Thursday, August 13, 2015

The need for eco-evolutionary biology in adaptive management under climate change

This is the talk that I gave at the 2015 Ecological Society of America Annual meeting in Baltimore, MD. (Will add images/graphs soon.) Thank you to Jesse Lasky for including me in the special session he organized, The Effects of Eco-Evolutionary Feedbacks on Communities, Ecosystems, and Responsee to Environmental Change.

----------

My talk, sitting in the last slot of today's session, draws on several presentations we have heard today. This final talk will put eco-evolutionary research and knowledge in the context of conservation and climate change.

Le me first remind you that climate change is a big deal with a large influence on biological systems. Under a business as usual emission scenario, temperatures on land are predicted to increase 5-6 deg C this century. The last time the world was this much cooler, Death Valley was covered with mesic forest, a habitat that has since been replaced with drought and temperature tolerant plants and animals. The last time the climate was this warm, there was a relative of the alligator living at the pole.

With this backdrop of large potential change, I'm arguing today four four points that form the basis of a commentary that my colleague Mike Pfrender and I wrote for the 25th anniversary issue of Conservation Biology (Hellmann and Pfrender 2011). I will illustrate these points with empirical data, from my lab or my close collaborators. Research that considers evolutionary processes in climate change biology is still rare but increasing everyday. We need to pick up this pace to inform good management and preserve a sizable portion of life as we know it.

In the language of evolutionary biology, climate change--in simplified form--looks something like this: A species "B" is composed of individuals with a particular distribution of phenotypes or traits. "A" depicts a simple fitness landscape and the mean of "B" sits at the optimum of that landscape. Climate change shifts the landscape to, say, "C." Most ecologists' attempts to predict B's response to change considers the distance between the current trait distribution and the new optimum. They also assume that "B" is fixed or that it is composed of a single fitness peak that matches a single fitness landscape per species. As a result, these predictions lack key eco-evolutionary features.



The first claim I will make is that we must consider variation within species in the reaction to climate change, variation due to processes such as drift and local adaptation. Two quick examples:

1. In common garden (translocation) experiments with a butterfly, we showed that populations at the northern edge of a species range react differently to temperature than central populations. When the RNA of experimental animals was hybridized to a microarray, we identified ~300 genes with localized expression in this butterfly species (O'Neil et al. 2014). 55 of those we can guess at their function thanks to annotation in related species. Many of them are oxidative and metabolic genes. Each line in this graph shows the expression level of a gene under central and peripheral temperature conditions; purple are individuals from peripheral populations and yellow are individuals from central populations. Nearly all of the 300 genes show an unbalanced response to temperature in this species: the two source regions/populations respond differently to peripheral, cooler conditions, but they respond similarly to central, warmer conditions. In this case, warming could be homogenizing force; still, localization is a strong signal in these data.



2. In another species of butterfly, we tested if modeling populations separately--as functionally distinct entities--versus together--as one uniform species--in in statistical niche modeling has an effect on predicted occupancy under climate change (Hallfors et al. in revision). In this species, populations in the eastern portion of the range occupy a distinct climate niche relative to western populations, as shown in this PCA of occupancy as a function of 2 principal components involving 19 different climate variables. This next slide shows where MaxEnt, a common statistical niche model, predicts Karner would live 15, 35, and 65 years from now. On the left, I show predictions based on the entire species as a single niche. On the right, I show predictions if we model the two ecotypes separately (black are current points; colors are predicted, future areas). Both the particular predicted occupied area and the total area occupied are different between the left and right panels. Whether we consider climatic differences between populations matters in generating predictions about where they might go in the future, or where we might put them.



My second claim is that we must consideration adaptive capacity when evaluating the effects of climate change. My previous slide involving the Karner blue butterfly assumes that species--or the two forms within the species--is static. The model assesses vulnerability in terms of the amount of change, called exposure, and how much it perturbs the insect, its sensitivity. But a third component of vulnerability, adaptive capacity, is not considered by most ecological assessments, including niche models. Adaptive capacity involves the ability of a species to adjust, physiologically, through dispersal, or through evolutionary change.

In a forthcoming paper, colleagues and I argue that we can think of adaptive capacity like we think of niche theory, that realized adaptive capacity is a constrained subset of fundamental adaptive capacity (Beever et al. in press). In this abstract graph, for example, consider two hypothetical dimensions of adaptive capacity: evolution and dispersal. The fundamental adaptive capacity is in purple, and the realized adaptive capacity is in green. Smart management can reduce constraints on realized adaptive capacity to increase the ability of species to cope with climate change. Here, for example, through habitat connectivity to release dispersal constraints or genetic mixing to release constraints on adaptive evolution.



My third and fourth points consider evolutionary aspects of management itself. If we aspire to create new management techniques or apply them in new ways because of climate change, we will need to consider eco-evolutionary dynamics in crafting these plans. Further, as all interventions have side effects, we must consider not only the ecological consequences of our actions--on population size, abating extinction, etc.--but also non-target *evolutionary* effects.

I'll tackle both of these points with the example of managed relocation. Managed relocation is an intervention technique aimed at moving species from area of historic occupancy to areas of future occupancy (Richardson et al. 2009). You might pursue managed relocation because it could prevent species or population losses, but it could likely also create new pests, put already-endangered populations at greater risk, and/or incur considerable costs.

Some key, unanswered questions for managed relocation include:
i) should introduced populations be sourced from multiple locales?
ii) how well matched is the source material to the current and future environment?
iii) as the environment changes further, how much genetic variation is desirable?
Each of these are eco-evolutionary questions.

Further, we recently asked published scientists their opinions on managed relocation and found considerable context-dependence (Javeline et al. 2015). For example, we asked about three hypothetical cases of managed relocation: one for an endangered butterfly to prevent species extinction, one for forest trees to enable timber production, and one for a symbiont of corals to prevent coral degradation and loss. We found variation among these cases in how necessary scientists felt managed relocation might be and how effective they thought it could be to address the stated goal. Eco-evolutionary dynamics would add another layer of complexity and uncertainty to an issue that experts already tell us will be different in different instances.



I've offered four essential areas where eco-evolutionary dynamics is critical to managing our conservation future under climate change. Fortunately, these issues are also incredibly interesting biologically. For both the future of biodiversity and humanity, and for the opportunity to study the inner-workings of nature, these four topics are worthy of your research attention as evolutionary biologists. I hope that the entirety of today's session inspires new, creative eco-evolutionary work.

References
Beever et al. In press. Improving conservation outcomes with a new paradigm for understanding species’ fundamental and realized adaptive capacity. Conservation Letters.
Hällfors et al. In revision. Addressing potential local adaptation in species distribution models: implications for conservation under climate change. Ecological Applications.
Hellmann and Pfrender. 2011. Future human intervention in ecosystems and the critical role for evolutionary biology. Conservation Biology 25: 1143.
Javeline et al. 2015. Expert opinion on extinction risk and climate change adaptation for biodiversity. Elementa. DOI 10.12952.
O’Neil et al. 2014. Gene expression in closely-related species mirrors local adaptation: consequences for responses to a warmer world. Molecular Ecology 23: 2686.
Richardson et al. 2009. Multidimensional evaluation of managed relocation. Proceedings of the National Academy of Sciences 106: 9721.

Friday, November 22, 2013

Branches of the Same Tree: Toward a Scientific Reflection upon Value

by Jessica J. Hellmann and Daniel John Sportiello

This essay was published in the NDIAS Quarterly, Fall 2013

[Scientific] thinking, born out of engineering and mathematics, implemented in computers, drawn from a mechanistic mind-set and a quest for prediction and control, leads its practitioners, inexorably I believe, to confront the most deeply human mysteries.
— Donella H. Meadows, Thinking in Systems: A Primer

The twenty-first century is likely to be remembered as the century of biology. We are gaining vast biological insights—and vast power over biological systems—because of high-throughput genetic sequencing technology and computer algorithms that handle vast amounts of data. These insights will give us not only the power to treat disease, but also the power to re-engineer the human body and even nature itself.

As science creates new opportunities, however, it also creates new challenges—ones, it seems, that we rarely anticipate. Those who invented rocketry surely never predicted the logic of mutually assured destruction. And those who invented the Internet thought to empower neither criminal syndicates nor child pornographers. This century will see technological and social changes that are equally profound—and we should think more about their consequences. But this can happen only if scientists—the people who study natural phenomena and invent technical solutions to human problems—are willing to confront the role that values play in determining the direction and application of their research.

Many scientists assume that they have nothing to do with questions of value. It is the task of the scientist, they assume, to reveal how things are—regardless of how the world wants them to be. But that misses the point. We live in a time of unprecedented scientific scope and power, and we can no longer pretend that our assumptions about which problems are most worthy of study and which solutions are most worthy of implementation are not rooted in value-laden judgements and decisions. If scientists intend to address the deepest needs of our world, they must play a role in the direction and application of scientific research—and doing this requires a discussion about social and scientific values. Moreover, since bringing values to bear on science is profoundly complex and potentially fraught with misunderstanding, scientists should engage in conversations with colleagues in the humanities, social sciences, and government. If scientists do not contribute to making decisions about values, others will do it for them, and these people may not have the public interest in mind. Also, non-scientists often lack the technical savvy to inform thoughtful dialogue about science and society. Medicine is a standout case, where sticky normative and technical issues demand scientific engagement with questions of value. Thanks to collaboration among geneticists, biochemists, and physicians, for example, recent biological research has yielded extraordinarily powerful and extraordinarily precise treatments for cancer. This work shows us that cancer is influenced by genetics, environmental pollutants, and infectious agents (Morris et al. 1995; Czene et al. 2002; Soto and Sonnenschein 2010), but it cannot tell how much money and effort should be devoted to cancer detection, how much to cancer treatment, and how much to environmental cleanup and pollution reduction. Yet, cancer researchers can contribute to conversations about these issues in important ways and thereby change the trajectory of public health. For example, scientists can gather data that link dollars spent on environmental cleanup to decreased cancer rates (e.g., Morisawa et al. 2007), and they can help see that the results from their research are applied in the broad public interest and not manipulated for the benefit of a few.

Such social engagement can reap benefits for science as well, including the opportunity to work on vexing and socially relevant issues. With relevancy can come greater funding, a larger stage for presenting scientific findings, and opportunities for large-scale collaboration. With greater social interest in research can come greater attention to scientific accuracy and greater attention to replication and validation (e. g., Fang et al. 2012). In addition, creativity and inspiration can stem from interdisciplinary conversations in which one field stimulates thinking in another—sometimes in a way that, thanks to the limitations of disciplinary language and disciplinary paradigms, would not have been possible otherwise. For example, the economist Thomas Malthus’s Essay on the Principle of Population, with its argument that the growth of any population is eventually checked by the scarcity of resources available to sustain it, had a definitive influence on the biologist Charles Darwin when he wrote his Origin of Species (Vorzimmer 1969).

This is not to say that all scientists must be engaged in conversations on the normative dimensions of their work or even have an interdisciplinary perspective. But some scientists—in fact, many—must be willing to confront normative considerations and work across disciplinary boundaries. This openness to questions of value also requires institutions that help scholars grapple with an increasing array of complex dilemmas and perennial questions about the human condition. After all, if the academy cannot help us to understand who we are and what we should do with the opportunities and constraints given to us, then it has little purpose at all.

Collaboration across the disciplines, in ways that bridge the descriptive and the normative, can produce significant results. Consider climate change. The climate is steadily but profoundly shifting due to the human emission of greenhouse gases. These climatic changes affect species and ecosystems worldwide, such that some species will decline and even go extinct. This risk to biodiversity should be an important motivation for reducing greenhouse gas emissions. But there might also be ways that humans, through ecosystem management, can help species and ecosystems deal with the effects of climate change even as concentrations of greenhouse gases steadily increase (Sala et al. 2008). All of these management strategies, however, come with costs, uncertainties, and possible side-effects, raising key questions about whether, when, and how to act. Scientists cannot answer these questions alone: they must participate in conversations about values to help society weigh the pros and cons of different courses of action and identify solutions for the greatest public good.

Scientists have learned a significant amount about species’ responses to climate change and what management strategies might be appropriate. For example, research shows that species shift geographically to account for changing climatic conditions; the historical records suggest that geographic response dominated over evolutionary change as the leading biological response to climate warming (Davis and Shaw 2001). Yet, not all species shift when the climate changes. Less abundant and geographically restricted species probably decline in numbers and go extinct with warming, and at least one case is known from the fossil record (e.g., Jackson and Weng 1999). Given the rapid rate of modern climate change and a landscape dominated by habitat loss and human modification (Haberl et al. 2007), decline and extinction is likely to be more prominent today than it has been in the past (Periera 2010).

Here is where a climate-change biologist first confronts the normative. The science suggests that climate change is likely to have significant consequences for biodiversity. But this sensitivity isn’t neutral—it arises from a change that humans caused, making culpability part of an otherwise scientific issue. Should we stand idly by and let nature (thanks to human-caused climate change) take its course, or should we intervene like doctors to try to achieve a particular outcome (an outcome that society values) and help species in their struggle with climate change?

Part of the answer to the question is technical. For example, we can evaluate the utility of different approaches. What strategies would be most effective, conferring the greatest benefit or incurring the least risk of negative side-effects? In exploring these “how” questions, however, we inevitably flirt with “why” questions, normative questions. For example, who should decide when action—such as moving a species to new areas—is appropriate? How much money should we spend in taking action? Can we defend inaction if critical biodiversity is lost from climate change? On the other hand, can we defend our actions if unanticipated consequences of those actions turn out to be significant?

The answers to these normative questions are unclear and may depend on the location, situation, and stakeholders in question, but they can be informed by scientific insight. Scientists can help decision-makers grapple with the uncertainty of nature, explaining the difference between noise and knowledge. They also can invite conversation about how we find ourselves in the climate-change predicament in the first place and help navigate complex decisions to the betterment of humanity and the environment. Finally, scientists are in a unique position to help society articulate the various steps it could take to protect the things it values (Hellmann et al. 2011).

Sometimes social values conflict, and science can play a role here as well. In the case of biodiversity and climate change, for example, few would argue that biodiversity has no value, but there are probably limits to its value. Should the government pass laws capping carbon emissions and levy taxes for expensive carbon-sequestration projects in order to protect biodiversity? Or should free markets and private property trump conservation objectives in some cases? Scientists can contribute two things to this kind of political debate. First, they can act as “measurers”: they can demonstrate the consequences of species extinction and the costs and benefits of different courses of action (e.g., see Millennium Ecosystem Assessment 2005). Such measurements will likely have some impact upon deciding what to do. Second, scientists can act as citizens of goodwill, ensuring that debates are not hijacked by those with interests other than those of the public at heart—by, for example, corporations intent on gaining profit, no matter the harm they do to the environment.

Abdicating decisions about how scientific theories are applied is itself a decision about values. It implicitly values democratic and free-market processes with little or no participation by scientists as the best way to make decisions, including decisions about how to act in response to scientific conclusions. Yet, scientists are both citizens and frequent recipients of public funding, giving them a duty to participate in these decisions. This duty is magnified by the fact that scientists are often the ones who most fully understand their own research.

While we have argued for scientific engagement in the normative, we recognize that scientists have quite a bit to lose when engaging in discussions about values because they can at times sacrifice data-based objectivity and adherence to the scientific method. We are certainly not arguing that scientists replace their worldview with another, more subjective, perspective. Instead, we claim that engaging with the normative can be a natural and necessary extension of scientific inquiry and that scientists should have a seat at the table, so to speak, when questions of value arise. Because it is critical to delineate the descriptive from the normative in conversations and decisions about values, it is wise, in our opinion, for scientists to approach such conversations in an interdisciplinary way—ideally, in collaboration with humanists. This approach helps scientists avoid the risk of seeming too subjective and gives them a broader perspective. With interdisciplinary collaboration, furthermore, scientists can become more effective at communicating their own work to society, which is critical for making informed decisions in our complex world.

References
Czene, K. C., P. Lichtenstein, and K. Hemminki. 2002. Environmental and heritable causes of cancer among 9.6 million individuals in the Swedish family-cancer database. International Journal of Cancer 99: 260-266.

Davis, M. B., and R. G. Shaw. 2001. Range shifts and adaptive responses to Quaternary climate change. Science 292: 673-679.

Fang, F. C, R. G. Steen, and A. Casadevall. 2012. Misconduct accounts for the majority of retracted scientific publications. Proceedings of the National Academy of Sciences 109: 16751-16752.

Haberl, H., K.-H. Erb, F. Krausmann, V. Gaube, A. Bondeau, C. Plutzar, S. Gingrich, W. Lucht, M. Fischer-Kowalski, 2007. Quantifying and mapping the human appropriation of net primary production in earth’s terrestrial ecosystems. Proceedings of the National Academy of Sciences 104, 12942-12947.

Hellmann, J. J., V. J. Meretsky, and J. S. McLachlan. 2011. Strategies for conserving biodiversity under a changing climate. Pages 363-288 In: Hannah, L., ed. Saving a Million Species: Extinction Risk from Climate Change. Island Press, Washington, DC.

Jackson, S. T., and C. Weng. 1999. Late Quaternary extinction of a tree species in eastern North America. Proceedings of the National Academy of Sciences 96: 13847-13852.

Millennium Ecosystem Assessment. 2005. Ecosystems and Human Well-Being: Synthesis. Island Press, Washington, DC.

Morisawa, S., T. Fukami, M. Yoshidac, M. Yoenda, and A. Nekayama. 2007. Applicability of Mathematical Model for Evaluating Cancer Mortality Risk. Journal of Risk Research 10: 853-869.

Morris, J. D. H., A. L. W. F. Eddelston, and T. Crook. 1995. Viral infection and cancer. The Lancet 346: 754–758.

Periera, H. M., P. W. Leadley, V. Proença, R. Alkemade, J. P. W. Scharlemann, J. F. Fernandez-Manjarrés, M. B. Araújo, P. Balvanera, R. Biggs, W. W. L. Cheung, L. Chini, H. D. Cooper, E. L. Gilman, S. Guénette, G. C. Hurtt, H. P. Huntington, G. M. Mace, T. Oberdorff, C. Revenga, P. Rodrigues, R. J. Scholes, U. R. Sumalia, and M. Walpole. 2010. Scenarios for global biodiversity in the 21st century. Science 330: 1496.

Sala, E. O., F. S. Chapin, J. J. Armesto, E. Berlow, J. Bloomfield, R. Dirzo, E. Huber-Sanwalkd, L. F. Huenneke, R. B. Jackson, A. Kinzig, R. Leemans, D. M. Lodge, H. A. Mooney, M. Oesterheld, N. L. Poff, M. T. Sykes, B. H. Walker, M. Walker, and D. H. Wall. 2000. Global biodiversity scenarios for the year 2100. Science 287: 1770-1774.

Soto, A. M., and C. Sonnenschein. 2010. Environmental causes of cancer: endocrine disruptors as carcinogens. Nature Reviews Endocrinology 6: 363-370.

Vorzimmer, P. 1969. Darwin, Malthus, and the Theory of Natural Selection. Journal of the History of Ideas 30, no. 4: 527–542.

Monday, November 4, 2013

A vision for research: revisited


This blog was originally written for Leopold Leadership 3.0.
A bit more than a year ago, my lab and I spent a day trying to figure out who we were and what we were about. We wanted to express this identity to ourselves—to help keep us on track and to give us purpose—and we wanted to express it to the outside world. I blogged about the process that we used in our self-exploration, and it’s been great to see other labs, like Chris Buddle’s, give it a try and share their wisdom.
From that process—the process of articulating a mission and vision for our research group—my students and I learned a number of things about ourselves. We learned that we all have different research questions (though the PI shares most of them!); we have different research methods and different stages of career. But we share a common objective. We work to see ecology and climate science inform decisions that protect people and nature. We also all strive for excellence in the work we do. Writing this shared vision down helped—a least for a little while—bring the lab together.
A vision statement should be something you want your organization to hope to achieve, something that reflects your goals and ambitions. A good vision statement should be something like Teach for America’s: “One day, all children in this nation will have the opportunity to attain an excellent education.” This vision doesn’t say a thing about the tasks that teachers do.
In making a statement for our lab, we brainstormed about how we want the world to be and how we want it to be changed or improved through our scientific work. In the day-to-day life of science, teaching, and research, we tend to emphasize productivity, mastery, and progress, the number of papers and grants. But a vision is the reason you do all of those things. Vision also is something that grad students, postdocs, undergrads, and even PIs don’t get to talk about and write down everyday.
Today, I find that we don’t reference our vision, or our mission, statement as much as we probably could or should. We mention it from time to time in lab discussion. We introduce it to new members of the lab. But I now think that group visioning should be a repeated exercise. The statement should be re-crafted from time to time. I also think that the activity of making the vision statement may be more important than having the statement itself, at least from the point of the view of group dynamics.
Our current vision statement does help me as a PI, however. As our group grows and the scope of our work steadily expands, there are more and more opportunities, different directions we could head, different projects we could initiate, and different students we could take on. I think frequently about whether a new project or a new collaboration will advance our vision, as much as I think about whether it will lead to good papers or new streams of funding.
So I think that visioning with a research group is a good idea, maybe not just once but periodically. It doesn’t have to be a formal process, and folks like Chris Buddle and Elena Bennett have a number of good ideas to share. Working with one’s research group to craft a collective mission and vision is just another way of stopping and taking stock. Taking stock provides clarity of purpose, and doing it as a group can elevate your collective endeavors to a new level.

Monday, June 17, 2013

Sharing the results of federally-funded research

The National Science Foundation requires that all grant recipients submit annual and final reports. I just submitted a final report for our grant, "Assessing temperature-related changes in introgression of hybridizing species across space and time." In the spirit of promoting the results of federally-funded research, I thought that I'd share my Project Outcomes Report. It is intended for a general audience. Publications on this research are forth-coming!


"Our research aimed to understand how two related organisms that interbreed (hybridize) have responded and will respond to climate change. Many studies have observed the movement of entire species in response to changing climate, but the movement of a entire species is just one possible reaction to climate change. In order to better understand the complexities of how organisms adapt to a changing environment, we must also look within species to see how traits and genes within a species’ range are moving and adjusting to climatic change. These traits and genes determine how an organism will be affected by climate change: where helpful genes are lacking, populations can decline; and where new genes arrive, evolutionary rescue can occur. Regions (or “zones”) of hybridization between related species are ideal places to study the movement of traits within and among species because, theoretically, genes can move across geography independent of the species that contain them. Our NSF-funded research focused on this lesser-studied aspect of climate change science: the movement of genes within and across species under climate change and changes in the flow of genetic information that results.

We examined the following questions: 1) Have traits (genes) moved across geography in two species of butterfly, Papilio glaucus and P. canadensis, in response to recent climatic change? 2) How do the parental forms of the two species and their hybrids perform under simulated climate and climate change? Papilio glaucus and P. canadensis are ideal candidates for answering these questions because they have been studied previously and they hybridize over a wide area in the upper Midwest US.

To answer the first question, we compared specimens of butterflies that were collected in the 1980s and again after 2007 from a wide area in Illinois and Wisconsin that spans the zone of hybridization. We measured these specimens for wing traits and genetic markers that are thought to be related to climate and for traits and markers that are likely not climate-related (control). We found that some traits appear to have moved northward in the last 30 years, while others are not. The ones that have shifted northward appear to be related to temperature and are associated with warming that has occurred most strongly in the southern portion of the hybrid zone. This result suggests that climate change can shape the geographic distribution of traits within species, but not alter others. In other words, climate change is altering the association of traits in these butterflies and the genetic composition of them.

To answer the second question, we performed an experiment with pure forms of each butterfly species and hybrid crosses that we created in the lab. We then exposed experimental animals to a range of climatic conditions that span the hybrid zone. For the northerly species and its hybrids, this experiment simulated a warming event. We found significant differences in the timing of life events in the two species and in the different types of hybrids, and these differences were affected by climate treatment. To make future projections with this information, we built a model that predicts the number of generations that will occur across Wisconsin under alternative climate scenarios. The number of generations per year affects the amount and direction of gene movement between the species and can be used to infer future changes in the northward flow of traits and genes.

Seven undergraduate students participated in this research project, one funded by the NSF on a supplement to this grant and six from other sources. One graduate student is earning a PhD with this project, and three full- or part-time research staff advanced their technical skills and career path with employment on this project. Work on this grant also enabled public presentations and outreach by the PI and the graduate student about climate change and its ecological consequences, and the project compiled a large and highly unique database of thousands of specimens from the hybrid zone. This resource will be made accessible to future researchers. Protocols and equipment from this project also informed research and conservation planning for other Lepidoptera, including one federally-listed endangered species. Most importantly, this work advances our understanding of how climate change can alter living systems, a crucial goal given the amount of warming that is projected to occur in the coming decades and centuries."

Wednesday, May 22, 2013

Guest blogs at Nature with Leopold colleagues

Several Leopold Leadership Fellows and I have a series of guests posts at the Soapboxscience blog at Nature this week. A series of three posts address #reachingoutsci and the opportunities and challenges that confront scientists in making their research understood by and useful to society. The series arises from a session at the 2013 AAAS symposium, The Beauty and Benefits of Escaping the Ivory Tower. (Check out #AAASbeit on Twitter.)
Highlights from the blogs...

Part I:
Bridging the Science-to-Society Gap
"This shift in what society needs—not just science for science’s sake, but to also using science to help recognize and solve societal problems—means that the goals of communicating science have to shift as well.  Society now needs information from scientists not just in the form of interesting facts assembled in hard-to-find places, but especially as recommendations about how to manage the biosphere to maintain what humans depend on for their physical, economic, and emotional well-being.  Scientists, after all, are the people paid to produce and collect the knowledge that is relevant to the world."

Part II:
The Twenty-fifth Hour of the Day: Finding Time for Outreach
"Is your career compromised if you spend time on outreach rather than science, or is engagement all that really counts in a world urgently in need of scientific leadership? Fortunately, new studies suggest that these tasks aren’t necessarily a conflict—those scientists who reach beyond the boundaries of traditional science-doing also appear to be the most productive scientists, probably because they find inspiration, cutting-edge ideas, and novel ways of working while directly engaging with society."

Part III:
Unclogging Institutional Conduits Between Research and Outreach
"Universities aren’t doing nearly enough to help or reward those who want to engage outside academe. While most institutions pay lip service to outreach, salary and promotion are usually determined by first considering “research productivity,” (i.e., numbers of publications and grants), and second by “teaching effectiveness,” (i.e., number of students and course evaluations). Highly focused pre-tenure faculty are particularly spread painfully thin. The connections needed for meaningful dialogue with decision-makers and the public take time to build, especially if you lack experience.  Collectively, we’ve spent hundreds of hours struggling with effects ways to incorporate outreach and engagement in our academic lives.  We believe that practical change must come—at least in part—from academic institutions in order to meaningfully expand the role of science outreach."

Monday, May 6, 2013

5th & 6th grader questions about climate change


The following came up after my presentation, "What is global warming?" to 5th and 6th graders at the Stanley Clark School, South Bend, IN. Thanks to the students for being so attentive and for their great follow-up questions!


1. What state produces the most CO2?
Wyoming releases the most greenhouse gases per person. The next are North Dakota, Alaska, and West Virginia--all are big states for oil or coal production. In total emissions, Texas emits the most, followed by California--these are both big states with quite a lot of people. Indiana is the 5th largest emitter of greenhouse gases in total and 11th based on emissions per person. Indiana does not have a lot of energy efficiency in place and relies heavily on coal to produce electricity. Burning coal releases quite a lot of CO2. You can see all the state rankings for yourself at: http://www.google.com/publicdata/explore?ds=z8cs5f2mcjthet_.

2. Will human civilization still be here in 20-30 years? Will climate change cause the end of the earth? Will the earth be too hot to live on? Will the world end, or will all life on earth die because of global warming?
A bunch of students asked this question, and it's a great one--and scary too. I don't think that global warming will destroy the planet. If you look back 2.5 (or more) million years ago, for example, you can find an atmosphere and a climate that is similar to the one that we creating today. So the planet will go on and some plants and animals that can adjust to the climate change will go on too. But that's not to say that climate change is not a big deal--it really is. We are creating an atmosphere unlike the one that has dominated for 800,000 or more years! And the threat of climate change is not to the planet but to us. It will likely cause many of the plants and animals that we use and enjoy to decline or go extinct (maybe 10-30% of them!). If we have a large amount of climate change--the amount that we are likely to get if we don't stop releasing greenhouse gases in the next 10 or 20 years--if will be difficult to feed all of the world's people and millions of people will loose their homes to rising seas. The question about global warming is: do we want to make it difficult for people around the world to feed themselves, to be happy and to be healthy?

3. What does you lab study at Notre Dame?
My lab studies the effects of climate on species and ecosystems, especially plants and insects. It is important to know how insects react to changing the climate because they play an important role in healthy ecosystems. We also study ways that people can manage species and ecosystems under climate change to try to preserve them for future generations. Check out our lab web page: http://www.nd.edu/~hellmann.

4. How much does deforestation affect global warming?
~15% of the greenhouse gases emitted that are causing global warming come from deforestation and forest degradation.

5. How long will it take for global cooling to come?
Global cooling isn't going to come for a long, long time, many thousands of years. The peak of the next ice age probably won't happen for about 80,000 years. The earth naturally goes in and out of ice ages based on variations of the earth's orbit. We are in one of the warm periods right now, called the Holocene, and we have been in this warm period for about 12,000 years. Interestingly, human emissions of greenhouse gases has pushed our climate way outside of the normal ups and downs that it experiences during and between the ice ages. So it's interesting question--one that scientists don't quite understand yet--if our changes to the climate will slow down or delay the start of the next ice age. When we talk about negative effects of global warming, however, we are usually thinking about how it will affect the next few generations of people, not our distant ancestors. 

6. Is there such thing as an ozone layer? How does it affect the environment?
The ozone layer is a really helpful part of the upper atmosphere where ozone tends to concentrate, and it helps to filter ultraviolet radiation that is harmful to living organisms in large doses. Some chemicals made by people, called CFCs, made their way into the upper atmosphere and broke down the ozone layer, creating the ozone hole. The ozone hole lets more UV reach the surface of the earth. Because many governments around the world passed laws outlawing CFCs, the growth in the ozone hole has slowed down. The ozone hole is a different problem than global warming, but the fact that we could stop growth in the ozone hole gives us some hope that we could also solve the problem of global warming. If society could just decide to take action through laws or other mechanisms, we can slow and stop the emission of greenhouse gases.

7. What causes acid rain?
Acid raid is caused by the release sulfur and nitrogen-based compounds from power plants and other things that burn fossil fuels. These compounds get in to the air and combine with water droplets to make the water acidic. So when those droplets fall from the air, they are "acid rain." The sources that make acid rain also release greenhouse gases, but these are different environmental problems. Learn more about acid raid at this EPA website: http://www.epa.gov/acidrain/what/index.html

8. If some of us start to stop releasing greenhouse gases, what effect will it have on the earth?
If some--or better yet many!--of us were to stop releasing greenhouse gases, we would slow down climate change. The more that the world emits, the more and the faster the climate changes. Eventually stopping emissions is the ultimate goal to stop the process of global warming. 

9. What is the strongest greenhouse gas?
Of the big three greenhouse gases, nitrous oxide is the most potent. Each molecule has ~300 times the heat trapping capacity of one molecule of carbon dioxide.  Each of the greenhouse gases, however, stays in the atmosphere a different length of time, so when thinking about the effect of each gas we have to think about how much we emit, how potent each molecule is, and how long it stays in the atmosphere. CO2 is the most important greenhouse gas because we emit so much of us and it stays in the atmosphere for a very long time.

10. How were there alligators in the Arctic?
In the early Eocene, about 50 million years ago, the Arctic was about 8 degrees C (or 14.5 degrees F) warmer than it is was before the humans started enhancing the greenhouse effect. At that time, northern parts of Canada had turtles, alligators, primates, and tapirs. Climate models tell us that if we keep on releasing more and more greenhouse gases to the atmosphere, like we have been doing the last 100 years, the Arctic could be that warm again by the end of this century.

11. Can we stop global warming completely?
Yes, if when we say "global warming" we mean the influence of people on the climate, we can stop that. All we need to do is stop adding carbon dioxide, nitrous oxide, methane, and other greenhouse gases to the atmosphere. To do that, we will need much greater energy efficiency than we have today--turn off those light bulbs when you don't need them and use energy-efficient appliances!--and we will need alternative energy sources that do not pollute the atmosphere, like solar and wind power.

12. Could the world ever be “fixed,” come back to its natural temperature?
If we could stop emitting more greenhouse gases to the atmosphere and take back the ones that we have already emitted, we could bring the earth back to the atmosphere that it would naturally have. It is going to be a lot easier to stop putting more greenhouse gases into the atmosphere, however, than it will be to remove the ones that we already put in. So we likely will have to live with some climate change from the gases that we have already emitted.

13. Will the government ever do something about global warming?
I'm afraid that this is one is hard to answer, and particularly hard for a scientist to answer. I think that people must have information about problems in order to want to do something about them, and I see that as my role--to help inform the public about an important problem. But there seem to be factors other than information that are holding politicians back. Some people are working hard to make sure that the government doesn't do anything because they benefit from the industries that release greenhouse gases. The way our political system works, it is also hard for politicians to make decisions that affect people today for the benefit of people in the future. Politicians are often more worried about getting reelected in 2 or 6 years than they are worried about what the climate will be like in 50 years. The only people who can get them to change their mind about that are citizens like you! 

14. What will happen to the ocean under global warming?
Global warming will cause the ocean to rise. First, the ocean will warm as it takes up some of the extra heat in the atmosphere and this will cause it to expand. Second, ice at the poles that is on land seems to be melting at a rapid rate under global warming, and this water will flow into the ocean. More water in the ocean means higher seas. 

15. What areas does global warming affect the most?
The largest amount of warming under global warming will take place at the poles and over land away from large bodies of water. The oceans will warm too, but we expect the average temperature over land to increase--at least within this century--more than the air over the ocean. You can see the patterns of warming on this map: http://www.ipcc.ch/publications_and_data/ar4/syr/en/figure-spm-6.html.

16. Is it true that global warming will happen anyway so there’s no need to try to stop it?
No, this is not true. We can stop global warming if we want to by stop releasing greenhouse gases to the atmosphere. If we keep releasing greenhouse gases, we continue to make global warming stronger and more severe. Some scientists are working on ways of taking out of the air some of the greenhouse gases that we already released. These technologies are probably a long way away, but they are important things to study.

17. How much CO2 does the average car release?
According to the US EPA, the average care releases 4.8 metric tons of CO2 equivalents per year. (CO2 equivalents allows one to think about all of the greenhouse gases coming out of a car together in one calculation.) Check out the EPA webpage for more calculations: http://www.epa.gov/cleanenergy/energy-resources/refs.html. We recently had a speaker visit Notre Dame (David Archer from the University of Chicago) who explained that each gallon of gasoline that we burn in our cars traps thousands-of-times more energy in the atmosphere than the energy value we get from burning the gas in the first place. 

Tuesday, April 30, 2013

Reflections on science communication & outreach--part of a blog carnival


On April 30, COMPASS published a commentary a paper in PLOS Biology on the journey from science outreach to meaningful engagement. This post is part of a series of reactions, reflections, and personal experiences to expand the conversation. Track the conversation by reading the summary or searching for #reachingoutsci.

I was a new assistant professor counting plants in the rain when I first truly realized that time was in short supply. The work was progressing slowly and my mood was soggy. I had to write a promised blog post for the class I was missing; I had a grant proposal due the next day that still needed to be routed through the research office; and I was having trouble with one of my field assistance who was going to need a heart-to-heart chat very soon. Don’t get me wrong. I had been busy and frantic before. Grad students are stressed; postdocs work hard; and I’ve never met an undergrad who hasn’t pulled at least one all-nighter. But I realized that this time constraint that I was facing wasn’t acute. It was chronic, and it was likely going to get worse because I only had more that I wanted to do.

One of the most important “more” that I wanted to do was engage with the people affected by my research. I realized that while standing in the rain, and I made a commitment to myself to try to be efficient and deliberate in my work choices. If I wanted to be accessible and relevant, for example, I might start by training someone else to stand in the rain counting plants. (Of course, every ecologists needs to spend at least some time in the rain to stay close to their study system.) My initial outreach and engagement attempts—once I had secured more field help—were initially targeted at the individuals who managed the land where we my students and I were performing research. I wanted to attend their planning meetings, have my grad students speak in their regional management conferences, and produce meaningful reports that helped them make decisions. I’m not sure that ever accomplished the latter, but we were able to draw regional attention to our research and the issues that we were studying.

Ten years later, my basic goals in outreach remain the same—help to make sure that what we are finding finds its way into the hands of someone who can use it and in a useful form—but the scope of my research has grown. Again, I’m faced with choices about how best to spend my time. I’m not so naïve to think that science by itself will change the world. In fact, if changing the world were my primary goal, I probably should have chosen another field. I chose to be an environmental scientist because I enjoy the mixture of discovery for the purpose only of knowing how nature works and the significance of those findings to society.

To achieve my outreach goals today, I have tried to implement a few things. First, I’ve tried to obtain more training, primarily through the Leopold Leadership Program and COMPASS but also through consultation with colleagues whose work in this area I really admire. Second, I’ve tried to kill as many birds as possible with one stone. For example, I’ve started using social media an outreach medium to talk about the scientific and science-social issues that I think are important, but I also use this medium to keep track of what is going on in my field and environmental news. In other words, I’ve switched from other modes of being informed to spend time in a place where I can also practice communication, accessibility, and transparency. And it’s quick. Third, I try not to let my worries take up too much of my time. I care deeply, for example, about the opinions of my peers and their evaluation of my scientific work. But that doesn’t mean that everything I do is intended for a peer audience, and I don’t need to continually fret about their opinions of my outreach and engagement (though I still do about promotion!).

I try to remember with some regularity that feeling that I had while standing in the rain. Over the life of a career, I know that I will feel that same sensation over and over again. But I’m trying to continually refine and redefine my priorities, make sure that my efforts are well-aligned with those priorities, and remember to seek help and assistance where my time and talents are not best invested. I’m grateful for a lab group to help me with all of this, and I hope that all of my students also have their rainy moment some day soon—and I hope that they become better scientists for it.