An Inorganic Chemistry Laboratory Course as Research - Journal of

Publication Date (Web): July 2, 2018 ... iteration), and this course is presented as one model from which individuals can develop CUREs at their own i...
0 downloads 0 Views 304KB Size
Article pubs.acs.org/jchemeduc

Cite This: J. Chem. Educ. XXXX, XXX, XXX−XXX

An Inorganic Chemistry Laboratory Course as Research Justin K. Pagano,*,†,§ Leslie Jaworski,‡ David Lopatto,‡ and Rory Waterman*,† †

Department of Chemistry, University of Vermont, Discovery Hall, 82 University Place, Burlington, Vermont 05404, United States Center for Teaching, Learning, and Assessment, Grinnell College, 102 Forum, 1119 Sixth Avenue, Grinnell, Iowa 50112, United States



Downloaded via KAOHSIUNG MEDICAL UNIV on July 2, 2018 at 22:16:56 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

S Supporting Information *

ABSTRACT: A research-based inorganic chemistry laboratory course is described. Using the defined protocol of a course-based undergraduate research experience (CURE), students undertake a self-designed research project to identify new catalysts for the dehydrogenation of ammonia borane. Students select ligands and metals, prepare and characterize catalysts precursors, develop protocols for catalysts, and screen catalyst precursors. These activities are designed to align with the five components of a CURE (research activities, discovery, relevance, collaboration, and iteration), and this course is presented as one model from which individuals can develop CUREs at their own institutions. The students were assessed with the CURE survey, which showed gains in both research activities and understanding the nature of science.

KEYWORDS: Upper-Division Undergraduate, Inorganic Chemistry, Curriculum, Hands-On Learning/Manipulatives, Inquiry-Based/Discovery Learning, Catalysis, Main-Group Elements, Metals

I

example, but the course is presented as a framework that could be adopted by any instructor.

t is widely accepted that research, as a high-impact practice, should be provided to as many undergraduate students as possible to afford positive learning outcomes and aid in meeting workforce objectives.1−4 Mentored research remains a mainstay intervention, but large numbers of STEM students have prompted innovative teacher−scholars to develop programs that provide research experiences for thousands of undergraduate students from the beginning of the curriculum, including the Center for Authentic Science Practice in Education (CASPiE), Freshman Research Institute (FRI), and Research Experiences to Enhance Learning (REEL), among others. What is already known about these large-scale efforts is that many of the benefits of mentored research experiences are realized,5 which is also consistent with data on smaller-scale research experiences contained within the curriculum (e.g., CURE, course-based undergraduate research experience).6 The development of CUREs has been successful, but adoption has been greater in the life sciences than in the physical sciences.7,8 There are an increasing number of CUREs in the chemical literature,9−20 but within chemistry, there is still limited awareness of many instructional practices,21 CUREs among them. Examples of both protocols22 as well as published data will improve awareness and thereby foster adoption of this practice.23 Beyond the benefit to students,5,6 it has also been suggested that cotaught CUREs provide excellent preservice professional development for graduate students or postdocs.8 This report outlines a CURE in inorganic chemistry with student data (n = 12) from fall 201624 as a specific © XXXX American Chemical Society and Division of Chemical Education, Inc.



LABORATORY MOTIVATION While interest in amine−boranes as precursors to main group polymers began more than 15 years ago, investigation of ammonia borane as a hydrogen storage material has ballooned over the past decade because of its relative stability and high weight percent of hydrogen.25 Ammonia and other amine boranes will liberate hydrogen under thermal conditions, but catalysts that provide controlled H2 release have been sought after for some time (eq 1).26−30 The discovery and understanding of simple, bench-stable catalysts31 for hydrogen evolution from ammonia borane will advance this molecular agent in hydrogen storage applications. catalyst

H3B−NH3 ⎯⎯⎯⎯⎯⎯→ nH 2 + B−N products

(1)

Course Objectives

The Department of Chemistry at UVM identified a set of learning objectives for students pursing BA and BS degrees with majors in chemistry.32 Among these goals, eight can be addressed through laboratory work and yet more fully explored through chemical research. In particular, the department places a specific premium on informed experimentation. Indeed, the specific learning outcome notes that students should be able to Received: October 22, 2017 Revised: May 15, 2018

A

DOI: 10.1021/acs.jchemed.7b00812 J. Chem. Educ. XXXX, XXX, XXX−XXX

Journal of Chemical Education



“describe the objective of a chemical experiment, execute the experiment correctly, and collect and analyze relevant data”.32 Traditional laboratory experiments can address this kind of objective, but such work does not drive at the understanding that faculty would like students to derive from academic laboratory work. When students design and explain their choices in experimentation, it is much easier to assess this objective and ensure that it is an effective educational experience. An additional objective is that students “adhere to procedures and regulations for the safe handling, use, and disposal of chemical reagents”.32 At minimum, this objective reflects that students can follow directions. When students are choosing reagents and producing potentially unique products and byproducts, they are addressing this objective more thoroughly, and safe handing is a highly transferable skill for developing chemists regardless of discipline. The undergraduate chemistry curriculum at UVM is in a period of change, and this laboratory “Inorganic Synthesis Laboratory” arose as a temporary course to fill a period in which certain objectives, synthesis in inorganic chemistry broadly defined, experimental observation of inorganic concepts, and additional exposure to spectroscopic techniques were not otherwise being met. The course was a clean slate from which to build a CURE. The protocol generated is shared herein as a framework meant to address some of the positive features of CUREs for both students and faculty.33−35

EXPERIMENTAL DESIGN To create an iterative cycle for the course (Table 1, item 5), a basic three-step procedure was implemented. The steps are (1) selection of target ligands and transition-metal compounds for preparation, (2) preparation and characterization of ligands and transition-metal compounds, and (3) evaluation of catalysts and controls through trial runs. For a weekly laboratory class, this cycle was completed twice to satisfy the need for iteration in the research process,6 though certain specific activities (e.g., catalytic runs and data collection) were repeated many times. Step 1: Selection of Target Ligands and Transition-Metal Compounds

Students were provided with a minimum set of criteria to choose ligands and metal compounds. On the basis of space constraints, we were unable to store air-sensitive compounds. For intellectual reasons, catalysis with earth-abundant metals is ideal.36 Students were directed to choose any 3d transition metal or metals and at least two ligands. One ligand needs to be a synthetic target (i.e., not commercial), and the other(s) could be a commercial reagent (e.g., pyridine). With the metal(s) and ligands, students needed to target between three and five transition-metal compounds for preparation. Literature compounds were encouraged to provide reliable data for isolation and accurate characterization. Students were given this prompt after two meetings that delivered central content for the laboratory course, including basic organometallic chemistry, catalysis, and a review of techniques. The lists of ligands, metals, and compounds were shared as “compound reports” that provided the selection, rationale, and references. These reports met two needs. First, they afforded opportunity to provide feedback to students on the choice of ligands and quality of their proposed syntheses. Second, they met the pragmatic need to stock necessary reagents for the proposed syntheses. After one iteration, it was clear that students were not fully conceptualizing their actual syntheses in advance. In the same semester and thereafter, compound plans were modified to include detailed experimental plans for each reaction, which also helped address potential hazards. After some feedback and approval of compound plans, students were prepared to conduct their syntheses. Replication. At the senior level, this step can be used exactly at suggested, but modifications are possible. Other laboratories may provide a set of specific ligands and avail any metal to allow for the preparation of new compounds, or the metal(s) may be specified with ligands as a choice. The ability to choose new ligands or design metal compounds is a higherorder activity. Students in lower-division laboratories may be provided with targets and be expected to identify preparations from literature searches.



PROCEDURE The course was built around supporting students engaging in a research experience. The research project was the synthesis of transition-metal compounds as potential catalysts for the dehydrogenation of ammonia borane. Because the course is for seniors, participant students are generally well-equipped to provide substantial input into the design and execution of the project. Therefore, class discussion was used to decide on specific details for design, data collection, and analysis. Fitting the CURE Model

By definition, CUREs are real research, and it has been noted that genuine research experiences have five common traits (Table 1).6 An experiment or course must demonstrate all of Table 1. Properties of a CURE (Course-Based Undergraduate Research Experience)6 and Relationship to This Laboratory Item 1

CURE Feature

2 3

Research activities Discovery Relevance

4

Collaboration

5

Iteration

Article

How Addressed Experimental design and execution, data collection, data analysis, and use of current instrumentation Identification of new catalysts Context: Hydrogen storage is a limitation in adoption of H2 as fuel. Students worked in pairs, and the class shared and considered all data. Cycle of compound preparation, catalyst screening, and analysis were repeated.

Step 2: Preparation and Characterization of Ligands and Transition-Metal Compounds

Students used commercial reagents and executed any necessary purifications themselves as part of their syntheses. Dry, air-free solvents were provided as needed. Students characterized their products by appropriate means, most commonly multinuclear NMR spectroscopy and infrared spectroscopy, but students also routinely utilized UV−vis spectroscopy and had electrochemistry and elemental analysis available. Though the compounds and the methods for assessing identity and purity were the choice of the students, the execution of this section of the course had some resemblance to a traditional, prescribed

these traits to be considered research and thereby a CURE. Therefore, careful consideration of the specific activities and how all five features would be addressed was made in course design. Table 1 presents a summary of these activities. Additional detail of the activities follow. B

DOI: 10.1021/acs.jchemed.7b00812 J. Chem. Educ. XXXX, XXX, XXX−XXX

C

Lab session 3

Lab session 4

Lab session 5 Presentations and shared results Lab session 6

5

6

7 8

Lab session 7 Lab session 8

Lab session 9

Lab session 10 Presentations No meeting

10 11

12

13 14 15

9

Lab session 1 Lab session 2

3 4

Description

Assignment(s) (Outcome/Observations) Required readings (Discussions in meeting 1 indicated high compliance with preclass readings.)

Compound plan 1 due (Plans were returned within approximately 2 days with feedback on proposed syntheses and safety.) Compound plans returned Initial compound plans lack sufficient detail in both syntheses and safety. Students were contacted with exact details for revisions prior to Lab session 1. Instructions for compound plans were revised accordingly. Chemistry Start Preparation of ligands All groups started on ligand syntheses and made progress Ligand preparation continued, start of metal compound preparation, Lab notebook check (Notebooks were of high quality, a result of emphasis in prior courses; all groups continued characterization of compounds to make good progress in ligand preparation.) Metal compound preparation and characterization continued, some groups started catalytic runs Any lingering preparation and characterization completed, all groups running Groups reminded to share data.f catalysis All groups completing catalysis Groups appeared to wait until this lab was complete to start report and presentation work. Report 1/Presentation 1 (Reports tended to lack comparison to group data and literature. Students were reminded of this aspect for the final report/presentation. The students independently created a class-wide plan for compounds in iteration 2.) Preparation of ligands, start of metal compound preparation Compound plan 2 (Compounds plans could be submitted on the day experimentation started because the presentation of data allowed the class to make decisions, which afforded time for planning.) Metal compound preparation, characterization of compounds Groups reminded to share data.f Metal compound preparation and characterization continued, catalytic runs started Any lingering preparation and characterization completed, all groups running catalysis All groups completing catalysis Groups appeared to wait until this lab was complete to start report and presentation work. Final presentations Final reports (Final reports had a stronger degree of comparison against results from other students and the literature as compared to the initial reports.)

Basic description of the course, syllabus, link to CURE presurvey, citations for required readings Organometallic overview,c catalysis overview,d discussion of ammonia borane dehydrcoupling,e constraints of project, description of assignment, review of safety plans and contracts Discussion of catalytic reaction protocol, NMR and e-chem training

See Supporting Information for the e-mail content. bLaboratory session was scheduled for 4 h. cOverview included oxidation state, electron counting, periodic trends, basic bonding, and simple organometallic reaction mechanisms (e.g., oxidative addition). dOverview included impact of catalysis on rate, catalysis nomenclature, catalytic cycles, and metrics of catalysis (e.g., turnover number). e Discussion included H2 storage needs and options, properties of ammonia borane, and brief review of catalysts from the literature. fStudents elected to share via Dropbox with tabulated data shared via Google. Campus learning software has the same capacities, but students chose these platforms.

a

Meeting 2

2

2

Meeting 1b

1

Event

E-maila

0

Week

Table 2. Week-by-Week Description of the Course

Journal of Chemical Education Article

DOI: 10.1021/acs.jchemed.7b00812 J. Chem. Educ. XXXX, XXX, XXX−XXX

Journal of Chemical Education



RESULTS The overall plan for the course, assignments, and instructor feedback on activities is presented in Table 2. The course was 14 4 h meetings. The first two meetings were utilized for intellectual and laboratory preparation, two were utilized for student presentations, and the balance were used for research work (preparation, characterization, and catalysis). Students were directed to use 3d metals and prepare compounds that could be isolated as air-stable products. In some cases, compound plan evaluation included redirecting students on what compounds would likely be air stable if not described in the literature report. For synthesized ligands, students were encouraged to identify simple ligands with highly effective reaction steps (e.g., condensations). Therefore, many groups synthesized chelating imines such as diimines and diiminopyridines as ligands. Most ligands were prepared and purified in two lab periods; metal compounds were prepared in similar or less time. Commercial ligands that yielded known LyMXz derivatives were pyridine or simple phosphines like triphenylphosphine. Some groups asked to start reactions earlier (i.e., a multihour reflux) to work up the reaction during the lab period or work up reactions outside of the lab period. All request were granted by arriving at a mutually agreeable time with the instructor. An instructor replicating or adapting this model could limit students to ligands and/or compounds by length of synthesis. In both 2016 (n = 12) and 2017 (n = 16), student gravitated toward a cobalt compound in a class-wide collaboration for the second iteration, though iron, nickel, and copper compounds were proposed and tested. In general, the compounds showed good activities, sufficient that results of the 2016 and 2017 cohorts are being compiled into a publication. The instructor has been collating the data and writing the main text during time beyond the course. Some instructors may profit from this model as a boost to their research productivity.7 Simple metal salts are generally available in the department’s stockroom as well as many basic reagents and all solvents. UVM has good turnaround for orders from major commercial suppliers. However, backordered or expensive compounds prompted a guided discussion on alternatives that an individual group could undertake. In two years, only one such instance occurred. Giving the students genuine carte blanche for ligand selection can result in expensive and untenable syntheses. If an instructor is constrained, it is easy to envision a strategy of prompting students to build ligands from a common scaffold or variants in a family (e.g., electron withdrawing versus donating or sterically encumbered versus not). Student primary data for the catalysis were primarily volume measurements over time. Catalysis conditions (1 h total reaction time, 5 mol % catalyst loading, concentration, temperature, etc.) were determined by student agreement after in-class discussion. The products of catalysis were determined primarily by 11B NMR spectroscopy, which was collected from the reaction mixture (THF solution) at ambient temperature. Student data are part of a forthcoming publication and not presented in detail here to avoid plagiarizing work for which students will receive acknowledgment as coauthors. Every group’s pure compounds exhibited catalytic activity that was better than that of the metal salt (typically cobalt chloride) alone. Interestingly, control reactions were critical, and basic issues such as solvent dryness needed to be investigated in detail. Sharing of digital data and

laboratory in that students created their own prescriptive pathway in advance with instructor feedback. Replication. The decisions made in Step 1 directly impact this step. Step 3: Evaluation of Catalysts and Controls through Trial Runs

The target of three to five transition-metal compounds was deliberate to ensure that each team would have at least two compounds to screen as potential catalysts, given that some syntheses would fail or some compounds would not be isolated in sufficient purity. During the initial set of synthetic preparations, the class took pause to discuss the conditions for catalytic reactions. That discussion led to a consensus set of conditions including concentration, catalyst loading, temperature, time, number of replications, etc., which were also informed by examples of this kind of catalysis from the literature. The students also determined a set of controls to run and a division of labor such that all controls were run and universal controls (e.g., no catalyst or metal salts without ligands) were run by several groups for comparison. Catalytic activity was measured by collecting and measuring the volume of hydrogen over time. The students deposited data (spreadsheets of time and volume entries) electronically such that all groups had access for analysis. Spectroscopic data (commonly 11 B NMR spectra of products) were also deposited to a central, shared location. Replication. Seniors have been able to synthesize the literature sufficiently to arrive at a reasonable set of conditions, recognize that sufficient controls are needed, and determine the full suite of experimental parameters that needed to be determined before starting. For lower-division courses, catalysis testing conditions can be supplied or more prompting can afford productive discussion. It is critical that a common set of conditions must be used to allow for comparison between groups. After an initial set of conditions were derived (first course offering), a discussion of conditions was held and students developed a set of conditions for the catalysis addressing features including temperature, concentration, reaction time, monitoring frequency, etc. Exact conditions were provided to the students with the explanation that the given conditions allow for comparable data between course offerings. The difference between the conditions derived by the students was not terribly substantial, but the homogeneity promotes comparisons between sections. Following each iteration, each group engaged in reporting and analysis of their and the entire lab’s data. The students shared that analysis in two forms, a presentation by the group to the lab and an individually written report. The purpose of the presentations was to prompt a class-wide discussion on the class results, which fueled the group’s choices for their next compounds.



Article

HAZARDS

Hazards vary with the specific compound and preparation, though the most common hazards were flammable and toxic reagents. For this work, investigating the catalytic dehydrogenation of ammonia borane, the key hazard is the evolution of hydrogen gas. Reaction were conducted in a fume hood, and students wore appropriate PPE (personal protective equipment). Catalytic reactions were prepared in a fume hood and run either in a fume hood or on a bench, and H2 was collected over water. Trapped H2 was vented into a fume hood. D

DOI: 10.1021/acs.jchemed.7b00812 J. Chem. Educ. XXXX, XXX, XXX−XXX

Journal of Chemical Education

Article

collected, except in the student evaluations, which did not directly address this point. However, two students specifically commented that they liked the “freedom” and “control” of the project in their course evaluations. Grades ranged from B to A in 2016 and C to A in 2017, which may be a function of the lager class size in 2017 (12 vs 16). The iterative nature of the course afforded substantially better second-round assignments. Also, students in 2016 were not penalized for the lack of specificity in compound plan request, and first compound plans in 2017 reflected better communication of the objectives on our part. From a research standpoint, all were successful. Each group was able to prepare and isolate at least one compound per iteration that was tested in catalysis. In the presentations, all students spoke articulately about the nature of the group’s choice for ligand and metal as well as the rationale for all catalysis conditions. However, it seems likely that even better gains would be realized by more specific instruction on the process and nature of research (e.g., Entering Research38).

archiving online allows for class data to be treated appropriately under data management protocols. Students completed the pre- and post-CURE survey developed by Lopatto that provides an assessment of research experience. These data were collected to evaluate this laboratory experience as a CURE and are therefore more relevant to this report than student data, which vary by the compound selected and are the subject of a forthcoming report. Survey data of this type have been used to evaluate the impact of research experiences.37



DISCUSSION Students completed individual laboratory reports that incorporated data from the entire class. These reports were completed after each iteration. The pairs of students each presented on their results and data while also including their analysis of the group’s shared data. Like the reports, presentations were made after each iteration. Anecdotally, student motivation was high. Many students sought to return to lab outside the class period to complete syntheses, make additional compounds, complete characterization, or run additional catalysis experiments. Discussion during presentations was lively and informed. Students at least appeared to have ownership of their projects. CURE survey data suggest that students had many positive outcomes from the course. Importantly, students reported strong gains in areas that overlap with departmental objectives explicitly (e.g., collect data, analyze data) or implicitly (e.g., understand how scientists work on real problems; Table 3).



CONCLUSIONS We present herein a simple model for a CURE laboratory in inorganic chemistry, of which there are few examples in the literature. While CUREs can present various complexities with respect to development and execution, the aim of this paper is to provide a model that can be adapted. Data from this course suggest that implementation of a CURE in inorganic chemistry has similar positive student effects as compared to those in other fields.



Table 3. Selected Self-Reported Learning Gains from the CURE Survey Data for the 2016 Cohorta Question

Score

Ability to analyze data and other information Learning laboratory techniques Ability to read and understand primary literature Understanding how scientists work on real problems Understanding how knowledge is constructed

3.27 3.55 3.09 3.82 3.55

S Supporting Information *

The Supporting Information is available on the ACS Publications website at DOI: 10.1021/acs.jchemed.7b00812. Syllabus, including descriptions of assessments, 2016 CURE Survey data omitting demographic information that may identify students, and sample compound plans (PDF)



a

The scale is 1 (strongly disagree) to 5 (strongly agree).

Overall, the student outcome data from the CURE survey support the notion that this course helped students meet learning objectives for their degree in chemistry (Table 4). Strong agreement with “this course was a good way of learning about” both subject and scientific research (4.55 and 4.82 out of 5.00, respectively) suggests that the perception of student interest in the course was genuine. The 2016 report, without potentially identifying demographic data, is presented in the Supporting Information. Written comments were not

Science is essentially an accumulation of facts, rules, and formulas I can do well in science courses Real scientists do not follow the scientific method in a straight line Explaining science ideas to others has helped me understand the ideas better

*E-mail: [email protected]. *E-mail: [email protected]. ORCID

Rory Waterman: 0000-0001-8761-8759 Present Address §

Chemistry Division, Los Alamos National Laboratory, Mail Stop J514, Los Alamos, New Mexico 87545. Notes

The authors declare no competing financial interest.



Precourse Postcourse 3.17

2.70

4.17 3.08

4.27 3.82

4.50

4.64

AUTHOR INFORMATION

Corresponding Authors

Table 4. Selected Results from Questions about Attitudes toward Science on the CURE Survey for the 2016 Cohorta Question

ASSOCIATED CONTENT

ACKNOWLEDGMENTS The student work was supported by the University of Vermont. We are grateful for the confidence of the Department of Chemistry chair, Chris Landry, in running this laboratory as a CURE and to Bill Geiger for his insight during planning. Part of this publication was supported by the U.S. National Science Foundation through CHE-1565658 to R.W. and by Research Corporation for Science Advancement.

a

The scale is 1 (strongly disagree) to 5 (strongly agree). E

DOI: 10.1021/acs.jchemed.7b00812 J. Chem. Educ. XXXX, XXX, XXX−XXX

Journal of Chemical Education

Article

(19) Winkelmann, K.; Baloga, M.; Marcinkowski, T.; Giannoulis, C.; Anquandah, G.; Cohen, P. Improving Students’ Inquiry Skills and Self-Efficacy through Research-Inspired Modules in the General Chemistry Laboratory. J. Chem. Educ. 2015, 92, 247. (20) Smith, T. L.; Gillmore, J. G.; Scogin, S. C. Incorporating Authentic Research in an Optional Component of the Second Semester Organic Laboratory Course. Chem. Educ. 2017, 22, 177. (21) Baker, L. A.; Chakraverty, D.; Columbus, L.; Feig, A. L.; Jenks, W. S.; Pilarz, M.; Stains, M.; Waterman, R.; Wesemann, J. L. Cottrell Scholars Collaborative New Faculty Workshop: Professional Development for New Chemistry Faculty and Initial Assessment of Its Efficacy. J. Chem. Educ. 2014, 91, 1874. (22) An excellent resource for examples of CUREs is the database of projects at the CUREnet Web page, https://curenet.cns.utexas.edu/. (23) Henderson, C.; Cole, R.; Froyd, J.; Khatri, R. Five Claims about Effective Propagation; 2012. (24) Fall 2016 was the first semester in which students completed the course and the CURE survey. (25) Staubitz, A.; Robertson, A. P. M.; Sloan, M. E.; Manners, I. Amine− and Phosphine−Borane Adducts: New Interest in Old Molecules. Chem. Rev. 2010, 110, 4023. (26) Staubitz, A.; Robertson, A. P. M.; Manners, I. Ammonia-Borane and Related Compounds as Dihydrogen Sources. Chem. Rev. 2010, 110, 4079. (27) Hamilton, C. W.; Baker, R. T.; Staubitz, A.; Manners, I. B-N compounds for chemical hydrogen storage. Chem. Soc. Rev. 2009, 38, 279. (28) Waterman, R. Mechanisms of metal-catalyzed dehydrocoupling reactions. Chem. Soc. Rev. 2013, 42, 5629. (29) Smythe, N. C.; Gordon, J. C. Ammonia Borane as a Hydrogen Carrier: Dehydrogenation and Regeneration. Eur. J. Inorg. Chem. 2010, 2010, 509. (30) Rossin, A.; Peruzzini, M. Ammonia−Borane and Amine− Borane Dehydrogenation Mediated by Complex Metal Hydrides. Chem. Rev. 2016, 116, 8848. (31) Pagano, J. K.; Stelmach, J. P. W.; Waterman, R. Cobaltcatalyzed ammonia borane dehydrocoupling and transfer hydrogenation under aerobic conditions. Dalton Trans. 2015, 44, 12074. (32) See Supporting Information. (33) Corwin, L. A.; Graham, M. J.; Dolan, E. L. Modeling CourseBased Undergraduate Research Experiences: An Agenda for Future Research and Evaluation. CBE Life Sci. Educ. 2015, 14, es1. (34) Brownell, S. E.; Kloser, M. J.; Fukami, T.; Shavelson, R. Undergraduate biology lab courses: Comparing the impact of traditionally-based ‘cookbook’ and authentic research-based courses on student lab experiences. J. Coll. Sci. Teach. 2012, 41, 36. (35) Shortlidge, E. E.; Bangera, G.; Brownell, S. E. Faculty Perspectives on Developing and Teaching Course-Based Undergraduate Research Experiences. BioScience 2016, 66, 54. (36) Schafer, L. L.; Mountford, P.; Piers, W. E. Earth abundant element compounds in homogeneous catalysis. Dalton Tans 2015, 44, 12027. (37) Denofrio, L. A.; Russell, B.; Lopatto, D.; Lu, Y. Linking Student Interests to Science Curricula. Science 2007, 318, 1872. (38) Branchaw, J.; Pfund, C.; Rediske, R. Entering Research: A Facilitator’s Manual; W. H. Freeman and Co.: Basingstoke, England, 2010.

The CURE survey is supported by the Howard Hughes Medical Institute.



REFERENCES

(1) Strengthening Research Experiences for Undergraduate STEM Students; National Academies of Sciences, Engineering, and Medicine, 2017. (2) Kuh, G. D. High-Impact Educational Practices: What They Are, Who Has Access to Them, and Why They Matter; Association of American Colleges & Universities: Washington, DC, 2008. (3) President’s Council of Advisors on Science and Technology. Engage to Excel: Producing One Million Additional College Graduates with Degrees in Science, Technology, Engineering, and Mathematics; 2012. (4) Project Kaleidoscope. Report on Reports: Recommendations for Action in Support of Undergraduate Science, Technology, Engineering and Mathematics; 2002. (5) Brownell, S. E.; Hekmat-Scafe, D. S.; Singla, V.; Chandler Seawell, P.; Conklin Imam, J. F.; Eddy, S. L.; Stearns, T.; Cyert, M. S. A High-Enrollment Course-Based Undergraduate Research Experience Improves Student Conceptions of Scientific Thinking and Ability to Interpret Data. CBE Life Sci. Educ. 2015, 14, ar21. (6) 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. (7) Heemstra, J. M.; Waterman, R.; Antos, J. R.; Beuning, P.; Bur, S.; Columbus, L.; Feig, A. L.; Fuller, A.; Gillmore, J. G.; Leconte, A.; Pomerantz, A.; Prescher, J.; Stanley, L. L. In Educational and Outreach Projects from the Cottrell Scholars Collaborative; Waterman, R., Feig, A. L., Eds.; ACS Symposium Series: Washington, DC, 2017. (8) Cascella, B.; Jez, J. M. Beyond the Teaching Assistantship: CURE Leadership as a Training Platform for Future Faculty. J. Chem. Educ. 2018, 95, 3. (9) Clark, T. M.; Ricciardo, R.; Weaver, T. Transitioning from Expository Laboratory Experiments to Course-Based Undergraduate Research in General Chemistry. J. Chem. Educ. 2016, 93, 56. (10) Kerr, M. A.; Yan, F. Incorporating Course-Based Undergraduate Research Experiences into Analytical Chemistry Laboratory Curricula. J. Chem. Educ. 2016, 93, 658. (11) Slade, M. C.; Raker, J. R.; Kobilka, B.; Pohl, N. L. B. A Research Module for the Organic Chemistry Laboratory: Multistep Synthesis of a Fluorous Dye Molecule. J. Chem. Educ. 2014, 91, 126. (12) Boyd-Kimball, D.; Miller, K. R. From Cookbook to Research: Redesigning an Advanced Biochemistry Laboratory. J. Chem. Educ. 2018, 95, 62. (13) Graham, K. J.; Schaller, C. P.; Johnson, B. J.; Klassen, J. B. Student-Designed Multistep Synthesis Projects in Organic Chemistry. Chem. Educ. 2002, 7, 376. (14) Iimoto, D. S.; Frederick, K. A. Incorporating Student-Designed Research Projects in the Chemistry Curriculum. J. Chem. Educ. 2011, 88, 1069. (15) Keller, V. A.; Kendall, B. L. Independent Synthesis Projects in the Organic Chemistry Teaching Laboratories: Bridging the Gap Between Student and Researcher. J. Chem. Educ. 2017, 94, 1450. (16) May, N. W.; McNamara, S. M.; Wang, S.; Kolesar, K. R.; Vernon, J.; Wolfe, J. P.; Goldberg, D.; Pratt, K. A. Polar Plunge: Semester-Long Snow Chemistry Research in the General Chemistry Laboratory. J. Chem. Educ. 2018, 95, 543. (17) Tomasik, J. H.; Cottone, K. E.; Heethuis, M. T.; Mueller, A. Development and Preliminary Impacts of the Implementation of an Authentic Research-Based Experiment in General Chemistry. J. Chem. Educ. 2013, 90, 1155. (18) Williams, L. C.; Reddish, M. J. Integrating Primary Research into the Teaching Lab: Benefits and Impacts of a One-Semester CURE for Physical Chemistry. J. Chem. Educ. 2018, 95, 928. F

DOI: 10.1021/acs.jchemed.7b00812 J. Chem. Educ. XXXX, XXX, XXX−XXX