Research at Predominantly Two-Year Campuses of Penn State

This chapter describes undergraduate research conducted at three Pennsylvania State University satellite campuses which do not offer a chemistry degre...
5 downloads 8 Views 1MB Size
Chapter 6

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

Research at Predominantly Two-Year Campuses of Penn State Lee J. Silverberg,1,* John Tierney,2 and Kevin C. Cannon3 1Pennsylvania

State University, Schuylkill Campus, 200 University Drive, Schuylkill Haven, Pennsylvania 17972 2Pennsylvania State University, Brandywine Campus, 25 Yearsley Mill Road, Media, Pennsylvania 19063 3Pennsylvania State University, Abington Campus, 1600 Woodland Road, Abington, Pennsylvania 19001 *E-mail: [email protected]

This chapter describes undergraduate research conducted at three Pennsylvania State University satellite campuses which do not offer a chemistry degree. An overview of both the challenges and the strategies associated with conducting research, most of it collaborative, on these two-year campuses is presented. Included in this chapter is a summary of the varied undergraduate projects that provide students an opportunity to obtain valuable research experience that will provide a solid foundation when they transfer to a degree-granting campus. Among the three authors, 29, 200+, and 43 undergraduates from the Abington, Brandywine, and Schuylkill campuses, respectively, have participated in research. The projects described, most of which are ongoing, have to date resulted in peer-reviewed publications with undergraduate co-authors from Abington (6), Brandywine (35), and Schuylkill (26) campuses. Undergraduate coauthors have reported that their published work has been a major topic in their interviews at professional schools or jobs that sets them apart from their peers. The majority of the students at these three campus locations are from low to middle-income socio-economic backgrounds, many of whom are first-generation college students. Approximately twenty percent of the students conducting research are minorities, and approximately fourteen © 2016 American Chemical Society Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

percent of the undergraduate co-authors are minorities. The projects, which range in focus from synthetic organic chemistry, pedagogical methodologies, and chemical history, have proven useful to the students in furthering their career goals while enabling faculty to meet research expectations at these two-year campuses.

Introduction The value of Science, Technology, Engineering and Math (STEM) research with undergraduates has been well-documented in recent years (1–12). For example, students are more satisfied and learn more (1), they integrate knowledge learned in courses (2), minority and female students are guided towards a scientific career (1), institutions attract more students (1), and students are better prepared for (2) and more likely to enroll in graduate school (3). College students in their first two years can also be effective researchers (13–17). At community colleges, where the teaching load for faculty is typically very high and there is no expectation of a research program, research in the sciences has been successfully performed (18–24). Even at the high school level, research has been reported (25–27). The Pennsylvania State University (PSU) is a multi-campus university with 24 locations. The University Park Campus offers the full array of 167 degree programs, with five other four-year locations offering a significant number of degree programs. There are fourteen predominantly two-year locations that offer a very limited number of four-year degrees, and four special mission campuses – a medical school, a law school, the affiliated Penn College of Technology and the Graduate Studies Center at Great Valley. Only three campus locations (University Park, Erie, and Berks) offer chemistry degree programs; none of the fourteen predominantly two-year locations offers an undergraduate chemistry degree, and students must transfer to one of the other locations to complete the degree after their sophomore year. Regardless of whether a degree program exists, chemistry faculty at all campus locations are required to maintain an active and productive research program as part of their evaluation towards promotion and tenure, and beyond. In this chapter, we discuss our experiences in conducting successful research programs with students on three of these non-degree campuses.

Issues and Advantages The authors of this chapter all conduct organic chemistry-based research as well as some pedagogically-focused research. The PSU two-year campuses are predominantly commuter campuses and have very limited equipment and personnel resources. This makes research that requires the use and maintenance of expensive equipment difficult to achieve, as will be discussed later in the chapter. 84 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

For example, at two of the three campuses the authors do not have a Nuclear Magnetic Resonance spectroscometer (NMR), which is a critical analytical instrument for organic chemists; the third campus, Brandywine, does have a 60 MHz Anasazi Instrument. To circumvent the lack of facilities and funding, many of the chemistry faculty at two-year campuses choose to do computational research to meet research expectations. Another challenge is to find a way to engage students in research while they are just beginning to learn chemistry and have usually not yet taken Organic Chemistry, a second-year course. Some students are still taking General Chemistry. Students therefore enter research having to quickly learn the chemistry related to their projects. Additionally, since most students leave after their second year at the campus, learning and productivity is typically compressed into two years or less, with potential extensions into summer semesters. The challenges associated with accelerated student learning and productivity that is necessitated by limited student time on campus are further exacerbated by limitations on the primary researcher’s availability. At our two-year campuses, the faculty course load is typically three courses each semester, similar to fouryear undergraduate institutions (28). In addition, the faculty may also have heavy advising loads.This leaves faculty with limited time to devote to the accelerated development of the undergraduate researchers that is often required. However, being a part of PSU brings advantages as well. As a large research university, it has vast library resources available, so journals and databases are readily available electronically. Computational facilities and faculty collaborations in this area are readily accessible within the university. For synthetic chemists such as ourselves, despite the lack of analytical equipment on campus, services are available through the University Park Campus, although we have to mail our samples and pay for the service.

Solutions Funding While articles in the literature about undergraduate research often discuss the use of significant external funding (11, 17–20, 22), such as NSF grants, we generally get our funding from internal campus grants, which are small (≤ $2000/ year). It is enough generally to buy needed materials, pay for analyses (even within PSU, there are charges, but less than externally), and pay publication fees for open-access journals. We usually don’t have enough money to pay students, although occasionally we can get some student support from industry. Finding Students There are five different ways through which students can become involved in our research: a) as volunteers; b) as paid work-study students; c) as an independent study course for credit; d) as part of an Honors course; and e) as part of Organic Chemistry Lab (Chem. 213 at PSU). 85 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

Volunteers are not hard to find. Most students are very happy for the opportunity to do research. There are not a large number of chemistry majors at a small campus, so we accept any students in our chemistry classes who have done well and are enthusiastic about science. Chemistry is “The Central Science” after all, so many future scientists/health care workers/engineers stand to benefit from doing chemistry research. We also have many students from our classes approach us, looking for the chance to do research. Most of the time, we accept them. Occasionally we obtain volunteers who live locally and can do research over the summer. This is when more experienced research students can be found or, in contrast, younger students such as a local high school student who conducted research over summer break. The students have the option of doing research for credit, in which case they are required to spend a specific amount of time in the lab for each credit earned. At the two-year campuses, chemistry classes are not usually large enough to have separate Honors sections. However, there is an active Honors program. Students submit a request to the professor to take the Honors Option in the course. At the Schuylkill campus, this option may include replacing the final exam with a semester-long research project culminating in a poster or oral presentation at the end. At Brandywine, the Honors Option approach is common and students work with the professor on an in-depth focus on a small part or topic in the course. The in-depth focus has resulted in a different set of exam questions for those individuals or a poster presentation of the work accomplished at the campus Exhibition of Undergraduate Research Enterprise and Creative Accomplishment (EURECA). Unlike most institutions, PSU’s Organic Chemistry Lab (Chem. 213) is not offered with the first semester of Organic Chemistry lecture (Chem. 210), but is introduced as a two credit course (six lab contact hours) with the second semester of lecture (Chem. 212). This results in students with a better theoretical background when they enter the lab and allows for the introduction of a mini-research project in the last three to four weeks of the semester of Chem. 213, which typically ends with a poster or oral presentation. The Chem. 213 research project has to draw on the laboratory skills and techniques already learned early on in the semester, plus the synthesis should be multistep rather than the regular one-step recipe type synthesis that students usually follow from a book or instruction sheets. In any research synthesis additional techniques, not already learned, can be introduced to the more advanced and adept students in the group.

Finding Appropriate Projects The authors have engaged students in four areas of research: synthetic chemistry, pedagogy, history of chemistry, and chemistry in science fiction literature. Introducing synthetic research into the classroom and laboratory produces excitement among the students, particularly when they realize the compounds they are preparing are novel. Each student in a group can be given a slight variation on a synthesis with regards to a substituent in the molecule, leading to a group of 86 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

related compounds in a series. The resulting compounds can then undergo further study using spectroscopic techniques or biological testing which encompasses further involvement on the part of the student with potential interdisciplinary collaborations.. The project becomes shared between the students, and the students start to experience the aspect of working in the environment of a scientific team rather than as individuals. Planning and approach are very important. The students, no matter how academically talented, are not going to have the level of expertise and experience to do projects that are overly complicated and advanced. Further, they should not be entrusted with projects where the potential hazards are high. So in planning the research to be carried out, it is important to keep in mind what they are going to be able to do, and able to do safely. The research must be relatively simple to perform, and in “bite-size” chunks, giving them a chance of accomplishing something in the limited time frame during which they conduct research. The synthetic path chosen should have a reasonable degree of success regardless of a student’s lab skills in order for the student not to become frustrated by failure. Ideally, each piece becomes publishable, if not by itself, as part of a larger effort. Rewarding research of a pedagogical nature can also be achieved with students in a classroom environment, and this research is easily geared to first-year students or non-majors. Again, novelty can be incorporated into the research. It is also a good idea to choose projects that the instructors have familiarity with, and it is not a good idea to start a completely new project where the instructor has not done a reasonable bit of groundwork. Researching the history of chemistry provides an opportunity to study chemistry from an alternative perspective. There are abundant opportunities for research in this field that will develop students’ chemical knowledge, critical thinking skills, and an understanding of the on-going evolution of chemical theories (29, 30). The study of chemical history provides students with a scientific view of the necessarily ephemeral nature of all chemical theory (31–35). Science fiction is typically defined as a genre of fiction which delves into the future and the past (often simultaneously) in an attempt to create an understanding of philosophical and scientific concepts. Science is used in a variety of ways: as subject, as plot, as setting or background, and as metaphor (31). The infusion of chemistry in science fiction has been reviewed according to individual authors (31–35) or to specific works (36–40) focusing on the accuracy of the chemistry or the chemical concepts presented. Likewise, the infusion of science fiction into the chemistry classroom has also been reported (41–43). A survey of academic chemists and physicists indicated that approximately 25% of the respondents made references to science fiction in their courses, and that a substantial number would be receptive to the use of science fiction as a supplement in the teaching of science (41). The goals of using science fiction may include increased enthusiasm about studies in chemistry, increased science literacy, or an in-depth examination of a specific topic in a particular work (44, 45). In line with all three of these goals, one author (KCC) has incorporated science fiction short stories into an ongoing research project targeted toward first and second year science students who would otherwise not participate in traditional undergraduate chemical research. Examples of these four types of research will be shared later in the chapter. 87 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

Training Students Attitude - faculty attitude - is the most important aspect. We view the research as a way to educate students. Yes, we need to publish, but the reality is that the students are not initially self-sufficient and may not be before they move on. So what are our goals? We teach them chemistry, often before they have the opportunity to learn it in the classroom, teach them how to be a researcher, teach them safety and proper lab techniques and practices, and hopefully generate some good data where the student has made a significant intellectual contribution that can result in publishable work. This means that we cannot just turn them loose in the lab and walk away. We need to be present with the students in the lab at all times, for the purposes of safety and consultation. The students benefit greatly from one-on-one interaction. The professors benefit because the students are doing publishable research in a more flexible teaching environment. Yes, the pace of research progress is often slower per student than if we did it ourselves, and it is slower than if we were doing one thing while they were doing another, but until they learn enough and are confident to work on their own, this is how it has to be, and that’s OK! The real point of undergraduate research is what we teach the students. The bonus is that before long, they become self-sufficient enough that we can have them work in a separate hood without us watching their every move, although of course we must still be present in the lab. When we have more than one student in the lab with us, the amount of work being done in the time we are spending becomes more than we could do by ourselves. In any case, when they go on to their next campus, they will be well trained and ready to do great things with the next advisor. We know this from experience and through feedback from our colleagues at the University Park campus. We are flexible and realistic. Students are very busy with classes and so are we. We are not going to get a lot of hours out of them during the semesters. They may only come in once a week for 2-3 hours. Our students know when we are available to go in the lab with them, and if they show up during those times, in we go. Sometimes, more than one student can come in at the same time, which maximizes the use of our time. If a procedure or analysis needs to be done when the student can’t attend lab, the faculty member or another student can do it. Everything is a collaborative effort.

Collaboration The three authors of this chapter collaborate with each other, and we each collaborate with scientists at other institutions as well. Research for us is a lesson in long-distance team building. With three synthetic organic chemistry groups, it becomes easier to effectively complete more experiments since we can combine our data for a publication. We have additional collaborators to obtain instrumental data, including NMR spectroscopy, infrared spectroscopy (IR), mass spectrometry (MS), and X-Ray crystallography. We collaborate with computational chemists on mechanistic and structural investigations. We work 88 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

with biologists and pharmacologists for biological testing. Simply put, we reach out in many different directions to put our ideas into practice. However, it must be understood that this need for collaboration means work will not get done quickly. Patience is important. In order to advance our work at a reasonable pace, we sometimes assume that an experiment has worked based on observation, thin layer chromatography (TLC), and melting points until we eventually obtain more indicative data such as an NMR spectrum. Sometimes those assumptions are wrong and we get a surprise (vide infra).

Case Studies Research Projects – History and Progression at Brandywine Organic Chemistry When one of the coauthors (JT) attempted to continue earlier gas phase kinetics (46) research for his initial campus-based research it became apparent that the maintenance of a vacuum rack and pump was not a viable proposition under the circumstances. The next thought was to complete a kinetic study (47) of the formation of substituted 2,3-diaryl-1,3-thiazolidin-4-ones from the reaction of imines with thioglycolic acid, a reaction originally demonstrated by Surrey (Figure 1) (48).

Figure 1. Synthetic scheme for the formation of substituted 2,3-diaryl-1,3-thiazolidin-4-ones. During this research it became apparent that there was a wide array of substituted 1,3-thiazolidin-4-ones that had not been prepared, but could be prepared by second-year organic chemistry students relatively easily. In addition, 89 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

physical organic chemistry studies utilizing IR, NMR and MS data could be also be achieved – further incorporating information that the students had learned in the lecture portion of the class. The move to a synthetic project proved to be much more doable under the equipment availability at the campus and fruitful for incorporation into the organic chemistry lab course. The spectroscopic data that was used initially tracked the chemical shift variation of the methine proton (at C2), methylene protons (at C5) and the 13C NMR chemical shift data at C2, C4 and C5 in the 1,3-thiazolidin-4-one ring (Figure 2) based on substituents in the phenyl rings. It was in this phase of the project, in the early 1990s, that undergraduate students were incorporated into the project as part of the Chem. 213 undergraduate lab.

Figure 2. Nomenclature and atom positions in 1,3-thiazolidin-4-one ring.

The starting materials for these syntheses are relatively cheap and readily available. The one downside is that the thioglycolic acid is somewhat malodorous and has to be handled carefully. However, it is pointed out to the students that thioglycolic acid is used in permanent waving solutions in women’s hairdressing salons, and some students recognize the odor. The first publication where undergraduate students were intimately involved in the work appeared in Magnetic Resonance in Chemistry in 1996 (49). Soon after, a steady stream of published data appeared on the different series of compounds shown in Figure 3 (50–56). At the same time additional features, such as computational work (57), have further been incorporated into the research, enhancing the students’ research experience.

90 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

Figure 3. Series of 1,3-thiazolidin-4-ones that have been synthesized by undergraduates.

On two occasions the reaction sequence shown in Figure 1 did not go according to plan – leading to very interesting new areas of investigation. In the first instance, when, in the presence of trichloroacetaldehyde, the amine used was switched from an arylamine (Series 7, Figure 3) to cyclohexylamine, the resulting product produced was an N-formyl compound instead of the corresponding thiazolidinone (58). However, use of an appropriate Lewis acid allowed for the formation of the thiazolidinone. The second reaction that gave an interesting, but yet to be characterized polymeric product, was the reaction of p-hydroxybenzaldehyde with aniline. On addition of the thioglycolic acid a thick glassy orange mass resulted. Initial work indicates the possible coupling of imine units with the loss of the hydroxyl group as water (59).

Pedagogy Research of a more pedagogical nature has also been completed with undergraduate students. The initial work was in a collaboration with an Exercise and Sports Science class where the pH of sweat was measured while students were working out in class (60). This work, at the request of two students, was further extended to include horses (61); however, with tightening internal regulations over the years, this type of research has been discontinued – even though it has shown real world applications of chemistry. Additional pedagogical work with undergraduates in non-science major classes with a focus on the periodic table has also been fruitful (62). Here, students build a portfolio of information on a group of elements, and use the trends that they discover to predict the properties of yet to be discovered elements in the same group. This exercise is carried out with the whole class from anywhere between 30 and 60 students.

91 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Research Projects – History and Progression at Abington

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

Synthetic Chemistry Upon joining PSU in 2003, one of the coauthors (KCC) initially attempted to incorporate undergraduate researchers into an ongoing collaboration with Professor Grant Krow at Temple University. The intention was to have Penn State undergraduates stereoselectively synthesize and purify precursors to conformationally constrained pyrrolidine analogs (methanopyrrolidines) which would be used by graduate students in Krow’s research group. The close proximity of Temple University to the Penn State Abington campus facilitated the collaboration and provided undergraduates exposure to graduate-level research. Penn State undergraduates were able to interact regularly with graduate students and eventually collaborated on a poster which was presented at a regional ACS meeting (63, 64). However, after two years into this collaboration the research and required lab techniques became too sophisticated for most undergraduate researchers, and so more viable projects focused on heterocyclic compounds were pursued. One synthetic organic chemistry project which served as a basis for several current synthetic projects was realized by collaboration with an ongoing research effort by another of the coauthors (JT). As described above, the synthesis of substituted 1,3-thiazolidin-4-ones could be prepared and purified by second-year organic chemistry students easily. Undergraduate researchers at Penn State Abington synthesized the 2-(m- and p-substituted phenyl)-3-cyclohexyl-1,3-thiazolidin-4-one series (Series 8, Figure 3) (65). These syntheses are currently incorporated into Chem. 213 as part of a practical examination. The 1,3-thiazolidin-4-ones produced in the class now serve as starting materials for expanded research projects described below. The 1,3-thiazolidin-4-ones have been used to produce organotin complexes which have demonstrated biological activity (66, 67). The 2-(m- and p-substituted phenyl)-3-cyclohexyl-1,3-thiazolidin-4-one series has been reacted with Ph3SnCl to produce 1:1 thiazolidin-4-one tin complexes in which the thiazolidin-4-one ligand is coordinated to the Sn metal center via the oxygen atom of the C4 carbonyl (Figure 4). Again, this is simple chemistry that the undergraduates can do.

Figure 4. Preparation of organotin complexes of 1,3-thiazolidin-4-ones. 92 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

Currently, physical organic chemistry studies utilizing IR and NMR data are being conducted, further incorporating material that the students have learned in the lecture portion of the class. As stated previously, spectroscopic data was used to track the chemical shift variation of the methine proton (at C2), methylene protons (at C5) and the 13C NMR chemical shift data at C2, C4 and C5 in the 1,3thiazolidin-4-one ring. Although this work has yet to be presented outside of the university, the authors anticipate several publications with the two undergraduate authors. The syntheses of the 3-cyclohexyl-1,3-thiazolidin-4-ones was expanded to include the o-substituted phenyl derivatives, which were then subjected to Oxone® oxidation to produce the corresponding sulfoxide or sulfone compounds (Figure 5) (68, 69).

Figure 5. Oxidation of 3-cyclohexyl-1,3-thiazolidin-4-ones.

This research originally started as a physical chemistry project with an undergraduate researcher synthesizing the o-substituted phenyl derivatives to determine the effect of these substituents on the thiazolidin-4-one ring conformation as determined by x-ray analysis performed in collaboration with Assistant Professor Mike Zdilla at Temple University. This investigation is still in progress. The scope of the investigation was further expanded to a synthetic chemistry project when undergraduates at both Penn State Abington and Penn State Schuylkill campuses showed that selective oxidation of thiazolidin-4-ones to the corresponding sulfoxides was realized using Oxone® at room temperature. The Abington group additionally showed that the 3-cyclohexyl-1,3-thiazolidin-4-ones could be further oxidized to sulfones selectively at high temperature by increasing the equivalents of Oxone®; the extent of this selectivity was affected by the substituent of the aromatic ring, and in several cases the sulfone was formed exclusively. Other oxidants evaluated included NaIO4, KMnO4, and MCPBA; NaIO4 favored sulfoxide formation under all conditions evaluated, while KMnO4 and MCPBA formed sulfones exclusively at room temperature for all thiazolidin-4-ones. This work was recently published, with two student coauthors, one undergraduate and one graduate (69). Work continues in this area to determine if substituted phenyl derivatives of 3-phenyl-1,3-thiazolidin-4-ones likewise exhibit variable Oxone®-based oxidation selectivities. Furthermore, the original x-ray crystallographic study of the thiazolidin-4-ones is being expanded to include the corresponding sulfoxides and sulfones as well. Another synthetic project in progress at Penn State Abington involves the synthesis of purported hydrazones having chromogenic groups; a series 93 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

of these compounds had been previously prepared by the condensation of isoquinolyl-1-hydrazine with several heterocyclic 2-carboxaldehydes (70). Although heterocyclic aldehydes typically show an overwhelming preference for E-hydrazone stereoisomers, both Z- and E-stereoisomers once formed can be isolated (Figure 6) (71).

Figure 6. Synthesis of chromogenic hydrazones. The series of isoquinolyl-1-hydrazones were all reported as Z-stereoisomers and complexed with iron(II), cobalt(II), nickel(II), and copper(II) (70). Metal complexes were characterized by color, molar absorptivity, and wavelength of maximum absorbance; however, no structural analyses were reported for either the hydrazones or the metal complexes. The initial intent of this research project was to investigate the structural possibilities of the isoquinolyl-1-hydrazones. The 2-pyrryl-isoquinolyl-1-hydrazone was initially targeted for synthesis and characterization because it seemed possible that hydrogen bonding between the proton on the pyrryl nitrogen and the nitrogen of the isoquinolyl fragment might lead to a preference for the Z-hydrazone (Figure 7).

Figure 7. Plan for synthesis of 2-pyrryl-isoquinolyl-1-hydrazone. The synthesis of the 2-pyrryl-isoquinolyl-1-hydrazone was repeated according to the published procedure (70) and a crystalline solid with a melting point of 200-201 °C was similarly produced in reasonable yield (63%). However, x-ray analysis done in collaboration with the Zdilla Research Group at Temple University of this solid determined that the product was a hydrazone tautomer, an E-amidrazoneazine, herein referred to as an azine (Figure 8) (72). 94 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

Figure 8. Synthesis of an E-amidrazoneazine. To determine the scope of this unprecedented azine formation via proton tautomerism, six additional hydrazone syntheses previously reported were repeated under identical reaction conditions (Figure 9). In four of the syntheses, azine formation was again the outcome as determined by x-ray analysis, with the products being the E-amidrazoneazine.

Figure 9. E-amidrazoneazines prepared. In two reproduced syntheses that involved isoquinolyl aldehydes (70), an unanticipated oxidation took place with both aldehydes to yield 1,2,4-benzotriazoles as the primary products, which were also characterized by x-ray analysis (Figure 10). Manuscripts which include at least one undergraduate author each detailing these results are currently being written.

Figure 10. 1,2,4-benzotriazoles prepared. In addition to an introduction to x-ray crystallography, students were able to apply knowledge of bond length/bond order relationships that they learned in lecture. The unexpected formation of the azine compounds enabled the expansion 95 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

of this undergraduate research project to include electronic structure calculations using density functional theory (DFT). This aspect of the project, overseen by Professor Ann Schmiedekemp of Penn State Abington, applied electronic structure calculations to six of the compounds originally reported to be hydrazones. For each compound, calculations were applied to multiple conformations (typically five) for both the azine and hydrazone tautomers in order to determine the relative energies of the possible conformations. The calculations were performed using the hybrid functional B3LYP and a 6-31G**++ basis set using the program Jaguar (73). The diffuse functions on the basis set were chosen to provide better modeling of the hydrogen bonding interactions (74). The structures of each conformation were optimized in geometry until the forces on all the atoms were minimized. The frequencies of the normal modes of vibration were calculated with the available Hessian matrix, and all frequencies were positive which showed a minimum stationary (optimized) state. From this, the relative Gibbs free energy in kcal/mol of each conformation was computed at T = 298 K. The electronic energy of each optimized conformation was then compared among the possible conformations, and it allowed a ranking of the structures according to energy. The lowest Gibbs free energy conformation should be the favored structure in the gas phase. The lowest energy conformation was assigned a value of 0 kcal/mol, and any energy difference less than 2 kcal/mol was considered equivalent due to the accuracy of the calculation. The calculated Gibbs free energies for the azine and hydrazone conformations predicted three of the six compounds to be significantly favored as azine tautomers. Calculations predicted two of the compounds may be favored as hydrazone tautomers, although at least one azine conformation was within 2 kcal/mol for each. One calculation showed values for azine and hydrazone tautomers that were nearly equivalent. These results were presented at both a regional and national ACS meeting by an undergraduate researcher (72).

Pedagogy One non-traditional method that is used as a vehicle for engaging students in organic chemistry lecture is group competitive exercises (75). This method combines cooperative learning with competition. During these exercises, students must work together in instructor-selected groups of three to five. As with other cooperative learning exercises, these exercises seek to promote peer learning, social skills, accountability, and a sense of self-worth. The group becomes a vehicle for support and feedback in both preparation for, and participation in, the competitive exercises. The nature of these exercises also requires frequent active interaction with the instructor and competitors in the form of discussion and debate. The goal of group competitive exercises is to make students more active learners and promote effective subject- and problem-based learning. The exercises also provide students with an opportunity for non-graded evaluation with which they can benchmark their progress. Using group competitive exercises for the intended positive outcomes must be tempered with the avoidance of pitfalls 96 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

associated with traditional classroom competitions that alienate students. These pitfalls include promoting a classroom environment that encourages winning over learning, student rivalries, and rankings or ratings that alienate students. A stressful classroom for students both in and out of the “winner’s circle” must also be avoided. As part of a research project that was customized for an undergraduate who was pursuing a dual major of a) Science and b) Psychology & Social Science, a survey was developed and distributed to students in undergraduate organic chemistry and biochemistry courses at Penn State Abington who participated in group competitive exercises. The intent of the survey was to evaluate student attitudes toward group competitive exercises applied in their respective science courses to determine if the intended positive outcomes and the potential pitfalls associated with classroom competitions were realized. The survey consisted of thirty-one statements to which respondents indicated strongly agree, agree, not sure, disagree, or strongly disagree (that is, a five point Likert response scale (76)). The survey also provided students the opportunity to expand upon their answers to the statements. This data was collected for research purposes and used to augment and enhance classroom activities. Generally respondents indicated that the group competitive exercises were a positive academic and social experience (75). Despite a general positive attitude that was initially indicated by the students toward these exercises, subsequent concern about the existence of potential gender differences or biases toward group competitive exercises merited further investigation. Gender issues in math, science, and technology education related to classroom environments and teaching methodologies which can lead to female students experiencing less meaningful learning have been identified in several reviews (77–79). Extensive research on gender differences in college classroom learning, including that which specifically targets science achievement, indicates that competitive activities encourage a gender gap detrimental to female achievement (80, 81). On the other hand, some studies suggest that females benefit especially by the use of active and group-based pedagogies, whereas males prefer working independently (82, 83). Since group competitive exercises employ both a competitive environment and collaborative learning, evaluation of this particular instruction method with respect to gender differences was undertaken as a second undergraduate project. Data collection using the survey of student attitudes toward group competitive exercises was continued and eventually included survey responses collected between Spring 2007 and Fall 2011. The data were then analyzed to determine if any gender biases associated with the competition existed. Because the sample sizes of men and women were not equal (51.5% male, 48.5% female, 154 total respondents), a Mann-Whitney U test was performed to see if there were any significant differences between male and female respondents (84). Survey responses did not change significantly throughout the collection period. Overall respondents indicated that the group competitive exercises were a positive academic and social experience. For example, survey results overwhelmingly indicate that students agree that the exercises increase student engagement within the classroom, and over 45% of respondents indicated increased interactions with classmates outside of class time. Favorable opinions 97 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

toward beneficial learning outcomes were also evidenced by positive responses toward increased subject knowledge and effective preparation for tests and exams. Most potential negative outcomes were not considered very significant by the students. One exception, however, was that nearly 17% of respondents indicated that the exercises contributed to classroom rivalries (84). Most of the statement responses showed no significant difference between genders. Responses to two statements showed significance at the 0.05 level: more females than males disagreed that competitions could have been better organized, but more importantly more females than males indicated that they looked forward to announced competitions. Responses to three additional statements showed significance at the 0.1 level: more females than males agreed that competitions made the subject more interesting, competitions were interesting, and competitions were a good preparation for a test or exam. Combined, responses indicate a more positive female attitude toward competition within group competitive exercises, not an aversion. The authors speculate that the combination of cooperative learning and competitive environment simultaneously present in group competitive exercises mitigates the gender-differentiated behaviors reported previously. For female students, the collaborative learning component of the exercises nullifies any aversion to a competitive environment. For male students, the competitive aspect introduced into cooperative learning may make it more palatable. Most importantly, there was no instance in which the female response to a statement was significantly more negative towards group competitive exercises than the male response. Group competitive exercises can therefore be applied as a vehicle for student engagement and learning without contributing to a gender gap in science classroom achievement. A third pedagogical-based project investigated the correlation of perceived value, student efficacy, and learning environment associated with organic chemistry courses taught by one of the present authors (KCC). Motivation is the personal investment that an individual has toward reaching a desired state or outcome. In learning, motivation influences the direction, intensity, persistence, and quality of learning behaviors in which students engage. If students are not properly motivated, the likelihood of academic success decreases substantially. Although many theories have been offered to explain motivation, most accept that there are two core concepts (85). These are: The subjective value for achievement-related activities and goals. Students must associate value with course performance. Sources of value may include attainment value (satisfaction gained from mastery or accomplishment, e.g., a good grade), intrinsic value (satisfaction gained from doing the task rather than from the outcome of the task), and instrumental value (an activity that helps accomplish other important goals). Outcome efficacies. Students must hold positive expectations that they are capable of performing specific actions that will achieve a desired outcome. In addition to value and efficacies, motivation is also affected by students’ perceptions of a supportive or unsupportive classroom environment; motivation is enhanced if students perceive the environment as supportive. Motivated student behavior is only realized when value is perceived, student efficacy is high, and a supporting learning environment exists (85). 98 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

As is the case at most institutions, organic chemistry has a reputation among students as a challenging, “gate-keeper” prerequisite course for multiple programs including but not limited to pre-medical, pre-dental, pre-pharmacy, biology, biochemistry, and chemistry. Many find organic chemistry challenging because skill sets such as spatial visualization, pattern recognition, and critical thinking must be applied by students who typically rely on memorization in other introductory science courses. Also, organic chemistry content is cumulative, and failure to understand previously presented concepts will significantly impact subsequent learning negatively. Therefore, the presence of proper motivation is critical to increase a student’s likelihood of academic success in this traditionally difficult course. Although individual factors that affect student motivation have been associated with academic performance in organic chemistry, correlation of the three combined factors has not been addressed. The correlation was evaluated using 120 student responses to statements pertaining to the factors in a twelve statement Likert survey conducted at the midpoint of four organic chemistry courses (86). Pair-wise correlations between student responses were calculated. Correlation was deemed significant at the 0.05 level (2 tailed). Correlations were expected among the statements measuring the same attribute (course value, self-efficacy, and class environment). Additionally, if students are to perform well, all three attributes must be present and that presence would lead to correlation between all statements. Responses to value perception statements suggested that multiple sources of value are operating in combination (86). Students overwhelmingly agreed (79.2%) that mastering organic chemistry will help accomplish other important goals, indicating that it has instrumental value. Perceived instrumental value was also supported when 76.7% of the respondents disagreed with the statement that the course material is not relevant to their intended field of study. Attainment value was indicated when nearly 62% of the respondents agreed that a good grade is their primary goal; only 15.8% disagreed with this contention. Only 22.5% of the respondents agreed with the statement that mastering the material in organic chemistry is less important than receiving a good grade. It is difficult to reconcile that 57.5% of the respondents disagreed that mastering the course material was less important than a good grade when 61.7% indicated agreement that a good grade was a primary goal unless mastery and grades are valued similarly. Among efficacy perception statements, students overwhelmingly indicated self-efficacy. Over 79% of the respondents agreed that they expected to receive a grade of B- or better (only 50% of the students completing the course realized such a grade), and over 76% agreed that they generally perform well in science courses. There was no attempt in this study to correlate performance outcomes with perceived motivational factors. Less than 22% indicated that they lack the ability to do well in organic chemistry. In regard to environment, responses indicated a supportive environment exists: 82.5% agreed classmates are helpful, 93.3% agreed they possess sufficient resources, 97.5% agreed classroom attendance is helpful, and only 21.7% indicated intimidation about asking questions in lecture, respectively. Interestingly, responses concerning the benefit of previously attended college courses may support the notion that organic chemistry is significantly different than other introductory science courses. To the statement, 99 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

"Previous courses in my college career have not sufficiently prepared me to learn organic chemistry," 34.8 % disagreed to some extent and 41.5 % agreed to some extent. Thus, only 34.8 % indicated that they felt previous classes prepared them for organic chemistry, including two semesters of general chemistry which are a prerequisite at PSU.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

Chemical History: Investigation into the Discovery of Carbon Monoxide and the Role of Joseph Priestley The late 18th Century was a period of great chemical discovery. Especially noteworthy was the discovery of gaseous elements and compounds which began in 1754 when Joseph Black identified what he called "fixed air" (now known as carbon dioxide). One prolific discoverer of gases during this period was Joseph Priestley who is credited with the identification of ammonia, hydrogen chloride, hydrogen sulfide, nitrous oxide, nitrogen dioxide, oxygen, silicon tetrafluoride, and sulfur dioxide. Another discovery that is often credited to Priestley is carbon monoxide. In collaboration with Ms. Mary Ellen Bowden of the Chemical Heritage Foundation in Philadelphia, an undergraduate research project was established to determine Priestley’s role in identifying carbon monoxide, and whether his role was sufficient to constitute “discovery.” This project first addressed the attribution of discovery. Two undergraduate researchers conducted a literature search from 1820 to the present to determine who is most often cited as the discoverer of carbon monoxide. In nearly 300 citations (books/journals/articles), Priestley was identified as either the sole discoverer or contributing discoverer in 76% of the citations. The next most frequently cited discoverers were J.M. François de Lassone and William Cruickshank in 17% and 7% of the citations, respectively. The published experiments and discussions related to carbon monoxide by Priestley, de Lassone, and Cruickshank were then reviewed by a group of six undergraduates, and it was decided to also include the work of James Woodhouse as well. Students had to familiarize themselves with antiquated terminology (both English and French) and experimental methodologies of the late 18th Century to competently interpret the literature in a modern chemical perspective. For example, carbon monoxide is not identified by this name until 1801 by Cruickshank even though it was first produced by Priestley in 1774 (87, 88). Priestley first characterized carbon monoxide as “an inflammable air … burns blue, and not at all like that which is produced from iron, or any other metal, by means of an acid (87).” Priestley eventually designated carbon monoxide as “heavy inflammable air” to distinguish it from “light inflammable air” (hydrogen, discovered by Henry Cavendish in 1766), which was produced by the reaction of metals with acids. In 1776, de Lassone characterized carbon monoxide produced by the reaction of charcoal and zinc oxide and Prussian Blue as a flammable gas distinct from hydrogen that “fired quietly without noise (89).” Students then reviewed the six volumes of Priestley’s Experiments and Observations on Different Kinds of Air to determine how Priestley characterized the physical and chemical properties of “heavy inflammable air” which he 100 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

typically produced via the reaction of charcoal with iron cinders (mixed ferrous and ferric oxides, Fe3O4) at high temperatures. By 1780, Priestley reported crude density experiments showed that while hydrogen gas was 1/10 the mass of common air (it is actually 1/14), carbon monoxide measurements varied between slightly heavier to half that of common air. He also reported that reaction of carbon monoxide with oxygen produced carbon dioxide, whereas reaction between hydrogen and oxygen produced water (Figure 11).

Figure 11. Experiments by Priestley.

The three volumes of Priestley’s Experiments and Observations relating to the Various Branches of Natural Philosophy with a Continuation of the Observations on Air are currently being reviewed for additional characterizations of carbon monoxide by Priestley. A literature search suggested that de Lassone ceased work in this field, but his work would be revisited by Lavoisian chemists to identify the heavy inflammable gas. What tarnishes the attribution of carbon monoxide’s discovery to Priestley was his refusal to accept that his heavy inflammable air was an oxide of carbon. Lavoisier and his followers eventually identified “light inflammable air” as hydrogen, and purported that it could only be produced by decomposing water. Priestley sought to use his anhydrous production of heavy inflammable air to discredit Lavoisier’s notion of connection between inflammable gases and water, and to likewise defend the existence of phlogiston (90–92). This spurred responses from several Lavoisian chemists including John Maclean, Louis-Bernard Guyton, Woodhouse, and Cruickshank who repeated and interpreted Priestley’s and de Lassone’s experiments. In 1801, Cruickshank published an article which correctly concluded that the reaction of iron cinders and charcoal produced two oxides of carbon: CO2 and CO (93). Priestley disputed these findings, claiming he could not reproduce them, despite publishing the nearly identical experiment previously (90–92). Priestley also challenged the notion that an oxide could be flammable, and doubted that two forms of carbon oxide could exist. The information collected from this investigation has been presented at the Joseph Priestley House in Northumberland, Pennsylvania on several occasions, 101 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

and the research findings has been forwarded to the Friends of the Joseph Priestley House to update their literature and presentations available to visitors at the museum. Students also gained an appreciation of challenges faced by researchers of this era, especially related to the lack of both available experimental apparatus and of a scientific knowledge foundation. Critical thinking skills were also challenged in defining what constitutes discovery, both in general terms and as applied to the discovery of a naturally occurring compound such as carbon monoxide.

Chemistry in Science Fiction: Is Chemistry’s Portrayal Changing with Time? The purpose of this ongoing project is to survey how chemistry is presented in science fiction and how its presentation may have changed in the last 32 years. This investigation uses Gardner Dozois’s The Year’s Best Science Fiction as the data base (94). Dozois has compiled this science fiction anthology, which averages approximately 30 short stories per volume every year since 1984. Dozois’s credentials in the science fiction field testify to his expertise in this genre. Dozois is the recipient of numerous awards, including multiple Hugo and Nebula awards as both an editor and as a writer; he was editor of Asimov’s Science Fiction magazine from 1984 to 2004 and was also inducted into the Science Fiction Hall of Fame in 2011. His long-running anthology therefore provides a reasonably consistent resource which enables students to track how chemistry is portrayed in short-story science fiction over the last three decades. The breadth of this survey requires the contribution of multiple undergraduate researchers; to date, five undergraduate researchers have participated in this project. Typically, each researcher will review a minimum of two volumes of the anthology over the course of two semesters. In order to make data collection among multiple researchers more consistent, an exercise is conducted at the beginning of the academic year in which researchers review a common short story individually and then evaluate it according to the criteria discussed below. As a group, researchers compared their individual evaluations to reach a common consensus as to what constitutes a chemical reference and how the criteria were to be applied. Researchers are then assigned volumes which they can review at their own pace. Throughout the academic year, the researchers meet monthly as a group to discuss progress and issues associated with the reviews. Near the end of the year, results from the individual researchers are collected and then analyzed. Researchers are required to record any reference to a chemical species or a chemical process that appear in the short stories of their assigned volumes. These citations are recorded by volume, story title, page, and author. From this data, students can track both the total number of citations and the percentage of short stories with chemistry citations in a particular volume. As data from multiple volumes are reviewed, trends in both the number and the distribution of citations as a function of time can be analyzed. Each individual citation is then further evaluated according to four criteria: How critical is the citation to the story’s plot? The chemistry is defined as either a major or a minor contribution to the plot. 102 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

Figure 12. Utilization of chemistry in the plots of science fiction stories. Is the chemistry portrayed in the story as useful or not useful? Is the chemistry presented in positive, negative, or neutral circumstances? Is the chemistry presently feasible or fictional? For example, the following passage from Video Star by Walter Jan Williams was cited as a chemical reference by a researcher (95): “The next day Ric went to the drugstore, where he purchased a large amount of petroleum jelly, some nasal mist that came in squeeze bottles, liquid bleach, a bottle of toilet cleaner, a small amount of alcohol-based lamp fuel, and a bottle of glycerin. They drove to a chemical supply store, where he brought some distilling equipment and some litmus paper.” This reference was deemed critical to the story plot. These materials were needed to produce a particular bomb required for a robbery which is the major event in the story. Useful: In the story, the chemicals served their intended purpose to make a functional bomb which was successfully detonated. Negative: The chemicals were used for a bomb to rob a hospital and kill conspirators. Feasible: Most of the chemicals cited in the story can be used as bomb ingredients. It is the evaluation of the last point where intended chemical education takes place. Even with a prior interest in science fiction, most students are not oriented toward the “hard science” category of the genre. Thus, by having students evaluate the chemical validity of these short stories, students will learn some new chemical concepts. More importantly, however, they may develop a greater interest in chemistry and/or become analytical toward the scientific issues such as global warming or fracking that they will encounter outside the classroom. The data from each volume are then compiled and analyzed to determine if any trends are evident. To date, only 13 of the 32 current volumes have been reviewed. Within this limited data set, science fiction is certainly not necessarily chemistry-related. The data shows that on average only 57% of the stories possess at least one chemical reference, and there does not appear to be any particular trend over time in the number of chemistry citations from the anthologies. For example, 103 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Figure 12 shows data related to the utilization of chemistry as a major component in the plot versus a minor component. The lack of trends in the targeted categories may be accounted for by the diversity in number and scientific background of the authors in each volume. There was no noticeable trend in the number of authors that used chemistry in the reviewed volumes, and only eight authors who used chemistry appeared in multiple volumes.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

Research Projects – History and Progression at Schuylkill Upon joining PSU 2009, one of the authors (LJS) opened two lines of synthetic research. The first was a progression from some research done while working as a process chemist in the pharmaceutical industry (96). In that work, the very reactive epoxides of (+)-2-carene had been studied. The goals of the new research were to study the aziridines and halonium ions of (+)-2-carene and of α- and β-pinene (Figure 13).

Figure 13. Cyclopropyl and cyclobutyl aziridines and halonium ions. X = NTs, Br+, Cl+ The results of these studies were positive, and undergraduates (thirteen in all) were able to do the work. However, there were issues that made these projects less than ideal. The aziridines were difficult to make, the yields were low, and they were oils. Some very interesting results were obtained with the bromonium ions, but a large graduate research group also found it and published it first (97, 98). 104 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

In order to make that publishable, many more experiments had to be done, so all in all, too much valuable time was used on this. This demonstrated the importance of not doing research in an area where graduate groups might be. Nonetheless, the research was productive enough that three students presented posters at the 2011 ACS MARM (99–101), the coauthor delivered two oral presentations at the 2013 National ACS meeting (102, 103), and two papers, with 11 student coauthors, were published (104, 105). The more fruitful line began in collaboration with the other present coauthors. The initial work was on preparation of a series of 2,3-diaryl-1,3-thiazolidin-4-ones (Figure 3, Series 3). This resulted in a publication (four students) (56). Then studies of S-oxidation of 2,3-diaryl-1,3-thiazolidin-4-ones (Figure 3, Series 3) and 3-benzyl-2-phenyl-1,3-thiazolidinones (Figure 3, Series 5) were done (Figure 14). In this project (eight students), the lack of easy access to an NMR at that time caused some surprises. Using Oxone® and judging by TLC, it was believed that by controlling the temperature (0 °C or room temp.) and the amount of Oxone® (1.5 or 3.0 equivalents) either the sulfoxide or sulfone could be produced. However, when the NMRs were eventually obtained, it turned out that all of the reactions had been highly selective towards the sulfoxide. The full series has not been published, but a paper detailing the crystal structures of one sulfide and its corresponding sulfoxide has been published (one student author) (106). As discussed earlier, the work has also been extended at Penn State Abington.

Figure 14. S-Oxidation of 1,3-thiazolidin-4-ones. The real success began when the work was branched out to making six-membered 2,3-diaryl-1,3-thiazin-4-ones (Figure 15).

Figure 15. 2,3-Diaryl-1,3-thiazin-4-ones The initial attempts focused on making 2,3-diaryl-2,3-dihydro-4H-1,3benzothiazin-4-ones. 2,3-Diphenyl-2,3-dihydro-4H-1,3-benzothiazin-4-one, which had been reported in the literature but prepared by other methods, did 105 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

not form to an appreciable extent when the imine was refluxed in toluene or xylenes with the thioacid, as is done to make the 1,3-thiazolidinone. This was disappointing but not surprising, as it had been documented in the literature that the six-membered rings were more challenging to make by this method than the five-membered rings and that the N-aryl rings were more difficult to prepare than the N-alkyl rings (47, 107, 108). Several promoters that had been successfully used in some examples in the literature (109–111) failed as well, but it was found that 2,4,6-tripropyl-1,3,5,2,4,6-trioxatriphosphorinane-2,4,6-trioxide (T3P) (112–115) worked very well (Figure 16). Taylor has also independently reported an example of an N-aryl-1,3-benzothiazin-4-one using T3P (115).

Figure 16. Synthesis of 2,3-diaryl-2,3-dihydro-4H-1,3-benzothiazin-4-ones.

This procedure has proven to be very general. Synthesis of Series 1 (R1 = H; = m- and p-F, Br, Me, OMe, CF3, NO2) is complete and Series 2 (R1 = m- and p-F, Br, Me, OMe, CF3, NO2; R2 = H) is nearly complete (20 students). A variety of other ring systems, including a seven-membered ring, have also been prepared (Figure 17). All of these compounds are solids, and in collaboration with Dr. Hemant Yennawar at the University Park campus, x-ray crystal structures have been obtained of all those in Figure 17, and some in Figure 16. These structural studies have become a major focus of the research. They give students exposure to growing high quality crystals, a definitive structure, and an exciting endpoint that becomes immediately publishable. Nine journal articles have been published so far on structural studies of 2,3-diaryl-1,3-thiaza-4-ones (106, 116–123). R2

106 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

Figure 17. 2,3-Diphenyl-1,3-thiaza-4-one heterocycles prepared.

Another example of the pitfalls that can occur when one does not have an NMR immediately available is the following. One of the present authors (JT) and students made a series of compounds by the reaction of anilines, benzophenone, and thioglycolic acid. These sat on a shelf, unanalyzed and unpublished, for a decade. Upon receiving a sample of the presumed 2,2,3-triphenyl-1,3-thiazolidinone, another author (LJS) grew crystals and obtained an x-ray crystal structure. It turned out that the product included no benzophenone units, two units from aniline, and two units of the thioacid, held together by a disulfide bond (Figure 18).

Figure 18. Outcome of Brandywine attempt at cyclizing of 2,2,3-triphenylimine.

107 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

Subsequently, the imine was prepared at Schuylkill, and this proved to be more difficult than usual because of steric hindrance. Typically, it takes less than 30 minutes in refluxing toluene to prepare a C-aryl-N-aryl imine (Figure 16). The triphenyl imine did not form in refluxing toluene until catalytic p-toluenesulfonic acid was added; two days later the reaction was still not complete. This suggests that the imine never formed when the earlier work (Figure 18) was done. Nonetheless, an acceptable amount of imine was isolated. However, an attempt to then prepare the 1,3-benzothiazin-4-one using T3P was unsuccessful, mainly returning unreacted imine. This is the first time this reaction has not worked for us (unpublished results, one student). Spectroscopic studies are also being carried out, even though the only instrument on campus is a UV/Vis spectrometer. NMR studies are being performed at the University Park campus by Carlos Pacheco, IR by Anthony Lagalante at Villanova University, and high resolution mass spectrometry (HRMS) by Phillip Smith at University Park. A paper compiling and comparing spectroscopic data of the compounds in Figure 17 has been published (four student coauthors) (124). Work is being completed (20 students) on two series of compounds in Figure 16 (Series 1: R1 = H. Series 2: R2 = H). Another aspect of the research that excites students is studies that have begun on biological activity. Anti-microbial testing has begun, with herbicidal and insecticidal planned. This is being done by Biology professor Rod Heisey on the Schuylkill campus. In fact, one of the chemistry research students is also doing the biological research. Along with the papers mentioned above and in progress, this work was presented orally at the Spring 2015 National ACS meeting (125) and the 2015 Science Symposium: Innovation of Science, Nanotechnology, Human Health and Environment for a Global Society at Coppin State University (126), in two posters at the 2014 Gordon Research Conference on Heterocyclic Compounds (127, 128), and in a poster at the 2014 XIII International Union of Crystallographers Congress (129). This research has been well received by the undergraduates. The chemistry is readily done using standard organic techniques. The compounds are novel and publishable, and will be further tested for possible applications. Since the interest is in the compounds themselves, yields are not critical. It has already been demonstrated that a wide diversity of heterocycles can be prepared using this methodology, and there is also much potential for performing reactions on the rings, as has been shown with the S-oxidations of 1,3-thiazolidin-4-ones. Synthesis of triphenyltin chloride complexes of the six-membered rings has already begun (two students) and one product’s identity has been confirmed by x-ray crystallography and published (Figure 19) (130). Pedagogical research can result from this research too. The imines used to make the heterocycles are usually prepared first (Figure 16). This is a nice reaction for the students to do and see as they can watch it proceed by collection of water in a Dean-Stark trap. One of the imine preparations done by a student was chosen and has been turned into a lab experiment for Chem. 213. The procedure has been accepted for publication (131). 108 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

Figure 19. Preparation of the triphenyltin chloride complex of 2,3-diphenyl-2,3,5,6-tetrahydro-4H-1,3-thiazin-4-one (130).

Conclusion Research at a two-year campus has its challenges for sure, but we look at it as a great opportunity to focus on educating younger students. They are still maturing, still deciding on their future, and faculty have a great opportunity to help them find their way. We probably won’t be advising them when they walk on the moon, but we will have been there on the launching pad. What a joy to have seen them off on their adventures!

References 1. 2. 3.

4.

5.

6.

7.

Lopatto, D. Survey of Undergraduate Research Experience (SURE): First Findings. Cell Biol. Educ. 2008, 3, 270–277. Undergraduate Professional Education in Chemistry; American Chemical Society: Washington, DC, 2015. Canaria, J.; Schoffstall, A. M.; Weiss, D. J.; Henry, R. M.; Braun-Saud, S. B. A Model for an Introductory Undergraduate Research Experience. J. Chem. Educ. 2012, 89, 1371–1377. Eagan, K. M., Jr.; Hurtado, S.; Chang, M. J.; Garcia, G. A.; Herrera, F. A.; Garibay, J. C. Making a Difference in Science Education: The Impact of Undergraduate Research Programs. Am. Educ. Res. J. 2013, 50, 683–713. Basu, A.; Lee, J.; Chapdelaine, A. The Advantages and Disadvantages of Embedding Undergraduate Research into the Curriculum. CUR Quarterly 2011, 32 (1), 3. Wenzel, T. J.; Larive, C. K.; Frederick, K. A. Role of Undergraduate Research in an Excellent and Rigorous Undergraduate Chemistry Curriculum. J. Chem. Educ. 2012, 89, 7–9. Brownell, J. E.; Swaner, L. E. Five High-Impact Practices: Research on Learning Outcomes, Completion, and Quality; Association of American Colleges and Universities, Washington, DC, 2010. 109 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

8.

9.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

10.

11.

12.

13.

14.

15.

16.

17. 18. 19. 20.

21.

22. 23.

24. 25.

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. Laursen, S.; Seymour, E.; Hunter, A. Learning Teaching and Scholarship: Fundamental Tensions of Undergraduate Research. Change: The Magazine of Higher Learning 2012, 44 (2), 30–37. Laursen, S.; Hunter, A.; Seymour, E.; Thiry, H.; Melton, G. Undergraduate Research in the Sciences: Engaging Students in Real Science; Jossey-Bass: San Francisco, CA, 2010. Randall, D. C.; Wilbur, F. H.; Burkholder, T. Two Models for an Effective Undergraduate Research Experience in Physiology and Other Natural Sciences. J. Adv. Physiol. Educ. 2004, 28, 68–72. Sadler, T. D.; Burgin, S.; McKinney, L.; Ponjuan, L. Learning Science through Research Apprenticeships: A Critical Review of the Literature. J. Res. Sci. Teach. 2010, 47, 235–256. Lindsay, H. A.; McIntosh, M. C. Early Exposure of Undergraduates to the Chemistry Research Environment: A New Model for Research Universities. J. Chem. Educ. 2000, 77, 1174–1175. Hutchison, A. R.; Atwood, D. A. Research with First- and Second-Year Undergraduates: A New Model for Undergraduate Inquiry at Research Universities. J. Chem. Educ. 2002, 79, 125–126. Ford, J. R.; Prudenté, C.; Newton, T. A. A Model for Incorporating Research into the First-Year Chemistry Curriculum. J. Chem. Educ. 2008, 85, 929–933. Lee, N. E.; Gurney, R.; Solzberg, L. Using Green Chemistry Principles as a Framework to Incorporate Research into the Organic Laboratory Curriculum. J. Chem. Educ. 2014, 91, 1001–1008. Carpenter, N. E.; Pappenfus, T. M. Teaching Research: A Curriculum Model That Works. J. Chem. Educ. 2009, 86, 940–945. Gaglione, O. G. Underground Existence of Research in Chemistry in TwoYear College Programs. J. Chem. Educ. 2005, 82, 1613–1614. Brown, D. R. Undertaking Chemical Research at a Community College. J. Chem. Educ. 2006, 83, 970–972. Brown, D. B.; Higgins, T. B.; Coggins, P. The Increasing Presence of Undergraduate Research in Two-Year Colleges. CUR Quarterly 2007, 28 (2), 24–28. Cejda, B. D.; Hensel, N. Undergraduate Research in Community Colleges: A Summary of the CUR/NCIA Conversations. CUR Quarterly 2008, 29 (1), 7–11. Brothers, M.; Higgins, T. Research Models that Engage Community College Students. CUR Quarterly 2008, 29 (1), 12–17. Brandt, L. S. E.; Hayes, J. L. Broader Impacts of Undergraduate Research at a Community College: Opening Doors to New Ideas. CUR Quarterly 2012, 33 (1), 17–21. Slezak, J. Research at Two-Year Colleges. J. Chem. Educ. 2007, 84, 1108. Diaz-de-Mera, Y.; Notario, A.; Aranda, A.; Adame, J. A.; Parra, A.; Romero, E.; Parra, J.; Muñoz, F. A. Research Study of Tropospheric Ozone 110 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

26.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

27.

28.

29. 30. 31.

32.

33. 34. 35.

36. 37. 38. 39. 40. 41.

42.

and Meteorological Parameters To Introduce High School Students to Scientific Procedures. J. Chem. Educ. 2011, 88, 392–396. Kennedy, B. J. Integrating Advanced High School Chemistry Research with Organic Chemistry and Instrumental Methods of Analysis. J. Chem. Educ. 2008, 85, 393–395. Pueyo, C. N.; Raub, A. G.; Jackson, J.; Metz, M. M.; Mount, A. C.; Naughton, K. L.; Eatin, A. L.; Thomas, N. M.; Hastings, P.; Greaves, J.; Blumberg, B.; Collins, T. J.; Sogo, S. G. Oxidation of Ethidium Using TAML Activators: A Model for High School Research Performed in Partnership with University Scientists. J. Chem. Educ. 2013, 90, 326–331. Ball, D. B.; Wood, M.; Lindsley, C.; Mollard, P.; Buzard, D. J.; Vivion, R.; Mahoney, M.; Taft, B. R. Research, Teaching, and Professional Development at a Comprehensive University. J. Chem. Educ. 2004, 81, 1796–1800. Jensen, W. B. Why Study History of Chemistry? Tecne, Episteme y Didaxis 2011, 29, 4–7. Mendelsohn, A. Why Study the History of Science, Medicine, and Technology? History Today 2003 (2), 53. Willis, C. Science in Science Fiction. In Chemistry and Science Fiction; Stocker, J. H., Ed.; American Chemical Society: Washington, DC, 1998; pp 21−32. Chastain, B. B. Beryllium, Thiotimoline, and Pate de Foie Gras. Chemistry in the Science Fiction of Issac Asimov. In Chemistry and Science Fiction; Stocker, J. H., Ed.; American Chemical Society: Washington, DC, 1998; pp 77−90. Schneider, T. Science & Fiction Isaac Asimov. Chemie in Unserer Zeit 2007, 41 (2), 80–85. Bataille, X.; Dupuy, L. The Extraordinary Chemistry of Jules Verne. Actualite Chimque 2007, 304, 30–36. O’Brien, J., Sherlock Holmes. The Eccentric Chemist. In Chemistry and Science Fiction; Stocker, J. H., Ed.; American Chemical Society: Washington, DC, 1998; pp 105−126. Ober, J.; Krebs, T. Chemical Elements in Fantasy and Science Fiction. J. Chem. Educ. 2009, 86, 114. Foster, N. Science and the Final Frontier: Chemistry and Star Trek. Chemistry & Industry 1992, 24, 947–949. Chaloner, P. A. In Chemistry and Science Fiction; Stocker, J. H., Ed.; American Chemical Society: Washington, DC, 1998; pp 173−204. Pence, H. In Chemistry and Science Fiction; Stocker, J. H., Ed.; American Chemical Society, Washington, DC, 1998; pp 241−249. Nanny, M. A. In Chemistry and Science Fiction; Stocker, J. H., Ed.; American Chemical Society: Washington, DC, 1998; pp 53−76. Murphy, C. J.; Mogus, M. A.; Crotty, P. H. In Chemistry and Science Fiction; Stocker, J. H., Ed.; American Chemical Society: Washington, DC, 1998; pp 241−250. Liberko, C. A. Using Science Fiction to Teach Thermodynamics: Vonnegut, Ice-nine, and Global Warming. J. Chem. Educ. 2004, 81, 509–512. 111 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

43. Frederking, V.; Schwedt, G.; Kometz, A. Chemistry, Language, and Literature: Trans-disciplinary Perspective of Chemistry and German Lessons. Naturwissenschaften im Unterricht Chemie 2013, 24 (138), 2–8. 44. Katz, D. A. Science and Science Fiction in the Classroom: an Update. Presented at National Meeting of the American Chemical Society, San Diego, CA, March 25−29, 2012. 45. Gelius, R. Chemistry and Science Fiction: Transplutum, Wroff, and Orbital Elevators. Chemie in Unserer Zeit 2012, 46 (5), 302–308. 46. Tierney, J.; Costello, F.; Dalton, D. R. Hydrochlorination of Alkenes. 3. Reaction of the Gases Hydrogen Chloride and (E)- and (Z)-2-Butene. J. Org. Chem. 1986, 51, 5191–5196. 47. Tierney, J. The Formation of 2,3-Disubstituted Thiazolidin-4-ones from Sα′-Aminomercaptoacetic Acid Derivatives. J. Heterocycl. Chem. 1989, 26, 997–1001. 48. Surrey, A. R. The Preparation of 4-Thiazolidones by the Reaction of Thioglycolic Acid with Schiff Bases. J. Am. Chem. Soc. 1947, 69, 2911–2912. 49. Tierney, J.; Houghton, G.; Sanford, K.; Mascavage, L.; McCoy, M.; Findeisen, A.; Kilburn, J. Effects and Conformational Analysis of Some Substituted 2,3-Diphenyl-1,3-thiazolidin-4-ones. Magn. Reson. Chem. 1996, 34, 573–576. 50. Issac, R.; Tierney, J.; Mascavage, L. M.; Findeisen, A. Synthesis, Spectroscopic and Structural Properties of Novel Substituted 2-Trichloromethyl-3-Phenyl-1,3-thiazolidin-4-ones. Heterocycl. Commun. 1996, 2, 227–232. 51. Tierney, J.; Sheridan, D.; Kovalseky, C.; Mascavage, L. M.; Sheely, C.; Schneider, K. Substituent Effects for Some Substituted 3-Benzyl-2-phenyl1,3-thiazolidin-4-ones Using 1H and 13C NMR. Heterocycl. Commun. 2000, 6, 105–111. 52. Tierney, J.; Mascavage, L. M.; Sheridan, D.; Gorbecheva, D.; Ripp, M.; Son, S. A Preliminary Study on Predicting the 13C Chemical Shifts for a Series of Disubstituted 2,3-Diphenyl-1,3-thiazolidin-4-ones. Heterocycl. Commun. 2005, 11, 215–222. 53. Tierney, J.; Colasante, N.; Eagles, J.; Kelly, S.; Lehmicke, G.; Lucas, O.; Mannikarottu, C.; Mehmood, O.; Nguyen, M. D.; Rai, V.; Tsai, J.; Mascavage, L. M.; Lagalante, A. F. A Comparison of Substituent Effects Using 13C Chemical Shifts for a Series of Substituted 2-Phenyl-3-pyridyl-1,3thiazolidin-4-ones Compared to Similarly Substituted 1,3-thiazolidin-4-ones. Heterocycl. Commun. 2007, 13, 335–342. 54. Tierney, J.; Koyfmann, V.; Cannon, K.; Mascavage, L. M.; Lagalante, A. F. A Second Study - Predicting the 13C Chemical Shifts for a Series of Substituted-3-(4-methoxyphenyl)-2-phenyl-1,3-thiazolidin-4-ones. Heterocyclic Commun. 2008, 14, 453–460. 55. Tierney, J.; Cannon, K.; Mascavage, L. M.; Lagalante, A. F.; Kistler, K.; Sonnett, P.; Yennawar, H. An Experimental and Theoretical Conformational Study of a Series of Substituted 3-Cyclohexyl-2-phenyl-1,3-thiazolidin-4ones. Int. J. Chem. (Toronto, Ontario, Canada) 2013, 5, 46–56. 112 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

56. Silverberg, L. J.; Bear, E. R.; Foose, K. N.; Kirkland, K. A.; McElvaney, R. R.; Cannon, K.; Tierney, J.; Lascio, S.; Mesfin, K.; Mitchell, D.; Sharkey, S.; So, L.; Treichel, J.; Waxman, M.; Lagalante, A. Verifying the Predictability of 13C Chemical Shifts for a Series of Substituted-2-(4-chlorophenyl)-3-phenyl1,3-thiazolidin-4-ones. Int. J. Chem. (Toronto, Ontario, Canada) 2013, 5, 121–127. 57. McGarity, D.; Tierney, J.; Lagalante, A. An Anomalous Hammett Correlation for a Series of Substituted 3-Benzyl-2-phenyl-1,3-thiazolidin-4ones. Int. J. Chem. (Toronto, Ontario, Canada) 2014, 6, 12–18. 58. Mascavage, L. M.; Tierney, J.; Sonnet, P. E.; Dalton, D. R. The Hoffmann-Schiff Dichotomy: On the Reaction Between Chloral and Amines. ARKIVOC 2010 (viii), 278–284. 59. Messick, M. A. Production and Analysis of 2-(p-Hydroxyphenyl)-3-phenyl1,3-thiazolidin-4-one and 3-(p-Hydroxyphenyl)-2-phenyl-1,3- thiazolidin-4one. Schreyer Honors Thesis, Pennsylvania State University, 1999. 60. Doran, D.; Tierney, J.; Varano, M.; Ware, S. A Study of the pH of Perspiration from Male and Female Subjects While Exercising in the Gymnasium: A Practical Challenge for Students in the Non-science Major Class. J. Chem. Educ. 1993, 70, 412–414. 61. Tierney, J.; Peacock, L.; Sorkin, E.; Ware, S. A Study of the pH of Sweat for Horses: A Student Devised Study by Science and Health Science Majors. J. Chem. Educ. 1997, 74, 1135–1137. 62. Tierney, J. Forecasting Periodic Trends: A Semester-Long Team Exercise for Nonscience Majors. J. Chem. Educ. 2008, 85, 1215–1217. 63. Cannon, K. C.; Mathews, A.; Garcia, N. L.; Patel, C. K.; Centafont, R. A.; Krow G. R. Neighboring Group Carbamate Participation in the Synthesis and Reactions of Dibromo- and Bromo-hydroxy-2-azanorbornanes. Presented at 15th Biennial Philadelphia Organic Chemists’ Club Day, Philadelphia, PA, April 21, 2005. 64. Cannon, K. C.; Mathews, A.; Garcia, N. L.; Patel, C. K.; Centafont, R. A.; Krow G. R. Neighboring Group Carbamate Participation in the Synthesis and Reactions of Dibromo- and Bromo-hydroxy-2-azanorbornanes. Presented at American Chemical Society Middle Atlantic Regional Meeting, New Brunswick, NJ, May 23, 2005. 65. Cannon, K.; Mascavage, L.; Tierney, J.; Kistler, K.; Yennawar, H.; Lagalante, A. An Experimental and Theoretical Conformational Study of a Series of Substituted 3-Cyclohexyl-2-phenyl-1-thiazolidin-4-ones. Int. J. Chem. (Toronto, Ontario, Canada) 2013, 5, 46–56. 66. Smith, F. E.; Hynes, R. C.; Tierney, J.; Zhang, Y. Z.; Eng, G. The Synthesis, Molecular and Crystal Structure of the 1:1 Adduct of Triphenyltin Chloride with 2,3-Diphenylthiazolidin-4-one. Can. J. Chem. 1995, 73, 95–99. 67. Eng, G.; Whalen, D.; Musingarimi, P.; Tierney, J.; DeRosa, M. Fungicidal and Spectral Studies of Some Triphenyltin Compounds. Appl. Organomet. Chem. 1998, 12, 25–30. 68. Lauro, S.; Cannon, K. C. The Synthesis of Ortho-Substituted 3-Cyclohexyl2-aryl-1,3-thiazolidin-4-ones. Presented at Philadelphia American Chemical Society Poster Session, Drexel University, Philadelphia, PA, February 21, 113 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

69.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

70.

71.

72.

73. 74.

75.

76. 77.

78. 79.

80. 81.

82.

83.

2013, and Sigma Xi Student Research Day, Thomas Jefferson University, Philadelphia, PA, April 17, 2013. Cannon, K. C.; Gandla, D.; Lauro, S.; Silverberg, L. J.; Tierney, J.; Lagalante, A. Selective Synthesis of Ortho-substituted 3-Cyclohexyl-2phenyl-1,3-thiazolidin-4-one Sulfoxides and Sulfones by S-Oxidation with Oxone®. Int. J. Chem. (Toronto, Ontario, Canada) 2015, 7, 73–84. Case, F. H.; Schilt, A. A. Synthesis of Some Isoquinoyl-1-hydrazones and Spectrophotometric Characterization of Some of Their Transition-Metal Chelates. J. Chem. Eng. Data 1986, 31, 503–504. Butler, R. N.; Johnston, S. M. Stereoisomerization in Heterocyclic Hydrazones Derived from 2-Acylpyridines and Their Oxidative Cyclization with Mercury(II)acetate and Lead Tetra-acetate to Fused 1,2,4-Triazoles and 1,2,3-Triazolium Systems. J. Chem. Soc., Perkin Trans. 1984, 1, 2109–2116. Cannon, K.; Lauro, S.; Schmiedekamp, A.; Wang, C. Unprecedented Azine Formation via Proton Tautomerization of Isoquinolyl-1-hydrazones. Presented at American Chemical Society Middle Atlantic Regional Meeting, UMBC, Baltimore, MD, June 2, 2012, and National Meeting of the American Chemical Society, Philadelphia, PA, August 20, 2012. Jaguar, version 7.6; Schrodinger, LLC: New York, NY, 2009. Pudzianowski, T. A Systematic Appraisal of Density Functional Methodologies for Hydrogen Bonding in Binary Ionic Complexes. J. Phys. Chem. 1996, 100, 4781–4789. Cannon, K. C.; Mody, T.; Breen, M. P. Student Perceptions of Group-Based Competitive Exercises in the Chemistry Classroom. Educ. Res. Rev. 2008, 3, 73–76. Likert, R. A Technique for the Measurement of Attitudes. Arch. Psychol. 1932, 140, 1–55. Sanchez, C. A.; Wiley, J. Sex Differences in Science Learning: Closing the Gap Through Animations. Learning and Individual Differences 2010, 20, 271–275. Thompson, J. Effects of Constructivist Mathematics Instruction on Gender Perception and Motivation, 2005. www.msu.edu. Wehrwein, E. A.; Lujan, H. L.; DiCarlo, S. E. Gender Differences in the Learning Style Preferences Among Undergraduate Physiology Students. Adv. Physiol. Educ. 2007, 31, 153–157. Lorenzo, M.; Crouch, C. H.; Mazur, E. Reducing the Gender Gap in the Physics Classroom. Am. J. Phys. 2006, 74, 118–122. Miyake, A.; Kost-Smith, L. E.; Finklestein, N. D.; Pollack, S. J.; Cohen, G. L.; Ito, T. A. Reducing the Gender Achievement Gap in College Science: A Classroom Study of Values Affirmation. Science 2010, 33, 1234–1237. Krupnick, C. G. Women and Men in the Classroom: Inequity and Its Remedies. http://isites.harvard.edu/fs/html/icb.topic58474/krupnick.html. Reprinted from Teaching and Learning: J. Harvard Danforth Center 1985, 1 (1), 18−25. Milgram, D. Gender Differences in Learning Style Specific to Science, Technology, Engineering and Math (STEM), 2004, Tech Equity Project of 114 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

84.

85.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

86.

87. 88. 89. 90. 91. 92. 93.

94. 95.

96. 97.

98.

99.

100.

California State University, Channel Islands. http://www.iwitts.org/provenpractices/retention-sub-topics/learning-style/297. Cannon, K. C.; Cannon, K. A.; Breen, M. P. Women Do Not Like Competitive Environments in the Classroom? Group Competitive Exercises Imply Not So! Int. J. Sci. Society 2013, 4 (2), 119–126. Ambrose S. A.; Bridges M. W.; DiPietro M.; Lovett M. C.; Norman M. K. How Learning Works: Seven Research-based Principles for Smart Teaching; Jossey-Bass: San Francisco, CA, 2010. Cannon, K. A.; Breen, M. P.; Cannon, K. C. Measuring Correlations Between Factors that Affect Student Motivation in Organic Chemistry. Brit. J. Education, Society & Behavioral Science 2015, 5, 426–432. Priestley, J. Experiments and Observations on Different Kinds of Air; London, 1774; pp 37−39. Priestley, J. Experiments and Observations on Different Kinds of Air; London, 1774; pp 129−133. De Lassone, J. M. F. D’Une Suite D’Experiences Nouvelles. Memoires de L’Academie Royale 1779, 686–696. Priestley, J. A. Reply to Mr. Cruickshank’s Observations in Defence of the New System of Chemistry. Medical Repository 1802, 5, 390–392. Priestley, J. Remarks on Mr. Cruickshank’s Experiments upon Finery Cinder and Charcoal. Medical Repository 1803, 6, 24–26. Priestley, J. Additional Remarks on Mr. Cruickshank’s Experiments upon Finery Cinder and Charcoal. Medical Repository 1803, 6, 271–273. Cruickshank, W. Some Observations on Different Hydrocarbonates and Combinations of Carbone with Oxygen. J. Natural Philosophy, Chemistry, and the Arts 1801, 5, 1–9. The Year’s Best Science Fiction; Dozois, G., Ed.; St. Martin’s Press: New York, NY, 1984−2015 (present). Williams, W. J. Video Star. In The Year’s Best Science Fiction, Fourth Annual Collection; Dozois, G. R., Ed.; St. Martin’s Press: New York, NY, 1987; pp 439−468. Silverberg, L. J.; Resnick, T. M.; Casner, M. L. U.S. Patent 6,867,335, 2005. Ganesh, V.; Sureshkumar, D.; Chandrasekaran, S. Tandem Ring Opening/ Cyclization of Vinylcyclopropanes: A Facile Synthesis of Chiral Bicyclic Amidines. Angew. Chem., Int. Ed. 2011, 50, 5878–5881. Ganesh, V.; Sureshkumar, D.; Chanda, D.; Chandrasekaran, S. BromeniumCatalysed Tandem Ring Opening/Cyclisation of Vinylcyclopropanes and Vinylcyclobutanes: A Metal-Free [3+2+1]/[4+2+1] Cascade for the Synthesis of Chiral Amidines and Computational Investigation. Chem. Eur. J. 2012, 18, 12498–12511. Rabb, J. M.; Silverberg, L. J. Cyclopropyl Aziridines: Solvolytic Reactions of the Tosyl Aziridine of (+)-2-Carene. Presented at American Chemical Society Middle Atlantic Regional Meeting, University of Maryland, College Park, MD, May 21, 2011; Abstract 238. Brobst, K. A.; Silverberg, L. J.; Yennawar, H. Reactions of the Bromonium Ion of (+)-2-Carene: Formation of a Chiral Cyclic Amidine by Reaction with Acetonitrile and Chloramine T. Presented at American Chemical Society 115 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

101.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

102.

103.

104.

105.

106.

107.

108. 109.

110.

111.

112. 113.

Middle Atlantic Regional Meeting, University of Maryland, College Park, MD, May 21, 2011; Abstract 239. Celani, N. M.; Rabb, J. M.; Silverberg, L. J. Improved Process for the Preparation of PhINTs. Presented at American Chemical Society Middle Atlantic Regional Meeting, University of Maryland, College Park, MD, May 21, 2011; Abstract 243. Silverberg, L. J.; Rabb, J. M.; Reno, J. M.; Anderson, J. B.; Rus, J.; Harbaugh, T. E.; He, G. Cyclopropyl Aziridines: Solvolytic Reactions of the Tosyl Aziridine of (+)-2-Carene. Presented at National Meeting of the American Chemical Society, New Orleans, LA, April 7, 2013; Abstract ORGN 26. Silverberg, L. J.; Celani, N. M.; Aurentz, D. J. Improved Synthesis and Solid-state NMR of Useful Iodoimines. Presented at National Meeting of the American Chemical Society, New Orleans, LA, April 7, 2013, Abstract ORGN 108. Silverberg, L. J.; Rabb, J. M.; Reno, J. M.; He, G. Cyclopropyl Aziridines: Solvolytic Reactions of the N-Tosylaziridines of (+)-2-Carene and (+)-3-Carene. Heterocycl. Commun. 2014, 20, 193–199. Silverberg, L. J.; Kistler, K. A.; Brobst, K. A.; Yennawar, H.; Lagalante, A.; He, G.; Ali, K.; Blatt, A.; Foster, S. B.; Grossman, D. B.; Hegel, S.; Minehan, M. J.; Valinsky, D. M.; Yeasted, J. G. Reactions of the Halonium Ions of Carenes and Pinenes: An Experimental and Theoretical Study. Eur. J. Chem. 2015, 6, 430–443. Yennawar, H. P.; Hullihen, P. D.; Tierney, J.; Silverberg, L. J. Crystal Structures of 2,3-bis(p-Chlorophenyl)-1,3-thiazolidin-4-one and trans-2,3bis(p-Chlorophenyl)-1,3-thiazolidin-4-one-1-oxide. Acta Crystallogr. 2015, E71, 264–267. Surrey, A. R.; Webb, W. G.; Gesler, R. M. Central Nervous System Depressants. The Preparation of Some 2-Aryl-4-metathiazanones. J. Am. Chem. Soc. 1958, 80, 3469–3471. Loev, B. 2,3-Dihydro-4H-1,3-benzothiazin-4-ones. J. Org. Chem. 1963, 28, 2160. Zarghi, A; Zebardast, T.; Daraie, B.; Hedayati, M. Design and Synthesis of New 1,3-Benzthiazinan-4-one Derivatives as Selective Cyclooxygenase (COX-2) Inhibitors. Bioorg. Med. Chem. 2009, 17, 5369–5373. Zhou, H.; Liu, A.; Li, X.; Ma, X.; Feng, W.; Zhang, W.; Yan, B. MicrowaveAssisted Fluorous Synthesis of 2-Aryl-Substituted 4-Thiazolidinone and 4Thiazinanone Libraries. J. Comb. Chem. 2008, 10, 303–312. Kamel, M. M.; Ali, H. I.; Anwar, M. M.; Mohamed, N. A.; Soliman, A. M. M. Synthesis, Antitumor Activity and Molecular Docking Study of Novel Sulfonamide-Schiff’s Bases, Thiazolidinones, Benzothiazinones and Their C-Nucleoside Derivatives. Eur. J. Med. Chem. 2010, 45, 572–580. Wissmann, H.; Kleiner, H. J. New Peptide Synthesis. Angew. Chem., Int. Ed. 1980, 19, 133–134. Dunetz, J. R.; Xiang, Y.; Baldwin, A.; Ringling, J. General and Scalable Amide Bond Formation with Epimerization-Prone Substrates Using T3P and Pyridine. Org. Lett. 2011, 13, 5048–5051. 116 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

114. Unsworth, W. P.; Kitsiou, C.; Taylor, R. J. K. Direct Imine Acylation: Rapid Access to Diverse Heterocyclic Scaffolds. Org. Lett. 2013, 15, 258–261. 115. Kitsiou, C.; Unsworth, W. P.; Coulthard, G.; Taylor, R. J. K. Substrate Scope in the Direct Imine Acylation of Ortho-substituted Benzoic Acid Derivatives: The Total Synthesis (±)-Cavidine. Tetrahedron 2014, 70, 7172–7180. 116. Yennawar, H. P.; Cali, A. S.; Xie, Y.; Silverberg, L. J. Crystal Structures of 2-(4-Nitrophenyl)-3-phenyl-2,3-dihydro-4H-1,3-benzothiazin-4-one and 2-(2-Nitrophenyl)-3-phenyl-2,3-dihydro-4H-1,3-benzothiazin-4-one. Acta Crystallogr. 2015, E71, 414–417. 117. Yennawar, H. P.; Singh, H.; Silverberg, L. J. Crystal Structure of N-[(2S,5R)-4-Oxo-2,3-diphenyl-1,3-thiazinan-5-yl]acetamide 0.375 hydrate. Acta Crystallogr. 2015, E71, 62–64. 118. Yennawar, H. P.; Tierney, J.; Silverberg, L. J. 2,3-Diphenyl-1,3-thiazolidin4-one. Acta Crystallogr. 2014, E70, o847. 119. Yennawar, H. P.; Singh, H.; Silverberg, L. J. 2,3-Diphenyl-2,3-dihydro-4Hpyrido[3,2-e][1,3]thiazin-4-one. Acta Crystallogr. 2014, E70, o638. 120. Yennawar, H. P.; Bendinsky, R. V.; Coyle, D. J.; Cali, A. S.; Silverberg, L. J. 2,3-Diphenyl-2,3-dihydro-4H-1,3-benzothiazin-4-one. Acta Crystallogr. 2014, E70, o465. 121. Yennawar, H. P.; Silverberg, L. J. 2,3-Diphenyl-2,3,5,6-tetrahydro-4H-1,3thiazin-4-one. Acta Crystallogr. 2014, E70, o133; Corrigendum: 2015, E71, e5. 122. Yennawar, H. P.; Silverberg, L. J.; Minehan, M. J.; Tierney, J. 2-(3Nitrophenyl)-3-phenyl-2,3-dihydro-4H-1,3-benzothiazin-4-one. Acta Crystallogr. 2013, E69, o1679. 123. Yennawar, H. P.; Silverberg, L. J. 6,7-diphenyl-5-thia-7-azaspiro[2.6]nonan8-one. Acta Crystallogr. 2013, E69, o1659. 124. Silverberg, L. J.; Pacheco, C. N.; Lagalante, A.; Cannon, K. C.; Bachert, J. T.; Xie, Y.; Baker, L.; Bayliff, J. A. Synthesis and Spectroscopic Properties of 2,3-Diphenyl-1,3-thiaza-4-one Heterocycles,”. Int. J. Chem. (Toronto, ON, Canada) 2015, 7, 150–162. 125. Silverberg, L. J.; Yennawar, H. P.; Tierney, J.; Pacheco, C. N.; Cannon, K. C.; Lagalante, A.; Bachert, J. T.; Baker, L.; Bayliff, J. A.; Bendinsky, R. V.; Cali, A. S.; Chen, L.; Cooper, A. D.; Coyle, D. J.; Dahl, J. R.; Minehan, M. J.; Mroz, C. R.; Singh, H.; Xie, Y. Studies on Cyclic Six- and Seven-Membered 2,3-Diaryl-1,3-thiaza-4-ones. Presented at National Meeting of the American Chemical Society, Denver, CO, Mar. 25, 2015; Abstract ORGN 456. 126. Silverberg, L. J. Studies on 1,3-Thiaza-4-one Heterocycles. Presented at 3rd Science Symposium: Innovation of Science, Nanotechnology, Human Health and Environment for a Global Society. Coppin State University, Baltimore, MD, Sep. 18, 2015. 127. Silverberg, L. J.; Yennawar, H. P.; Cannon, K.; Pacheco, C.; Singh, H. Studies on Cyclic Six- and Seven-Membered 2,3-Diaryl-1,3-thiaza-4-ones. Presented at Gordon Research Conference on Heterocyclic Compounds, Newport, RI, June 15−20, 2014. 128. Silverberg, L. J.; Yennawar, H. P.; Tierney, J.; Pacheco, C.; Bendinsky, R. V.; Cali, A. S.; Coyle, D. J.; Minehan, M. J.; Singh, H.; Cooper, A. D.; 117 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by UNIV OF CALIFORNIA SAN DIEGO on December 12, 2016 | http://pubs.acs.org Publication Date (Web): November 22, 2016 | doi: 10.1021/bk-2016-1231.ch006

Chen, L. Studies on 2,3-Diaryl-2,3-dihydro-4H-1,3-thiazin-4-ones with a Fused Arene. Presented at Gordon Research Conference on Heterocyclic Compounds, Newport, RI, June 15−20, 2014. 129. Yennawar, H.; Silverberg, L.; Tierney, J.; Cannon, K.; Bendinsky, R.; Cali, A.; Coyle, D.; Minehan, M. Crystal Structures of 1,3-Thiaza-4-one Heterocycles. Presented at XIII International Union of Crystallographers Congress, Montreal, Canada, August 2014. Abstract published: Acta Crystallogr. 2014, A70 (Supplement), C1777. 130. Yennawar, H. P.; Fox, R.; Silverberg, L. J. Crystal structure of the 1:1 adduct of 2,3-diphenyl-2,3,5,6-tetrahydro-4H-1,3-thiazin-4-one with triphenyltin chloride. Acta Crystallogr. 2016, E72, 276–279. 131. Silverberg, L. J.; Coyle, D. J.; Cannon, K. C.; Mathers, R.; Richards, J. A.; Tierney, J. Azeotropic Preparation of a C-Phenyl N-Aryl Imine: An Introductory Undergraduate Organic Chemistry Laboratory Experiment. J. Chem. Educ., DOI: 10.1021/acs.jchemed.6b00056.

118 Murray et al.; The Power and Promise of Early Research ACS Symposium Series; American Chemical Society: Washington, DC, 2016.