scientists advised that archaeogenomic research teams should follow, but critically examine NAGPRA’s outlined regulations before proceeding with fieldwork or destructive analysis on ancestral remains and other culturally significant artifacts. Research teams must also prepare a detailed plan to be presented to the appropriate governing board of research before proceeding.

Another guideline suggested by this committee was to make all resulting data accessible and publishable. This applies to cases where sequencing data for certain genes or reference genomes from previous biobank studies may already be available, eliminating the need to conduct repeated ancient DNA extraction. There is only a finite number of skeletal samples in the world (the real number of available samples is unknown due to the lack of a formal documentation system). These remains contain a plethora of ancient DNA that we can learn from, but they also hold great historical and cultural significance. Because these remains cannot be

reconstructed after use in genomic studies, it is imperative that researchers minimize destructive harm to samples during analytic research.

Advancements in genetic sequencing technology have enabled anthropologists, archaeologists, and geneticists to unveil groundbreaking findings in the study of ancient cultures and civilizations. However, conducting these projects without regard for upholding ethical standards in research may unintentionally cause more harm to the involved communities rather than allow them to benefit from the research, as seen in the removal of ancestral remains by the AMNH in Chaco Canyon. Although research teams often evade punishment for these ethical violations due to the ambiguity of legal policy in this field, meaningful consultation or discussion of these research studies with tribal leaders and local Native American communities should be encouraged, or perhaps required. In fact, researchers themselves may benefit from learning a given sample’s significance, as they would acquire a greater understanding of the sociocultural context of the samples.

Often, ancient DNA studies reflect a hierarchical dynamic between genetic sequencing and archaeological interpretations, hailing the former as a “true” revelation of history and discrediting the latter as a discipline prone to imprecision and inaccuracy. In truth, both fields are self-encompassing on their own, and simultaneously rely on each other to uncover the past. Another pressing concern among archaeologists is that a genetics-based approach to history may offer “sanitized versions of the past” that perpetuate misleading Eurocentric views and avoid engagement with violent experiences.10 Without placing both fields on equal footing, these studies may unintentionally project a historical narrative ridden with antiquated ideologies.

Archaeogenomics, ideally, should transform into a more conscientious field to Native American communities when conducting fieldwork or sequencing ancient DNA. Research teams can avoid future incidents reminiscent of that of Chaco Canyon, and take both the interests of scientists and of Native American communities into account.

References

1. Cortez, Amanda et al. “An Ethical Crisis in Ancient DNA Research: Insights From the Chaco Canyon Controversy as a Case Study.” Journal of Social Archaeology, vol. 21 (2), 2021, pp. 157-178.

2. Balter, Michael. “Ancient DNA Yields Unprecedented Insights into Mysterious Chaco Civilization.” Scientific American, 2017. https:// www.scientificamerican.com/article/ancient-dna-yields-unprecedented-insights-into-mysterious-chaco-civilization.

3. United States, House of Representatives. U.S. Code 25 Ch. 32, Native American Graves Protection and Repatriation Act.1990. U.S. Government Printing Office. https://uscode.house.gov/view.xhtml?path=/prelim@title25/chapter32&edition=prelim.

4. Samorodnitsky, Eric et al. “Evaluation of Hybridization Capture vs. Amplicon-Based Methods for Whole-Exome Sequencing.” Human Mutation, vol 36 (9), 2015, pp. 903-14.

5. Daley, Jim. “Ancient ‘Chewing Gum’ Reveals a 5,700-Year-Old Microbiome.” Scientific American, 2019. https://www.scientificamerican. com/article/ancient-chewing-gum-reveals-a-5-700-year-old-microbiome1/

6. “Human Genome Project Completion: Timeline.” National Human Genome Research Institute, https://www.genome.gov/humangenome-project/timeline.

7. Knapp, Michael and Michael Hofreiter. “Next Generation Sequencing of Ancient DNA: Requirements, Strategies and Perspectives.” Genes (Basel), vol 1 (2), 2010, pp. 227-43. doi: 10.3390/genes1020227.

8. Alpaslan-Roodenberg, Söngul et al. “Ethics of DNA Research on Human Remains: Five Globally Applicable Guidelines.” Nature, vol. 599 (7883), 2021, pp. 41-46.

9. Arbuckle, Benjamin and Zoe Schwandt. “Ancient Genomes and West Eurasian History.” Science, vol 377 (6609), 2022, pp. 922-923. doi: 10.1126/science.add9059.

10. Furlong, Rebecca. “Waking the Dead: Sequencing Archaic Hominin Genomes.” Nature, 2021, https://www.nature.com/articles/d42859020-00112-6.

16

HUMAN GENE EDITING: WHEN TO SAY STOP?

Lizzie Huffaker is a second-year Philosophy and Astrophysics major at the University of Chicago. She loves investigating the intersection of science and the humanities, both in her studies and outside of school. Outside of SISR, Lizzie is a Co-Music Director of the Ransom Notes, one of UChicago’s competitive a cappella groups, and plays bass in the Symphony Orchestra. She also loves baking, watching shows with friends, and her weekly Trader Joes run.

The field of gene editing is one of great controversy, both promising miracle-esque solutions to human diseases and presenting great ethical dangers. What if scientists could edit human genes to eliminate cancer? To cure blindness? To increase intelligence or athleticism? To build the “perfect human?” The potential is incredible— and incredibly frightening.

Our debate began in the years following the Holocaust, when the field of “bioethics” began to emerge, born from an urgent need to prevent the atrocities, specifically those emerging from genetics being used to support supremacist ideologies in WWII, from happening again. Though the prospect of directly editing a human gene seemed fantastical at the time, a new era of eugenics was on the horizon, and scientists, politicians, ethicists— and even the public— were ready to get involved.

Around the 1950s and 60s, there was a shift in ethics ideology, led by both scientists and politicians who wanted to defend the field of eugenics while still distancing themselves from the German eugenics that led to the Holocaust. A leader in this movement, biologist Robert Shinseimer, argued that desirable human traits depended not on race or ethnicity, but on intelligence, strength, and virtuosity, which incentivized scientists of the age to discover a technique to directly manipulate the genetic composition of humans.1

Although scientists made little progress towards this goal in the late 1900s, the debate continued regarding the moral implications of human gene editing. Around this time, geneticists made an important distinction between somatic and germline editing. Somatic editing, also called “somatic gene therapy,” refers to the manipulation

17

of an individual’s non-reproductive cells during their lifetime in a way that doesn’t affect their reproductive cells. On the other hand, germline editing affects the reproductive cells of a person, with therefore longer-lasting effects, impacting not only the manipulated individual, but also their future children. Generally, the eugenicists of the 20th century were concerned with germline editing as a means to affect the entire human race, aiming to “edit out” mental and physical disabilities, as well as susceptibility to diseases such as HIV.

Opponents of eugenics argued for restricting human gene editing to purely somatic purposes, given the longevity of the unknown consequences in the case of germline editing. They believed that if we allow for germline editing, it becomes harder to set a moral boundary. This belief won out for most of the 20th century and even into the early 2000s, primarily because the public itself was so invested in creating clear restrictions on this kind of research, especially as the prospect of actual genetic manipulation became increasingly realistic. For example, one prominent figure opposing eugenics was the Jewish philosopher Hans Jonas, whose experience escaping Nazi Germany shaped his later bioethical beliefs in regard to eugenics. He had such a big influence on the field of bioethics that he convinced George W. Bush’s Council on Bioethics to retract federal funding from stem cell research in 2001.2

However, most could not deny the long list of potential benefits to such research, and in 2005, zinc finger nucleases

(ZFNs) entered the biotechnology research scene as a new method of selecting and copying strands of DNA to an unprecedented level of accuracy. ZFNs, along with transcription activator-like effector nucleases (TALENs), discovered in 2010, work together to cut two strands of DNA at a precise point, triggering an alarm system which forces the DNA to repair itself. Scientists can then manipulate the DNA to repair itself in a new way, without the target gene. However exciting and effective this discovery was, the process of designing and synthesizing the ZFNs and TALENs proved to be lengthy and expensive, making it an unrealistic option for future use.3

As the field of biotechnology flourished in the early 2010s, the issue of bioethics fell away from the public sphere as other hot-button issues like abortion became more prominent.4 This allowed much of biotechnology research to go uncontested, paving the way for the introduction of CRISPR-Cas9 in 2012.

The story of CRISPR, or “Clustered Regularly Interspaced Short Palindromic Repeats,” actually begins much earlier in 1987, when Yoshizumi Ishino discovered an intriguing phenomenon in his laboratory.5 The Japanese student was sequencing a gene in E. Coli bacteria when he noticed an unusual pattern of 29 repeated base pairs of genes separated by unique DNA. This seemingly small observation proved revolutionary in the world of genetics, as scientists continued to research CRISPR in the late 1900s and early 2000s. This phenomenon works in tandem with Cas9, a dual RNA-guided protein, meaning the Cas9 enzyme is

18

guided to a select strand of viral RNA and bound to that strand using another piece of RNA, resulting in the problematic strand being cut out. Once combined with the Cas9 protein, CRISPR would change the field of genetics forever. The revolutionary part of CRISPR-Cas9 is the way the DNA repairs itself: once the Cas9 has cut out the target strand of DNA, it can receive a “suggested” replacement strand to repair the damages, which can be manipulated by scientists to contain any number of variations on the original strand. This provides an efficient and effective way of editing certain strands of DNA.

Regarding safety, the introduction of CRISPR-Cas9 reduces much of the previously held worries surrounding gene editing and specifically germline editing. While the CRISPR-Cas9 technique does still pose the safety risk of long-term consequences from unintentional gene edits, the level of precision at which it can operate is incredibly promising for the future of gene editing and especially for the increased safety of gene therapy operations.

The development of CRISPR-Cas9 revitalized the otherwise dormant debate on the use of biotechnology for somatic and germline editing practices, this time with a new urgency as many saw the previously dystopian (or perhaps utopian, depending on who was asked) landscape becoming an ever-nearer reality. The primary justifications for setting the boundary at somatic vs. germline editing in the mid-1900s were for both safety and maintaining autonomy. Safety was a concern at the time because the field of human gene editing was still mostly foreign, and many were worried about the unknown long-term consequences of germline editing. Autonomy was relevant because while somatic editing allowed for knowing consent from the altered individual, germline editing would affect those who did not have

the option to choose. However, as technology improved and the socio-political landscape shifted, these justifications became harder to defend as reasons to hold the somatic/germline border.

Though the argument of autonomy was originally claimed by the anti-germline side of the debate, the ruling of Roe v. Wade in 1973 and the growing relevance of the abortion debate changed the moral landscape and forced many in the bioethics community to reconsider how they used autonomy to justify the constraint on human gene editing. While proponents of “consent of the altered,” argue that germline editing is wrong because those potentially affected cannot consent, a new argument arose that a couple should have the autonomy to choose whether their own reproductive cells are edited. As expected, this debate continues to prevail in the bioethics community as the question of abortion does in the socio-political sphere.

CRISPR-Cas9 is one large step in the field of genetics, presenting both inspiring results for curing diseases and frighteningly long-lasting consequences. Now, the task remains to continue refining this research while maintaining an open discussion about the ethics of the practice. This incredible potential to better human medicine cannot be ignored, and it has motivated bioethicists to work on redefining the boundary of what is “ethical” in gene editing.

References

1. Evans, John H. “Setting Ethical Limits on Human Gene Editing after the Fall of ... - PNAS.” Institute for Practical Ethics, 22 Sept. 2020, https://www.pnas.org/doi/10.1073/pnas.2004837117.

2. Bush, George W. “President Discusses Stem Cell Research.” National Archives and Records Administration, National Archives and Records Administration, 9 Aug. 2001, https://georgewbush-whitehouse.archives.gov/news/releases/2001/08/20010809-2.html.

3. Luthra, Ritika, et al. “Applications of CRISPR as a Potential Therapeutic.” Life Sciences, Pergamon, 25 Aug. 2021, https://www. sciencedirect.com/science/article/pii/S002432052100895X?via%3Dihub.

4. Gabel, Isabel, and Jonathan Moreno. “Genome Editing, Ethics, and Politics.” AMA Journal of Ethics, U.S. National Library of Medicine, 1 Dec. 2019, https://www.ncbi.nlm.nih.gov/pmc/articles/PMC7001785/.

5. Ishino, Yoshizumi. History of CRISPR-Cas from Encounter with a Mysterious Repeated Sequence to Genome Editing Technology. Journal of Bacteriology, 1 April 2018, https://journals.asm.org/doi/10.1128/JB.00580-17.

19
The Triple Helix UChicago

Turn static files into dynamic content formats.

Create a flipbook
Issuu converts static files into: digital portfolios, online yearbooks, online catalogs, digital photo albums and more. Sign up and create your flipbook.