Developing an Integrated Research-Teaching Model - ACS Publications

Publication Date (Web): May 2, 2018 .... and initial literature leads are also posted to the course website, allowing the students time choose a topic...
0 downloads 0 Views 710KB Size
Chapter 6

Developing an Integrated Research-Teaching Model

Downloaded by MIAMI UNIV on May 13, 2018 | https://pubs.acs.org Publication Date (Web): May 2, 2018 | doi: 10.1021/bk-2018-1275.ch006

Robert E. Bachman* Department of Chemistry, The University of the South, 735 University Avenue, Sewanee, Tennessee 37383, United States *E-mail: [email protected].

The goal of this education project is to convert a traditional skill-building laboratory sequence within an inorganic chemistry course into an authentic research experience. This approach has developed a natural teaching-research nexus that provides students with valuable intellectual growth and faculty with recognition for both their teaching and scholarship. Given that research experiences involve an integrated intellectual exploration, students in the course learn an array of valuable skills, such as literature research, literature reading, teamwork, and scientific communication, and both synthetic and analytical techniques. The faculty member teaching the course has gained new leads for their overall research program and initial results that have even led to public presentation.

Introduction Over the last four decades, faculty across higher education have come to recognize that Undergraduate Research (UR) experiences are an advantageous educational touchstone for undergraduate students. One of the earliest recognitions of this learning paradigm was the launching of the Council on Undergraduate Research (CUR) in 1978 by a small community of chemistry leaders. CUR developed one of the earliest definitions of undergraduate research: “An inquiry or investigation conducted by an undergraduate student that makes an original intellectual or creative contribution to the discipline.” Interestingly, the American Chemical Society Committee on Professional Training describes a similar definition but also notes the importance of faculty mentorship/teaching © 2018 American Chemical Society Gourley and Jones; Best Practices for Supporting and Expanding Undergraduate Research in Chemistry ACS Symposium Series; American Chemical Society: Washington, DC, 2018.

Downloaded by MIAMI UNIV on May 13, 2018 | https://pubs.acs.org Publication Date (Web): May 2, 2018 | doi: 10.1021/bk-2018-1275.ch006

providing intellectual growth of students that are engaged in research (1). The educational value of undergraduate research has been examined extensively, especially by Lopatto (2, 3). As part of the overall expansion of UR activity, many faculty at Primarily Undergraduate Institutions (PUIs) moved from focusing their work almost entirely on teaching toward a more even balance of teaching and research efforts. This transition was worthwhile because it produced a significant amount of valuable scientific knowledge and provided many young potential scientists the opportunity to gain research experience earlier in their educational career. Increased national recognition of UR as a “high impact practice” (4) has led faculty and administrators at large research-intensive doctoral institutions to develop UR opportunities for students, in many cases these schools have launched a campus-wide UR office and program. While the overall recognition of UR has provided many students an important opportunity, the growth of this practice has reached a capacity problem in many settings. At the larger schools, it is frequently obvious that all students in a chemistry program cannot find a home in a faculty member’s research group. Even at smaller schools, this capacity problem can occur as the number of majors grows. At one point, I attempted to manage eight students with effectively different small projects in a given semester. Not surprisingly, this approach was challenging for the students and myself. Scheduling the individual learning time needed for each student was the most significant challenge. The students often had to attempt a procedure or technique for the first time with only a short introductory explanation. Individually teaching and training so many students created a natural stress in terms of balancing teaching and scholarship goals, especially since the mentorship/supervision of research students is not part of the teaching load at Sewanee. At a few institutions, the acknowledgement of work balance related to UR mentorship has provided some level of load compensation; however, faculty at many institutions routinely support UR without official compensation because their research program benefits from having research assistants and because they find internal contentment as a mentor to future scientists. Given the obvious capacity problems and the related work stress, I began to explore how the traditional segmented landscape of courses and research could be bridged in the inorganic chemistry course (Figure 1), which contains both three lecture blocks and a weekly afternoon laboratory. I hoped that formally joining these two realms would decrease the perceived divide between class-related knowledge and “real” science felt by many students and faculty, increase student UR capacity, and decrease faculty workload stress. As an added benefit, this integrated approach would provide a way to naturally link the “High Impact Practice” (HIP) (4–6) of UR to other HIPs such as collaborative projects, capstone experiences, and writing-intensive (WI) communication. The inorganic course was already designated as a required major-level writing-intensive course so it was expected that integrated a research experience into the writing process would provide a variety of communication opportunities. For example, students would be able to present initial research as a poster at our campus research symposium and develop a journal style manuscript reporting their progress on a research question. 84 Gourley and Jones; Best Practices for Supporting and Expanding Undergraduate Research in Chemistry ACS Symposium Series; American Chemical Society: Washington, DC, 2018.

Downloaded by MIAMI UNIV on May 13, 2018 | https://pubs.acs.org Publication Date (Web): May 2, 2018 | doi: 10.1021/bk-2018-1275.ch006

Figure 1. A visual approach of how research and teaching can be integrated.

In the research milieu, a variety of oral communication learning experiences are available, such as in-lab idea sharing, collaborative group-meeting research conversations, and discussing their research progress via a poster in a public venue.

Course Development The inorganic chemistry laboratory began originally with traditional “cookbook” laboratories based on well-tested and repetitive experiments. This time-honored approach began with classic experiments in areas such as coordination complex syntheses (synthesis of cis/trans isomers of [Co(en)2Cl2], reactivity studies (nitro to nitrito isomerization), and solid-state synthesis (1-2-3 Superconductor) (7–10). An early step toward the research approach began with labeling the course as a major-related WI course; students at Sewanee are required to complete at least one WI course in the major. This first step led to asking students to write all “lab reports” as parts of a typical ACS journal article and utilizing review and revision techniques. The details of this approach are described below. Additionally, I decided to replace one or two traditional experiments with an open-ended inquiry experiment over multiple weeks. This approach refocused the student work away from cookbook thinking. As you can see in Table 1, this small move toward the complete research-focused course began by inserting a two-week research experience in weeks 12 and 13 in the 2010 course. That initial project was connected to my ongoing primary research focus at that point, with the students making simple derivatives of platinum-centered liquid crystalline compounds (11). When this two week period of inquiry was introduced, it was immediately visible that the students found tremendous excitement about the idea of possibly making a novel material that had never existed before. Even though every one of the students’ trials technically failed, they all indicated that the work gave them a significant insight into the world of research—delving into the literature to find a new idea or approach, the value of trial and error, and the willingness to push through a challenge. In their own way, these research failures were a success because an initial idea was explored within the safety-net of the teaching realm, a place where success is often measured by knowing that students understand and value a learning experience.

85 Gourley and Jones; Best Practices for Supporting and Expanding Undergraduate Research in Chemistry ACS Symposium Series; American Chemical Society: Washington, DC, 2018.

Week

Example from 2010

2015 to 2017

1

The Three R’s: Reading wRiting and Research

Discussion of Physical Techniques

2

Acylation of Ferrocene

The Three R’s: Reading wRiting and Research

Develop an Abstract for a Literature Paper

3

Separation of Acylated Ferrocene Products

Computational Chemistry and Choice of Project Framework

Results and Discussion of Computational Experment

4

Synthesis of Superconductive Oxide

Synthesis and Characterization of Metal Complexes

Initial Experimental Section (ACAC)

5

Characterization and Analysis of Superconductive Oxide

Synthesis and Characterization of Metal Complexes

Introduction (ACAC)

6

Computational Chemistry

Synthesis and Characterization of Metal Complexes and Research Proposal

Proposal for Research

7

Synthesis of Geometric Isomers

Data Analysis and Writing Workshop

Full Experimental Section as well as Results and Discussion (ACAC)

8

Reactivity and Characterization of Geometric Isomers

Research Project

9

Synthesis of Linkage Isomers

Research Project

10

Reactivity and Characterization of Linkage Isomers

Research Project

11

Synthesis of ACAC Complexes

Research Project

12

Synthesis of a Metallomesogen

Research Project

86

Downloaded by MIAMI UNIV on May 13, 2018 | https://pubs.acs.org Publication Date (Web): May 2, 2018 | doi: 10.1021/bk-2018-1275.ch006

Table 1. Weekly Schedules for the Inorganic Laboratory and Writing 2015-2017 Writing

Initial Experimental Section

Initial Results and Discussion

Gourley and Jones; Best Practices for Supporting and Expanding Undergraduate Research in Chemistry ACS Symposium Series; American Chemical Society: Washington, DC, 2018.

Example from 2010

2015 to 2017

2015-2017 Writing

13

Characterization of a Metallomesogen

Research Project and Poster Preparation

Introduction and Conclusions

14

Presentations

Research Project and Poster Preparation

Poster Development

15

Presentations

Poster Presentation at Scholarship Sewanee

Final Manuscript

87

Downloaded by MIAMI UNIV on May 13, 2018 | https://pubs.acs.org Publication Date (Web): May 2, 2018 | doi: 10.1021/bk-2018-1275.ch006

Week

Gourley and Jones; Best Practices for Supporting and Expanding Undergraduate Research in Chemistry ACS Symposium Series; American Chemical Society: Washington, DC, 2018.

Downloaded by MIAMI UNIV on May 13, 2018 | https://pubs.acs.org Publication Date (Web): May 2, 2018 | doi: 10.1021/bk-2018-1275.ch006

To build on this partial success, I endeavored to convert the laboratory portion of the course into a full semester “internship” organized much like the developmental steps many mentors use with students when they join their research group, both during the summer and in the academic year. The initial portion of the 14-week laboratory period (Table 1) was converted into a series of group meetings to expand, explore and discuss several useful skills: (a) the wide array of characterization methodologies available and how they are used, (b) how to search and read the scientific literature, (c) how to write scientific communication, and (d) how computational chemistry can aid and expand experimental science. The next steps for the students are to develop their synthetic and characterization skills, and to imagine and design a small part of a larger research effort. After the initial orientation week of the analytical tools that will be available, an initial orientation to literature searching and the planned research frameworks are initially presented in the second lab meeting (“Reading, wRiting and Research”). Two example frameworks derived from my laboratory, and a third framework that the co-instructor of the course in 2017 are presented in Scheme 1. One reason multiple frameworks are developed is so that the students can discover a project in an area of potential intrinsic interest; the other is that a variety of any instructors’ novel research ideas can be explored and tested.

Scheme 1. Three Potential Project Frameworks Derived from the Instructors’ Laboratories: (a) Selective bromination of a [M(acac)3] complex and Suzuki coupling to a naphthalenedimide dye. (b) Potential synthesis routes for an array of potential drug derivative compounds related to KP1019—[Ru(ind)2Cl4]–. (c) Synthesis of new catalysts for amide preparation. 88 Gourley and Jones; Best Practices for Supporting and Expanding Undergraduate Research in Chemistry ACS Symposium Series; American Chemical Society: Washington, DC, 2018.

Downloaded by MIAMI UNIV on May 13, 2018 | https://pubs.acs.org Publication Date (Web): May 2, 2018 | doi: 10.1021/bk-2018-1275.ch006

Generalized representations such as those shown in Scheme 1 are used in the initially framework presentations so that the students see the overarching ideas and goals and engage in an open discussion about both the possible underlying questions and potential “bite-sized” projects that can be explored for the research portion of the course. Short project synopses of these discussions and initial literature leads are also posted to the course website, allowing the students time choose a topic of interest. Here are a few examples of how the frameworks in Scheme 1 connect lead toward the instructors’ research goals. The framework in Scheme 1a is to develop a new class of inorganic-organic conjugate dyes capable of harvesting solar energy. The first goals of this work have been (a) to develop a more useful monobromination of metal acetylacetonate complexes (12) and (b) to conjugate the resulting complex to a naphthalenediimide photodye via a Suzuki coupling (13). The hope is that this project appeals to students interested in materials and environmental questions. While the bromination approach has proved difficult, a series of Suzuki conditions using other useful model compounds (for example, 5-bromophenanthroline) have led to preliminary results being moved into my research laboratory. The framework in Scheme 1b focuses on the development of derivatives associated with a ruthenium-centered anti-cancer drug, usually referred to as KP1019, that reached early clinical trials but did not move beyond that point (14). At this point the project is focused on synthesis; however, it could be easily expanded in terms of reactivity studies or even biological assays if two faculty begin to collaborate in the research realm and/or the project might be integrated into a potentially related course; for example, biochemistry or molecular biology. This project helps students see the relationships between chemical synthesis, pharmacology, and medical care. The framework in Scheme 1c is aimed at developing new catalytic processes capable of preparing amides in gentle and environmentally useful ways (15). In the first year of utilizing this framework students have improved the proposed synthetic methodologies aimed to create potential catalysts, and even attempted to test catalytic behavior for the new compounds. Once this initial couple of orientation weeks are carried out, the students begin their experiential learning with an in-house written expanded version of the traditional syntheses of metal acetylacetonate complexes. To make this experiment more inquiry-driven, the instructor(s) specifically do not fully state what the complexes’ structures are, and pose questions such as that shown in Scheme 2, “How does the acetylacetonate ligand bond to the metal?” Each pair of students begins by choosing two metals from the listed options (typically, Cr, Mn, Fe, Co, and Al) and then carries out the appropriate syntheses provided (16–18). Each complex has a slightly different historical methodology, and as a result the students gain extra hands-on experience by following multiple protocols simultaneously. Each complex is usually prepared by at least two student teams so that the students can compare the outcomes. Once the samples are made, each team begins by checking purity (i.e., by melting point determination) and developing a purification method (i.e. recrystallization), if needed. Next comes a full battery of analytical characterization—IR, UV-Vis, NMR (diamagnetic and paramagnetic Evans methodology), and solid-state magnetometry, using a Gouy 89 Gourley and Jones; Best Practices for Supporting and Expanding Undergraduate Research in Chemistry ACS Symposium Series; American Chemical Society: Washington, DC, 2018.

Downloaded by MIAMI UNIV on May 13, 2018 | https://pubs.acs.org Publication Date (Web): May 2, 2018 | doi: 10.1021/bk-2018-1275.ch006

balance. The development of a computational model using Spartan is suggested as an optional exercise to examine the bonding question raised earlier (Scheme 2). This allows the students to return to the computational skills they began to gain in the initial computational session a few weeks earlier.

Scheme 2. Hypothetical Metal-Ligand Bonding Options Once the initial analytical work and group comparisons are complete, the students and the instructor(s) work together to consolidate data for all the metal acetylacetonate complexes in an analysis and writing workshop. This discussion provides an opportunity to practice data sharing and replicate checking. The consolidated spectroscopic information also allows the students to understand structure-property relationships and spectral analysis trends more completely. The group discussion provides teaching moments about how to tabulate data in a style often used in the synthetic chemistry literature for experimental reporting and how to develop a narrative of the results and their meaning, both spoken and written. The gathered data is also brought into the lecture portion of the course to explore the intellectual understanding of how transition metal complexes, and matter of all kinds, interacts with electromagnetic radiation in ways that provide information about structure and properties.

Writing Aspects of the Course The course has been designated as a writing-intensive (WI) requirement within the chemistry major, and as such the students are offered the opportunity to resubmit edited versions of their writing after an initial review by the instructor. In the initial years of developing this approach, the use of peer review was attempted, but the students repeatedly indicated that peer review was overly challenging and time-consuming. Upon moving back to instructor review and requested revision, 90 Gourley and Jones; Best Practices for Supporting and Expanding Undergraduate Research in Chemistry ACS Symposium Series; American Chemical Society: Washington, DC, 2018.

Downloaded by MIAMI UNIV on May 13, 2018 | https://pubs.acs.org Publication Date (Web): May 2, 2018 | doi: 10.1021/bk-2018-1275.ch006

the use of both a writing contract and a writing rubric have been implemented. In all approaches, the main goal was to guide the students to the realization that rewriting and revision are important skills in scientific discourse. Throughout the first seven weeks of the course, the students begin their development of scientific writing skills by creating individual sections of a “full paper” manuscript in a somewhat unusual order (see Table 1)—abstract, experimental, results and discussion, and introduction. This order begins the writing process by formulating a summary that explains what was done, what was discovered, and what is “big picture” of a research effort. Students repeatedly indicate that the “big picture” writing needed for a well-framed introduction is the most difficult part of scientific writing, which is why that is left to the end. It is also hoped that this order helps the students gain more insight into the research framework they have chosen, and the small research goals they have developed. The first writing assignment (see Table 1) begins with reading and discussing a short literature paper, typically a communication, with its abstract excised so that each student writes their own abstract for the paper without referring the one found on the publication website or SciFinder. The goal of this initial exercise is having students note key features of a shorter literature work and develop a clear summary of the outcomes. Once they have completed the initial assignment, the next step is to formulate their understanding of how experimental findings can be communicated. First, they draft a “results and discussion” section for the initial computational experiment they have explored, which is focused on questions related to the formation of molecular orbitals and molecular structure. This writing exercise allows the students to evaluate their computed results by connecting the ideas to theoretical topics under discussion in the lecture portion of the course and by comparing their own results to literature sources. Once the students begin the metal acetylacetonate complex synthesis, the next writing goal is drafting an experimental section in the traditional third-person and past passive voice used in many journals. As the analytical and spectroscopy data on the complexes is collected, the students explicitly revise the experimental section and add the accumulated spectral information in the format used in many synthetic manuscripts (e.g. “1H NMR (CDCl3, ppm): 4.55 (1H), 2.15 (6H)….”). In the last two weeks of the initial pre-research period, the students are asked to draft two assignments. The first is an introduction to present the metal acetylacetonate complex synthesis. The second assignment is developing a research proposal that maps out the procedures they will attempt in the research weeks and the goals of this work as they seem them. Both of these exercises help the student to consider questions such as “What is the work focused on?” and “Why does the reader care?” As the research project part of the course begins, the students are asked to assemble the sections developed for their metal acetylacetonate complexes as a complete manuscript, and to complete as many revision suggestions as possible. They know that they will need to develop a public poster presentation and write a full manuscript about their research work at the end of the course. As such, they value the relatively easy opportunity to assemble this manuscript while primarily focusing on their research. 91 Gourley and Jones; Best Practices for Supporting and Expanding Undergraduate Research in Chemistry ACS Symposium Series; American Chemical Society: Washington, DC, 2018.

Downloaded by MIAMI UNIV on May 13, 2018 | https://pubs.acs.org Publication Date (Web): May 2, 2018 | doi: 10.1021/bk-2018-1275.ch006

The last week of the course caps the research process by preparing a poster for our campus symposium, “Scholarship Sewanee.” Being ready to present the poster and discuss their work strongly motivates the students to compile all of their research results, accepting the good, bad, and ugly data as part of the process. To reexamine and analyze their research data, they convert the poster information into a manuscript. While the poster is authored by a pair, or in a few cases two pairs working on a related question, the final manuscripts are solitary writing tasks. Interestingly, the instructor(s) have noticed that each student of a team usually chooses to focus their manuscript on different points of discussion related to the overall project outcomes. For example, one team member focuses on how the spectral data defines the product made and the other focuses on unexpected results, positive or negative.

Outcomes Two survey-related tools were utilized to examine the potential impacts of this integrated teaching-research model. The more straightforward instrument was the on-campus teaching evaluation, which offers open questions related to topics like “was the course intellectually stimulating” and allows optional questions such as “was the laboratory experience valuable.” The Classroom Undergraduate Research Experience (CURE) survey developed by Lopatto (19, 20), which uses a traditional 5-point scale, was utilized to hopefully discover more detailed insights about what intellectual and experiential growth the students have gained in this course. The CURE survey asks the students about their perceptions in three areas—course elements, intellectual growth benefits, and attitudes about science. The course elements portion uses a pre vs. post self-reflection approach to assess both prior knowledge and learning gains in the course. While the questions’ wordings are identical in both sections, the answer scales provided for the students are different. The “preflection” examines prior experience by using a scale ranging from 1 = “no experience/inexperience” to 5 = “significant previous experience.” The post-survey focuses on the learning gains by using a scale ranging from 1 = “no to little gain” and 5 = “large gain.” In contrast to the pre vs. post approach of the course element portion, the intellectual growth benefits portion only measures post-experience expectations, ranging from career exploration to problem solving and analysis. Tables 2 and 3 examine the averaged data of the three years (2015 to 2017) of CURE data accumulated since the course redesign. A rationale for using an average is to more accurately compare the small data sets at Sewanee (4 to 12 submitted surveys) with the nation-wide data. A weighted averaging of the oncampus data was utilized to minimize any potential skewing of the average by the smallest class’s averages. That said, examination of an individual course year’s data indicates that there are no significant outliers from year to year. Moreover, the averaging of both the local data (24 participants) and the national data (>27K participants) generate relatively small standard deviations, for the national data, less than ±0.1; for the local data, less than ±0.5. As such, comparison of the local 92 Gourley and Jones; Best Practices for Supporting and Expanding Undergraduate Research in Chemistry ACS Symposium Series; American Chemical Society: Washington, DC, 2018.

and national data should provide insight with regard to what impacts the course might provide. Because of the pre vs. post process of the CURE course elements portion, I was intrigued with the possibility that calculating a “differential value” between the students’ “learning gains” and their “prior experience” (differential = learning gain – prior experience) would highlight additional insights into the potential educational impacts of this integrated research-teaching model. Both the local (Sewanee) and national data for these differential values, as well as the established CURE data, are compared in Table 2.

Downloaded by MIAMI UNIV on May 13, 2018 | https://pubs.acs.org Publication Date (Web): May 2, 2018 | doi: 10.1021/bk-2018-1275.ch006

Table 2. Self-Reported Experiences and Learning Gains Element

Prior Experience

Learning Gains

Post-Pre Differential

Sewanee / National

Sewanee / National

Sewanee / National

4.13

3.45

2.36

3.30

-1.77

-0.16

A

Scripted lab or project where students know outcome

B

Lab or project where only instructor knows outcome

3.75

3.32

3.08

3.42

-0.68

0.09

C

Lab or project where no one knows the outcome

3.04

2.52

4.06

3.41

1.01

0.89

2.46

3.01

4.17

3.87

1.71

0.86

D

A project where students have input into process or topic

E

A project entirely of student design

2.00

2.55

3.56

3.57

1.56

1.03

F

Work individually

3.17

3.61

2.93

3.40

-0.23

-0.20

G

Work as a whole class

2.42

3.14

2.89

3.24

0.48

0.10

H

Work in small groups

4.00

3.85

4.17

3.91

0.17

0.06

I

Become responsible for a part of the project

3.75

3.81

4.13

3.94

0.38

0.13

J

Read primary scientific literature

3.04

3.11

3.92

3.62

0.87

0.51

K

Write a research proposal

1.54

2.52

4.29

3.50

2.75

0.98

L

Collect data

3.29

3.67

4.17

3.94

0.88

0.27

M

Analyze data

3.29

3.60

4.25

4.05

0.96

0.45

N

Present results orally

2.46

3.15

3.76

3.64

1.31

0.48

Continued on next page.

93 Gourley and Jones; Best Practices for Supporting and Expanding Undergraduate Research in Chemistry ACS Symposium Series; American Chemical Society: Washington, DC, 2018.

Table 2. (Continued). Self-Reported Experiences and Learning Gains

Downloaded by MIAMI UNIV on May 13, 2018 | https://pubs.acs.org Publication Date (Web): May 2, 2018 | doi: 10.1021/bk-2018-1275.ch006

Element

Prior Experience

Learning Gains

Post-Pre Differential

Sewanee / National

Sewanee / National

Sewanee / National

O

Present results in written papers or reports

3.21

3.42

4.04

3.76

0.83

0.33

P

Present posters

2.34

2.93

4.08

3.34

1.75

0.41

S

Critique work of other students

2.54

2.95

3.16

3.31

0.62

0.36

T

Computer modeling

2.37

2.32

3.04

3.12

0.67

0.80

I hypothesized that the differential values for these course elements would correlate to both the students’ understanding of how the course is structured and their own educational growth. Figure 2 provides easy visualization of the relative differential values between all the local and national data (right column in Table 2). As can be seen in Figure 2, the differential values for most of the survey items are slightly, or significantly, higher for the local data relatively to the national. Of specific note, the differential values for “student input” and “student design” are approximately twice as high as their national differentials. Additionally, all the items related to communication show differential values between two to four times larger than the national data. Consequently, it appears that this course design leads to significant learning gains in important research traits—exploration of the unknown and communication of discovery.

Figure 2. Differential Between Self-reported Prior Experience and Learning Gains. 94 Gourley and Jones; Best Practices for Supporting and Expanding Undergraduate Research in Chemistry ACS Symposium Series; American Chemical Society: Washington, DC, 2018.

Downloaded by MIAMI UNIV on May 13, 2018 | https://pubs.acs.org Publication Date (Web): May 2, 2018 | doi: 10.1021/bk-2018-1275.ch006

Interestingly, this differential value approach led to the intriguing discovery that only two items in the national data, and three in the Sewanee data, (others that focus on the lecture portion of the course have been omitted) have a negative differential value—scripted labs, working individually, and project where the instructor knows the outcome. This set of correlations suggests that shifting to a research-based experience has made an impact on student perspectives relating to the purpose of a laboratory experience. These negative differentials suggest that students perceive little to no gains in these course elements, which are to some degree inverse to the research experience goals. For example, the differential value for the “scripted lab” is significantly different between the local and national data (-1.77 vs. -0.16), which suggests that students do not generally sense the partially scripted experiences in the course even with some partial cookbook procedures present early in the course. In a similar vein, the differential for “a lab or project where only the instructor knows the outcome” also is negative for the local data, while the national is slightly positive (-0.68 vs. 0.09). This difference highlights that the students in the course understand they are working on research that has no prior instructor insight and see the work as not containing traditional “cookbook” learning. The intellectual growth benefits portion of the survey unfortunately does not provide a pre vs post comparison; it is only contains a post-experience survey. As can be seen in Table 3, again comparing the averaged data for three years, the local and national data are relatively similar in most cases. The national data for each year is reported with a standard deviation of approximately ±1, which suggests that students across all participating institutions have a wide array of different views. Interestingly, the three-year averages for both the local and national data have significantly smaller deviations (approximately 0.1), which suggests views from year to year are relatively stable overall. The Sewanee data are slightly higher than the national information in many categories, but it is unlikely most of the differences between the two datasets is statistically significant given the relatively small deviations for the three-year averaging. The one item with a notable distinction is “skill in science writing.” The local data is 0.65 higher than the national data, which makes sense given the integration of the inorganic course into our campus-wide WI program. Other smaller differences (approximately 0.2 higher) can be seen for “Understanding how scientists work on real problems,” “Self-confidence,” “Understanding how scientists think,” and “Learning to work independently.” These general trends hint that the course is providing creative and experiential learning in the research realm. The CURE survey also includes some overall assessment from the student view, Table 4. As with the “Intellectual Growth” the report from Lopatto’s team reports an annual standard deviation of approximately ±1, suggesting the student beliefs about an overall class are somewhat varied but that the overall view of the classes is favorable.

95 Gourley and Jones; Best Practices for Supporting and Expanding Undergraduate Research in Chemistry ACS Symposium Series; American Chemical Society: Washington, DC, 2018.

Table 3. The Self-Reported Intellectual Growth Benefits

Downloaded by MIAMI UNIV on May 13, 2018 | https://pubs.acs.org Publication Date (Web): May 2, 2018 | doi: 10.1021/bk-2018-1275.ch006

Learning Element

Sewanee

National

Skill in interpretation of results

3.65

3.62

Tolerance for obstacles faced in the research process

3.62

3.59

Readiness for more demanding research

3.62

3.51

Understanding how knowledge is constructed

3.60

3.53

Understanding the research process

3.65

3.57

Ability to integrate theory and practice

3.57

3.55

Understanding how scientists work on real problems

3.86

3.67

Understanding that scientific assertions require supporting evidence

3.76

3.71

Ability to analyze data and other information

3.65

3.79

Understanding science

3.76

3.69

Learning ethical conduct

3.23

3.30

Learning laboratory techniques

3.81

3.79

Ability to read and understand primary literature

3.62

3.48

Skill in how to give an effective oral presentation

3.05

3.29

Skill in science writing

4.10

3.45

Self-confidence

3.52

3.35

Understanding how scientists think

3.67

3.51

Learning to work independently

3.65

3.43

Becoming part of a learning community

3.67

3.55

Clarification of a career path

2.94

3.08

Confidence in my potential as a teacher

3.11

3.05

As previously stated, three years of this data was averaged to attempt a clearer comparison. Given the typical annual standard deviation for the national data, it is not surprising that the local data is not significantly different than the national. The slightly lower value related to the course being a “good way of learning about the subject” probably occurs because the student career demographics vary widely in the small sample (24 students) that have taken the course. However, the small increases in “learning about the process of scientific research” and being “able to ask questions…get helpful responses” might hint that students have valued the experiential and collaborative learning environment that the course is focused on. 96 Gourley and Jones; Best Practices for Supporting and Expanding Undergraduate Research in Chemistry ACS Symposium Series; American Chemical Society: Washington, DC, 2018.

Our open-ended course evaluation questions provide a qualitative comparison to the quantified student perception measures in the CURE survey (Table 4). Table 5 collates testimonials from students in all three years of the new lab format.

Downloaded by MIAMI UNIV on May 13, 2018 | https://pubs.acs.org Publication Date (Web): May 2, 2018 | doi: 10.1021/bk-2018-1275.ch006

Table 4. Self-Reported Views About Value of the Course Sewanee

National

This course was a good way of learning about the subject

3.7

4.1

This course was a good way of learning about the process of scientific research

4.4

4.1

This course had a positive effect on my interest in science

4.0

3.9

I was able to ask questions in this class and get helpful responses

4.4

4.1

Table 5. Student Testimonials Provided in Course Evaluations “I really liked doing the research project in the lab. It made me think about what I was doing and why as opposed to just blindly following instructions given to me.” “This a perfect example of how to teach a student to be independent.” “I really liked the research based class since it gives a more applicable way of teaching students how to actually use the knowledge learned in class in a lab setting.” “I really liked doing the KP1019 lab and how it was a lot like a research lab. However, I think it would be beneficial to have more time to work on it” “I learned a lot in this course, one of my favorite things was the way in which the lab was run, giving us what felt like an actual taste of what research is like in real life.”

These quotes indicate that several students deeply valued the opportunity to experience the research world within the course. Interestingly, while many students did not provide this type of positive feedback, none really provided strong negative comments. Rather, they made small requests such as the one above that asks for more research time. Toward that end, the instructor(s) have intentionally pared down the scripted exercises in the courses. For example, the first year of the course included a second two-week “practice” laboratory, the synthesis of KP1019. Removing that exercise provided both an additional week of research time and the data analysis workshop in the current form of the course. 97 Gourley and Jones; Best Practices for Supporting and Expanding Undergraduate Research in Chemistry ACS Symposium Series; American Chemical Society: Washington, DC, 2018.

Annual course modifications like this, as well as updating the research frameworks, have benefited not only the students but also the faculty members involved. Expanding the research period has helped provide fuller partial answers for the instructors’ original research questions. These partial answers both refine the presented frameworks and the instructor’s own research goals, helping to forge a link between teaching and research ambitions.

Downloaded by MIAMI UNIV on May 13, 2018 | https://pubs.acs.org Publication Date (Web): May 2, 2018 | doi: 10.1021/bk-2018-1275.ch006

Conclusions This inorganic chemistry course has become a link between teaching and research, both for the students and the faculty member. It helps address the capacity issue associated with undergraduate research by providing all chemistry majors on campus with a “mini-REU” research experience. The CURE data itself, the differential value approach being explored here, and the individual course evaluation testimonials strongly indicate that the students perceive the laboratory to provide educationally beneficial and intellectual growth. Examining the primary student outputs—lab writing, poster presentations, verbal communication in the laboratory—it is also clear that the students are learning a range of laboratory techniques, spectroscopic analysis, and important communication skills through this educational laboratory experience. From a faculty perspective, this approach has provided an integrated link between both the work of teaching and the work of scholarship. Several new research ideas initiated in the course have grown and flourished in my research lab over the last three years. Moreover, my co-instructor for the last year gained new initial data for their research program. Additionally, this link has lead to some of the valuable educational impacts for students that Astin (6) noted as essential years ago: (a) quality and quantity of student interactions with faculty outside the classroom and (b) level of student involvement. This course also reconnects to Boyer’s long-term goal of reconsidering scholarship (21) and hopefully balancing two of the priorities in faculty work, teaching and research.

Acknowledgments Dr. Evan Joslin has worked on this project as a co-instructor and by providing a research framework for several students in the most recent year of the course. Dr. Bridget Gourley and Dr. Rebecca Jones are also thanked for providing valuable discussion focused on the integration of research and teaching at several American Chemical Society national meetings.

98 Gourley and Jones; Best Practices for Supporting and Expanding Undergraduate Research in Chemistry ACS Symposium Series; American Chemical Society: Washington, DC, 2018.

References 1. 2. 3. 4.

Downloaded by MIAMI UNIV on May 13, 2018 | https://pubs.acs.org Publication Date (Web): May 2, 2018 | doi: 10.1021/bk-2018-1275.ch006

5.

6. 7.

8. 9. 10.

11. 12. 13. 14. 15. 16.

17.

18. 19. 20.

Undergraduate Professional Education in Chemistry; American Chemical Society: Washington, DC, 2015; p 16. Lopatto, D. Undergraduate Research as High-Impact Student Experience. Peer Rev. 2010, 12, 27–30. Lopatto, D. Science in Solution: The Impact of Undergraduate Research on Student Learning; Research Corporation: Tuscon, AZ, 2009. Kuh, G. D. High-impact educational practices: What they are, who has access to them and why they matter; Association of American Colleges and Universities: Washington, DC, 2008. Kuh, K. D.; O’Donnell, K. Ensuring Quality & Taking High-Impact Practices to Scale; Association of American Colleges and Universities: Washington, DC, 2008. Astin, A. W. What Matters in College: Four Critical Years Revisited. JosseyBass: San Francisco, CA, 1993. Girolami, G. S.; Rauchfuss, T. B.; Angelici, R. J. Synthesis and Technique in Inorganic Chemistry: A Laboratory Manual, 3rd ed.; Univeristy Science Books: Sausalito, CA, 1999; pp 17−26. Work, J. B. Inorg. Synth. 1946, 2, 221. Broomhead, J. A.; Dwyer, F. P.; Hogarth, J. W. Inorg. Synth. 1957, 6, 183. Richen, D. T.; Glidewell, C. Linkage Isomerism: An Infrared Study in Inorganic Experiments: Second, Completely Revised and Enlarged Edition; Woolins, J. D., Ed.; Wiley VCH Verlag: Weinheim, Germany, 2003; pp 31−33. Cocker, T. M.; Bachman, R. E. Mol. Cryst. Liq. Cryst. 2004, 408, 1–19. Collman, J. P.; Moss, R. A.; Maltz, H.; Heindel, C. C. J. Am. Chem. Soc. 1961, 83, 531–534. Pirtsch, M.; Paria, S.; Matsuno, T.; Isobe, H.; Resier, O. Chem. Eur. J. 2012, 18, 7336–7340. Cebrian-Losantos, B.; Reisner, E.; Kowol, C. R.; Roller, A.; Shova, S.; Arion, V. B.; Keppler, B. K. Inorg. Chem. 2008, 47, 6513–6523. Zhang, Y.; Chen, C.; Ghosh, S. C.; Li, Y.; Hong, S. H. Organometallics 2010, 29, 1374–1378. Glidewell, C. Metal Acetylacetonate Complexes: Preparation and Characterization in Inorganic Experiments; Woollins, J. D., Ed.; Wiley VCH: Weinheim, Germany, 2003; pp 149−159 Szafran, Z.; Pike, R. M.; Singh, M. M. Microscale Inorganic Chemistry, A Comprehensive Laboratory Experience; John Wiley and Sons, Inc.: New York, 1991; pp 224−227. Shalhoub, G. M. J. Chem. Ed. 1980, 57, 525–526. Denofrio, L. A.; Russell, B.; Lopatto, D.; Lu, Y. Linking student interests to science curricula. Science 2007, 318, 1872–1873. Lopatto, D.; Alvarez, C.; Barnard, D.; Chandrasekaran, C.; Chung, H.-M.; Du, C.; Eckdahl, T.; Goodman, A. L.; Hauser, C.; Jones, C. J.; Kopp, O. R.; Kuleck, G. A.; Mcneil, G.; Morris, R.; Myka, J. L.; Nagengast, A.; Overvoorde, P. J.; Poet, J. L.; Reed, K.; Regisford, G.; Revie, D.; 99

Gourley and Jones; Best Practices for Supporting and Expanding Undergraduate Research in Chemistry ACS Symposium Series; American Chemical Society: Washington, DC, 2018.

Downloaded by MIAMI UNIV on May 13, 2018 | https://pubs.acs.org Publication Date (Web): May 2, 2018 | doi: 10.1021/bk-2018-1275.ch006

Rosenwald, A.; Saville, K.; Shaw, M.; Skuse, G. R.; Smith, C.; Smith, M.; Spratt, M.; Stamm, J.; Thompson, J. S.; Wilson, B. A.; Witkowski, C.; Youngblom, J.; Leung, W.; Shaffer, C. D.; Buhler, J.; Mardis, E.; Elgin, S. C. Undergraduate research. Genomics Education Partnership. Science 2008, 322, 684–685. 21. Boyer, E. L. Scholarship Reconsidered: Priorities of the Professoriate The Carnegie Foundation for the Advancement of Teaching; John Wiley: New York, 1990.

100 Gourley and Jones; Best Practices for Supporting and Expanding Undergraduate Research in Chemistry ACS Symposium Series; American Chemical Society: Washington, DC, 2018.