Implementing Course-Based Undergraduate ... - ACS Publications

Course-based undergraduate research experiences (CUREs) provide significant benefits to .... (3 hours) per week. An undergraduate teaching assistant (...
1 downloads 10 Views 419KB Size
Chapter 3

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

Throwing Away the Cookbook: Implementing Course-Based Undergraduate Research Experiences (CUREs) in Chemistry Jennifer M. Heemstra,*,1 Rory Waterman,*,2 John M. Antos,3 Penny J. Beuning,4 Scott K. Bur,5 Linda Columbus,6 Andrew L. Feig,7 Amelia A. Fuller,8 Jason G. Gillmore,9 Aaron M. Leconte,10 Casey H. Londergan,11 William C. K. Pomerantz,12 Jennifer A. Prescher,13 and Levi M. Stanley14 1Department of Chemistry, University of Utah, 315 S. 1400 E, Salt Lake City, Utah 84112, United States 2Department of Chemistry, University of Vermont, 82 University Place, Burlington, Vermont 05045, United States 3Department of Chemistry, Western Washington University, 516 High Street, Bellingham, Washington 98225, United States 4Department of Chemistry and Chemical Biology, Northeastern University, 360 Huntington Avenue, Boston, Massachusetts 02115, United States 5Department of Chemistry, Gustavus Adolphus College, 800 West College Avenue, Saint Peter, Minnesota 56082, United States 6Department of Chemistry, University of Virginia, McCormick Road, Charlottesville, Virginia 22904, United States 7Department of Chemistry, Wayne State University, 5101 Cass Avenue, Detroit, Michigan 48202, United States 8Department of Chemistry & Biochemistry, Santa Clara University, 500 El Camino Real, Santa Clara, California 95053, United States 9Department of Chemistry, Hope College, 35 E. 12th Street, Holland, Michigan 49423, United States 10W.M. Keck Science Department of Claremont McKenna, Pitzer, and Scripps Colleges, 925 N. Mills Avenue, Claremont, California 91711, United States 11Department of Chemistry, Haverford College, 370 Lancaster Avenue, Haverford, Pennsylvania 19041, United States 12Department of Chemistry, University of Minnesota, 207 Pleasant Street SE, Minneapolis, Minnesota 55455, United States 13Department of Chemistry, University of California, 1102 Natural Sciences 2, Irvine, California 92697, United States

© 2017 American Chemical Society Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

14Department

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

of Chemistry, Iowa State University, 2415 Osborn Drive, Ames, Iowa 50011, United States *E-mails: [email protected] (J.M.H.); [email protected] (R.W.).

Course-based undergraduate research experiences (CUREs) provide significant benefits to students compared to prescriptive (“cookbook”) laboratory curricula. However, carrying out research on the scale of an undergraduate course can present logistical challenges. Fortunately, the CURE format provides significant flexibility to tailor curricula to meet the needs of various class sizes, disciplines, and student groups, and to fit with the resources available in the institutional environment. Here we present a diversity of experiences and perspectives on the implementation of CUREs, with the goal of offering examples and practical advice to current or prospective CURE practitioners while highlighting the numerous approaches available for incorporating authentic research into undergraduate laboratory courses.

Introduction The recent Engage to Excel PCAST report lists “replacing standard laboratory courses with discovery-based research courses” as one of the top five recommendations for preparing tomorrow’s STEM workforce (1). This is supported by assessment data demonstrating that CUREs provide significant benefit to both students and faculty members (2–4). Despite the wide acceptance that discovery-based curricula are superior to “cookbook” labs, significant barriers to implementation of this evidence-based practice still remain. Key barriers cited by faculty include logistics, time investment, increased cost, research problem selection, and the chance of failed experiments (3). While some of these challenges are general across disciplines, others are specific to the subject area. This chapter aims to offer practical advice for implementing CUREs in the field of chemistry, and demonstrates the wide range of approaches that are available for successful implementation of a CURE.

What Is a CURE? A broad spectrum of laboratory curricula formats exists, including standard prescriptive experiments, inquiry-based experiments, and authentic research activities. Recognizing the need to delineate between these different formats, the 2012 CUREnet working group sought to address the question of what constitutes a CURE (2). This working group proposed that CURE curricula are characterized by the inclusion of five elements: (1) scientific practices such as formulating 34 Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

hypotheses, collecting and analyzing data, coping with the “messiness” of real world data, and communicating findings; (2) discovery of new knowledge or capabilities that were previously unknown to both the students and instructor; (3) research that is broadly relevant and has potential to impact society or the scientific community; (4) collaboration between students; (5) an iterative process in which data from one experiment are used to design or guide a subsequent experiment. Arguably, all laboratory curricula include at least some of the elements listed above. For example, even “cookbook” laboratory courses typically require students to collect and analyze data and collaborate with other students. Taking a step beyond this, inquiry-based laboratory courses are characterized by their requirement that students discover knowledge that is unknown to them (but is typically known to the instructor), and students frequently have the opportunity to utilize data from one experiment to design a subsequent experiment. However, in contrast to “cookbook” and inquiry-based laboratory curricula, CUREs are characterized by inclusion of all five of the elements listed above, which are nearly identical to the characteristics of traditional research internships in the university setting.

Challenges and Opportunities CUREs represent a unique opportunity to dramatically increase student exposure to authentic research. However, scaling the research experience from that found in a faculty research laboratory up to the size of a course can present challenges. There is a rapidly growing body of literature demonstrating that CUREs increase student engagement and learning gains, and thus provide many of the same benefits of traditional research internships (5–12). Additionally, Brownell and coworkers recently conducted an investigation into faculty attitudes and experiences with teaching CUREs. Through a series of interviews with faculty in the biological sciences who had recently taught a CURE, key themes were identified regarding the perceived challenges and opportunities associated with this curriculum format. The three most frequently reported challenges were logistics, time investment, and financial constraints, while key opportunities included connecting education and research, increased enjoyment of teaching, and the ability of the research from the CURE to enhance the faculty member’s scholarly productivity (3). While a number of resources are available to aid in the implementation of CUREs, these are primarily focused on the biological sciences. Because the challenges associated with implementing a CURE are primarily practical in nature, they often require solutions that are specific to a given discipline. Thus, many of the resources currently available may be difficult to translate to the chemical sciences. A working group of Cottrell Scholars has formed to address this need by generating resources aimed specifically at CURE implementation in the physical sciences. This chapter serves as an initial work product of that effort, with the goal of sharing our own experiences and perspectives in the implementation of CUREs in the chemical sciences. The examples below 35

Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

highlight that just as there is a virtually infinite number of ways to carry out research, CUREs offer significant flexibility in their design, and thus can be customized to accommodate factors such as class size, subject area, resources available, and institutional environment. Our goals in sharing these perspectives are to disseminate ideas that may inspire additional faculty to “throw away the cookbook” and incorporate research into their curricula, and to highlight the many options that are available for doing so.

Perspectives and Experiences in Teaching CUREs Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

Andrew Feig, Wayne State University, Department of Chemistry In conjunction with the ReBUILDetroit Program, a consortium also including University of Detroit Mercy and Marygrove College Course title: Research Coordination Network Laboratories Level: Freshman Approximate enrollment: 75 students on three campuses

Lecture/Laboratory Format – – –

Biology 4 credit combined lecture and lab where the lab is in the CURE format Chemistry 1 credit lab taken during the same term as 3 credit lecture course Health Disparities, 1 credit add-on taken during the same term as a 3 credit introduction to public health course

Summary of the Course The Chemistry Lab course focused on the analytical chemistry of urban gardens and issues related to soil contamination. In the first year, the team looked at garlic as a model system for understanding the uptake of metal from contaminated soil using commercial soil (purchased from a garden store), sampled ground soil, and doped samples. Students developed standard operating procedures (SOPs) for the analysis of several metal contaminants. They then grew garlic plants in the coded soil samples and measured speciation of the metal contaminants in different parts of the plant and the garlic bulb. The biology project, depending on the site, either participated in the HHMI SEA-PHAGE project (9) or the Barcode of Life Project (13, 14). In SEA-PHAGE, students isolate novel phage from soil samples in the first term and then sequence one of the novel phage species. In the second semester, they annotate the genome looking for novel features. The Barcode project focused on water samples from a local watershed, identifying invertebrate species in the sample using modern DNA methods and investigating the health of the waterway based on the diversity 36

Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

of species isolated. Attempts are being made to go further and look at the microbiomes within the GI tracts of the invertebrates to look for pathogens that are being transmitted through these vectors. The health disparities work has focused on issues related to food choices and the impact on public health using a mixed methods approach. That project has led to a less focused result due to consortium-related issues. The course is taught in different majors on the three campuses and the projects are not well-synchronized yet. They currently function more as individual projects than a concerted whole, and structural issues are being addressed to try to bring this project into better alignment with the cross-consortium goals for curriculum alignment.

Why did you decide to teach the course as a CURE? The CUREs are part of an integrated first year curriculum for the BUILD Scholars program (15). These students participate in a pre-freshman year summer program, a fall research methods class and then a winter term CURE. The entire curriculum is based on cohort building and development of self-efficacy in research, such that students can picture themselves as researchers. The long-term goal of the program is to enhance retention of underrepresented groups in biomedical research.

What lessons did you learn and what changes did you make (or would you make in the future)? Faculty professional development is essential for running the courses effectively. This was a team project involving more than 10 faculty across three campuses in Detroit and multiple disciplines, but with common learning outcomes for all sections. While it was easy to put the learning outcomes on paper, getting the discipline-specific actionable pedagogy in place was more difficult. The goal was to have the projects be very culturally aware and informed, but the faculty research interests of the instructors involved did not always facilitate the easy development of one common project that all members could rally around.

What do you view as the most important outcome from your CURE? The most important outcome is associated with science identity. Students completing the course show very large changes in their self-perceptions of themselves as scientists and their readiness to enter research laboratories for mentored research experiences at the end of their first year in college.

37 Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Aaron Leconte, Keck Science Department of Claremont McKenna, Pitzer, and Scripps Colleges Course title: Introduction to Biological Chemistry (IBC) Level: Freshman Approximate enrollment: 50 students

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

Lecture/Laboratory Format The course blends the first semester of Introductory Chemistry and Introductory Biology and counts as two courses. Lecture meets for 6 hours/week. Lab meets for 8 hours/week. The labs have undergraduate student teaching assistants who will be researchers in my laboratory.

Summary of the Course Introduction to Biological Chemistry (IBC) is a course co-developed by Biology and Chemistry faculty members in the Keck Science Department with the support of the S.D. Bechtel Jr. Foundation. The course is designed to use the intersections of modern biology and chemistry to reinforce and enrich the core ideas in these subject areas. In this course, students should expect: –



– –

A challenging and exciting introductory-level course that intertwines and enhances the core material of a typical Introductory Biology and Introductory Chemistry course. Exposure to modern research that combines chemistry and biology such as personalized medicine, epigenetics, protein engineering, synthetic biology, and more. To apply core course concepts to modern research questions on the current leading edges of biology, chemistry, and all things in between! A meaningful laboratory experience that emphasizes skills development while pursuing a faculty-guided, student-driven modern research project.

Why did you decide to teach the course as a CURE? Early research experiences are well-known to be crucial to development of young research scientists, increasing retention, GPA, and pursuit of graduate studies. Importantly, these opportunities are particularly important for underrepresented groups in science such as women and minority groups. While our department offers quite a few CUREs, they are all part of upper-level courses for more advanced students. I wanted to give first-year students the opportunity to get some exposure to research during their first semester. IBC, which serves as both an Introductory Biology and Introductory Chemistry course seems like an ideal place in our curriculum to begin developing CUREs for first-year students. 38 Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

What do you view as the most important outcome from your CURE? Overall, I want students to gain an appreciation for and understanding of the scientific process. There are two major aspects of a research experience that I hope that students in my class will obtain. Specifically, students should i.) experience the intellectual excitement of research by creating and characterizing new proteins and ii.) experience the culture of research by working on team-focused research initiatives and interfacing with more senior undergraduate student researchers.

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

John M. Antos, Western Washington University, Department of Chemistry Course title: CHEM 356 – Organic Chemistry Laboratory II for Life Sciences Level: Sophomore/Junior Approximate enrollment: 24–48 students

Lecture/Laboratory Format 1–2 sections (24 students per section) are offered per academic quarter. Each section has one lecture (50 mins) and one laboratory session (3 hours) per week. An undergraduate teaching assistant (TA) assists the faculty instructor during lab sessions. All other responsibilities (lectures, grading) are handled by the faculty instructor.

Summary of the Course CHEM 356 is an advanced undergraduate Organic Chemistry Lab course. The goal of this course is to instruct students in the theory and practice of experimental organic chemistry, including reaction setup using air and moisture sensitive materials, separation and purification techniques, and practical spectroscopy (IR and NMR). This course is also intended to meet the curricular needs of pre-health, biochemistry, and life sciences majors; therefore, course content is designed to emphasize the relevance of organic chemistry to these disciplines. In its current iteration, approximately half of the course is taught as a CURE in which students are tasked with developing inhibitors of a therapeutically relevant enzyme family. Other lab exercises consist of stand-alone modules focused on organocatalysis, synthesis of over-the-counter pharmaceuticals, and enzyme-catalyzed reactions. With respect to the CURE component, students engage in a multi-week research project in which they design, synthesize, and test structurally diverse small molecule chalcones for their ability to serve as covalent inhibitors of bacterial sortase enzymes. Sortases have an established role in anchoring virulence factors to the bacterial cell wall, and therefore represent authentic targets for drug development. Furthermore, the chalcone scaffold has been largely unexplored for sortase inhibition, and therefore student effort on this project will generate new knowledge for the field. Specific activities in the CURE begin with computer-aided structure design (covalent docking). 39

Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

Students are first given a set of available chalcone precursors. They then model the interactions of multiple chalcones with the sortase enzyme in order to select a lead inhibitor structure. Students then synthesize their chalcone of choice via aldol condensation using commercially available acetophenone and benzaldehyde building blocks. Chalcone structures are verified using standard characterization techniques (melting point, IR, 1H NMR). Finally, students test the inhibitory activity of their chalcone in vitro using recombinant sortase A and model peptide substrates. In total, the CURE project covers four weeks and exposes students to topics in synthetic organic chemistry, computational modeling, and enzymology. Moreover, the CURE project is intended to be iterative, with the results of current students informing and guiding the activities of subsequent class sections.

Why did you decide to teach the course as a CURE? My decision to incorporate a CURE into this course arose from a desire to expose more undergraduates to research and to foster increased student engagement in our organic chemistry curriculum. By offering an authentic research problem to course participants, I’m able to include significantly more students in the scientific process than I’m able to accommodate in my independent research group. Furthermore, by offering each student the opportunity to make genuine discoveries, my hope is that students will take ownership of the project and feel as though they have a stake in the experimental outcome. One additional benefit of the CURE model is that it provides a great way to infuse my scholarly interests into the teaching curriculum. This fact, coupled will the continually evolving nature of the CURE project, ensures that the course will remain fresh and relevant for both instructor and students.

What lessons did you learn and what changes did you make (or would you make in the future)? The fact that this CURE arose from my independent research has made the implementation of the project fairly straightforward. However, a lesson I have learned is that the individual CURE activities do require time for development and optimization. An effective strategy to address this is to gradually phase in new components of the CURE each quarter. For example, the first instance of this CURE did not include computer modeling, a feature which has since been added to the project. Looking ahead, I plan to continue expanding the CURE and would like to have the project encompass the entire 10-week academic quarter. An additional change that I would like to make concerns transfer of results and information between course sections. The CURE is intended to be iterative, and so facilitating the transfer of data between class sections is critical. A way that I plan to address this is by maintaining a database of previous results and then introducing a writing assignment at the outset of the CURE project that requires students to evaluate and discuss this information as a point of departure for their work. 40

Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

What do you view as the most important outcome from your CURE?

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

Given the fact that we are still in the early stages of implementing this CURE at Western Washington University, the most concrete positive outcome is simply the increase in number of undergraduate students exposed to research. Going forward, this CURE will be performed with a larger cohort of students across multiple class sections, and we will be employing the Grinnell College CURE Survey as a means to more rigorously assess student experiences (16). It is my hope that this CURE will enhance students’ confidence in their abilities, provide them with an understanding of the scientific process, and ultimately strengthen student interest in pursuing a career in STEM. Amelia Fuller, Santa Clara University, Department of Chemistry & Biochemistry Course title: Organic Chemistry II Laboratory Level: Sophomore Approximate enrollment: 16 students

Lecture/Laboratory Format I teach one section of approximately ten offered in each term, and it is populated primarily by chemistry and biochemistry majors. The course is 10 weeks long and meets once per week for 4 hours. I personally instruct the laboratory, and one undergraduate TA assists me with preparation of reagents and materials, helps students use instrumentation (e.g., acquisition of NMR spectra) and assists with troubleshooting experiments in the laboratory.

Summary of the Course I have developed a 10-week curriculum to provide an authentic, discovery-based research experience for many of the chemistry and biochemistry majors inspired by the model of “Distributed Drug Discovery” pioneered by Profs. William Scott and Martin O’Donnell at Indiana University-Purdue University, Indianapolis (17). Drug discovery is frequently initiated by screening large compound libraries for desired biological activity. Because of the large number of students involved, beginning organic chemistry students can substantially contribute to the diversity of potentially bioactive molecules available. Both the students and I are inspired and motivated by the exciting prospect of making and identifying a bioactive molecule. In this course, each student designs a combinatorial library and carries out parallel six-step syntheses of six compounds based on the “arylopeptoid” molecular scaffold. Although robust methods to prepare analogs with this molecular scaffold have been detailed in the literature, our examples expand the scope of this chemistry; molecules prepared in this course have never been 41

Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

reported in the chemical literature. Thus, in addition to preparing diverse examples for biological evaluation, analysis of synthetic data enables us to assess the scope of these reactions and to optimize reaction conditions efficiently. Crude reaction products are analyzed by LC-MS to evaluate purities. Following their synthesis, the inhibition of bacterial growth by these molecules is evaluated in my own research laboratory or at Community for Open Antimicrobial Drug Discovery (CO-ADD), which offers free evaluation of synthetic compounds’ ability to kill important pathogens. In their responses on simple surveys, students are highly enthusiastic about the research-based approach and self-report gains in conceptual understanding and in their confidence.

Why did you decide to teach the course as a CURE? I have two motivations for teaching this course as a CURE. First, I want to advance my own scholarship. The CURE structure allows me to do this by obtaining more experimental results by engaging more students in projects that interest me. Moreover, I am able to identify excellent students early in their undergraduate careers and train them with skills that will enable smooth transition to contributing to related projects in my independent research laboratory. Second, we simply cannot accommodate all of the students who wish to do independent research in our individual faculty laboratories. The CURE course structure allows us to train more students in research.

What lessons did you learn and what changes did you make (or would you make in the future)? A significant challenge as I implement my CURE course is to calibrate how much independence to give students. Two representative considerations are highlighted here. First, I charge students with the preparation of their own reagent solutions to mimic what they would do in a research laboratory. Although this slows down the process and introduces an added level of variability to consider when analyzing student results, I feel that the experience and skills developed are more important. Second, I have allowed students to design what compounds they want to prepare given a set of combinatorial reagents. Because of the somewhat random nature of which molecules are made, we have “gaps” in the reaction space that has been covered. In future iterations of the laboratory, I intend to structure this more closely to be more systematic about the targets to be synthesized while still enabling some student participation in experiment design.

What do you view as the most important outcome from your CURE? I think the most important outcome from students’ participation in the CURE course is that they are more confident in their ability to contribute to scientific research. In this course, we take small steps toward solving big, important 42 Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

research problems, and we make those steps accessible to novice scientists. Students emerge from the course with a greater appreciation of how they can contribute among the many hands and many hours that go into making meaningful scientific discoveries. Scott Bur, Gustavus Adolphus College, Department of Chemistry, William C. K. Pomerantz, University of Minnesota, Department of Chemistry

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

Course title: Organic Chemistry Laboratory II Level: Sophomore Approximate enrollment: 80 students

Lecture/Laboratory Format Two sections meet two times per week with two TAs.

Summary of the Course This course is a redesign of a traditional skill-focused curriculum to incorporate discovery-based research experiences into an undergraduate Organic Chemistry Laboratory. The laboratory is designed around discovery of bio-active small molecules, based on multi-step reaction design and “fragment-based” synthesis, and using a readily analyzed biomolecular nuclear magnetic resonance spectroscopy (NMR) experiment. Fragment-based synthesis with protein NMR analysis has yet to be incorporated in an organic chemistry laboratory, but offers a route for independent molecular design, tractable syntheses, and exposure to macromolecule:small-molecule analysis. One method to reduce the challenge of small molecule discovery for protein interfaces is to reduce the size and complexity of small molecules that are employed, a technique called fragment-based screening. However, a sensitive detection method such as NMR is required to quantify the interaction. The Pomerantz laboratory screened 508 low complexity small molecules called fragments and characterized their interactions along the protein surface (18). Due to the low complexity of the molecules, students in the CURE course at Gustavus Adolphus College (GAC) will design analogs of these small organic molecules, each taking ownership of their individual scaffold and fostering creativity in the design process. As a skill building exercise, students will first carry out a three-step synthesis of a known ligand. This molecule will also be used at the end of the semester for conducting a modern organometallic coupling reaction. Derivatives of their first small molecule will then be proposed by the students, selecting from a matrix of different building blocks to yield a total of 80 target molecules. Small molecule fragments can be made in short two to four step syntheses with only a small amount of material needed for testing (1–5 mg), making multi-step synthesis tractable in an undergraduate laboratory. Small molecules will be tested against the Pomerantz laboratory’s fluorinated proteins. 43

Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

The 19F NMR spectrum is acquired in a few minutes and requires low amounts of protein. All students will receive the NMR spectral data to analyze and assess the success of their designs. The learning objectives to be met by this research experience in an undergraduate laboratory are designed to teach students to plan a multistep synthesis, troubleshoot reaction conditions using literature precedent rather than follow a prescribed set of reaction conditions, analyze biomolecular NMR data for comparing experimental results to a positive control, identify important parts of small molecules for binding, and develop new synthesis proposals for these small molecules. Assessment will include pre-and post-knowledge and attitudinal surveys developed with the UMN Center for Educational Innovation, and online assessment using the Classroom Undergraduate Research Experience (CURE) survey (16). Finally, analysis of two reflective writing assignments on the students’ perception of their research experience (pre-, and post semester) will be used to assess the change in of their impression of the scientific process (e.g., What is an experiment, a theory, a hypothesis, and research?).

Why did you decide to teach the course as a CURE? At both the University of Minnesota (UMN) and primarily undergraduate institution GAC, over 40% of chemistry majors fail to receive an independent research experience in their home department. Engaging a large number of students in a discovery-based research experience in the undergraduate laboratory can be scaled to help meet this need. Although challenging, successful cases such as the Science Education Alliance Phage Hunters Advancing Genomics and Evolutionary Science (SEA-PHAGES) course, “The Chemistry and Biology of Everyday Life (CBEL),” and the University of Texas at Austin’s Freshmen Research Initiative have garnered enthusiasm. Fragment-based synthesis with protein NMR analysis has yet to be incorporated in an organic chemistry laboratory, but offers a route for independent molecular design, tractable syntheses, and exposure to macromolecule:small-molecule analysis. A winter “J-term” pilot course of 17 students at GAC in 2015 and follow-up research experience this summer with three students hosted jointly by UMN and GAC, have generated considerable enthusiasm, prompting further curriculum development.

What lessons did you learn and what changes did you make (or would you make in the future)? The lessons still need to emerge as we are in the preplanning stages. As our baseline survey data is starting to come in, we will have a better idea for curricular changes.

44 Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

What do you view as the most important outcome from your CURE? The most important outcome that we hope to gain from this course is to see how we can reshape student perception and engagement with their major. The ultimate hope is to increase retention in STEM fields based on this early research experience, while providing a more realistic exposure to the scientific method.

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

Levi Stanley, Iowa State University, Department of Chemistry Course title: Laboratory in Organic Chemistry II (primarily for chemistry majors) Level: Sophomore Approximate enrollment: 25 students

Lecture/Laboratory Format There are 3 sections. Each section meets 2 times for a total of 6 hours. The course utilizes 4 TAs with one TA per section and a “head” TA that coordinates the research project.

Summary of the Course Our CURE is incorporated into our standard second-semester chemistry laboratory for chemistry majors and students pursuing a research career. This course is a 16-week course split into approximately 4 weeks of traditional “cookbook-type” labs, four weeks of inquiry-guided experiments, and eight weeks built around a course-based undergraduate research project. In 2012, we implemented a 4-week research experience (8 laboratory periods of 3 hours) into this course. Students were tasked with the multistep synthesis of a novel fluorous dye molecule as the main goal of the research experience. Although students were generally positive about the project and appreciate the introduction to advanced laboratory practices such as Schlenk techniques, affinity chromatography, and microwave-assisted chemistry, student motivation consistently wanes toward the end of the project. Assessment shows that students do not connect the project to a real-world research problem, and thus lose interest as the project moves forward without the potential for an end-game payoff. To improve student motivation and investment in our CURE, we redesigned the research experience by 1) providing target molecules that stimulate independent thought and are potentially important to the research community, 2) providing a framework for students to generate publishable data, and 3) providing a gateway project to enter research in faculty laboratories. Students are tasked with the synthesis of a derivative of a 2,3-dihydro-1H-pyrroloindole. Students are required to research known methods to access 2,3-dihydro-1H-pyrroloindoles and present a proposed route to the class three weeks prior to the start of the project. Students are asked to focus on developing flexible routes that enable 45

Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

the installation of a variety of substituents on the 2,3-dihydro-1H-pyrroloindole core. The laboratory instructor and TAs provide feedback on the feasibility of the route (in an undergraduate laboratory) and suggest modifications to the proposed routes at this time. Students whose synthetic routes are not appropriate for an undergraduate laboratory or the duration of the project will be asked to consider an alternative route that has been pre-validated by the PI’s laboratory. The 2,3-dihydropyrroloindoles that are successfully synthesized are then screened for biological activity in collaboration with scientists at the College of Veterinary Medicine at Iowa State University.

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

Why did you decide to teach the course as a CURE? The importance of incorporating research opportunities into the undergraduate experience has become a driver for improvement in undergraduate curriculum over the past twenty years. However, simply incorporating a research experience into an undergraduate course is not enough to generate positive student outcomes. Studies show that students benefit most from research experiences in laboratory courses in which fundamental concepts are emphasized in the context of current research applications and problems. These types of laboratory course-based research experiences have been shown to improve student interest, enhance independent thinking, and motivate students to search for answers beyond textbooks and faculty.

What lessons did you learn and what changes did you make (or would you make in the future)? I learned that progress is never as fast as one anticipates in a CURE. Our undergraduate students do not like to fail in the laboratory and have a hard time adjusting to unsuccessful experiments as an important part of research. Thus, significant time to explain the goals of the CURE is necessary to motivate students. I also found that anticipation of proposed research directions and timely feedback to students are critical in CUREs that provide significant freedom for students to design synthetic routes. Through three iterations of the course, we have scaled back our expectations for student progress. We have accommodated this change by providing students with more advanced starting materials to ensure that more students “finish” the project.

What do you view as the most important outcome from your CURE? I view three outcomes as significant from our CURE: 1.

Students learn for the first time to design and execute experiments without a provided procedure. 46

Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

2. 3.

Students learn to approach the chemical literature and adapt existing synthetic methods to make new molecules. Students learn to deal with unsuccessful experiments in the laboratory and find solutions to achieve their project goals.

Jason G. Gillmore and Traci L. Smith (and all Hope College organic chemistry lab faculty since 1971), Hope College, Department of Chemistry

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

Course title: CHEM 256B Organic Chemistry Laboratory Independent Synthesis Projects Level: Sophomore Approximate enrollment: 20–35 students

Lecture/Laboratory Format We run 4 sections of 4–10 students each (condensed from 5 sections of 16–20 students each of CHEM 256A Organic Chemistry Lab II in the first half of the semester) meeting for 7 weeks, second half of Spring Semester. Students in the course typically spend 5–10 hours per week over 7 weeks (2 weeks in library training and preparing proposals, 5 weeks in execution). This is a 1 credit, 5–6 hour per week, half-semester laboratory course, requiring a faculty member (ca. 1/2 – 2/3 FTE for half semester, or net 1/4 – 1/3 FTE for the semester teaching load) and one undergraduate TA per section of up to 10 students. The undergraduate TA helps with trouble-shooting, safety, lab supervision, instrumentation, etc., but a faculty member is present at all times. Also required are chemical equipment, supplies, and instrumentation (primarily NMR, GC/MS, FTIR, and melting point apparatus), mostly common to the organic teaching laboratories or faculty research laboratory but some specifically purchased for the projects. The support of an organic lab coordinator and supporting student workers is essential in gathering required chemicals and helping with ordering, as well as helping cover necessary "open lab" hours when projects do not fit neatly into laboratory periods. Time spent outside direct teaching required in this implementation include 30–40 hours by the organic laboratory coordinator, including preparing for library training, syllabi, an hour lecture on project structure and requirements, and mostly the gathering and ordering of necessary chemicals. An additional 30–40 student worker hours assisting are also required, along with an hour of science reference librarian time and 1–2 hours of time from each participating faculty member to delineate possible projects related to their research. As this has been an ongoing evolution but part of our curriculum for 45+ years, it is already built into the department budget and not much more costly than having a full semester Organic II Lab would be. Hope’s HHMI grant and Gillmore’s NSF CAREER grant have helped fund CURE assessment of this project (and we gratefully acknowledge the help of Stephen C. Scogin, Professor of Biology and Education at Hope College, for help with the analysis of CURE survey data). 47

Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

Summary of the Course This 7-week section is an optional continuation of CHEM 256A (half semester Organic Chemistry Laboratory II) to be taken in the same semester. In Chem 256B, students search the chemical literature, write a proposal, and execute an independent synthetic project concluding with a final report. Students grow in independence and autonomy and gain appreciation for authentic research while developing new skills such as reaction design, spectroscopic analysis, and the purification and characterization of mixtures. Offered last half of the semester. CHEM 231 Organic Chemistry II lecture is a required pre- or co-requisite course, as is CHEM 256A. Our 6 hour/week two-semester organic laboratory sequence has long included a half-semester three-step independent synthesis project in the second semester. Over the past two decades this has evolved from a mandatory, closely controlled suite of projects tied to a central ‘theme’ that varied annually, to an elective and increasingly independent set of projects tied to a wide variety of ongoing faculty research programs across chemistry and beyond. This has allowed faculty to leverage this course to advance their research as well as students’ learning, and has drawn a diversity of organic chemical targets from most of the faculty in the chemistry department and some from other STEM disciplines. Students gain exposure to and connection with authentic faculty research, providing some a springboard to future research engagement. All have the satisfaction of contributing to ongoing work rather than to a waste container. Projects begin with a literature search workshop on using SciFinder and other library resources. Students identify a target molecule and develop a synthetic plan. They craft a research proposal including lists of chemicals, hazards, and required equipment. Students execute the multistep synthesis with complete purification and characterization of intermediates over five weeks, and prepare a comprehensive written report. Two years of CURE survey data indicate students are successfully achieving desired learning gains from this course-based research experience at a level that is comparable with summer undergraduate research experiences. Moreover, targets prepared in these projects have contributed to the work of 21 faculty including 18 papers and 4 funded research proposals since 2000, which speaks to the authenticity of the research in this CURE.

Why did you decide to teach the course as a CURE? This course has had an undergraduate research-like experience since the 1970s, as a way to expose students to research early in their undergraduate careers, long before there was documented evidence of the effectiveness of such experiences. The shift to a genuine research focus over the past decade or two was to better leverage resources (faculty teaching time & department dollars, in what is a somewhat “expensive,” in both resources, endeavor) to advance ongoing externally funded publishable student-faculty collaborative research.

48 Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

What lessons did you learn and what changes did you make (or would you make in the future)? This course transitioned from being required of all Organic II students to an elective primarily taken by slightly more than half of all Chemistry majors (most of those pursuing ACS certified degrees, plus nearly half of those pursuing less rigorous chemistry degree tracks). This allowed us to manage increasing organic enrollment and the increased demands as we moved from a closely controlled suite of projects to a more diverse range of projects. This change was also somewhat student driven, as it allowed those with research in their home department or those not interested in additional organic chemistry the ability to opt out. To aid in registration logistics at the behest of our Registrar, we’ve recently moved from a variable credit model to a separate course number model, with little direct impact on us but much to the relief of the Registrar’s office.

What do you view as the most important outcome from your CURE? 1.

2.

3.

Hope College’s CHEM 256B advances ongoing faculty-student research programs by leveraging teaching time and dollars and the manpower of a fairly large course. The course expands capacity for giving students an early taste of research (enticing them to pursue organic research or research in other chemical sub-disciplines or different scientific disciplines altogether), or helping them realize that research is not something they want to pursue further. On the CURE survey, students who participate in the lab demonstrate substantial self-reported learning gains and scientific attitude changes more commensurate with, or even exceeding, a full-time 8-10 week summer research experience – we rationalize this based on the “full cycle” nature of searching the literature, preparing a proposal, gathering the resources, executing the plan, analyzing results, and preparing a report (even if this is far from a complete research project, they do all of the elements.)

Jennifer A. Prescher, University of California-Irvine, Department of Chemistry Course title: Chem 128L: Introduction to Chemical Biology Laboratory Level: Junior/Senior Approximate enrollment: 80–120 students

Lecture/Laboratory Format Chem 128L comprises one lecture (1 hour) and one laboratory session (4 hours) per week. We offer 8–10 laboratory sections each year, with one TA assigned per section. There is also a head TA assigned to the course. 49 Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

Summary of the Course Chem 128L is an undergraduate laboratory course designed to teach basic principles and techniques in chemical biology. This course is often a student’s first exposure to biology-related laboratory skills, as most chemistry graduates do not enroll in an advanced biochemistry course (offered by a different department). Thus, a large part of Chem 128L is devoted to mastering laboratory fundamentals: pipetting, gel electrophoresis, etc. To provide these students with more than just a set of techniques, I developed experiments that not only teach basic skills, but also incorporate an element of research. The most popular of these experiments involves a high-throughput screen of small molecules for anti-cancer properties. UCI is home to a collection of 3000+ molecules synthesized by various research laboratories on campus. The molecules are arrayed over 96-well plates and provided to researchers courtesy of the UCI High-Throughput Synthesis and Screening Facility (HTSS). For the lab exercise, the students screen the compound library using luciferase-labeled cancer cells. Cellular readouts are obtained using luminescence screening equipment, with a loss of signal correlating with cell death. Students work in teams to collect data and present their findings (via classroom PowerPoint presentations). Additionally, undergraduates interested in follow up analyses of compound “hits” are eligible for summer research opportunities. The experiment has thus far been a success: over 250 students have screened a total of >1200 compounds (in triplicate), and compounds with anti-cancer properties have been identified. Students have consistently praised the “real world” nature of the screening exercise in written evaluations. Importantly, the experiment also provides training in key areas relevant to chemical biology: compound screening, pharmacokinetics, assay development, and data analysis.

Why did you decide to teach the course as a CURE? One of my major educational goals involves exposing students to interdisciplinary science via “real world” experiences. Significant coursework is already required for a bachelor’s degree in chemistry at UCI, leaving little time for extra classes. So, rather than develop a suite of new specialized courses, I decided to leverage our existing curriculum and integrate authentic research projects into UCI laboratory courses. These experiences enhance the undergraduate experience and expose students to modern and interdisciplinary science.

What lessons did you learn and what changes did you make (or would you make in the future)? Several lessons were learned: 1.

Research-based inquiry can be integrated into large laboratory sections with careful planning. 50

Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

2. 3. 4.

Research-based classroom learning (in large courses) can lead to additional undergraduate research opportunities in the department. Students are particularly responsive and enthusiastic about UCI (home institution)-driven projects and data. Research-based inquiry in courses can build critical team working and data dissemination skills.

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

What do you view as the most important outcome from your CURE? While I’m not sure this is the most important outcome, the success of research-based laboratory exercises in Chem 128L is changing how other instructors approach their courses. Research-based exercises are not “impossible” to execute in large laboratories, and I have consulted with faculty at UCI and beyond about how to replicate some of the exercises in their own courses. Penny Beuning, Northeastern University, Department of Chemistry and Chemical Biology Course title: Chemical Biology for Chemists Level: Junior/Senior Approximate enrollment: 20–40 students

Lecture/Laboratory Format This is a 3 semester-hour lecture course with 1–2 lab sections (1 semester hour) in fall and spring that meet 3.5 hours per week There is one TA per not more than 18 students. The lab has several research projects that together comprise half of the semester.

Summary of the Course Chemical Biology for Chemists is an upper-level course that introduces students to biochemistry and chemical biology concepts. The course enrolls primarily Chemistry and Biochemistry majors. The lecture section includes active learning and presentations by students on papers from the primary research literature. The accompanying laboratory section provides opportunities for students to build skills in biochemistry along with completing research projects. The class satisfies the chemistry major requirement for Biochemistry and the laboratory is writing intensive. There are two small research projects and one multiple-week research project as part of this laboratory that integrate teaching skills with research. In the process of teaching sterile technique, we have the students carry out a phenotypic screening experiment using a zone-of-inhibition assay. Students are given a wild-type bacterial strain and the same strain with a gene of interest deleted. 51

Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

We chose different deletion strains of interest depending on current research in my laboratory. There are numerous knockout strains available from stock centers, so obtaining bacterial strains for this experiment is straightforward. In a second one-week experiment inspired by Computational Bridges to Experiments (COMBREX) (19, 20), the students are given a protein sequence that represents a protein of known function and asked to identify other proteins of similar sequence from sequenced genomes. Each student chooses a different protein from their list of similar sequences and then searches for a structure in the Protein Data Bank (www.pdb.org) or generates a homology model. From the structure or model, the students use a predictor that was developed at Northeastern to identify the functionally important residues (www.pool.neu.edu) (21). The students then compare the predicted functionally important residues of the known protein and their protein of interest in order to make predictions about protein function. In the multi-week experiment, students design and construct site-directed mutations in a protein. The proteins are chosen to be relevant to work in my research laboratory and, ideally, relevant to the research of the TA(s) for the course. The students are given an introduction to the protein either by the TA or by another student who is working on the protein as part of thesis or dissertation research. We usually guide the students to choose from a list of potential residues to mutate, along with why these residues are of interest. Past projects have included cancer-associated mutations, mutations of residues predicted by students in a molecular modeling class to be important for protein-protein interactions, or residues predicted by POOL to be important for activity. The students choose the mutation to make, design DNA primers, carry out the site-directed mutagenesis reaction and subsequent molecular biology steps, purify the plasmid DNA, carry out a restriction digest of their DNA and have the DNA sequenced. After the proteins are purified by TAs or others, we generally have the students carry out assays of their predicted effects. This aspect of the laboratory changes periodically depending on the protein and whether assays are amenable to a teaching laboratory. The students must include a rationale for their chosen mutation and discuss the effects on activity in their reports.

Why did you decide to teach the course as a CURE? Our students take this class with varying degrees of prior biology and biochemistry courses, and so one motivation for teaching this course as a CURE was to avoid duplicating laboratory experiences some students may have already had, while also ensuring that students learn fundamental skills. Northeastern follows a cooperative education model in which students alternate periods of full-time work with periods of attending classes. Therefore, another motivation for designing this course as a CURE was to make students more aware of biotechnology or biochemistry careers, as well as more competitive for co-op positions in those fields after having a more in-depth laboratory experience.

52 Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

What lessons did you learn and what changes did you make (or would you make in the future)? We have found that the skills of pipetting and sterile technique learned in the first week have to be reinforced throughout the semester. For many students, this is the first experience in research and we found that we need to emphasize that we expect new results to be obtained. So, for example, in the mutagenesis laboratory, mutations should be well justified and designed to give stable proteins if possible. We have also developed rubrics for grading reports and the TAs spend considerable time discussing writing in pre-lab lectures with the class. This course requires regular maintenance as we choose new “known” proteins for the bioinformatics laboratory every semester and every few semesters we choose new strains for phenotypic screening and proteins for site-directed mutagenesis. We also are careful not to use examples in which our need for results is critical; these experiments are ideally suited to gathering very preliminary data for new or longer-term projects in the research laboratory.

What do you view as the most important outcome from your CURE? There have been a number of important outcomes from this Chemical Biology laboratory. One is that some students find a new interest in carrying out chemical biology or biotechnology research. This class is writing intensive and some students have reported that they have come to appreciate the importance of writing for understanding material. They also report gains in confidence in using computer modeling. There is a major benefit to the teaching assistants who have taught the course, as they gain extensive teaching experience and in many cases, acquire data relevant to their projects. Ideally, TAs who have an interest in teaching as a career are sought to serve as TAs for this course, as they have the opportunity to redesign parts of the course periodically, providing a deep teaching experience and building their teaching portfolios. Linda Columbus, University of Virginia, Department of Chemistry Course title: Biological Chemistry Laboratory I and II Level: Junior/Senior Approximate enrollment: 90 students

Lecture/Laboratory Format The students meet for 50 minutes on Monday for the laboratory “lecture”, which is in an active learning classroom with round tables, multiple projectors, and a throwable microphone. The laboratory meets in six sections with no more than 18 students and is facilitated by a graduate student teaching assistant. The course is team taught with a departmental staff member and a research active faculty member. 53 Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

Summary of the Course The Biochemistry Laboratory course is designed as a full year, two semester curriculum (22). Students are assigned to study a protein which has a known structure and either a putative enzymatic function or a confirmed enzymatic function. Depending on which protein they are studying, their goal is to determine the function, which they are able to narrow down through the use of bioinformatics and their knowledge of the structure-function relationship, or, in the case of proteins with confirmed function, to design a mutation which will alter the specificity of the enzyme without eliminating activity. By creating a year-long curriculum, we are able focus our teaching in the first semester on the biochemical techniques and theory that the students need to know to conduct their second semester research. They then work much more independently in the second semester, applying what was learned in the first semester to study their protein. Students can take their project as far as time and their interests allow. Most groups are able to design assays that yield kinetic parameters for their enzyme and many groups go further, including doing substrate screens, assessing the pH and temperature range for the enzymatic activity, or determining the effect of inhibitors. Students enjoy the freedom they are afforded in this course to take their research in the direction that interests them. In addition to learning hands-on biochemistry techniques, the course is designed to give the students real-world experiences. These experiences include collaboratively working in groups, writing scientifically, reading primary literature, and communicating their work both orally during group meetings and at poster presentations and in writing in the final course paper, a publication style manuscript.

Why did you decide to teach the course as a CURE? The desire to educate scientists, not science majors, is at the heart of our redesign. While students at the University of Virginia, and certainly other institutions, excel at learning facts and learning about theories, many biochemistry graduates are deficient in their problem-solving abilities, and in their ability to design experiments. Many laboratory courses are focused primarily on integrating principles taught in lecture courses into the laboratory, using textbook experiments in an attempt to transfer information. Implementation of a CURE curriculum encourages students to be self-directed, while also learning the team skills necessary for most post-undergraduate pursuits. The curriculum does not require sacrificing the student’s mastery of basic techniques, but rather incorporates the learning of these techniques into real, bona fide research projects.

54 Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

What lessons did you learn and what changes did you make (or would you make in the future)? The main challenge for this course is the cost and number of personnel required to run the course. The development, instruction, and maintenance of the course was completed with a part-time PhD level instructor, two research-active faculty, teaching assistants (TAs), and undergraduate students. With the curriculum (mostly) established, the course requires a TA for each section (we have five – six for ~ 90 students), a part-time PhD level instructor, and one faculty instructor. The team aspect of the development was critical so that the research-active faculty could still maintain their research group. Training teaching assistants both in pedagogy and content remains a challenge. Most incoming graduate students have not seen the content or been challenged in experimental design as this course encourages. The TA could in essence be the principal investigator of the research that the laboratory is conducting; however, expectedly, they are not ready for this type of responsibility, mentoring, or teaching.

What do you view as the most important outcome from your CURE? First, all of our developed materials are distributed and shared with instructors through Biochemistry Laboratory Education (BioLEd) a resource available at http://biochemlab.org. Second, assessment indicates that 79% of the students learn the necessary understanding and skills by the end of the year course (using pre- and post-tests) (22). Using Student Assessment of their Learning Gains (SALG) (23), ~90% of the students positively rated their learning of biochemistry and experimental design at the end of the year compared to less than 30% before the class started. Finally, the number of students now impacted by this curriculum is over 500 and growing.

Jennifer Heemstra, University of Utah, Department of Chemistry Course title: Advanced Chemical Biology Laboratory Level: Junior/Senior Approximate enrollment: 50–60 students

Lecture/Laboratory Format This is a 7.5-week course during which the students have 2 hours/week of lecture and 8 hours/week of laboratory. Students are divided into 5 sections with an enrollment cap of 12 students each. Every section has one graduate student TA.

55 Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

Summary of the Course In this course, all of the students work together to generate new knowledge that can be beneficial to researchers in the area of biological chemistry. Part of the inspiration from this model comes from the widely-known publication in J. Org. Chem. that lists all of the NMR chemical shifts for common solvent impurities (24). While the research behind this paper is by no means as “flashy” and “innovative” as other work that is published in the journal, this paper consistently ranks as the most accessed article in the journal, and serves as a key reference for innumerable organic chemistry students and researchers. Thinking about this paper helped me realize that there is much research that can be of great value to the community, but is straightforward enough to be completed in the context of a CURE. Thus, the overarching theme for this CURE is to find reactions that are used by a large number of scientists, but where systematic studies of reaction parameters are lacking. In the first two iterations of the course, the curriculum has focused on exploring the effect of buffers, solvents, and surfactants on the kinetics of strain-promoted azide-alkyne cycloaddition (SPAAC) (25). This reaction is widely used in biological research, yet systematic studies of reaction kinetics were lacking in the literature. The 2014 student cohort made two interesting discoveries: (1) surfactants can be used to catalyze SPAAC between hydrophobic azides and cyclooctynes; (2) the kinetics of SPAAC are highly tolerant of buffer identity, pH, and ionic strength, but are impacted by the presence of organic co-solvents. From this cohort, we have published two papers, with all students who contributed usable data listed as co-authors (26, 27). Building upon the results from the 2014 cohort, the 2015 curriculum explored SPAAC in the context of cyclooctyne-functionalized DNA. The students tackled questions regarding the effect of DNA sequence context on reaction rate, and the ability of the micelle catalysis phenomenon to be extended to cyclooctyne-functionalized biomolecules. At the conclusion of the course, I offer students the opportunity to continue with the research project as members of my laboratory. Of the ~50 students enrolled in the course, 4–5 have volunteered for this opportunity each year. This aspect of the course has proven to be highly successful, as it essentially doubles my laboratory’s capacity for hosting undergraduate research students. Importantly, this increased capacity is made possible by the fact that the students have already learned the techniques they need to carry out their research project. Additionally, having a sub-group of students continue the project has proven to be essential to our goal of publishing the research in peer-reviewed journals, as these students can repeat experiments as needed or gather additional data to expand on the most exciting discoveries.

Why did you decide to teach the course as a CURE? I was asked to teach this as a new course in my department, and therefore there was no existing curriculum that I could adopt. I decided that it would not be much more work to establish the course as a CURE instead of creating a “cookbook” 56 Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

curriculum, but that the benefits of the CURE format to the students could be significant. Also, I was excited by the prospect to use this course as a means to explore interesting scientific problems that aren’t a good fit for my research group because either: (1) we don’t have federal funding for the research, or (2) the study would require that similar experiments be repeated a large number of times, and this would become too mundane for one person to tackle alone. I have come to appreciate that while both of these factors make the research poorly suited for my research laboratory to pursue, they are ideal for a CURE, as the students can contribute information to the community that might not otherwise be gathered, and the need for a large number of similar experiments is very well-suited to the scale of a CURE.

What lessons did you learn and what changes did you make (or would you make in the future)? The main lesson I learned is that CUREs are not immune to the fact that research often doesn’t work on the first try and takes much longer than expected. I’ve now learned to set more reasonable goals regarding what can be accomplished in a term. I also learned that while the majority of students were highly invested in the idea of the course being research-based, some just wanted to get through the course so that they could graduate. The outcome of this is that while most students were careful with their experiments and turned in carefully acquired and analyzed data, some students produced data that were clearly flawed. This would normally have caused problems for our goal of publishing the research, but having a few dedicated students continue to work on the research during the following semester successfully averted this problem.

What do you view as the most important outcome from your CURE? Survey data from the course show that it increased students’ desire to pursue graduate studies or a career in the sciences, and students reported higher gains in research-related skills from the CURE compared to a “cookbook” lab course. Many of the students also commented to me on how excited they were to be able to contribute knowledge to the scientific community via publication. One student even told me that her mother printed out the journal article (on which the student was a co-author) and proudly displayed it on her refrigerator! Finally, showing that CUREs can lead to real research productivity – in the form of both data and publications – has increased the interest of my colleagues in adopting this curriculum format. Rory Waterman, University of Vermont, Department of Chemistry Course title: Synthetic Inorganic Laboratory Level: Senior Approximate enrollment: 12 students 57

Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Lecture/Laboratory Format One 4-hour afternoon laboratory section per week supported by the instructor and one TA.

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

Summary of the Course The course is meant to be an experience in experimental and synthetic inorganic chemistry, broadly defined. The outline for the course is that students build and characterize compounds that are then tested as catalysts for a particular reaction. To make the experience research, the process is repeated (compound selection, synthesis, characterization, and catalyst testing) with data from the previous iteration. The division of these efforts (synthesis, 2 weeks; characterization, 1 week; catalysis, 2 weeks) only allows for two iterations, but a third iteration would be possible and no less valuable. As preparation, the students are supplied with a limited set of initial resources for reading, some parameters for compounds, and a few suggestions for ligands. Our first two meetings cover some essential content relating to the course and the research problem as well as critical laboratory administration. The students are provided with the minimum requirements for characterization (two forms of spectroscopy and elemental analysis) and are supervised in the collection of electrochemical data for all compounds. The students then use their pure compounds in catalysis. The parameters for the catalysis (concentration, temperature, length of reaction, controls, etc.) were determined by consensus through a guided group discussion. Assessment arises from a range of components. Students write two significant reports, one after their first iteration of compounds and catalysis and the second at the end of the semester. The students also provide brief written pieces on proposed compounds and preparations, which allow for feedback on the choices made. The students also give brief presentations on their data. Because the students are working on the same problem, the mid-semester presentations are essential for the group to provide the students who have identified successful catalyst(s) feedback and suggestions, and it is important for the groups that did not identify active catalysts to garner ideas and receive help for their second round of compounds. Students are assessed on their laboratory notebooks as well. In this term, we have been investigating base metal catalysts for the dehydrogenation of ammonia borane. This is a good catalytic reaction because we can monitor activity by H2 evolution rather than spectroscopy, which makes scaling the class more straightforward. Their objective is to make one ligand and prepare three to five compounds as potential catalysts, which allows for some syntheses to fail, even after repeated effort but still have a candidate catalyst(s) to test. Their limitations are to use 3d transition metals and that their compounds should be air stable for convenience.

58 Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Why did you decide to teach the course as a CURE?

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

I wanted the students to gain more from the experience than a traditional laboratory. As seniors, I know they can follow a set of directions and make a single compound or set of compounds. What I sought was the tinkering that comes in research, but even more than that, I wanted students to do the thinking that was required. When problems come up (as they invariably do), I wanted students to realize that the solution would not be in a lab manual or a book but that the solution would only come from them.

What lessons did you learn and what changes did you make (or would you make in the future)? This iteration has provided two majors lessons, one about the students and one more mechanical. About the students, we’ve clearly done too good of a job teaching them to expect that laboratories are formulaic! Though the students had found and summarized preparations for their ligands and compounds, they had little conception of how they would actually work in laboratory in the absence of a lab manual. Thus, students arrived and floundered a bit initially. After the first lab session, my TA had the excellent idea of requesting work plans for each week from the students, and then providing feedback. The activity of summarizing their plans transformed the lab sessions in a short time into much more efficient and task oriented times. The mechanical lesson came from giving perhaps too much choice to the students for the design of their compounds, though some suggestions and constraints were provided. Many of them shot the moon intellectually. While this is validating from an education standpoint, it did not work well in the lab. From a practical view, I would have added yet more constraint on compound design to allow some creativity on the part of the students, but not obligate us to redesign any student’s plans.

What do you view as the most important outcome from your CURE? The students spent a lot of their time planning, thinking, and problem solving. There’s not much else I would want them to be doing!

Casey Londergan, Haverford College, Department of Chemistry Course title: Laboratory in Chemical Structure and Reactivity (“Superlab”) Level: Junior Approximate enrollment: 20–30 students

59 Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Lecture/Laboratory Format Two meetings per week for 4.5 hours, roughly including 1 hour of classroom activity and 3–3.5 hours of lab; senior undergrad TAs used as needed (mainly for instrumental experiments) and the departmental instrument technician is “on call” or training students throughout the course. The course runs for the entire junior year and is taught by four instructors, sometimes together and sometimes separately, depending on projects and research interests.

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

Summary of the Course “Superlab” is an integrated replacement for all of our majors-level lab courses (i.e. anything past organic chemistry). It was conceived and first executed in the late 1960s during the “instrumental revolution” in chemistry. The content of Superlab is highly variable year-to-year and depends on the instructors, who run some combination of 7- to 14-week projects in their areas of research expertise. The main learning goals of Superlab are process-based (experimental design, problem assessment, literature adoption, scientific writing) rather than skill-based. Besides the chemistry-only Superlab, similar courses and curricular structures exist at Haverford in Biology and in Biochemistry (where a Superlab module may be team-taught by faculty across departments). The 7- to 14-week integrated projects frequently include experiments and investigations motivated by current literature that have not been executed previously anywhere.

Why did you decide to teach the course as a CURE? Superlab was designed to be training for independent research; Haverford has a senior thesis requirement and all students do independent research as seniors. In our current context, many students are already involved in independent research as juniors, but this experience provides them with breadth of experience, collaborative opportunities with a much larger group, and process skills.

What lessons did you learn and what changes did you make (or would you make in the future)? Superlab changes every year and every semester, and instructors frequently run brand new projects in Superlab even when others have worked well before. One lesson from many years of this model is that training in research process, rather than specific skills or questions, is the most important outcome. Creativity and openness to trying new things are the best instructor attributes.

60 Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

What do you view as the most important outcome from your CURE? Training in research process, rather than skills, is the most important outcome. Our alumni frequently come back to tell us that “everything important that I learned as an undergraduate was in Superlab.”

References

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

1.

2.

3.

4. 5.

6.

7.

8.

9.

Engage to Excel: Producing One Million Additional College Graduates with Degrees in Science, Technology, Engineering, and Mathematics; President’s Council of Advisors on Science and Technology Report, 2012. Auchincloss, L. C.; Laursen, S. L.; Branchaw, J. L.; Eagan, K.; Graham, M.; Hanauer, D. I.; Lawrie, G.; McLinn, C. M.; Pelaez, N.; Rowland, S.; Towns, M.; Trautmann, N. M.; Varma-Nelson, P.; Weston, T. J.; Dolan, E. L. Assessment of Course-Based Undergraduate Research Experiences: A Meeting Report. CBE Life Sci. Educ. 2014, 13, 29–40. Shortlidge, E. E.; Bangera, G.; Brownell, S. E. Faculty Perspectives on Developing and Teaching Course-Based Undergraduate Research Experiences. BioScience 2016, 66, 54–62. Hanauer, D. I.; Graham, M. J.; Hatfull, G. F. A Measure of College Student Persistence in the Sciences (PITS). CBE Life Sci. Educ. 2016, 15, 1–10. Bascom-Slack, C. A.; Arnold, A. E.; Strobel, S. A. Student-Directed Discovery of the Plant Microbiome and Its Products. Science 2012, 338, 485–486. Brownell, S. E.; Hekmat-Scafe, D. S.; Singla, V.; Seawell, P. C.; Conklin Imam, J. F.; Eddy, S. L.; Stearns, T.; Cyert, M. S. A High-Enrollment CourseBased Undergraduate Research Experience Improves Student Conceptions of Scientific Thinking and Ability to Interpret Data. CBE Life Sci. Educ. 2015, 14, 1–14. Brownell, S. E.; Kloser, M. J.; Fukami, T.; Shavelson, R. Courses: Comparing the Impact of Traditionally Based “Cookbook” and Authentic Research-Based Courses on Student Lab Experiences. J. Coll. Sci. Teach. 2012, 41, 36–45. Harrison, M.; Dunbar, D.; Ratmansky, L.; Boyd, K.; Lopatto, D. Classroom-Based Science Research at the Introductory Level: Changes in Career Choices and Attitude. CBE Life Sci. Educ. 2011, 10, 279–286. Jordan, T. C.; Burnett, S. H.; Carson, S.; Caruso, S. M.; Clase, K.; DeJong, R. J.; Dennehy, J. J.; Denver, D. R.; Dunbar, D.; Elgin, S. C.; Findley, A. M.; Gissendanner, C. R.; Golebiewska, U. P.; Guild, N.; Hartzog, G. A.; Grillo, W. H.; Hollowell, G. P.; Hughes, L. E.; Johnson, A.; King, R. A.; Lewis, L. O.; Li, W.; Rosenzweig, F.; Rubin, M. R.; Saha, M. S.; Sandoz, J.; Shaffer, C. D.; Taylor, B.; Temple, L.; Vazquez, E.; Ware, V. C.; Barker, L. P.; Bradley, K. W.; Jacobs-Sera, D.; Pope, W. H.; Russell, D. A.; Cresawn, S. G.; Lopatto, D.; Bailey, C. P.; Hatfull, G. F. A Broadly Implementable Research Course in Phage Discovery and Genomics for First-Year Undergraduate Students. mBio 2014, 5, e01051-13. 61

Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

10. Kloser, M. J.; Brownell, S. E.; Chiariello, N. R.; Fukami, T. Integrating Teaching and Research in Undergraduate Biology Laboratory Education. PLoS Biol. 2011, 9, e1001174. 11. Rhode Ward, J.; Clarke, H. D.; Horton, J. L. Effects of a Research-Infused Botanical Curriculum on Undergraduates’ Content Knowledge, STEM Competencies, and Attitudes toward Plant Sciences. CBE Life Sci. Educ. 2014, 13, 387–396. 12. Sanders, E. R.; Hirsch, A. M. Immersing Undergraduate Students into Research on the Metagenomics of the Plant Rhizosphere: A Pedagogical Strategy to Engage Civic-Mindedness and Retain Undergraduates in STEM. Front. Plant Sci. 2014, 5, 157. 13. Henter, H. J.; Imondi, R.; James, K.; Spencer, D.; Steinke, D. DNA Barcoding in Diverse Educational Settings: Five Case Studies. Philos. Trans. R. Soc., B 2016, 371, 20150340. 14. Marcus, J. M.; Hughes, T. M.; McElroy, D. M.; Wyatt, R. E. Engaging FirstYear Undergraduates in Hands-On Research Experiences: The Upper Green River Barcode of Life Project. J. Coll. Sci. Teach. 2010, 39, 39–45. 15. Andreoli, J. M.; Feig, A.; Chang, S.; Welch, S.; Mathur, A.; Kuleck, G. A Research-Based Inter-institutional Collaboration to Diversify the Biomedical Workforce: ReBUILDetroit. Biomed. Central Proc. 2016under revision. 16. Lopatto, D. Survey of Undergraduate Research Experiences (SURE): First Findings. Cell Biol. Educ. 2004, 3, 270–277. 17. Scott, W. L.; Denton, R. E.; Marrs, K. A.; Durrant, J. D.; Samaritoni, J. G.; Abraham, M. M.; Brown, S. P.; Carnahan, J. M.; Fischer, L. G.; Glos, C. E.; Sempsrott, P. J.; O’Donnell, M. J. Distributed Drug Discovery: Advancing Chemical Education through Contextualized Combinatorial Solid-Phase Organic Laboratories. J. Chem. Educ. 2015, 92, 819–826. 18. Gee, C. T.; Koleski, E. J.; Pomerantz, W. C. Fragment Screening and Druggability Assessment for the CBP/p300 KIX Domain through Protein-Observed 19F NMR Spectroscopy. Angew. Chem., Int. Ed. 2015, 54, 3735–3739. 19. Roberts, R. J. Identifying Protein Function--A Call for Community Action. PLoS Biol. 2004, 2, E42. 20. Anton, B. P.; Chang, Y. C.; Brown, P.; Choi, H. P.; Faller, L. L.; Guleria, J.; Hu, Z.; Klitgord, N.; Levy-Moonshine, A.; Maksad, A.; Mazumdar, V.; McGettrick, M.; Osmani, L.; Pokrzywa, R.; Rachlin, J.; Swaminathan, R.; Allen, B.; Housman, G.; Monahan, C.; Rochussen, K.; Tao, K.; Bhagwat, A. S.; Brenner, S. E.; Columbus, L.; de Crecy-Lagard, V.; Ferguson, D.; Fomenkov, A.; Gadda, G.; Morgan, R. D.; Osterman, A. L.; Rodionov, D. A.; Rodionova, I. A.; Rudd, K. E.; Soll, D.; Spain, J.; Xu, S. Y.; Bateman, A.; Blumenthal, R. M.; Bollinger, J. M.; Chang, W. S.; Ferrer, M.; Friedberg, I.; Galperin, M. Y.; Gobeill, J.; Haft, D.; Hunt, J.; Karp, P.; Klimke, W.; Krebs, C.; Macelis, D.; Madupu, R.; Martin, M. J.; Miller, J. H.; O’Donovan, C.; Palsson, B.; Ruch, P.; Setterdahl, A.; Sutton, G.; Tate, J.; Yakunin, A.; Tchigvintsev, D.; Plata, G.; Hu, J.; Greiner, R.; Horn, D.; Sjolander, K.; Salzberg, S. L.; Vitkup, D.; Letovsky, S.; Segre, D.; 62

Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.

21.

22.

Downloaded by UNIV OF FLORIDA on December 22, 2017 | http://pubs.acs.org Publication Date (Web): October 24, 2017 | doi: 10.1021/bk-2017-1248.ch003

23.

24.

25.

26.

27.

DeLisi, C.; Roberts, R. J.; Steffen, M.; Kasif, S. The COMBREX Project: Design, Methodology, And Initial Results. PLoS Biol. 2013, 11, e1001638. Tong, W.; Wei, Y.; Murga, L. F.; Ondrechen, M. J.; Williams, R. J. Partial Order Optimum Likelihood (POOL): Maximum Likelihood Prediction of Protein Active Site Residues Using 3D Structure and Sequence Properties. PLoS Comput Biol 2009, 5, e1000266. Gray, C.; Price, C. W.; Lee, C. T.; Dewald, A. H.; Cline, M. A.; McAnany, C. E.; Columbus, L.; Mura, C. Known Structure, Unknown Function: An Inquiry-Based Undergraduate Biochemistry Laboratory Course. Biochem. Mol. Biol. Educ. 2015, 43, 245–262. Seymour, E.; Wiese, D. J.; Hunter, A.-B. In Creating a Better Mousetrap: On-Line Student Assessment of Their Learning Gains; National Meeting of the American Chemical Society, San Francisco, CA, 2000. Gottlieb, H. E.; Kotlyar, V.; Nudelman, A. NMR Chemical Shifts of Common Laboratory Solvents as Trace Impurities. J. Org. Chem. 1997, 62, 7512–7515. Sletten, E. M.; Bertozzi, C. R. Bioorthogonal Chemistry: Fishing for Selectivity in a Sea of Functionality. Angew. Chem., Int. Ed. 2009, 48, 6974–6998. Anderton, G. I.; Bangerter, A. S.; Davis, T. C.; Feng, Z.; Furtak, A. J.; Larsen, J. O.; Scroggin, T. L.; Heemstra, J. M. Accelerating Strain-Promoted Azide-Alkyne Cycloaddition Using Micellar Catalysis. Bioconjugate Chem. 2015, 26, 1687–1691. Davis, D. L.; Price, E. K.; Aderibigbe, S. O.; Larkin, M. X.-H.; Barlow, E. D.; Chen, R.; Ford, L. C.; Gray, Z. T.; Gren, S. H.; Jin, Y.; Keddington, K. S.; Kent, A. D.; Kim, D.; Lewis, A.; Marrouche, R. S.; O’Dair, M.; Powell, D. R.; Scadden, M. H. C.; Session, C. B.; Tao, J.; Trieu, J.; Whiteford, K. N.; Yuan, Z.; Yun, G.; Zhu, J.; Heemstra, J. M. Effect of Buffer Conditions and Organic Co-Solvents on the Rate of Strain-Promoted Azide-Alkyne Cycloaddition. J. Org. Chem. 2016, 81, 6816–6819.

63 Waterman and Feig; Educational and Outreach Projects from the Cottrell Scholars Collaborative Undergraduate and Graduate ... ACS Symposium Series; American Chemical Society: Washington, DC, 2017.