Teaching chemistry with forensic science 9780841234826, 0841234825

2,357 201 12MB

English Pages 0 [243] Year 2019

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Teaching chemistry with forensic science
 9780841234826, 0841234825

Table of contents :
Introduction to teaching chemistry with forensic science --
Chemistry and crime : investigating chemistry from a forensic science perspective --
Incorporating forensic science throughout the undergraduate analytical curriculum : from nonmajors through instrumental analysis --
Using forensic science to engage nontraditional learners --
Teaching introductory forensic chemistry using open educational and digital resources --
On utilizing forensic science to motivate students in a first-semester general chemistry laboratory --
Interdisciplinary learning communities : bridging the gap between the sciences and the humanities through forensic science --
Interdisciplinary learning activity incorporating forensic science and forensic nursing --
Drugs and DNA : forensic topics ideal for the analytical chemistry curriculum --
From DUIs to stolen treasure : using real-world sample analysis to increase engagement and critical thinking in analytical chemistry courses --
Integration of forensic themes in teaching instrumental analysis at Pace University --
Using expert witness testimony with an illicit substance analysis to increase student engagement in learning the GC/MS technique --
Generative learning strategies and prelecture assignments in a flipped forensic chemistry classroom.

Citation preview

Downloaded via 109.94.175.201 on November 5, 2019 at 02:10:48 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Teaching Chemistry with Forensic Science

Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

ACS SYMPOSIUM SERIES 1324

Teaching Chemistry with Forensic Science Amanda S. Harper-Leatherman, Editor Department of Chemistry & Biochemistry, Fairfield University Fairfield, Connecticut, United States

Ling Huang, Editor Department of Chemistry, Hofstra University Hempstead, New York, United States

Sponsored by the ACS Division of Chemical Education

American Chemical Society, Washington, DC Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Library of Congress Cataloging-in-Publication Data Names: Harper-Leatherman, Amanda S., editor. | Huang, Ling, 1978- editor. Title: Teaching chemistry with forensic science / Amanda S. Harper-Leatherman, editor ; Ling Huang, editor. Description: Washington, DC : American Chemical Society, [2019] | Series: ACS symposium series ; 1324 | Includes bibliographical references and index. Identifiers: LCCN 2019033067 (print) | LCCN 2019033068 (ebook) | ISBN 9780841234833 (hardcover) | ISBN 9780841234826 (ebook other) Subjects: LCSH: Chemistry--Study and teaching (Higher)--United States--Case studies. | Chemistry, Forensic--Study and teaching--United States--Case studies. | Forensic sciences--Study and teaching--United States--Case studies. Classification: LCC QD47 .T43 2019 (print) | LCC QD47 (ebook) | DDC 540.71/173--dc23 LC record available at https://lccn.loc.gov/2019033067 LC ebook record available at https://lccn.loc.gov/2019033068

The paper used in this publication meets the minimum requirements of American National Standard for Information Sciences—Permanence of Paper for Printed Library Materials, ANSI Z39.48n1984. Copyright © 2019 American Chemical Society All Rights Reserved. Reprographic copying beyond that permitted by Sections 107 or 108 of the U.S. Copyright Act is allowed for internal use only, provided that a per-chapter fee of $40.25 plus $0.75 per page is paid to the Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. Republication or reproduction for sale of pages in this book is permitted only under license from ACS. Direct these and other permission requests to ACS Copyright Office, Publications Division, 1155 16th Street, N.W., Washington, DC 20036. The citation of trade names and/or names of manufacturers in this publication is not to be construed as an endorsement or as approval by ACS of the commercial products or services referenced herein; nor should the mere reference herein to any drawing, specification, chemical process, or other data be regarded as a license or as a conveyance of any right or permission to the holder, reader, or any other person or corporation, to manufacture, reproduce, use, or sell any patented invention or copyrighted work that may in any way be related thereto. Registered names, trademarks, etc., used in this publication, even without specific indication thereof, are not to be considered unprotected by law. PRINTED IN THE UNITED STATES OF AMERICA

Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Foreword The purpose of the series is to publish timely, comprehensive books developed from the ACS sponsored symposia based on current scientific research. Occasionally, books are developed from symposia sponsored by other organizations when the topic is of keen interest to the chemistry audience. Before a book proposal is accepted, the proposed table of contents is reviewed for appropriate and comprehensive coverage and for interest to the audience. Some papers may be excluded to better focus the book; others may be added to provide comprehensiveness. When appropriate, overview or introductory chapters are added. Drafts of chapters are peer-reviewed prior to final acceptance or rejection. As a rule, only original research papers and original review papers are included in the volumes. Verbatim reproductions of previous published papers are not accepted. ACS Books Department

Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Chapter 1

Introduction to Teaching Chemistry with Forensic Science

Downloaded via 46.161.56.31 on October 6, 2019 at 16:06:52 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Amanda S. Harper-Leatherman1,* and Ling Huang2 1Department of Chemistry & Biochemistry, Fairfield University, 1073 North Benson Road, Fairfield, Connecticut 06824, United States 2Department of Chemistry, Hofstra University, 106 Berliner Hall, South Campus, Hempstead, New York 11549, United States *E-mail: [email protected].

Forensic science is an interdisciplinary field that attracts a lot of students to potentially begin a career in the field as well as to simply learn chemistry through the use of this applied and popular field. Chemistry is a key part of the field of forensic science, but chemistry can be taught using forensic science to enhance the interest, motivation, and skill development of students as part of any degree program. This introductory chapter outlines the motivation for this book and many of the advantages of teaching chemistry with forensic science. Chapters in this book provide examples on incorporating forensic science into chemistry courses and laboratory experiments as well as other innovative pedagogies. Instructors can use this book as a resource to find a multitude of ways to incorporate this important area of science into their chemistry curricula.

Background on Forensic Science and the Role of Chemistry within Forensic Science Forensic science is an interdisciplinary field that “in its broadest definition is the application of science to law” (1). Forensic science applies to law with a wide variety of the scientific disciplines including, but not limited to, chemistry, biology, physics, engineering, medicine, computer science, and psychology. The forensic science field has attracted an increasing number of students in recent years. According to Data USA which reports data from the National Center for Education Statistics of the U.S. Department of Education, the total Forensic Science and Technology degrees awarded in 2016 were 3243, which represents a growth of 2.95% in one year across 216 schools offering programs in this field (2). As stated by the Bureau of Labor Statistics, there were 15,400 forensic science technicians employed in 2016, and the Bureau predicts that jobs will grow 17% between 2016 and 2026 in this profession (3). Within the broad, interdisciplinary field of forensic science, chemistry plays a key role. According to the Occupational Information Network which is sponsored by the US Department of Labor/Employment and Training Administration, one of the key knowledge areas for forensic © 2019 American Chemical Society Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

science technicians is chemistry (4). Specifically mentioned is, “knowledge of the chemical composition, structure, and properties of substances and of the chemical processes and transformations that they undergo. This includes uses of chemicals and their interactions, danger signs, production techniques, and disposal methods” (4). Chemistry courses and associated laboratories make up the largest percentage of the required natural science core classes for undergraduate Forensic Science programs with an emphasis on chemistry, biology, or toxicology, according to the accreditation standards established by the American Academy of Forensic Sciences or AAFS (5). Sixteen semester hours coming from general chemistry and organic chemistry are recommended. In addition, a minimum of 12 semester hours of advanced coursework in chemistry or biology is recommended such as courses in biochemistry, inorganic chemistry, analytical and quantitative chemistry, physical chemistry, or instrumental analysis (5). In addition to the interdisciplinary field of forensic science including a lot of chemistry, the more specific area of forensic chemistry is also offered at many colleges and universities as its own degree, minor, or concentration. There are currently five universities offering Forensic Science Education Programs Accreditation Commission (FEPAC under AAFS)–accredited Bachelor of Science degrees in forensic chemistry including Buffalo State SUNY (6), University of Mississippi (7), Ohio University (8), Towson University (9), and West Chester University of Pennsylvania (10). Examples of other universities also offering forensic chemistry bachelor’s degrees through their American Chemical Society–approved chemistry departments are University at Albany SUNY (11), Sam Houston State University (12), and Lake Superior State University (13). As an alternative, some chemistry departments offer chemistry bachelor’s degrees with concentrations in forensic chemistry or forensic science such as those at Hampton University (14), Appalachian State University (15), and Fort Valley State University (16). Chemistry departments at some schools (such as Olivet College and Western Oregon University) specifically offer minors in forensic science to combine with their chemistry degrees (17, 18). In addition, some universities offer forensic science minors that can be combined with a chemistry major or any other related major such as George Mason University (19), Sam Houston State University (20), Syracuse University (21), and Southern Connecticut State University (22). Of course, there are chemistry departments that just offer single courses in forensic science or forensic chemistry as well, and many other colleges and universities that offer a variety of chemistry courses with forensic topics included. Many examples of these types of courses will be highlighted in this book. Clearly, chemistry is an essential part of forensic science, which makes this book relevant from a forensic scientist’s perspective.

Using Forensic Science To Teach Chemistry While chemistry is an important field within forensic science, teaching chemistry within the context of forensic science has become a popular teaching method whether the chemistry course is in a forensic science program or not. In the Journal of Chemical Education, there were at least 21 articles, laboratory experiments, or activities published in the past five years related to forensic science in some way (23–43). In the Chemical Educator, there have been at least eight articles, laboratory experiments, or activities published in the past five years related to forensic science (44–51). Of these 29 publications, 3 of them were related directly to teaching forensic science students. The others explored the use of forensic science to motivate students and showed applications in a variety of areas such as analytical chemistry, general chemistry, organic chemistry, biochemistry, and even in high school or middle school science courses. Many of the articles also mention the applicability of the 2 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

published lab, activity, or idea to a variety of courses or teaching situations. Therefore, educators are accordingly making use of forensic science to teach chemistry in a wide range of courses. It is well-published that using real-world applications to give some relevance to the chemical concepts that students are learning is a very motivating pedagogical technique (52–57). Students can easily relate to forensic science, and thanks to the high number of television programs and other media products involved with reporting on forensic science, students have some familiarity with the field. This familiarity can help draw students in when discussing chemical concepts and techniques. Students not only learn the chemistry in this way, but they also get some exposure to an applied field adding value to the taught material. In addition, many instructors have found that simply setting an experiment in the context of a mock murder investigation, or other mock mystery to be solved, naturally adds motivation for the activity (24, 25, 29, 41, 42, 44, 58–60). Due to the interdisciplinary, applied nature of forensic science, using it to teach chemistry helps students develop critical thinking skills given the clear questions and problems needed to be solved. Students have the opportunity to learn logical reasoning, data interpretation and presentation, error analysis, team work, and formal report writing. Not unlike other applications of chemistry, the interest that forensic science develops in students helps in the development of these valued skills. In fact, it is a requirement for chemistry departments at colleges and universities to demonstrate that they are teaching skills such as problem-solving, communication, and team skills for American Chemical Society approval purposes (61). Owing to the popularity of using forensic science to teach chemistry and due to the many advantages of this teaching approach, we organized the symposium, “Teaching Chemistry in the Context of Forensic Science,” held as part of the 25th Biennial Conference on Chemical Education in 2018 at the University of Notre Dame. Many presenters at the symposium and other chemistry professionals have contributed to this book and each have given their perspectives on the use of forensic science to teach chemistry. The chapters describe the latest teaching techniques, courses, pedagogies, laboratories, and other innovations related to teaching chemistry with forensic science. The first chapters are related most closely to teaching non-majors chemistry, teaching general chemistry with forensic science, or applying forensic science across the chemistry curriculum. The later chapters are related most closely to teaching upper level chemistry courses with forensic science. We hope this book will be a valuable resource in the chemical education field, and we are happy that it describes how many faculty members are utilizing this popular area of science to teach chemistry so that other faculty members may be encouraged and inspired to do the same.

Overview of Chapters Related to Using Forensic Science To Teach Lower Level Chemistry or to Applying Forensic Science across the Chemistry Curriculum In Chapter 2, Kaplan gives an extensive overview on how forensic science can be used across the chemistry curriculum, from general chemistry through biochemistry, to motivate students to learn concepts that many have perceived as dull or unconnected to their lives (62). The author draws from his long-term expertise on teaching chemistry with forensic science to give specific case studies and examples that show how forensic science can be used to teach non-science majors courses, general chemistry, organic chemistry, analytical chemistry, inorganic chemistry, and biochemistry. In Chapter 3, Beussman shows how applicable incorporating forensic science in different courses across the chemistry curriculum can be with an emphasis on the use of forensic-science-themed laboratory experiments (63). Beussman outlines all of the laboratory experiments used in a nonscience-majors forensic science course including analysis of crime scenes, blood alcohol 3 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

concentrations, ink, DNA fingerprinting, drugs, fibers, blood, poison, glass, and paint. For both the paint experiment and drug overdose experiment, Beussman gives the full student laboratory manual to help instructors more clearly plan how to incorporate experiments like these into their own courses. Beussman also highlights two very popular forensic-science-themed laboratory experiments used in an analytical chemistry laboratory (high performance liquid chromatography drug overdose experiment) and an instrumental analysis laboratory (inductively coupled plasma optical emission spectroscopy analysis of matches) with the detailed student laboratory manual for each experiment included in the chapter. Pajski describes a five-week accelerated forensic science laboratory course specifically designed for nontraditional, nonscience major students in Chapter 4 (64). Forensic science is used as a topic to specifically teach problem-solving, critical thinking, and the scientific method to students who will go on to a variety of majors and careers. Pajski outlines many case studies and experiments used throughout the course. Although a laboratory-based course, it can be taught in any type of space, which adds flexibility to the course design and curriculum. An introductory forensic chemistry course taught at a community college using open educational and digital resources is described by Taylor in Chapter 5 (65). Taylor reviews the advantages to open educational resources, and the chapter references the course website that includes a lab manual and links to a variety of online resources in forensic chemistry. The topics in forensic chemistry covered in this course include the metric system, density, refractive index, organic analysis, drug analysis, thin layer chromatography, inorganic analysis, spectroscopy, serology, arson analysis, and DNA analysis. The chemistry content differentiates the course from a similar, but different, introduction to forensic science course also offered at the college. Assignments including article review assignments are outlined, and all of the experiments used in this course are green chemistry experiments in addition to being forensic chemistry experiments. Two chapters give uses of forensic science in general chemistry. In the first of the two (Chapter 6), Testa provides a way to incorporate forensic science to motivate students to learn about oxidationreduction reactions through a mystery-based general chemistry laboratory experiment (66). The advantages to using forensic science in general chemistry are overviewed including a review of other published ways of incorporating forensic science into the chemistry curriculum. Testa comprehensively outlines the many different published forensic science scenarios, types of physical evidence, and analytical methods and instrumentation that have been used to teach chemistry with forensic science. In the second of the two chapters focused on general chemistry (Chapter 7), Garrett describes how to link the second semester of General Chemistry with a Forensics in Literature English course through what are known as Interdisciplinary Learning Communities at Belmont University (67). The linked theme adds depth to both courses and helps students build scientific and communication skills within an applied context. Garrett outlines unique aspects of the General Chemistry course including nontraditional texts, nontraditional experiments, and a final project on writing an original fictional mystery that includes an accurate overview of the forensic chemical evidence and analysis. In another interdisciplinary chapter, Harper-Leatherman and Roney describe an interdisciplinary extracurricular learning activity focused on forensic science in the healthcare setting that helped to supplement the coursework of a group of nursing major and science major students by going beyond what the nursing and science curricula provided in Chapter 8 (68). Throughout a onetime, three-hour voluntary event, students learned background information on forensic science and emergency room care for patients and participated in the care of a mock crime victim. In addition,

4 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

they collected mock fiber evidence and analyzed the evidence using three chemistry techniques that are described in detail.

Overview of Chapters Related to Using Forensic Science To Teach Upper Level Chemistry The chapters related to using forensic science to teach upper level chemistry courses start with Chapter 9, in which Huang focuses on how the forensic topics of drugs and DNA can be used to enhance the analytical chemistry curriculum (69). Many job opportunities exist in the fields of drug analysis and DNA analysis making these topics very relevant to students. The analytical techniques used in drug analysis and recommended by the Scientific Working Group for the Analysis of Seized Drugs are reviewed including many separations and spectroscopic techniques. Huang discusses several suggested experiments, assignments, and analytical chemistry curriculum topics related to the analysis of over-the-counter drugs, controlled substances, and drug metabolites. Huang also outlines and reviews the process of DNA analysis, which shows how many important analytical chemistry concepts can be highlighted through the discussion of DNA in analytical chemistry courses such as extraction, quantification, and separation. In Chapter 10, Gray summarizes how forensic science is used to enhance engagement and critical thinking in an instrumental analysis laboratory course and in an upper level elective course in forensic science for junior and senior chemistry and biochemistry majors (70). In the instrumental analysis course, students are given a case scenario, and then over the course of five to six laboratory periods, they are assigned to analyze evidence from the case using a variety of spectroscopic and separations techniques. Students build critical analytical skills when considering the significance of the results and comparing the instrumental methods used. Gray outlines each experiment and discusses information on what was most difficult for students. In the forensic science elective course, a crime scenario is also used to engage students as they use crime scene processing, fingerprinting, trace evidence fiber analysis, drug analysis, questioned documents examination, forensic serology, and firearms and impression analysis to draw conclusions about the case using higher order thinking skills. Mojica et al. introduce and give background information on the forensic science curriculum at Pace University and provide many examples of the use of forensic science topics in instrumental analysis and other upper level courses in Chapter 11 (71). The authors emphasize how their forensic science program developed from existing science departments, which provides rigor to their program. Sample course sequences for the bachelor’s degree and master’s degree in forensic science are given. Some examples of forensic topics included in instrumental analysis and other related courses at Pace University are chiral separations of ephedrines, pseudoephedrines, amphetamines and methamphetamines, chiral analysis using spectroscopy, and arson accelerant analysis. The authors also discuss the success of using an interactive online eTextbook that includes forensic topics and assignments. In Chapter 12, Dukes et al. discuss developing an authentic forensic analysis experience for students by combining gas chromatography mass spectrometry analysis of a suspected illicit substance with student testimony of these results in a simulated courtroom (72). The procedure for a simulated gas chromatography mass spectrometry drug analysis used in an instrumental analysis laboratory is outlined including sample results and the use of a realistic drug analysis worksheet similar to that used by forensic scientists. Students also gain experience with presumptive colorimetric tests for illicit compounds within the experiment. Finally, the authors discuss how each 5 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

student testifies in front of a volunteer prosecutor and a volunteer defense attorney to build scientific communication skills. Since this activity has built-in authenticity and value beyond the course, this experiment and the courtroom experience can build motivation for learning the chemistry involved in the analysis. In Chapter 13, Legron-Rodriguez relates the successful use of generative learning strategies, pre-lecture assignments, and a flipped classroom pedagogy in an upper level course entitled, “The Forensic Analysis of Controlled Substances” (73). The course is a subject-specialized, upper level required course in the forensic science bachelor’s degree program at the University of Central Florida. Legron-Rodriguez describes how the course instructional strategies evolved over the course of three years resulting in increased student final exam grades. The strategies discussed include concept maps, drawing, practice testing, and summarizing. These techniques proved useful in this content-heavy upper level course but could also be useful to enhance student learning in any course according to research presented in the chapter.

Conclusion There are a wide range of ways to teach chemistry with forensic science as reported in the chapters within this book. The main theme is to show how enriching it is for students to learn chemistry in the context of a field that is directly applicable and valuable to people’s lives. From nonmajors to chemistry majors to students studying to work in forensic science laboratories, all students benefit through the incorporation of forensic science into chemical education. Students are motivated to gain chemical knowledge and can build important overarching skills such as problemsolving, critical thinking, communication, and team work. Hopefully all instructors can find at least one teaching technique, innovation, or experiment from this book that could be applied within their own teaching context. Students and instructors alike will find added value by learning and teaching chemistry with forensic science.

References 1. 2. 3.

4. 5. 6. 7. 8.

Saferstein, R. Criminalistics: An Introduction to Forensic Science, 12th ed.; Pearson: Boston, MA, 2018. Forensic Science & Technology: Data USA. https://datausa.io/profile/cip/430106/#intro (accessed March 22, 2019). Forensic Science Technicians: Occupational Outlook Handbook: U.S. Bureau of Labor Statistics. https://www.bls.gov/ooh/life-physical-and-social-science/forensic-science-technicians. htm#tab-1 (accessed March 22, 2019). 19-4092.00 – Forensic Science Technicians. https://www.onetonline.org/link/summary/194092.00 (accessed Mar 22, 2019). FEPAC Standards 07302014. http://fepac-edu.org/sites/default/files/FEPAC% 20Standards%2007302014.pdf (accessed March 22, 2019). Forensic Chemistry, B.S. Program, Chemistry – SUNY Buffalo State. https://chemistry. buffalostate.edu/forensic-chemistry-bs-program (accessed June 5, 2019). Academic Programs – B.S. in Forensic Chemistry – University of Mississippi. https://olemiss.edu/ academics/programs/liberal-arts/bs-for-chem (accessed June 5, 2019). Forensic Chemistry (B.S.) – Ohio University. https://www.ohio.edu/cas/chemistry/ undergraduate/forensic-chemistry (accessed June 5, 2019). 6 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

9. 10. 11. 12. 13. 14. 15. 16.

17.

18. 19. 20.

21. 22. 23.

24.

25.

26.

Forensic Chemistry Major – Towson University. https://www.towson.edu/fcsm/departments/ chemistry/undergrad/forensic.html (accessed June 5, 2019). B.S. Forensic and Toxicological Chemistry – West Chester University. https://www.wcupa.edu/ sciences-mathematics/chemistry/forensicChem.aspx (accessed June 5, 2019). Forensic Chemistry – Department of Chemistry – University at Albany SUNY. https://www.albany. edu/chemistry/initiative.shtml (accessed June 6, 2019). Bachelor of Science in Forensic Chemistry – Sam Houston State University. https://www.shsu.edu/ programs/bachelor-of-science-in-forensic-chemistry/index.html (accessed June 6, 2019). Forensic Chemistry – Lake Superior State University. https://www.lssu.edu/school-physicalsciences/forensic-chemistry/ (accessed June 6, 2019). Forensic Chemistry Concentration – Hampton University Department of Chemistry. https://science.hamptonu.edu/chem/programs/forensic.cfm (accessed June 6, 2019). Appalachian State University / Chemistry (BS) – Forensic Science. https://www.appstate.edu/ academics/majors/id/chemistry-forensic-science (accessed June 6, 2019). Forensic Science Concentration (Chemistry) – Fort Valley State University. https://www.fvsu.edu/ about-fort-valley-state-university/academics/college-of-arts-and-sciences/chemistrydepartment/chemistry-concentrations/ (accessed June 6, 2019). Chemistry Major with Forensic Science Minor – Olivet College https://www.olivetcollege.edu/ undergraduate/academics/list-of-majors/natural-physical-sciences/chemistry-with-forensicscience/ (accessed June 5, 2019). Forensic Science (Chemistry Majors) Minor, Academics – Western Oregon University. https://www. wou.edu/academics/forensic-science-chemistry-majors-minor/ (accessed June 5, 2019). Forensic Science Minor – George Mason University. https://catalog.gmu.edu/colleges-schools/ science/forensic-program/forensic-science-minor/ (accessed June 6, 2019). Minor in Forensic Science – Sam Houston State University. http://catalog.shsu.edu/ undergraduate/colleges-academic-departments/criminal-justice/forensic-science/minorforensic-science/ (accessed June 6, 2019). Program: Forensic Science Minor – Syracuse University – Acalog ACMS. http://coursecatalog.syr. edu/preview_program.php?catoid=3&poid=853&returnto=253 (accessed June 6, 2019). Forensic Science Minor Home – Southern Connecticut State University. https://www2.southernct. edu/academics/schools/arts/departments/forensicscience.html (accessed June 6, 2019). Adams-McNichol, A. L.; Shiell, R. C.; Ellis, D. A. Accurate, Photoresistor-Based, Student Built Photometer and Its Application to the Forensic Analysis of Dyes. J. Chem. Educ. 2019, 96, 1143–1151DOI:10.1021/acs.jchemedu.8b00862. Ferreiro-González, M.; Amores-Arrocha, A.; Espada-Bellido, E.; Aliaño-Gonzalez, M. J.; Vázquez-Espinosa, M.; González-de-Peredo, A. V.; Sancho-Galán, P.; Álvarez-Saura, J. A.; Barbero, G. F.; Cejudo-Bastante, C. Escape ClassRoom: Can you Solve a Crime Using the Analytical Process? J. Chem. Educ. 2019, 96, 267–273. Testa, S. M.; Selegue, J. P.; French, A.; Criswell, B. Permanganate Oxidation of DNA Nucleotides: An Introductory Redox Laboratory Framed as a Murder Mystery. J. Chem. Educ. 2018, 95, 1840–1847. Porter, S. E. G.; Barber, A. E.; Colella, O. K.; Roach, T. D. Using Biological Organisms as Chemical Sensors: The MicRoboCop Project. J. Chem. Educ. 2018, 95, 1392–1397.

7 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

27. Parker, P. D.; Beers, B.; Vergne, M. J. What Is in Your Wallet? Quantitation of Drugs of Abuse on Paper Currency with a Rapid LC–MS/MS Method. J. Chem. Educ. 2017, 94, 1522–1526. 28. Cresswell, S. L.; Loughlin, W. A. A Case-Based Scenario with Interdisciplinary Guided-Inquiry in Chemistry and Biology: Experiences of First Year Forensic Science Students. J. Chem. Educ. 2017, 94, 1074–1082. 29. Valente Nabais, J. M.; Costa, S. D. A Forensic Experiment: The Case of the Crime at the Cinema. J. Chem. Educ. 2017, 94, 1111–1117. 30. Hamnett, H. J.; Korb, A-S. The Coffee Project Revisited: Teaching Research Skills to Forensic Chemists. J. Chem. Educ. 2017, 94, 445–450. 31. Thompson, R. Q. Forensic Chemistry and Its Flip Side. J. Chem. Educ. 2016, 93, 1677–1678. 32. Volpi, G. Demonstrating the Presence of Cyanide in Bitter Seeds while Helping Students Visualize Metal–Cyanide Reduction and Formation in a Copper Complex Reaction. J. Chem. Educ. 2016, 93, 891–897. 33. Contakes, S. M. Misconduct at the Lab? A Performance Task Case Study for Teaching Data Analysis and Critical Thinking. J. Chem. Educ. 2016, 93, 314–317. 34. Zuidema, D. R.; Herndon, L. B. Using The Poisoner’s Handbook in Conjunction with Teaching a First-Term General/Organic/Biochemistry Course. J. Chem. Educ. 2016, 93, 98–102. 35. Cresswell, S. L.; Loughlin, W. A. An Interdisciplinary Guided Inquiry Laboratory for First Year Undergraduate Forensic Science Students. J. Chem. Educ. 2015, 92, 1730–1735. 36. Kanu, A. B.; Pajski, M.; Hartman, M.; Kimaru, I.; Marine, S.; Kaplan, L. J. Exploring Perspectives and Identifying Potential Challenges Encountered with Crime Scene Investigations when Developing Chemistry Curricula. J. Chem. Educ. 2015, 92, 1353–1358. 37. Friesen, J. B. Activities Designed for Fingerprint Dusting and the Chemical Revelation of Latent Fingerprints. J. Chem. Educ. 2015, 92, 505–508. 38. Friesen, J. B. Forensic Chemistry: The Revelation of Latent Fingerprints. J. Chem. Educ. 2015, 92, 497–504. 39. Smith, M. J.; Vale, I. C.; Gray, F. M. Identification of Paper by Stationary Phase Performance. J. Chem. Educ. 2014, 91, 1679–1683. 40. Ahrenkiel, L.; Worm-Leonhard, M. Offering a Forensic Science Camp to Introduce and Engage High School Students in Interdisciplinary Science Topics. J. Chem. Educ. 2014, 91, 340–344. 41. Marle, P. D.; Decker, L.; Taylor, V.; Fitzpatrick, K.; Khaliqi, D.; Owens, J. E.; Henry, R. M. CSI–Chocolate Science Investigation and the Case of the Recipe Rip-Off: Using an Extended Problem-Based Scenario To Enhance High School Students’ Science Engagement. J. Chem. Educ. 2014, 91, 345–350. 42. Meyer, A. F.; Knutson, C. M.; Finkenstaedt-Quinn, S. A.; Gruba, S. M.; Meyer, B. M.; Thompson, J. W.; Maurer-Jones, M. A.; Halderman, S.; Tillman, A. S.; DeStefano, L.; Haynes, C. L. Activities for Middle School Students to Sleuth a Chemistry “Whodunit” and Investigate the Scientific Method. J. Chem. Educ. 2014, 91, 410–413. 43. Ravgiala, R. R.; Weisburd, S.; Sleeper, R.; Martinez, A.; Rozkiewicz, D.; Whitesides, G. M.; Hollar, K. A. Using Paper-Based Diagnostics with High School Students To Model Forensic Investigation and Colorimetric Analysis. J. Chem. Educ. 2014, 91, 107–111.

8 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

44. Johnson, B. O.; Blackwell, A.; Patel, M.; Robins, A. Mystery of the Missing Medicine: Forensic Drug Identification of Pseudoephedrine in a Non-Science Majors Course. Chem. Educator 2018, 23, 112–118. 45. Sinclair, K.; Mirjafari, A.; Coticone, S. R. A Tale of Two Spices: Differential Discrimination between the Chemical Properties of Saffron and Turmeric in a Forensic Chemistry Laboratory. Chem. Educator 2018, 23, 143–144. 46. Davidson, K.; Chasten, V. D.; Pinder, T.; Wellman, S.; Byrd, G.; Wilson-Kennedy, Z. S.; Fakayode, S. O. Analysis of Footwear Co-Polymer Compositions by FTIR Spectroscopy and Principal Component Analysis: Sophomore Immersion Program in Research and Academics Program. Chem. Educator 2018, 23, 149–158. 47. Deng, Y.; Zhang, A.; Kandah, J.; Akins, S.; Jawad, K. Forensic Analysis in the Instrumental Laboratory: A Nondestructive Ultrasound-enhanced Aqueous Extraction Coupled with GCMS Analysis of Cocaine on US Bandnotes. Chem. Educator 2017, 22, 197–200. 48. Cote, J.; Hilbert, L.; Coticone, S. R. Differential Indentification of Six “Mystery” Substances Using Three Distinct Analytical Techniques in a Forensic Science Laboratory. Chem. Educator 2016, 21, 115–118. 49. Kanu, A. B.; Kaplan, L. J. The Quest for Confirmatory Data in Crime Scene Investigations. Chem. Educator 2016, 21, 231–239. 50. Garcia, O.; Keeney, C.; Coticone, S. Detection of Coumarin in Artificially Adulterated Vanilla Bean Extracts in a Forensic Science Laboratory. Chem. Educator 2015, 20, 227–228. 51. Rouse, B. D.; Schneider, R. L.; Smith, E. T. Presumptive and Confirmatory Tests using Analogs of Illicit Drugs: An Undergraduate Instrumental Methods Exercise. Chem. Educator 2014, 19, 70–72. 52. Morra, B. The Chemistry Connections Challenge: Encouraging Students to Connect Course Concepts with Real-World Applications. J. Chem. Educ. 2018, 95, 2212–2215. 53. Thomas, R.; Baker, M.; Cross, C.; Miehl, M. Value of Using STEM Professionals in the K12 Classroom: Connecting Chemistry to the Real World. In Citizens First? Democracy, Social Responsibility and Chemistry; Maguire, C. F., Sheardy, R. D. , Eds.; ACS Symposium Series 1297; American Chemical Society: Washington, DC, 2018; pp 33–41. 54. Urban, S.; Brkljača, R.; Cockman, R.; Rook, T. Contextualizing Learning Chemistry in FirstYear Undergraduate Programs: Engaging Industry-Based Videos with Real-Time Quizzing. J. Chem. Educ. 2017, 94, 873–878. 55. Forest, K.; Rayne, S. Thinking Outside the Classroom: Integrating Field Trips into a First-Year Undergraduate Chemistry Curriculum. J. Chem. Educ. 2009, 86, 1290–1294. 56. Clauss, A. Real-World Topics: Medicinal Chemistry. J. Chem. Educ. 2008, 85, 1655–1657. 57. Jones, M. B.; Miller, C. R. Chemistry in the Real World. J. Chem. Educ. 2001, 78, 484–487. 58. Charkoudian, L. K.; Heymann, J. J.; Adler, M. J.; Haas, K. L.; Mies, K. A.; Bonk, J. F. Forensics as a Gateway: Promoting Undergraduate Interest in Science, and Graduate Student Professional Development through a First-Year Seminar Course. J. Chem. Educ. 2008, 85, 807–812. 59. Tarr, M. A. Solving a Mock Arsenic-Poisoning Case Using Atomic Spectroscopy. J. Chem. Educ. 2001, 78, 61–62. 60. Harper-Leatherman, A. S. Making Connections to the Liberal Arts College Mission: Exploring Identity and Purpose in a Chemistry Course. In Liberal Arts Strategies for the Chemistry

9 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

61.

62.

63.

64.

65.

66.

67.

68.

69.

70.

71.

Classroom; Kloepper, K. D., Crawford, G. L. , Eds.; ACS Symposium Series 1266; American Chemical Society: Washington, DC, 2017; pp 185–194. ACS Guidelines & Supplements – American Chemical Society. https://www.acs.org/content/acs/ en/about/governance/committees/training/acs-guidelines-supplements.html (accessed June 5, 2019). Kaplan, L. J. Chemistry and Crime: Investigating Chemistry from a Forensic Science Perspective. In Teaching Chemistry with Forensic Science; Harper-Leatherman, A. S., Huang, L., Eds.; ACS Symposium Series 1324; American Chemical Society: Washington, DC, 2019; Chapter 2, pp 13–34. Beussman, D. J. Incorporating Forensic Science Throughout the Undergraduate Analytical Curriculum: From Non-Majors through Instrumental Analysis. In Teaching Chemistry with Forensic Science; Harper-Leatherman, A. S., Huang, L. , Eds.; ACS Symposium Series 1324; American Chemical Society: Washington, DC, 2019; Chapter 3, pp 35–57. Pajski, M. L. Using Forensic Science to Engage Nontraditional Learners. In Teaching Chemistry with Forensic Science; Harper-Leatherman, A. S., Huang, L. , Eds.; ACS Symposium Series 1324; American Chemical Society: Washington, DC, 2019; Chapter 4, pp 59–77. Taylor, B. Teaching Introductory Forensic Chemistry using Open Educational and Digital Resources. In Teaching Chemistry with Forensic Science; Harper-Leatherman, A. S., Huang, L., Eds.; ACS Symposium Series 1324; American Chemical Society: Washington, DC, 2019; Chapter 5, pp 79–91. Testa, S. M. On Utilizing Forensic Science to Motivate Students in a First-Semester General Chemistry Laboratory. In Teaching Chemistry with Forensic Science; Harper-Leatherman, A. S., Huang, L. , Eds.; ACS Symposium Series 1324; American Chemical Society: Washington, DC, 2019; Chapter 6, pp 93–108. Garrett, M. D. Interdisciplinary Learning Communities: Bridging the Gap Between the Sciences and the Humanities Through Forensic Science. In Teaching Chemistry with Forensic Science; Harper-Leatherman, A. S., Huang, L. , Eds.; ACS Symposium Series 1324; American Chemical Society: Washington, DC, 2019; Chapter 7, pp 109–136. Harper-Leatherman, A. S. Interdisciplinary Learning Activity Incorporating Forensic Science and Forensic Nursing. In Teaching Chemistry with Forensic Science; A. S. Harper-Leatherman, A. S., Huang, L., Eds.; ACS Symposium Series 1324; American Chemical Society: Washington, DC, 2019; Chapter 8, pp 137–153. Huang, L. Drugs and DNA: Forensic Topics Ideal for the Analytical Chemistry Curriculum. In Teaching Chemistry with Forensic Science; Harper-Leatherman, A. S., Huang, L., Eds.; ACS Symposium Series 1324; American Chemical Society: Washington, DC, 2019; Chapter 9, pp 155–167. Gray S. E. C. From DUIs to Stolen Treasure: Using Real-World Sample Analysis to Increase Engagement and Critical Thinking in Analytical Chemistry Courses. In Teaching Chemistry with Forensic Science; Harper-Leatherman, A. S., Huang, L., Eds.; ACS Symposium Series 1324; American Chemical Society: Washington, DC, 2019; Chapter 10, pp 169–202. Mojica, E.-R. E.; Marvin, R.; Evans, N.; Reilly, L.; Mendoza, D.; Karpadakis, S.; Cusumano, C.; Athanasopoulos, D.; Dai, Z. Integration of Forensic Themes in Teaching Instrumental Analysis at Pace University. In Teaching Chemistry with Forensic Science; Harper-Leatherman, A. 10 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

S., Huang, L., Eds.; ACS Symposium Series 1324; American Chemical Society: Washington, DC, 2019; Chapter 11, pp 203–219. 72. Dukes, III, A. D.; Hollifield, J. M.; Gardner, D. E. Using Expert Witness Testimony with an Illicit Substance Analysis to Increase Student Engagement in Learning the GC/MS Technique. In Teaching Chemistry with Forensic Science; Harper-Leatherman, A. S., Huang, L. , Eds.; ACS Symposium Series 1324; American Chemical Society: Washington, DC, 2019; Chapter 12, pp 221–231. 73. Legron-Rodriguez, T. Generative Learning Strategies and Pre-Lecture Assignments in a Flipped Forensic Chemistry Classroom. In Teaching Chemistry with Forensic Science; HarperLeatherman, A. S., Huang, L. , Eds.; ACS Symposium Series 1324; American Chemical Society: Washington, DC, 2019; Chapter 13, pp 233–241.

11 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Chapter 2

Chemistry and Crime: Investigating Chemistry from a Forensic Science Perspective Downloaded via UNIV OF GEORGIA on October 6, 2019 at 16:08:15 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Lawrence J. Kaplan* Department of Chemistry, 47 Lab Campus Drive, Williams College, Williamstown, Massachusetts 01267, United States *E-mail: [email protected].

Forensic chemistry is one of the most interesting and integrative subdivisions of chemistry. Most students find its study challenging but very rewarding. Forensic chemistry can best be described as applied analytical chemistry but also finds use in most of the other subdivisions of chemistry. Forensic chemistry is distinguished from other types of chemistry by the legal context in which it must be conducted. Over the past few decades, the public profile of forensic science has dramatically increased, especially when discussed with regard to both historical and present day criminal cases. However, many chemistry educators do not appreciate the impact forensic science can have in motivating students to study and appreciate the concepts in courses ranging from general chemistry to biochemistry. This chapter will explore the broad way in which the fascination with crime detection (finding out “who dun it”) can be used in virtually all chemistry courses to stimulate student interest and motivate a willing, even enthusiastic, response from students challenged to study what is frequently viewed as boring and irrelevant. An overview of the concepts in which forensic science can be helpful is discussed for each of the courses. This is followed by a number of specific examples that illustrate how the case studies or the forensic topics can stimulate interest. Reference to both texts and reference books, as well as the original literature, is provided to guide further exploration.

Introduction College students today have grown up in a digital age in which information is only a click away. Professors have been increasingly challenged to engage these students and to provide relevance between classroom content and real-life situations. Unfortunately, most students do not view science in general, and chemistry in particular, as being relevant to their lives. As a biochemistry professor, I am constantly amazed at how (even premedical) students find little relevance in the material discussed in a typical undergraduate biochemistry class. And if it is difficult for students (even © 2019 American Chemical Society Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

chemistry majors) to appreciate the relevance of biochemistry, how much more difficult is it for them to see the relevance of oxidation/reduction or stoichiometry in general chemistry or the classical name reactions of organic chemistry. In short, many students see this material as intimidating and with little relevance. In fact, students report that what they learn in school is boring and has little relation to everyday life outside of the classroom. They admit that they are more interested in and learn more from their own explorations with the media and the Internet outside of school than they learn in their content classes (1). It is difficult to know how much students actually learn from their own explorations in informal settings or how students relate the concepts they learn in school to their knowledge and skills acquired through interactions with media or digital texts such as websites, television programs, video games, and trade books (1). When this author was asked in the early 1980s to teach a course for non-science majors, many of those courses were being taught in a formulaic way. The professor would choose a topic, such as environmental (chemistry) science, and then teach an introductory chemistry course adding a few examples drawn from environmental science. Usually these examples were added as an afterthought. The courses were disappointing for the instructor because the students were not interested, and they were disappointing for the students because they did not get the kind of course they were expecting. Since this author had a long-standing interest in law and a strong interest in forensics, forensic science seemed like an ideal course for non-scientifically oriented students. And a signature of pedagogical approach was not to teach a standard introductory chemistry course covering all of the usual topics, but to introduce only those chemistry topics needed to understand the forensic science concepts. Forensic chemistry is one of the most interesting and integrative subdivisions of chemistry. Most students find its study challenging, but very rewarding. Forensic chemistry can best be described as applied analytical chemistry. It finds use in most of the other subdivisions of chemistry. Forensic chemistry is distinguished from other types of chemistry by the legal context in which it must be conducted. In fact, as mentioned above, this author showed in the early 1980s that, with the fascination with crime detection (finding out “who dun it”), forensic science is an extremely effective vehicle for promoting student interest, engagement, and achievement by creating a relevant learning experience (2, 3). Others have demonstrated this effect and expressed it this way: “To address this problem, we use forensic science to positively motivate students and increase their interest level (4),” and “Educators have observed that forensic activities have resulted in students’ increased engagement in science activity and enthusiasm for science” (1). Furthermore, a focus on forensic science can foster analytical thinking and problem solving—ultimately one of the primary goals of an undergraduate education. Finally, forensic science can be used as a gateway to promote undergraduate interest in science (5). Over the past few decades, the public profile of forensic science has dramatically increased. However, many chemistry educators do not appreciate the impact forensic science can have on motivating students to study and appreciate the concepts taught in courses ranging from general chemistry to biochemistry. Indeed, forensic science can be used to introduce and explain concepts in courses ranging from chemistry for the non-science major and basic chemistry for the chemistry major to biochemistry. This chapter will present some of the many ways in which the fascination with crime detection can be used to stimulate student interest and motivate students challenged to study what is frequently viewed as difficult, uninteresting, and irrelevant. For each of the courses, an overview of the concepts for which forensic science can be helpful is discussed. This is followed by a number of specific examples that illustrate how the case studies or the forensic topics can stimulate the students’ interest. 14 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Brief History of Forensic Science Teaching In the early 1980s, the present author noticed a disappointing situation. There were few forensic science courses taught at the undergraduate level and, while there were several graduate level programs, they were having difficulty growing and attracting new interest. As Peterson and Angelos (6) noted at this time, “forensic science faculties…were not highly regarded by academics in other areas,” and Gaensslen and Lee (7) noted, “the number of academic forensic science programs actually declined in the 1980s to approximately 15, down from a previous high of 25 in 1978.” Also, before the developments in the O. J. Simpson case were broadcast live daily on television and before the television show CSI, there were very few books on forensic science. While books by Califano and Levkov (8), Goddard (9), and O’Brien and Sullivan (10) were available, the two books most widely used were by Saferstein (11) and DeForest et al. (12). The book by Saferstein is now in its 12th edition (13), and a relatively new edition of the other book has been published by Gaensslen et al. (14). By the mid-1990s, forensic science had gained enormous popularity in the wider academic world as expressed by Madhusoodanan (15): The recognition of science’s importance to the practice of law—coupled with the popularity of crime shows on television—led to a booming interest in forensic science and academic programs at both the undergraduate and graduate levels. Specifically, the burgeoning number of aspiring forensic scientists has driven at least 55 U.S. academic institutions to offer undergraduate concentrations in forensics, and several to offer master’s degrees or programs specializing in forensics. In 2003, the American Association of Forensic Sciences created the Forensic Science Education Programs Accreditation Commission to ensure that the training for the students met standards set by the National Institute of Justice. With the tremendous increase in popularity of forensic science, there has been a huge increase of the number of forensic science textbooks. Books for use in forensic chemistry courses as well as more advanced forensic courses include the following: Bertino and Bertino (16), Eckart (17), Elkins (18), Fisher et al. (19), Girard (20), Houck (21), Houck and Siegel (22), James et al. (23), and Siegel (24). A few texts are more geared toward undergraduate forensic chemistry courses: Johll (25), Bell (26), and Siegel and Mirokovits (27). Two books deal with the crime scene but also contain many of the basic concepts contained in the other texts: Fisher and Fisher (28) and Houck et al. (29). A number of general laboratory manuals have been produced: Erickson (30), Kaplan (31), Kubic and Petraco (32), and Meloan et al. (33). A variety of other resources are available, such as the books that provide numerous case studies, including many valuable pictures to illustrate presentations: Kaye (34) and Owen (35).

The Locard Exchange Principle One of the most important concepts that guides the work of forensic scientists is the Locard Exchange Principle (36). While it is usually expressed simply as “every contact results in an exchange” or every contact leaves a trace,” Locard actually said: “Toute action de l’homme, et a fortiori, l’action violente qu’est un crime, ne peut pas se dérouler sans laisser quelque marque.” Translated to English, this means “Any action of an individual, and obviously the violent action constituting a crime, cannot occur without leaving a trace.” Kirk (37) stated it best:

15 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

“Wherever he steps, whatever he touches, whatever he leaves, even unconsciously, will serve as a silent witness against him. Not only his fingerprints or his footprints, but his hair, the fibers from his clothes, the glass he breaks, the tool mark he leaves, the paint he scratches, the blood or semen he deposits or collects. All of these and more bear mute witness against him. This is evidence that does not forget. It is not confused by the excitement of the moment. It is not absent because human witnesses are. It is factual evidence. Physical evidence cannot be wrong, it cannot perjure itself, it cannot be wholly absent. Only human failure to find it, study and understand it, can diminish its value.” One should keep this principle in mind as the theoretical and experimental aspects of various areas of chemistry and their application to forensic science are discussed.

Forensic Science for the Non-Science Major or the Chemistry Major The topics covered and the related chemical concepts are fairly obvious for a course in forensic science, whether at the level of a chemistry major or a non-science major. The major difference is the level of sophistication of the material. The author has taught a course for non-science majors for over 35 years and has been able to elicit a surprising level of sophistication and willingness in these students to discuss basic science concepts. The approach is what some colleagues years ago called “science taught backward.” This author, in fact, realized that the way general chemistry was being taught was really “backward.” So, the model adopted was to stimulate the students’ interest by discussing specific cases, the resolution of which required a certain understanding of basic chemical concepts. For example, in 1982, seven people died in Chicago as the result of tampering with the over-the-counter medicine, Tylenol (38). In short order, it was determined that some bottles of the medicine contained capsules contaminated with potassium cyanide (KCN). In an attempt to identify the source of the KCN, its elemental composition was analyzed (39). In order to understand the chemistry of the analysis of the elemental composition of the KCN, one must have a basic knowledge of atomic absorption (and related atomic emission) spectroscopy. This would not be possible without a basic understanding of atomic structure and the arrangement and energy of the electrons. However, for the purposes of explaining these spectroscopic procedures, one need not spend an extensive amount of time on the historical development of the current view of atomic structure, but rather develop the model (usually the Bohr model is used under these circumstances) and then explain how those spectroscopic methods can be understood with that atomic structure model. It is important to move quickly back to a discussion of the case and how the basic science introduced can lead to a better understanding of the case. Another example of this instructional approach is the explanation of the analysis of the Shroud of Turin (40). The Shroud has been proposed to be the burial cloth of Jesus and, in an attempt to determine its authenticity, the age of the linen cloth was determined by radiocarbon dating (41). Of course, before one can understand radiocarbon dating, one must have an understanding of atomic structure, with an emphasis on the nucleus. In this case, the formation of radioactive nitrogen in the atmosphere and its conversion to radioactive carbon (14C) explains the essentials of both the formation of the radioactive carbon and its subsequent detection with simple decay detection equipment. Employing mass spectrometric detection of the total 14C atoms (not just the ones decaying) introduces the students to the issue of sampling, conservation of evidence samples, and sensitivity of various detection methods. Of course, for a complete explanation of radiocarbon 16 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

dating, the formation of 14CO2 and its incorporation into plant matter with the transfer to animal matter also would be discussed. Before we close the discussion of the Shroud, the following should be noted: Three laboratories independently radio-dated samples from the shroud by the accelerator mass spectroscopy (AMS) method and reported a reasonably precise 14th century date, in apparent agreement with the known historic record. Unfortunately, the recommended detailed sampling protocol that would assure both precision and accuracy was not followed. Subsequent FT-IR and scanning electron microprobe data showed that the samples taken for radiocarbon dating were not representative of the bulk of the non-image portions of the Shroud (42). For a more thorough discussion of the importance of sampling, see the discussion below on the analysis of the Kennedy bullet fragments. As mentioned above, many experiments and more extensive laboratory programs have been developed for these courses (30–33). Others have used some of the relatively new pedagogical approaches such as Nabais and Costa (43), who have developed experiments for analyzing fiber (using infrared spectroscopy) and lipstick (using thin layer chromatography) samples collected at “the Crime at the Cinema.” Echoing the comments above, they note the following about their scenario: [It]takes advantage of the motivation and interest shown by the students for the forensic subject to promote the learning process. The problem-based learning (PBL) method that uses forensic topics is a proven method to motivate students and a tool to raise critical thinking skills….The described methodology involves a guided experimental learning tool, which increases long-term knowledge, provides higher satisfaction, and improves performance more efficiently than traditional instructions. In addition, Creswell and Loughlin (44, 45) have developed an interdisciplinary guided inquiry laboratory program for first year forensic science students.

The Role of Forensic Science in the Arts and in Outreach Programs A review of the role forensic science plays in teaching chemistry for the non-science major would not be complete without mentioning its application to authenticating works of art (46–48) and its role in the artistic worlds of theater (49, 50) and literature (51, 52). Of particular note in the world of literature are the classic works of Dorothy L. Sayers (53), Agatha Christie (54), and Arthur Conan Doyle (55). More modern writers such as Patricia Cornwell (56) and Kathy Reichs (57) have contributed new excitement to this genre. Also important are outreach programs (58, 59) for younger students. These provide still additional avenues for stimulating student interest in chemistry.

General/Introductory Chemistry A wide range of topics usually discussed in general/introductory chemistry may be introduced with a forensic science theme. As an introduction to a variety of chemical concepts and forensic science cases to introduce them, the collection of articles in the Journal of Chemical Education called “The Chemical Adventures of Sherlock Holmes” is a terrific place to start. From 1989 until 2004, Waddell and Rybolt authored 15 of these mysteries and then Shaw added 4 more from 2008 17 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

until 2012. Jacobsen summarized the collection in 2011 (60). As examples of their applicability to general chemistry, “The Case of the Stoichiometric Solution” (61) deals with stoichiometric calculations, “The Hound of Henry Armitage” (62) deals with physical properties and balancing chemical equations, and “The Blackwater Escape” (63) deals with oxidation-reduction processes. Of course, the cases mentioned under Forensic Science courses would be applicable to general chemistry courses as well. One example drawn from this author’s instruction involves a somewhat complicated stoichiometric calculation. Consider the reaction that takes place in the Breathalyzer (64), in which a vial containing a solution of potassium dichromate reacts with the alcohol in the subject’s breath. As the reaction takes place the yellow color decreases, as some of the potassium dichromate is converted (reduced) to chromium(III) sulfate, which is bluish in color, while the ethanol is oxidized to acetic acid. The students are asked how much potassium dichromate would be used if the blood alcohol concentration was 0.08 (the legal limit while driving in most jurisdictions). To answer this question, the students have to know how much potassium dichromate is in the vial, how much breath is captured by the breathalyzer, the relationship of the blood alcohol concentration to the breath alcohol concentration, the stoichiometric ratios, etc. Details of the function of the Breathalyzer are discussed on the website “The Alcohol Pharmacology Educational Partnership” (65). A final example, which has application to general chemistry, organic chemistry, and even biochemistry involves the oxidation of DNA nucleotides to introduce redox concepts. In a recent article, Testa and colleagues pointed out the following: Effective chemistry teaching requires attention to general pedagogical issues, including how to maximize student engagement and how to most productively support the learning of chemical concepts. For student engagement, research has shown that creating authentic or simulated contexts for learning generates a sense of relevance for students, which can have a positive effect on their intrinsic motivation and enhance their learning capacity. Such contexts, though, must be designed to adequately account for the conceptual difficulties inherent in the chemical concepts they are meant to address. Their article details an experiment to solve a murder mystery employing the permanganate oxidation of DNA nucleotides (66). Testa further discusses the use of forensic science to motivate general chemistry students in a separate chapter of this ACS Symposium Series (67).

Organic Chemistry A variety of the concepts presented in organic chemistry can be illustrated using forensic examples. Some of these examples have been developed for the organic chemistry lab (68). It can be noted, for example, that the same classical organic technique for distinguishing alcohols is used in the Breathalyzer for breath alcohol detection (the reaction of potassium dichromate with alcohol mentioned above) and, when introducing infrared spectroscopy, it would be engaging to discuss that the Intoxilyzer, also for breath alcohol detection, is based on the infrared spectrum of ethanol (69). One of the primary examples showing the relationship between organic chemistry and forensic science involves the application of basic organic chemical principles to drug molecules containing various organic functional groups that gives rise to their polarity and resultant solubility in hydrophobic and hydrophilic solvents. This, of course, allows for a ready understanding about the methods of detecting different drugs using thin-layer chromatography or gas chromatography-mass 18 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

spectrometry. Adding the effect pH has on the drugs introduces the reason and effect of the different methods of ingestion of drugs in different ionic states. These concepts are discussed after drug related cases such as the deaths of Marilyn Monroe, John Belushi, and Janice Joplin are introduced (70). In general, drugs are detected first by a presumptive method and then confirmed by an evidentiary method (71, 72). Presumptive methods for drug identification include color tests (either in a spot plate or with the prepackaged NIK System (73)) or by thin-layer chromatography. The confirming, evidentiary tests include either infrared spectroscopy or gas chromatography-mass spectrometry (GC-MS) (74). For drugs in urine (75), a presumptive test such as the competitive antibody binding assay in the STAT Cassette from Micro-Distributing (76) is quick and easy and the GC-MS evidentiary test is conducted after extraction of the drug from the urine. While some are concerned about the availability of illicit drugs in the undergraduate teaching laboratory, a number of common compounds provide a false positive when tested with the presumptive color tests. The author has found that diphenhydramine yields a false positive for cocaine, Patchouli oil (allow a small quantity it to be absorbed on a spice such as oregano, basil, or parsley) yields a false positive for marijuana, Mucinex DM (which contains dextromethorphan) gives a false positive for heroin and morphine, and Dr. Bronner’s Magic Soap (which contains hemp oil) gives a false positive for gamma hydroxybutyric acid (GHB) (77, 78). For analysis of drugs in the urine, Sigma Chemical Company and Cerilliant sell many drugs in 1 mg/1 mL ampules in either methanol or acetonitrile (available without a Drug Enforcement Administration license). In either case, these drugs can be added to synthetic urine at a rate of 1 mL per 50 mL to give an ideal solution to test with the STAT cassette or from which to extract the drug and analyze on the GC-MS.

Analytical Chemistry As Rose and Fitzgerald (79) stated, “The challenge for any analytical chemistry instructor is to teach the fundamental methods and techniques of analytical chemistry while maintaining the enthusiasm of the students.” Some instructors introduce their students to an extensive range of equipment with little integration—essentially discussing each piece of equipment as an isolated unit—while presenting each instrument as one of a number of tools employed in an investigation is far more effective. This way, the strengths and limitations of each instrument as well as its relationship to the results of other equipment can be emphasized. Using forensic science is an ideal way to achieve this integration. Once again, it is effective to introduce the investigation with a crime scenario and emphasize the use of the various instrumentation as the pieces of evidence are analyzed both presumptively and evidentiarily. Both Shulman (80) and Thompson (81) have written laboratory manuals for this type of course. More advanced monographs of the various instrumentation are available, such as that by Yinon on mass spectrometry (82). A monograph like this provides the background for the use of mass spectrometry in a wide range of situations that could suggest a variety of crime scenarios. The applications include the following: analysis of body fluids and hair for drugs of abuse, drug testing in sports, analysis of accelerants in fire debris, detection of hidden explosives in luggage and mail, identification of explosives in post-explosion debris, examination of evidential materials (paints, fibers, synthetic polymers), authentication of regulated products (flavoring substances, fruit juices), and protection of industrial products by isotopic signature. Other monographs at this level include: Fundamentals of Fingerprint Analysis (83), Scientific Protocols for Fire Investigation (84), as well as the comprehensive Instrumental Data for Drug 19 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Analysis (85). The Forensic Science Handbooks (86–88) and the Forensic Chemistry Handbook (89) provide advanced articles on a wide range of topics by “recognized experts well versed in the practices of their chosen areas of expertise.” Finally, publications in the original literature are constantly updating material that would be very valuable for these courses, such as two recent articles that deal with the use of vibrational spectroscopy to revolutionize areas for forensic investigation (90) and recent developments in trace evidence analysis (91). To provide some concrete examples, consider the John F. Kennedy assassination (92, 93) and the elemental analysis of the bullet fragments removed from President Kennedy and the limousine in which he was riding. The main controversy in resolving the assassination was this: did Lee Harvey Oswald have one or more accomplices shooting from the grassy knoll? It was assumed that if the elemental analysis indicated that all of the bullet fragments came from Oswald’s MannlicherCarcano rifle, no accomplices were implicated. In 1979, Guinn (94) performed such elemental analysis, and while there was considerable variation, concluded that the bullet fragments represented only two bullets, both fired from that rifle. Not until 2006 did Randich and Grant (95) point out serious problems with Guinn’s data and conclusions. They noted that, although Guinn worked in good faith, he did not understand elemental bullet metallurgy. They claim that Guinn was wrong that Western Cartridge Company (WCC) bullets were unique because the trace levels of Sb varied. The levels of antimony in WCC bullets do vary, but so do Sb levels in many bullets manufactured by other companies that, like WCC shells, are jacketed. It is the non-jacketed bullets that have consistent levels of trace components. Randich and Grant also disproved another key Guinn contention that there is little variation in Sb levels within a single bullet. Using exquisite micrographs showing WCC bullets cut in cross section, Randich and Grant demonstrated that WCC bullet lead exhibits a “crystalline” type structure, with Sb tending to “microsegregate” around crystals of lead. The crystals are large enough that a sample taken from one portion of a bullet might easily have an Sb level one or two orders of magnitude higher or lower than one taken from another portion of the same bullet. They claim that Guinn found Sb matches within the WCC bullets he tested because he measured bits taken from only a very small portion of his test bullets, which said nothing about what he would have found had he sampled an entirely different area of the bullet. Thus, fragments with similar antimony levels could have come from one bullet or more than one, and those with different antimony levels could have come from but a single bullet. Taken from this discussion of the Kennedy bullet fragments, one can broaden the discussion to the general area of “sampling” when analyzing a piece of evidence or any object under investigation. One exercise has been to provide students with a large beaker of pennies and ask them to determine the density of a penny. They are provided with a graduated cylinder, a balance, and water. They usually do not know that the composition of the penny was changed in 1982 from all copper to copper-clad zinc. The beaker contains a layer of pre-1982 pennies with another layer of post-1982 pennies on top. (Pennies made in early 1982 are 95% copper and weigh approximately 3.11 g; pennies made after mid-1982 are 97.5% zinc and weigh 2.5 g.) So as the students scoop pennies from the beaker for analysis, the first groups get mostly post-1982 pennies and the next groups get various mixtures of the two types of pennies and finally the last few groups get mostly pre-1982 pennies. After their experimental determination, the reported density of a penny varies considerably from group to group. In over 35 years of having students perform this experiment, no one has mixed the pennies to ensure a homogeneous mixture before removing a sample. These concepts have relevance in many areas of forensic analysis as well as analytical chemistry in general. Already mentioned was the analysis of the linen from the Shroud of Turin. But also 20 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

consider what one must do when analyzing a brick of suspected illicit drug. Should the analysis be of a sample scraped from a corner of the brick or of a core through the brick to be sure the analyses are representative of the entire sample? Let us return to the discussion about drug testing and propose a wonderful experiment involving the determination of the presence of cocaine on paper money (74, 96). Due to the extensive drug market in the United States, it is claimed that all paper currency is contaminated with cocaine as well as a number of other illicit drugs. If a bill is rolled up and placed in a test tube with methanol for brief time, the drugs will be extracted from the paper. Evaporating some of the methanol concentrates the drugs and analysis with GC-MS will invariably show a peak for cocaine. Many more examples could be presented, but it is fair to say that an entire course in analytical chemistry/instrumental analysis could be based on a forensic science theme (52, 97, 98).

Inorganic Chemistry Some of the examples cited in previous sections would be applicable in an inorganic chemistry course. Also, the development of an inkless fingerprinting system and the cause of the death of Napoleon are particularly suited for discussion in inorganic chemistry. About 40 years ago, a new method for taking fingerprints was developed by Identicator Corporation (73). This method, called inkless fingerprinting, does not use ink but involves the use of two reagents, which, when mixed produced a blue-black color. One reagent, a ferric chloride (FeCl3) solution is dispensed from a Poralin disk contained in a plastic housing. The other reagent, n-propylgallate, is impregnated in a paper fingerprint “card.” When the finger is pressed onto the Poralin pad, the ridges of the finger pick up some FeCl3. When the finger is rolled on the propylgallate paper, a blue-black print is formed as the Fe3+ complexes with three molecules of n-propylgallate (99). The structure of the complex is shown in Figure 1.

Figure 1. Structure of complex formed in the inkless fingerprinting system. It is well known that gallol and many functionalized derivatives, such as the deprotonated npropylgallate, are effective metal chelators. Metal ions that prefer octahedral geometry, such as Fe2+ and Fe3+, can coordinate up to three gallate groups. Because of this, it might be expected that polyphenols with gallol groups would always bind iron in a 3:1 fashion. However, since polyphenol compounds are so structurally varied and the complexes formed are pH dependent, they often exhibit variable coordination modes. Despite pKa values in the range of 7–9 for the most acidic 21 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

phenolic hydrogen, polyphenols are easily deprotonated at or below physiological pH in the presence of iron and form very stable complexes (100, 101). The deprotonated gallate ligands behave as hard Lewis bases and give rise to particularly large metal-binding stability constants with the hard Lewis acid, Fe3+. When one Fe3+ion reacts with three molecules of n-propylgallate, a complex is formed with octahedral coordination geometry (100). With this as an entree, one can discuss a variety of inorganic chemical concepts such as coordination complexes formed from the interaction of various metal ions with a range of ligands (charged or neutral, saturated or unsaturated), the crystal field stabilization that arises from specific electronic configurations, and so forth. In addition, this material could lead to a discussion of the fact that iron is a primary cause of reactive oxygen species generation in vivo, and because it plays such a pivotal role in contributing to oxidative stress, DNA damage, and cell death, iron has been the target of many antioxidant therapies. Because of their ability to coordinate iron, polyphenols are one large class of antioxidants that have been extensively examined for treatment and prevention of conditions associated with iron-generated reactive oxygen species and oxidative stress (100). It also is interesting to point out that while the inkless fingerprinting system was developed for use in forensic science, it has now found use in a wide variety of applications where authentication is required (102). Napoleon was defeated in the Battle of Waterloo in June 1815 by Wellington. He died six years later, in 1821, on the island of St. Helena off the coast of Africa, where he had been exiled by the British. Some say he died of arsenic poisoning, some say of stomach cancer. In order to test the rumors that had persisted of arsenic poisoning, in 1961, samples of Napoleon’s hair were subjected to neutron activation analysis for arsenic (103). The result of that analysis was that a significant amount of arsenic was present in Napoleon’s hair. The obvious conclusion was that the emperor had been murdered by poisoning. But the question remained—who was responsible? Many years later the British chemist David Jones asked the interesting question on a radio show, “Does anyone know the color of Napoleon’s wall paper on St. Helena?” One of the listeners called and said she knew the color of the wallpaper and, in fact, had a piece of it, which she obtained from a trip to the island. When Jones was told that the paper had a green color, he was eager to test the paper for arsenic. He speculated that the green compound, Scheele’s green (cupric arsenite, CuAsHO3), might have been used in the wallpaper. He performed a scanning electron microscope–energy dispersive xray (SEM-EDX) analysis and, upon microscopically analyzing the green pigment on the wallpaper, confirmed that it contained arsenic (104). The question still remained, how did the arsenic from the wallpaper get into Napoleon? The answer comes from the climate on St. Helena, which is hot and wet and promotes the formation of mold. It had been shown as early as 1815, that cases of arsenic poisoning were caused by wallpaper containing such arsenic compounds as Scheele’s green. The mechanism was first thought to be the ingestion or inhalation of dust particles that had flaked off the wallpaper; but, when poisoning occurred with fresh paper, that theory was abandoned. After realizing that the rooms where symptoms occurred had a garlic odor, the symptoms were ascribed it to a volatile arsenic compound called arsine (H3As) produced by molds on damp arsenic-pigmented wallpaper (105, 106). While there is evidence of arsenic in the remains of the Emperor, there also is evidence of stomach cancer. That evidence stemmed from weight loss studies, an autopsy, eyewitness accounts, and medical histories of family members and is not relevant to inorganic chemistry and will not be discussed here. 22 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

While the methods used for the detection of arsenic in the hair (neutron activation analysis) and wallpaper (SEM-EDX) are modern techniques, the original, reliable method for determining arsenic, especially that used as the “inheritance powder,” was the Marsh test (107, 108). In fact, the arsenic in question was generally not the pure chemical element itself but rather its oxide, As2O3 or diarsenic trioxide. Marsh (a chemist working at the Royal Arsenal in Woolwich, England) was frustrated as a result of his inability to help obtain a conviction of a man accused of poisoning his grandfather even though a selective precipitation test had indicated arsenic in the grandfather’s coffee. Unfortunately, the precipitate had decomposed and was unavailable for presentation at trial. As a result of the frequent need for arsenic analyses, Marsh developed his novel test. It depended on the fact that when arsenical material came in contact with hydrogen freshly liberated by the action of zinc and dilute acid, hydrogen gas containing a small amount of arsine was evolved. The mixed gases were then ignited and the flame was allowed to play on a cold plate of window glass, on which was then deposited a metallic mirror consisting largely of metallic arsenic. The method could be applied not only to inorganic materials but to biological substances as well, and hence could be used for toxicological and forensic purposes (108). A thermodynamic and kinetic analysis of the actual reactions involved in the Marsh test present puzzling chemical questions. In fact, the diarsenic trioxide has been shown to be composed of discrete molecules of As4O6, tetraarsenic hexaoxide. The steps originally proposed are shown in Figure 2.

Figure 2. The steps originally proposed for the reactions involved in the Marsh test for arsenic. However, it was not clear how the overall reaction would proceed since while the first step and the last step are thermodynamically favorable, the second step is not. In fact, it was known that As4O6(s) could not be reduced to H3As(g) using molecular hydrogen. This was at odds with the fact that the reaction took place in the Marsh apparatus and led to the speculation that the actual reactive specie was monoatomic hydrogen. This led to the development of the “nascent state” concept, which proposed that freshly generated gases are initially monoatomic and that the transient, free atoms later combine to form the diatomic gas (107).

Biochemistry In the field of biochemistry, the techniques of DNA profiling provide terrific examples of the modern genetic manipulation of DNA. These have recently exploded on the public scene with so many eager to learn their genetic background employing commercially available kits. However, let us start with the simple detection of blood at a crime scene since it perhaps is the most visible illustration of the Locard Exchange principle. Luminol, leuchomalachite green, phenolphthalein, Hemastix, Hemident, and Bluestar are all used as presumptive tests for blood. A suspected sample of blood is gently swabbed and the swab is tested with the reagent, which contains a peroxide. The formation of a color is indicative of the presence of blood, although not conclusive. The color results from the reaction of the reagent with the heme group in the hemoglobin (which has

23 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

peroxidase-like activity and produces oxygen free radicals from the peroxide). The oxygen free radical converts the reduced form of the reagent to the oxidized form, which is colored. To perform this test, one needs either real blood or a synthetic blood that is formulated to give a positive test. Much of the synthetic (or fake) blood commercially available is fine for blood spatter patterns but will not yield a positive presumptive blood test. Arrowhead Forensics (109) supplies a specially formulated “synthetic” or “mock” non-biological blood that will give a positive presumptive test. Perhaps the simplest and least expensive way to perform the test is to use Hemotest tablets, available at a local pharmacy. A more professional approach (one that the students have seen on television) involves using a product such as Quick Check (110). Blood typing would be the next step in the characterization of the blood sample (111). Once again, a synthetic, non-biological system is available for blood typing (112). The kit contains synthetic blood of all different types and “antisera” for type A cells, type B cells, and Rh+ cells. With this kit, one can conduct the same type of procedure performed in forensic and clinical labs for blood typing. Since the typing process involves the binding of antibodies to the carbohydrate portion of proteins bound in the red cell membrane, this provides the opportunity to discuss antibody-antigen interaction as the red blood cells undergo agglutination. And, of course, a case is needed—the O. J. Simpson case. When the Simpson case developed in the mid-1990s, DNA profiling was still in its infancy and standard blood typing was widely used. Kolata (113) described the initial blood typing results as follows: One test determines blood type by examining the ABO blood proteins: the drop of blood at the townhouse was Type A blood, the same type shared by O. J. Simpson and his former wife [Nicole Brown Simpson]. Ronald L. Goldman, who was killed along with Mrs. Simpson, was Type O. These results mean that the drop could not have come from Mr. Goldman. The second test looked at esterase D, or ESD, an enzyme that has three variants. In this case, everyone had the same type as the bloodstain—Type 1—so this test did not exclude anyone. The third test looked at phosphoglucomutase, or PGM, which has 10 different forms. In the test performed by the prosecutors, the bloodstain was PGM Type 2+ 2- and so was Mr. Simpson’s. Mrs. Simpson and Mr. Goldman were of different types, so the dried blood did not come from them. Of course, DNA profiling is now the most widely used method (and the most widely accepted) method for individualization (even more than fingerprinting, which recently has been challenged in court). DNA profiling was developed in 1984 by Jeffreys (114, 115) employing the restriction fragment length polymorphism method with subsequent identification of the fragment lengths by electrophoresis. The technique has grown and matured and is now routinely used for forensic as well as non-forensic cases where individualization is needed. Elkins and Kadunc (116) have reported on the amplification of human DNA using short tandem repeats (STR) single locus primers by real-time polymerase chain reaction (PCR) with SYBR Green detection, and Jackson et al. (117) have reported on the profiling of the D1S80 locus. The current technique involves amplification of 20 STRs in the nuclear DNA by PCR. The resultant pieces of DNA are analyzed by capillary electrophoresis. For mitochondrial DNA, PCR is used to amplify the hypervariable regions and the resultant DNA is subjected to direct sequencing or next generation sequencing (118). For the latest developments in the forensic analysis of DNA, see the article by McCord et al. (119).

24 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

While many cases could be used as examples of the value of DNA profiling for individualization, here we will discuss the classic case of the identification of the Romanov bones and the identification of Anastasia. Then, to bring DNA profiling up-to-date, the next example will be use of “ultra-deep next generation [DNA] sequencing” to differentiate between “identical” twins. In 1918, Tsar Nicholas II, his wife, and five children, along with three family servants and the family physician, were executed by a firing squad. The Tsar’s daughter, Anastasia, was 17 at the time of execution. After killing them, the bodies were transported by truck to a forest and buried. In 1921, a woman claiming to be Anastasia appeared in Germany, eventually moving to the United States and calling herself Anna Anderson. She married Richard Manahan. She died in 1984 and was cremated (120). In 1979, graves were located in Russia that were attributed to the Romanovs, but the bodies were not exhumed until 1991. Based on the skeletal remains (presumably of the Romanovs), Dr. Sergei Abramov claimed that he identified Nicholas, Alexandra, their physician, their three servants, and three of the five children - Olga, Tatiana, and Anastasia. Alexis and Marie were missing. Later, Dr. William Maples identified Nicholas, Alexandra, their physician, their three servants, and three of the five children—Olga, Tatiana, and Marie. Anastasia and Alexis were missing (121). Since mitochondrial DNA is passed through a maternal line, it would be possible to determine if the bones were truly Romanov and to determine if Anna Anderson was Anastasia if there was a female descendent (or first generation male descendent) of the Romanovs for comparison. For comparison of the alleged Romanov bones, the samples were compared to biological material from Romanov descendants, Xenia Sfiris, James G. A. B. Carnegie, and Prince Philip of England. The bones were determined to be Romanov, but the question remained as to whether some of the bones were those of Anastasia (122). Although Anna Anderson’s remains were cremated, samples of her hair and medical samples from a biopsy were available for analysis. DNA profiling showed these samples did not have a common origin with the Romanov remains or living Romanov relatives. Analysis of Anderson’s mitochondrial DNA proved that she was not Anastasia. The DNA profile did match that of Karl Maucher, a great-nephew of Franziska Schanzkowska, a Polish factory worker (123). In 2007, an additional grave was excavated just meters away from the earlier one. More bones were discovered and, through mitochondrial DNA, shown to be Romanov. Profiling nuclear DNA using STRs demonstrated that bones from all of the Tsar’s children, including Anastasia, were present (124). While DNA profiling may be the standard for bringing criminals to justice, when it comes to identical twins, standard testing cannot tell the difference. When in 2014, a crime scene DNA sample showed a match to a suspect (Dwayne McNair) in two rape cases in Boston in September 2004, it also showed a match to his twin brother (Dwight McNair). The prosecutors requested the court to allow them to submit evidence distinguishing the twins using “next generation,” or “massive parallel,” sequencing, which enables scientists to map the entire genome of each twin. The goal was to find the relatively few mutations that might have arisen during cell division, in both twins, during which most of the three billion base pairs in the genome are faithfully copied (125). The prosecutors were aware of a paper by the German company Eurofins (126) that used ultradeep next generation sequencing to differentiate between identical twins. As part of their paper’s research (having nothing to do with this crime), they reported on their DNA analysis a child, his mother, and a set of monozygotic twins—one of whom was the child’s father and the other the child’s uncle. They tested the blood of the child, the buccal mucosa of the mother and each of the twins, and the blood and the sperm of the father and the uncle. They found mutations yielding five 25 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

single nucleotide polymorphisms present in the twin father and child but not in the uncle, thus identifying which of the twins was the father. When, in late 2013, the scientists at Eurofins subjected the saliva from Dwight and Dwayne McNair to this test, they found nine differences between their genomes. After establishing the genetic differences, Eurofins compared both DNA profiles from the saliva with the DNA of the semen found at the site of the rape. Analysts found seven locations where the DNA of Dwayne McNair’s saliva showed a common origin with the semen and two where his brother Dwight showed a common origin. In addition, scientists found two of the mutations in Dwayne’s saliva also were present in the semen sample. In Dwight’s case, none were present. On that basis, Eurofins concluded Dwayne McNair was 2 billion times more likely than Dwight to be the source of the semen and therefore the rapist. Dwayne was indicted on charges of rape, based on these results, and on armed robbery (127). A few months after, Superior Court Judge Linda Giles ruled the technique was not admissible because it had not been replicated or peer reviewed and was not accepted by the general genetics community. While this, strictly speaking, was consistent with the criteria for evidence to be admitted in court, the manuscript for a paper presenting the details of the “ultra-deep next generation sequencing” (the one discussed above) had been submitted and accepted for publication, after peer review, in October 2013. It was published in early 2014. While most scientists are concerned with publication of their research, forensic scientists also have to be concerned with the admissibility of their results in court. For many years, the Frye (128) ruling was the standard by which evidence was evaluated for admissibility in courts. The Frye, or general acceptance ruling, in 1923 was handed down in a case involving the admissibility of lie detector evidence. It has been replaced, in federal cases, with the 1993 Supreme Court ruling in Daubert v. Merrell Dow Pharmaceuticals (129). In this ruling, the Supreme Court set out a number of conditions to be considered when seeking admissibility of relatively new techniques and procedures. Under the Daubert standard, the factors that may be considered in determining whether the methodology is valid are the following: (1) whether the theory or technique in question can be and has been tested; (2) whether it has been subjected to peer review and publication; (3) its known or potential error rate; (4) the existence and maintenance of standards controlling its operation; and (5) whether it has attracted widespread acceptance within a relevant scientific community. However, the Supreme Court emphasized that these factors were not to be considered a checklist and all need not be present for the methodology to be deemed legally admissible. In January 2018, the Suffolk Superior Court jury found Dwayne McNair guilty of abducting and sexually assaulting two women within nine days of each other. While part of the evidence used against him was the DNA profile (even though it also “matched” his twin), the primary evidence was the testimony of Anwar Thomas, an accomplice, who testified at trial that he could distinguish between the two brothers and that it was Dwayne McNair who had taken part in the assaults with him.

The Importance of Diversity and Inclusion in the Forensic Sciences With all of discussion and examples of the role forensic science can play in a wide variety of courses, let us consider a topic that seems to be on most minds. The issues of diversity and inclusion are gaining a lot of attention and many mechanisms are being put into place to help ensure that all are afforded equal opportunities. In 2018, the National Institute of Justice posted an article on its website (130) discussing the importance of diversity and inclusion in the forensic sciences. The authors stated, “Having people 26 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

of different races, ethnicities, genders, socioeconomic status, and backgrounds within a workplace can drive innovation, problem-solving, and competitiveness,” and “Increasing diversity in thought, perspectives, and backgrounds allows for new and more complex research questions and problems to be addressed.” They made a compelling argument that, “As the forensic sciences continue to evolve, it is critical that we leverage the skills and expertise of people from all backgrounds to provide innovative solutions to complex issues.” They also point out the primary areas of forensic science (forensic biology and DNA, medicolegal death investigation, forensic toxicology, controlled substances, fire and arson investigation, impression and pattern evidence, firearms and tool marks, blood stain pattern analysis, questioned documents, trace evidence, crime scene investigations, and digital evidence) require study and competence in diverse scientific areas (computer science, statistics, mathematics, geology, physics, spectroscopy, all aspects of chemistry, and many aspects of biology, including molecular biology and genetics). The point here is that the multidiscipline and interdiscipline area of forensic science is ideally suited to provide the vehicle to drive an increase of participation in the STEM fields and to help seal the “leaky pipeline.” Individuals from all backgrounds, races, ethnicities, genders, and socioeconomic status can find some area of forensic science that is attractive and compelling and in which each can excel. As stated earlier, all seem to be fascinated by the challenge of “who dun it.”

Conclusion Hopefully this article has been convincing in demonstrating the enormous potential forensic science has in stimulating student interest in many aspects of chemistry, in particular, and science, in general. Besides the examples given and the numerous quotations of colleagues who have found forensic science effective in their classrooms, the author has conducted 23 weeklong workshops in forensic science for college and university professors. These workshops were sponsored by grants from the National Science Foundation and coordinated by the Center for Workshops in the Chemical Science (CWCS, later rebranded, Chemistry Collaborations, Workshops and Communities of Scholars, cCWCS). The details of these workshops have been summarized in two recent papers (131, 132). Approximately 400 faculty participated in these workshops and developed curricular materials (including demonstrations, experiments, modular units, whole courses, and even programs in forensic science) as tools to introduce a new generation of students to the fascination of chemistry.

References 1. 2. 3. 4. 5.

6.

Guzzetti, B. Thinking Like a Forensic Scientist: Learning with Academic and Everyday Texts. J. Adolesc. Adult Literacy 2009, 53 (3), 192–203. Kaplan, L. J. Chemistry and Crime: From Sherlock Holmes to Modern Forensic Science—A Science Course for Non-Science Majors. Crime Lab. Dig. 1992, 19, 107–132. Kaplan, L. J. Forensic Science: Crime in the Chemistry Curriculum. J. Chem. Educ. 1993, 70, 574. Funkhouser, J.; Deslich, B. J. Integrating Forensic Science. Sci. Teach. 2000, 67, 32–35. Charkoudian, L. K.; Heymann, J. J.; Adler, M. J.; Haas, K. L.; Mies, K. A.; Bonk, J. F. Forensics as a Gateway: Promoting Undergraduate Interest in Science and Graduate Student Professional Development through a First–Year Seminar Course. J. Chem. Educ. 2008, 85, 807–812. Peterson, J. L.; Angelos, S. A. Characteristics of Forensic Science Faculty within Criminal Justice Higher Education Programs. J. Forensic Sci. 1983, 28, 552–559. 27 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30.

Gaensslen, R. E.; Lee, H. C. Regional Cooperation and Regional Centers Among Forensic Science Programs in the United States. J. Forensic Sci. 1988, 33, 1069–1070. Califano, A. L.; Levkov, J. Criminalistics for the Law Enforcement Officer; McGraw-Hill: New York, NY, 1978. Goddard, K. Crime Scene Investigation; Reston: Reston, VA, 1977. O’Brien, K. P.; Sullivan, R. Criminalistics: Theory and Practice; Allyn & Bacon: Boston, MA, 1972. Saferstein, R. Criminalistics: An Introduction to Forensic Science; Prentice Hall: Upper Saddle River, NJ, 1976. DeForest, P. R.; Gaensslen, R. E.; Lee, H. C. Forensic Science: An Introduction to Criminalists; McGraw-Hill: New York, NY, 1983. Saferstein, R. Criminalistics: An Introduction to Forensic Science, 12th ed.; Pearson Education: Upper Saddle River, NJ, 2017. Gaensslen, R. E.; Harris, H. A.; Lee, H. C. Introduction to Forensic Science and Criminalists; McGraw-Hill: New York, NY, 2008. Madhusoodanan, J. Forensic Science: The Promise and Perils of Using Science in the Courtroom; American Chemical Society: Washington, DC, 2017. Bertino, A. J.; Bertino, P. Forensic Science: Fundamentals and Investigations, 2nd ed.; Cengage Learning: Boston, MA, 2015. Eckert, W. G. Introduction to Forensic Sciences, 2nd ed.; CRC Press: Boca Raton, FL, 1992. Elkins, K. M. Introduction to Forensic Chemistry; CRC Press: Boca Raton, FL, 2019. Fisher, B. A. J.; Tilstone, W. J.; Woytowicz, C. Introduction to Criminalistics: The Foundation of Forensic Science; Elsevier: Burlington, MA, 2009. Girard, J. E. Criminalistics: Forensic Science, Crime, and Terrorism; Jones and Bartlett Learning: Burlington, MA, 2017. Houck, M. M. Forensic Chemistry; Academic Press: San Diego, CA, 2015. Houck, M. M.; Siegel, J. A. Fundamentals of Forensic Science, 2nd ed.; Academic Press: Burlington, MA, 2010. James, R. E.; Nordby, J. J.; Bell, S. Forensic Science: An Introduction to Scientific and Investigative Techniques, 4th ed.; CRC Press: Boca Raton, FL, 2014. Siegel, J. A. Forensic Chemistry: Fundamentals and Applications; Wiley-Blackwell: Hoboken, NJ, 2015. Johll, M. E. Investigating Chemistry: Introductory Chemistry From A Forensic Science Perspective, 3rd ed.; W. H. Freeman: New York, NY, 2013. Bell, S. Forensic Chemistry; Pearson/Prentice Hall: Upper Saddle River, NJ, 2012. Siegel, J. A.; Mirakovits, K. Forensic Science: The Basics, 3rd ed.; CRC Press: Boca Raton, FL, 2015. Fisher, B. A. J.; Fisher, D. R. Techniques of Crime Scene Investigation, 8th ed.; CRC Press: Boca Raton, FL, 2012. Houck, M. M.; Crispino, F.; McAdam, T. The Science of Crime Scenes; Academic Press: San Diego, CA, 2012. Erickson, E. Criminalistics Laboratory Manual: the Basics of Forensic Investigation; Routledge: New York, NY, 2013. 28 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

31. Kaplan, L. J. Chemistry and Crime: From Sherlock Holmes to Modern Forensic Science: A Laboratory Manual for An Introduction to the Crime Lab; Williams College: Williamstown, MA, 2017. 32. Kubic, T.; Petraco, N. Forensic Science: Laboratory Experiment Manual and Workbook, 3rd ed.; CRC Press: Boca Raton, FL, 2009. 33. Meloan, C. E.; James, R. E.; Thomas, B.; Saferstein, R. Lab Manual for Criminalistics: An Introduction to Forensic Science, 11th ed.; Pearson: Upper Saddle River, NJ, 2014. 34. Kaye, B. H. Science and the Detective; VCH Publishers: New York, NY, 1995. 35. Owen, D. Hidden Evidence, 2nd ed.; Firefly Books: Buffalo, NY, 2009. 36. Locard, E. La Police et les Méthodes Scientifiques; Rieder: Paris, France, 1934. 37. Kirk, P. L. Crime Investigation: Physical Evidence and the Police Laboratory; Interscience Publishers, Inc.: New York, NY, 1953. 38. Chicago Tylenol Murders. https://en.wikipedia.org/wiki/Chicago_Tylenol_murders (accessed Jan 10, 2019). 39. Wolnik, K. A.; Fricke, F. L.; Bonnin, E.; Gaston, C. M.; Satzger, R. D. The Tylenol Tampering Incident—Tracing the Source. Anal. Chem. 1984, 56, 466A–474A. 40. Shroud of Turin Website. https://www.shroud.com/menu.htm (accessed Jan 6, 2019). 41. Damon, P. E.; Donahue, D. J.; Gore, B. H.; Hatheway, A. L.; Jull, A. J. T.; Linick, T. W.; Sercel, P. J.; Toolin, L. J.; Bronk, C. R.; Hall, E. T.; Hedges, R. E. M.; Housley, R.; Law, I. A.; Perry, C.; Bonani, G.; Trombore, S.; Woelfli, W.; Ambers, J. C.; Bowman, S. G. E.; Leese, M. N.; Tite, M. S. Radiocarbon Dating of the Shroud of Turin. Nature 1989, 337, 611–615. 42. Orna, M. V. Doing Chemistry at the Art/Archaeology Interface: 1996 Norris Award Address. J. Chem. Educ. 1997, 74, 373–376. 43. Nabais, J. M. V.; Costa, S. D. A Forensic Experiment: The Case of the Crime at the Cinema. J. Chem. Educ. 2017, 94, 1111–1117. 44. Cresswell, S. L.; Loughlin, W. A. An Interdisciplinary Guided Inquiry Laboratory for First Year Undergraduate Forensic Science Students. J. Chem. Educ. 2015, 92, 1730–1735. 45. Cresswell, S. L.; Loughlin, W. A. A Case-Based Scenario with Interdicsciplinary GuidedInquiry in Chemistry and Biology: Experiences of First Year Forensic Science Students. J. Chem. Educ. 2017, 94, 1074–1082. 46. Harmon, K. J.; Miller, L. M.; Millard, J. T. Crime Scene Investigation in the Art World: The Case of the Missing Masterpiece. J. Chem. Educ. 2009, 86, 817–819. 47. King, A. G. Research Advances: Improved Methods for Detecting Art Forgeries and Blood Types. J. Chem. Educ. 2010, 87, 1284–1285. 48. Nivens, D. A.; Padgett, C. W.; Chase, J. M.; Verges, K. J.; Jamieson, D. S. Art, Meet Chemistry; Chemistry, Meet Art: Case Studies, Current Literature, and Instrumental Methods Combined To Create a Hands-On Experience for Nonmajors and Instrumental Analysis Students. J. Chem. Educ. 2010, 87, 1089–1093. 49. Andre, J. P. Opera and Poison: A Secret and Enjoyable Approach To Teaching and Learning Chemistry. J. Chem. Educ. 2013, 90, 352–357. 50. Harper-Leatherman, A. S.; Miecznikowski, J. R. O True Apothecary: How Forensic Science Helps Solve a Classic Crime. J. Chem. Educ. 2012, 89, 629–635.

29 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

51. Kloepper, K. D. Bringing in the Bard: Shakespearean Plays as Context for Instrumental Analysis Projects. J. Chem. Educ. 2015, 92, 79–85. 52. Lucy, C. A. Analytical Chemistry: A Literary Approach. J. Chem. Educ. 2000, 77, 459–470. 53. The Official Site of the Renowned English Crime Writer Dorothy L. Sayers. https://www.sayers.org. uk/ (accessed Jan 10, 2019). 54. The Home of Agatha Christie. https://www.agathachristie.com/ (accessed Jan 10, 2019). 55. The Official Site of the Sir Arthur Conan Doyle Literary Estate. https://www.arthurconandoyle. com/ (accessed Jan 10, 2019). 56. Patricia Cornwell. http://www.patriciacornwell.com/ (accessed Jan 10, 2019). 57. Kathy Reichs. https://kathyreichs.com/ (accessed Jan 10, 2019). 58. Koehler, B. G.; Park, L. Y.; Kaplan, L. J. Science for Kids Outreach Programs: College Students Teaching Science to Elementary School Students and Their Parents. J. Chem. Educ. 1999, 76, 1505–1509. 59. Meyer, A. F.; Knutson, C. M.; Finkenstaedt-Quinn, S. A.; Gruba, S. M.; Meyer, B. M.; Thompson, J. W.; Maurer-Jones, M. A.; Halderman, S.; Tillman, A. S.; DeStefano, L.; Haynes, C. L. Activities for Middle School Students to Sleuth a Chemistry “Whodunit” and Investigate the Scientific Method. J. Chem. Educ. 2014, 91, 410–413. 60. Jacobsen, E. K. Sherlock Holmes Goes Virtual. J. Chem. Educ. 2011, 88, 368–369. 61. Waddell, T. G.; Rybolt, T. R. Stoichiometric Calculations. J. Chem. Educ. 1993, 70, 1003–1005. 62. Waddell, T. G.; Rybolt, T. R. The Hound of Henry Armitage. J. Chem. Educ. 1994, 71, 1049–1051. 63. Waddell, T. G.; Rybolt, T. R. The Blackwater Escape. J. Chem. Educ. 2003, 80, 401–406. 64. Labianca, D. A. The Chemical Basis of the Breathalyzer: A Critical Analysis. J. Chem. Educ. 1990, 67, 259–261. 65. The Alcohol Pharmacology Educational Partnership. https://sites.duke.edu/apep/module-4alcohol-and-the-breathalyzer-test/ (accessed Jan 10, 2019). 66. Testa, S. M.; Selegue, J. P.; French, A.; Criswell, B. The Permanganate Oxidation of DNA Nucleotides: An Introductory Redox Laboratory Framed as a Murder Mystery. J. Chem. Educ. 2018, 95, 1840–1847. 67. Testa, S. M. On Utilizing Forensic Science to Motivate Students in First Semester General Chemistry Laboratory. In Teaching Chemistry with Forensic Science; Harper-Leatherman, A. S., Huang, L. , Eds.; American Chemical Society: Washington, DC, 2019. 68. Boucher, M. A.; Specht, K. M. A Forensic-Themed Case Study for the Organic Lab. J. Chem. Educ. 2009, 86, 847–848. 69. Fiandach and Fiandach. http://www.nydwi.com (accessed Jan 6, 2019). This site of a law firm was selected since they discuss other issues related to breath alcohol testing. 70. Noguchi, T. T.; DiMona, J. Coroner; Simon and Schuster: New York, NY, 1983. 71. Schurter, E. J.; Zook-Gerdau, L. A.; Szalay, P. Analysis of a Suspected Drug Sample. J. Chem. Educ. 2011, 88, 1416–1418. 72. Szalay, P.; Zook-Gerdau, L. A.; Schurter, E. J. A Multi-Technique Forensic Experiment for a Nonscience-Major Chemistry Course. J. Chem. Educ. 2011, 88, 1419–1421.

30 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

73. Safariland Group, NIK. https://www.safariland.com/products/forensics/field-drug-tests/ nik-drug-tests-and-kits/ (accessed Jan 10, 2019). 74. Heimbuck, C. A.; Bower, N. W. Teaching Experimental Design Using a GC-MS Analysis of Cocaine on Money: A Cross-Disciplinary Laboratory. J. Chem. Educ. 2002, 79, 1254–1256. 75. Cerilliant. Surine Negative Urine Control. https://www.cerilliant.com/shoponline/Item_ Details.aspx?itemno=1cedba31-c7f4-4ab1-ba9f-a02f3328d204&item=S-020 (accessed Jan 10, 2019). 76. Micro-Distributing, STAT Cassette. http://www.micro-distributing.com/stat_cass.cfm (accessed Jan 10, 2019). 77. Anderson, C. Presumptive and Confirmatory Drug Tests. J. Chem. Educ. 2005, 82, 1809–1810. 78. Hasan, S.; Bromfield-Lee, D.; Oliver-Hoyo, M. T.; Cintron-Maldonado, J. A. Using Laboratory Chemicals to Imitate Illicit Drugs in a Forensic Chemistry Activity. J. Chem. Educ. 2008, 85, 813–816. 79. Rose, R. E.; Fitzgerald, N. Incorporating a Forensic Science Problem-Solving Activity into an Analytical Chemistry Laboratory Course. J. Chem. Educ. 2011, 16, 1–4. 80. Shulman, M. Chemical Instrumentation; Saint Mary’s College of California: Moraga, CA, 2009. 81. Thompson, R. Q. Instrumental Investigations: A Labortory Manual of Forensic Analytical Chemistry; Oberlin College: Oberlin, OH, 2016. 82. Yinon, J. Forensic Applications of Mass Spectrometry, 2nd ed.; CRC Press: Boca Raton, FL, 1994. 83. Daluz, H. M. Fundamentals of Fingerprint Analysis, 2nd ed.; CRC Press: Boca Raton, FL, 2018. 84. Lentini, J. J. Scientific Protocols for Fire Investigation, 3rd ed.; CRC Press: Boca Raton, FL, 2018. 85. Mills, T., III; Roberson, J. C.; Matchett, C. C.; Simon, M. J.; Burns, M. D.; Ollis, R. J., Jr. Instrumental Data for Drug Analysis, 3rd ed.; CRC Press: Boca Raton, FL, 2005. 86. Saferstein, R. Forensic Science Handbook, 2nd ed.; Prentice Hall: Upper Saddle River, NJ, 2001; Vol. 1. 87. Saferstein, R. Forensic Science Handbook, 2nd ed.; Pearson Prentice Hall: Upper Saddle River, NJ, 2004; Vol. 2. 88. Saferstein, R. Forensic Science Handbook, 2nd ed.; Pearson Education: Upper Saddle River, NJ, 2009; Vol. 3. 89. Kobilinsky, L. F. Forensic Chemistry Handbook; Wiley: Hoboken, NJ, 2011. 90. Muro, C. K.; Doty, K. C.; Bueno, J.; Halamkova, L.; Lednev, I. K. Vibrational Spectroscopy: Recent Developments to Revolutionize Forensic Science. Anal. Chem. 2015, 87, 306–327. 91. Mistek, E.; Fikiet, M. A.; Khandasammy, S. R.; Lednev, I. K. Toward Locard’s Exchange Principle: Recent Developments in Forensic Trace Evidence Analysis. Anal. Chem. 2019, 91, 637–654. 92. JFK Assassination Records—National Archives. https://www.archives.gov/research/jfk (accessed Jan 9, 2019). 93. Assassination of John F. Kennedy. https://www.history.com/topics/us-presidents/jfkassassination (accessed Jan 9, 2019). 94. Guinn, V. P. JFK Assassination: Bullet Analyses. Anal. Chem. 1979, 51, 484A–493A. 95. Randich, E.; Grant, P. M. Proper Assessment of the JFK Assassination Bullet Lead Evidence from Metallurgical and Statistical Perspectives. J. Forensic Sci. 2006, 51, 717–728. 31 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

96. Parker, P. D.; Beers, B.; Vergne, M. J. What is in Your Wallet? Quantitation of Drugs of Abuse on Paper Currency with a Rapid LC-MS/MS Method. J. Chem. Educ. 2017, 94, 1522–1526. 97. Brewer, W. E.; Lambert, S. J.; Morgan, S. L.; Goode, S. R. ComConf ’98: Teaching Forensic Analytical Chemistry. https://terpconnect.umd.edu/~toh/ChemConference/ChemConf98/ forensic/ChemConf.htm (accessed Jan 13, 2019). 98. Frederick, K. A. Using Forensic Science to Teach Method Development in the Undergraduate Analytical Laboratory. Anal. Bioanal. Chem. 2013, 405, 5623–5626. 99. Ponce, A.; Brostoff, L. B.; Gibbons, S. K.; Zavalij, P.; Viragh, C.; Hooper, J.; Alnemrat, S.; Gaskell, K. J.; Eichhorn, B. Elucidation of the Fe(III) Gallate Structure in Historical Iron Gall Ink. Anal. Chem. 2016, 88, 5152–5158. 100. Perron, N. R.; Brumaghim, J. L. A Review of the Antioxidant Mechanisms of Polyphenol Compounds Related to Iron Binding. Cell Biochem. Biophys. 2009, 53, 75–100. 101. Perron, N. R.; Wang, H. C.; DeGuire, S. N.; Jenkins, M.; Lawson, M.; Brumaghim, J. L. Kinetics of Iron Oxidation upon Polyphenol Binding. Dalton Trans. 2010, 39, 9982–9987. 102. FingerPrintPads.com. https://www.fingerprintpads.com/category/true-inkless-fingerprintingsystem/ (accessed April 3, 2019). 103. Forshufvud, S.; Smith, H.; Wassen, A. Arsenic Content of Napoleon I’s Hair Probably Taken Immediately after His Death. Nature 1961, 192, 103–105. 104. Jones, D. E. H.; Ledingham, K. W. D. Arsenic in Napoleon’s Wallpaper. Nature 1982, 299, 626–627. 105. Arsenic: Medical and Biological Effects of Environmental Pollutants; National Academy of Sciences: Washington, DC, 1977. 106. Kalia, K.; Joshi, D. N. Detoxification of Arsenic. In Handbook of Toxicology of Chemical Warfare Agents; Gupta, R. C. , Ed.; Elsevier Inc.: Cambridge, MA, 2009; pp 1083–1100. 107. Jensen, W. B. The Marsh Test for Arsenic. Museum Notes 2014May/June, 1–6. 108. Webster, S. H. The Development of the Marsh Test for Arsenic. J. Chem. Educ. 1947, 24, 487–490. 109. Arrowhead Forensics. https://www.arrowheadforensics.com/a-1219b-synthetic-blood-8oz. html (accessed Jan 10, 2019). 110. Lynn Peavey Company. QuickCheck Bloodstain Green. https://lynnpeavey.com/product/ quickcheck-bloodstain-green/ (accessed Jan 10, 2019). 111. Rose, N. L.; Palcic, M. M.; Evans, S. V. Glycosyltranferases A and B: Four Critical Amino Acids Determine Blood Type. J. Chem. Educ. 2005, 82, 1846–1852. 112. Carolina Biological. ABO-Rh Blood Typing with Synthetic Blood Kit. https://www.carolina.com/ blood-typing/carolina-abo-rh-blood-typing-with-synthetic-blood-kit/FAM_700101.pr (accessed May 20, 2019). 113. Kolata, G. New York Times: The Simpson Case: The Details; Testing the Blood 3 Ways. https://www.nytimes.com/1994/07/09/us/the-simpson-case-the-details-testing-theblood-3-ways.html (accessed Jan 10, 2019). 114. Gill, P.; Jeffreys, A. J.; Werrett, D. J. Forensic Application of DNA ‘Fingerprints.’. Nature 1985, 318, 577–579. 115. Jeffreys, A. J.; Wilson, V.; Thein, S. L. Hypervariable ‘Minisatellite’ Regions in Human DNA. Nature 1985, 314, 67–73. 32 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

116. Elkins, K. M.; Kadunc, R. E. An Undergraduate Laboratory Experiment for Upper-Level Forensic Science, Biochemistry, or Molecular Biology Courses: Human DNA Amplification Using STR Single Locus Primers by Real-Time PCR with SYBR Green Detection. J. Chem. Educ. 2012, 89, 784–790. 117. Jackson, D. D.; Abbey, C. S.; Nugent, D. DNA Profiling of the D1S80 Locus: A Forensic Analysis for the Undergraduate Biochemistry Laboratory. J. Chem. Educ. 2006, 83, 774–776. 118. Sultana, G. N. N.; Sultan, M. Z. Mitochondrial DNA and Methods for Forensic Identification. J. Forensic Sci. Criminal Inves. 2018, 9, 1–4. 119. McCord, B. R.; Gauthier, Q.; Cho, S.; Roig, M. N.; Gibson-Daw, G. C.; Young, B.; Taglia, F.; Zapico, S. C.; Mariot, R. F.; Lee, S. B.; Duncan, G. Forensic DNA Analysis. Anal. Chem. 2019, 91, 673–688. 120. ThoughtCo. Biography of Anastasia Romanov, Doomed Russian Duchess. https://www. thoughtco.com/anastasia-romanov-biography-4173902 (accessed Jan 11, 2019). 121. Saul, T. Death of a Dynasty: How the Romanovs Met Their End. https://www. nationalgeographic.com/archaeology-and-history/magazine/2018/07-08/romanovdynasty-assassination-russia-history/ (accessed Jan 12, 2019). 122. Gill, P.; Ivanov, P. L.; Kimpton, C.; Piercy, R.; Benson, N.; Tully, G.; Evett, I.; Hagelberg, E.; Sullivan, K. Identification of the Remains of the Romanov Family by DNA Analysis. Nat. Genet. 1994, 6, 130–135. 123. Stoneking, M.; Melton, T.; Nott, J.; Barritt, S.; Roby, R.; Holland, M.; Weedn, V.; Gill, P.; Kimpton, C.; Aliston-Greiner, R.; Sullivan, K. Establishing the Identity of Anna Anderson Manahan. Nature Genetics 1995, 9, 9–10. 124. Coble, M. D.; Loreille, O. M.; Wadhams, M. J.; Edson, S. M.; Maynard, K.; Meyer, C. E.; Niederstatter, H.; Berger, C.; Berger, B.; Falsetti, A. B.; Gill, P.; Parson, W.; Finelli, L. N. Mystery Solved: The Identification of the Two Missing Romanov Children Using DNA Analysis. PLoSOne 2009, 4, e4838. 125. Bidgood, J. New York Times: New DNA Test Sought in Identical Twins Rape Case. https://www. nytimes.com/2014/09/16/us/new-dna-test-sought-in-identical-twins-rape-case.html (accessed Jan 11, 2019). 126. Weber-Lehman, J.; Schilling, E.; Gradl, G.; Richter, D. C.; Wiehler, J.; Rolf, B. Finding a Needle in the Haystack: Differentiating “Identical” Twins in Paternity Testing and Forensics by Ultra-Deep Next Generation Sequencing. Forensic Sci. Int. Genet. 2014, 9, 42–46. 127. Boeri, D. Standard DNA Testing Can’t Differentiate Between Identical Twins. A New Test Challenges That. https://www.wbur.org/news/2017/03/07/twin-dna-crime-tech (accessed Jan 11, 2019). 128. Frye v. United States. 293 F. 1013 (D.C. Cir.), 1923. 129. Daubert v. Merrell Dow Pharmaceuticals Inc. 509 U.S. 579, 1993. 130. Wagstaff, I. R.; LaPorte, G. The Importance of Diversity and Inclusion in the Forensic Sciences. https://www.nij.gov/journals/279/Pages/importance-of-diversity-and-inclusion-inforensic-sciences.aspx (accessed Jan 9, 2019). 131. Kanu, A. B.; Kaplan, L. J. The Quest for Confirmatory Data in Crime Scene Investigations. Chem. Educator 2016, 21, 231–239.

33 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

132. Kanu, A. B.; Pajski, M.; Hartman, M.; Kimaru, I.; Marine, S.; Kaplan, L. J. Exploring Perspectives and Identifying Potential Challenges Encountered with Crime Scene Investigations when Developing Chemistry Curricula. J. Chem. Educ. 2015, 92, 1353–1358.

34 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Chapter 3

Downloaded via MIAMI UNIV on November 5, 2019 at 02:16:00 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Incorporating Forensic Science throughout the Undergraduate Analytical Curriculum: From Nonmajors through Instrumental Analysis Douglas J. Beussman* St. Olaf College, Department of Chemistry, 1520 St. Olaf Avenue, Northfield, Minnesota 55057, United States *E-mail: [email protected].

The incorporation of forensic science, a very popular field, into the chemistry curriculum can lead to increased student interest and engagement. Having seen forensic science applications in popular culture, even if not accurately portrayed, allows students to directly understand the potential application of what they are studying. Students also like to solve the mystery aspect of forensic scenarios. While entire lecture courses based on forensic science have been developed, the laboratory is the most natural fit for forensic science applications. This chapter presents lab experiments from a nonmajors introductory forensic science course as well as from labs designed for chemistry majors. Since many of the methods used in forensic science are analytical chemistry techniques, it is natural to incorporate forensic-themed experiments into the analytical chemistry curriculum. A highperformance liquid chromatography lab designed to analyze a simulated blood sample from a deceased subject for the presence of cardiac drugs is presented. Students need to both qualitatively and quantitatively analyze the blood sample in order to determine if the death was likely due to a drug overdose and, if so, whether it was likely accidental or intentional. This experiment is regularly taught in the junior-level majors Analytical Chemistry course. A second experiment taught in the upper-level majors Instrumental Analysis course is also presented. In this experiment, students analyze the atomic content of different types of matches to determine if they can differentiate the match types in a simulated arson analysis. All of the labs presented are routinely cited by the students as their favorite and the ones they most remember, indicating at least some advantage to incorporating forensic scenarios into the chemistry lab curriculum.

© 2019 American Chemical Society Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Introduction Due in part to the popularity of forensic science in recent pop culture (1), the incorporation of forensic science into the undergraduate curriculum is a natural choice. This incorporation can easily manifest as a complete stand-alone course, as individual lab experiments in a course focused on other areas, or even as just examples of how scientific concepts can be applied to answer important forensic questions. Student interest can be used to capture attention while teaching concepts that might otherwise be perceived as hard, boring, or not important. Although it might be seen as a gimmick, anything that can be done to increase student engagement and buy-in is ultimately a good teaching strategy. There are three distinct courses at St. Olaf College that utilize forensic science concepts or scenarios. A general education, nonmajors course in forensic science was developed several years ago in part to stimulate student interest in science courses. A course in forensic science was seen as a way to likely increase the number of students enrolling in a chemistry course. Obviously, since the entire focus of the course is forensic science, the lectures and labs are all focused on the application of scientific concepts and techniques to solve forensic questions. Students seem more willing to study and learn scientific concepts, from acid-base chemistry to separation science to spectroscopic analysis, when they are taught in the context of how these concepts are applied to solve forensic questions and not just as something that must be learned in order to pass the course. This course was designed as part of the general education liberal arts curriculum and was not meant to be taken by science majors, but there have been several science majors who have taken the course due to interest in the forensic applications. While not intended as a gateway into a career in science, there have been a few students who after taking the nonmajors forensic course decided to continue studying science, eventually graduating from St. Olaf with biology or chemistry degrees. Since the field of forensic science can be considered an applied subfield of analytical chemistry, it makes sense to incorporate labs designed around forensic scenarios into the junior-level (Analytical Chemistry) and senior-level (Instrumental Analysis) analytical chemistry courses. While most of the students that take these courses are not focused on future careers in forensic science, they appreciate the application of analytical methods to forensic questions and often comment that these labs were their favorite. The labs described here could just as easily be done with nonforensic scenarios with the same analytical content presented, but the incorporation of the forensic applications has led to increased student engagement in these labs as observed by instructors and reported by students on end-of-semester surveys. A quick literature search using the keyword “forensic” finds 80 articles in the Journal of Chemical Education alone, dating back to 1973 (2). A number of forensic-focused laboratory experiments have also been described. These include the analysis of pen inks (3), the analysis of suspected drug samples (4, 5), collection of fingerprints (6), and the analysis of mitochondrial DNA (7). Clearly, the incorporation of forensic science into the undergraduate curriculum is not new. In fact, a number of different textbooks have been published for the teaching of forensic science at an introductory level (8–10). As these textbooks can provide good frameworks for lecture courses, this chapter will focus on the lab components of the three St. Olaf courses and on how forensic applications have been built into the lab curriculum of these courses. Specific descriptions of the forensic-focused lab experiments developed and taught at St. Olaf College in these three courses are given below.

36 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Nonmajors or Introductory Chemistry Liberal arts institutions, such as St. Olaf College, require students take courses in a wide variety of subject areas. This often includes one or more science-focused courses. For students whose academic interests lay in other disciplines, this requirement can be stressful or at least something they do not look forward to completing. Courses that resonate with students, that relate to topics they are familiar with, or even that provide some level of entertainment while teaching basic scientific concepts and the scientific method can be more popular choices for these non-science majors. Forensic science courses are excellent examples of such courses. Virtually all students have read a crime novel or watched a crime movie or television show. They hear about criminal acts and prosecutions on the news. Even if they have little interest in science, they often are interested in the outcome of these real or fictitious crime scenarios. This stimulates their interest in forensic science, even if they do not realize it is a science-filled field. They want to know “who did it” and this becomes the perfect opportunity to introduce some level of scientific concepts into their education. General student interest in forensic science drove the development of an Introduction to Forensic Science course at St. Olaf College almost 15 years ago. This course was designed as a general education course intended primarily for non-science majors, although there have been a handful of science majors and even a few chemistry majors that have taken the course due to their interest in the topic. As this course was being developed, attendance at a Forensic Science workshop hosted by Larry Kaplan at Williams College as part of the National Science Foundation funded Center for Workshops in the Chemical Sciences provided invaluable ideas about course content, especially lab experiments for possible incorporation into our new course. This course remains a very popular choice and is always filled to capacity. The course consists of a 50-minute lecture three times a week and one 3-hour lab each week. We ultimately developed the lab program to have 11 different experiences. A number of these were based off of the labs developed by Larry Kaplan. A brief description of each lab is given below, with more detailed descriptions of some of the experiments developed at St. Olaf College following. Labs are not necessarily listed in the order in which they are performed over the semester. The Crime Scene Processing lab is always the lead-off lab and the DNA Fingerprinting lab is always the final lab, but the other labs are done on a rotation basis, with several labs being run simultaneously. This is done because of limited instrumentation for many of the experiments. We are unable to have all nine lab groups use the single gas chromatograph–mass spectrometer (GC-MS) or reflection spectrometer, for example. Students work in groups of three, with groups rotating each week through the various experiments. Each group must submit a lab report containing a description of the possible crime, the evidence analyzed, a detailed description of the analysis methods used, a data analysis section, and a results and conclusion section where they state what they found. These lab reports are incorporated into their overall course grade for the class. Lab 1: Crime Scene Processing Students are presented with a crime scene involving a kidnapping. Various items of potential forensic value are strewn about the room including a ransom note, broken glassware, a ripped lab coat, overturned chairs, and muddy footprints. The students need to process the crime scene, documenting the scene and collecting the possible evidence. A number of these items are used later in the semester in subsequent labs that investigate these types of evidence. For example, the ink used in the ransom note is analyzed and compared to several ink pens found in various suspects’ possession and the ripped portion of lab coat is compared to coats obtained from suspects. 37 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Lab 2: Alcohol Lab Students investigate several different methods used for the detection and quantitation of blood alcohol levels. There are three different parts to this experiment. In the first part, students use a saliva swab (ALCO Screen, www.chematics.com) to detect alcohol. In the second portion or the lab, students use a solid-state breath sensor, such as the one shown in Figure 1 (AlcoHAWK Pro), to determine the amount of alcohol in a person’s breath. For both of these experiments the consumption of alcohol is simulated by spraying a breath freshener into a student’s mouth prior to testing. Not only does this show students positive alcohol results for these methods of detection, it also allows for the discussion of false positives and the need to ensure nothing is placed into a suspect’s mouth for several minutes prior to administering these tests in the field. The final portion of this lab has students determine the concentration of alcohol in simulated blood samples using a chemical reaction between alcohol and dichromate followed by detection using absorbance spectroscopy at 440 nm. Students need to create calibration solutions by diluting a stock alcohol solution prior to reacting the solutions with dichromate. They can then construct a calibration curve and using the best fit line determine the concentration of alcohol in the unknown blood sample.

Figure 1. A solid-state breath alcohol detector. Lab 3: Ink Lab Students extract the ink from the previously collected ransom note as well as examples of ink from several pens by soaking the paper containing the ink in methanol. After concentrating the sample by evaporating some of the methanol, students analyze the ink samples using thin layer chromatography (TLC) and can compare the ransom note ink with the ink pens collected from suspects. Lab 4: DNA Fingerprinting Lab This experiment uses a commercial DNA fingerprinting kit available from BioRad (Forensic DNA Fingerprinting Kit). The kit contains six DNA samples, one from the crime scene and five from suspects, as well as a restriction endonuclease mixture. Students mix the DNA samples with the endonuclease mix and during the digestion incubation they cast agarose gels which they then use to separate the six different digested DNA samples. After staining and destaining the gels, they can compare the DNA from the crime scene with that from the five suspects, ultimately identifying the perpetrator. 38 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Lab 5: Drug Lab Like the alcohol lab, this experiment lets the students use three different methods to detect the presence of controlled substances. Students investigate powdered substances using the narcotics identification kit shown in Figure 2. This consists of different plastic pouches containing different vials of chemicals for each drug to be investigated. Students can place a small amount of powder into the pouch, seal it, and after cracking open the vial(s) of chemical reagents look for a color test indicating a positive result for the presence of the drug that the test pouch is designed to detect. We routinely have students investigate two different powders, one suspected to be a barbiturate and the other suspected to be cocaine, using two different pouches. They are given a powdered sample of phenobarbital and of diphenhydramine HCl. Diphenhydramine HCl gives a false positive in the cocaine test, which allows for a discussion of false positives. In the second portion of the lab, students use a dipcard (STATDIP, www.micro-distributing.com) to detect drugs from a simulated urine sample. These dipcards wick sample up into the card where immunoassays detect the presence of various drugs. Colored water can be spiked with analytical drug standards, available without a Drug Enforcement Administration license from Cerilliant or Sigma Aldrich. The final part of this lab has students analyzing the simulated urine sample using GC-MS (11). They extract the drug from the urine sample using liquid-liquid extraction and introduce the sample in the GC-MS. Having already tentatively identified the drug present using the presumptive dipcard test, they can look for the known molecular mass of the drug. They also compare the mass spectrum they collect to a database spectrum for the identified drug in order to reach a more conclusive answer.

Figure 2. Narcotics identification kit pouch used for the detection of cocaine. Lab 6: Fingerprinting Lab In this lab students collect their own fingerprints using an inkless fingerprinting pad. To avoid any possible privacy issues, they maintain control of their own fingerprint card. They can then examine their prints for the presence of aches, loops, and whorls. They also collect fingerprints deposited on glass beakers by dusting as well as superglue fuming. Lastly, they attempt to collect fingerprints from the ransom note paper using a ninhydrin reaction. Lab 7: Fiber Lab Students analyze the lab coat fibers found at the staged crime scene along with fibers collected from several suspects. They first use two different fabric stains to differentially stain the fibers. These stains are sold by Testfabrics, Inc. and result in differently colored stains for fibers of different 39 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

chemical composition. This allows students to get an idea of the chemical composition of the various fiber samples they have obtained. In order to provide a more concrete identification of the fiber composition, students follow up the staining analysis with an infrared (IR) absorption analysis using a Fourier Transform IR (FTIR) spectrometer (Thermo Scientific Nicolet iS5) equipped with an attenuated total reflectance (ATR) source, as shown in Figure 3. This lets students obtain the IR spectrum for the crime scene fibers as well as those from suspects that were tentatively identified as being chemically consistent based on the staining analysis. Direct comparison of the spectra allows students to determine if in fact the fiber samples are made from the same type of cloth or not.

Figure 3. A cloth sample being analyzed using an ATR-FTIR instrument.

Lab 8: Blood Lab In this experiment students use either simulated blood or bovine blood. In the first part of the experiment students use a kit (Ward’s Science ABO and Rh Blood Typing Kit) to determine the blood type of a simulated blood sample by mixing it with simulated anti-serum and looking for agglutination. They compare a crime scene blood sample to blood samples collected from several potential suspects, and by determining if the samples are type A, B, AB, or O and if they are Rh positive or negative, they can identify which suspect(s) could have deposited the blood at the crime scene. They then test dried blood spots, made from dried bovine blood, to determine if the spots could be blood by swabbing the spots and then adding a few drops of a phenolphthalein solution followed by a few drops of a hydrogen peroxide solution and looking for the pink color indicative of a presumptive identification of blood (12). Finally they investigate the shape of blood drops and determine the angle at which the drops hit the target by measuring the length and width of spots dropped at different angles onto the targets.

40 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Lab 9: Poison Lab Students analyze a simulated blood extract for the presence of elemental poisons arsenic, lead, or mercury. They utilize a flame atomic absorbance spectrometer (AAS) to perform this analysis (Buck Scientific model 210VGP), shown in Figure 4. They are given a simulated blood extract and a standard sample (1000 ppm) of one of the elemental poisons and told the working range for that element on the AAS. They are also told the toxic blood concentration of their element. As the standards are relatively high concentration solutions, students are reminded to wear goggles and gloves while handling the solutions. They must dilute the standard to make calibration solutions within the working range of the instrument and then introduce the standards and blood sample into the AAS, recording the observed absorption. They can create a calibration curve from the standards, determine the best fit curve and then calculate the concentration of the poison in the blood sample. Using this information, they can determine if the victim died from poisoning.

Figure 4. A flame AAS used to analyze for the presence of elemental poisons. Lab 10: Glass and Paint Lab For the glass portion of this lab, students investigate the density of various glass samples using the flotation technique where two liquids of different densities are mixed such that the glass floats in the middle of the liquid (13). After analyzing the glass samples, students analyze several different paint samples that all appear to be the same shade of blue to the human eye. These paint samples have been created by adding varying small amounts of white paint to a common blue paint base. By using a reflectance spectrometer, students can obtain spectra and quantitative color values for the different paint samples, allowing them to determine if they are indeed the same color or if they only look to be the same color. This paint analysis lab was developed at St. Olaf College (14). The student lab manual is presented below.

Paint Analysis While two different paint samples may appear to the human eye to be the same, subtle differences may exist. These differences can be detected using optical spectroscopy. It is generally not possible to measure the absorbance through a paint sample, but the light reflected off the surface of a paint sample can be analyzed. When the entire visible spectrum of light is directed onto a paint sample, all wavelengths are reflected except those which are absorbed by the paint molecules. Since 41 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

different paints consist of different chemical compositions they will have different absorbance or reflectance spectra. By taking advantage of this fact we can compare different paint samples and determine if they have a common origin. Procedure 1. Obtain paint samples, either dried paint evidence or in liquid form. 2. If you have received liquid paint samples, apply a spot (1 in. × 1 in. is sufficient) of each paint to a piece of paper and allow to dry in the hood. It is best to apply 3–4 different spots of the same paint to different areas on the page. This will help take into account differences in paint thickness. 3. On the computer, open the SI 400 Series Colorimetry software. Make sure the spectrometer is powered up and initialize the software. Set the wavelength range to be 380–780 nm with a crossover wavelength of 379 nm. 4. Obtain a piece of printer paper and fold in half. Holding the reflectance probe squarely against the white paper, collect a “Blank” spectrum. This will likely take a few minutes. Make sure that the probe surface remains flat against the white printer paper until the “Blank” has been collected. 5. Place the reflectance probe squarely onto a dry paint spot and collect a “Sample” spectrum. This may also take a few minutes. 6. Once the “Sample” has been collected, record the CIELAB variables (L*, a*, b*), Tristimulus Values, and Chromaticity Coordinates. Collect 3-4 readings on each of the applied paint spots of a single color sample in order to gain an understanding of the variability in individual readings for the same paint sample. CIELAB Color Space The CIELAB color space is one of several methods of characterizing the color of an object. Figure 5 shows how each CIELAB variable relates to color. The more positive the “a” variable, the more red the color, and the more negative the “a” variable, the more green. The more positive the “b” variable, the more yellow, while the more negative the “b” variable, the more blue the color. The “L” variable defines the lightness of the color analyzed. Combining all three variables allows the entire color space to be defined. Continue this analysis for all spots of all paint samples. Calculate the average values for each of the CIELAB variables in order to reach a conclusion regarding the similarity of the paint found at the crime scene compared to the evidence collected from the suspect. Look at how much the values for the multiple spots of the same paint sample vary and compare this to how much the values vary between different paint samples.

42 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Figure 5. The CIELAB Color Space. Reproduced with permission from reference (13). Copyright 2007 American Chemical Society. Lab 11: Overdose Lab This experiment allows students to use UV spectroscopy to determine the concentration of diphenhydramine HCl, an over-the-counter drug, in a deceased victim’s blood. Students are given a standard of the drug and a simulated blood extract. They must create dilutions of the standard and then obtain absorbance measurements for the standards and the blood sample. They create a calibration curve and use the results to determine the concentration of the diphenhydramine HCl in the victim’s blood. Based on the weight of the victim and the amount of blood in an average person of that size, the students can determine if there was enough diphenhydramine HCl in the victim’s blood to be fatal. This experiment was also developed at St. Olaf and the student manual is given below.

Overdose Lab Many drugs, including illicit drugs as well as prescription and over the counter medications, contain aromatic functional groups. These functional groups have a strong absorbance in the UV region around 254 nm. This wavelength can be used to determine the concentration of these species in a solution. It should be noted that quantitative analysis is only possible if the solution contains a single aromatic species, which would not likely be the case in a real blood sample without significant chemical separation steps. Diphenhydramine HCl, structure given in Figure 6, is a common ingredient in cold medications (e.g., Benedryl) and sleeping pills.

43 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Figure 6. Chemical structure of diphenhydramine HCl. It thus has a potential to be abused and, if taken in high enough quantities, can be fatal. In this experiment you will determine if a victim’s death can be ruled a suicide by ingestion of diphenhydramine HCl. Theory The basis for any colorimetric or spectrophotometric method of analysis is Beer’s Law, which relates the amount of incident light absorbed by a solution to the number of species (molecules, atoms, or ions) in the solution that is absorbing the light:

where A is the absorbance, ε is the extinction coefficient, b is the path length, and c is the concentration of the absorbing species. Notice that this equation expresses the relationship that the amount of light absorbed by a solution is proportional to the number of absorbing species. The number of absorbing species is dependent on the concentration (c) of those species and the distance the light travels (b) through the solution. To make the proportionality an equality, the proportionality constant ε is used. In practice, Beer’s Law forms the basis for two different methods of determining the concentration of a solute in a solution such as blood. One method relies on the fact that an extinction coefficient is known for the absorbing species at the specific wavelength. If this is true, then the concentration can be determined by simply dividing the absorbance at that wavelength by the extinction coefficient. After appropriate calculations to correct for dilution, the actual solute concentration is determined. The second method is used when an extinction coefficient is not known. This experiment actually involves using a series of standard solutions of known concentration as well as the samples assumed to contain the chemical in question and in a “reagent blank.” The absorbance of each solution is measured in the spectrophotometer. The absorbance of each of the standards is used to prepare a reference graph of analyte concentration versus absorbance. The unknown analyte concentration of the sample is then determined by extrapolating from the experimentally determined absorption to the concentration on the graph. Procedure You will be provided with a stock solution of diphenhydramine HCl, at a concentration of 1 g/ L, as well as a sample from a potential suicide victim. You will need to use the stock solution to make diluted calibration standards with concentrations in the range of 1 g/L to 10 mg/L using de-ionized water to make the dilutions. You should make a minimum of 5 diluted calibration standards.

44 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Before actually making the dilutions, discuss the procedure that you will use to make these dilutions, including the calculations, with your laboratory instructor. All dilutions will be made with de-ionized water. Briefly calculations can be made using the equation C1V1 = C2V2, where C1 is the concentration of the stock solution, V1 is the volume of stock solution needed, C2 is the desired concentration, and V2 is the final total volume. You will want to make 10 mL of each calibration standard. Obtain a quartz cuvette from your instructor. Be careful, as these are fragile and expensive! Fill the cuvette three-quarters of the way full with de-ionized water to be used as a “Blank.” On the spectrometer (Shimadzu model UV-1700), shown in Figure 7, if the wavelength is not already set to 254 nm, click on the GOTO WL button, type in 254, and click ENTER. You will hear the instrument motor adjust to the correct wavelength. Place the cuvette in the front sample holder, making sure that the clear surfaces of the cuvette are in the light path (right to left). Make sure not to touch the clear surfaces of the cuvette, as your fingerprints will cause errors in the data. Close the spectrometer lid.

Figure 7. UV/VIS spectrometer for collection of absorbance data, 7a closed sample chamber, 7b open chamber showing placement of cuvettes. Zero the instrument on the water “Blank.” This will provide a background reading that will be subtracted from subsequent readings, which will subtract out the absorbance due to the cuvette and water. Put the water “Blank” in the front sample holder and press the AUTO ZERO button. After a few seconds, the instrument will beep and the absorbance reading should be 0.000. If you do not see this, let your instructor or teaching assistant know. Once the instrument has been zeroed, do not re-zero unless the instrument is shut off and restarted. Remove the cuvette and replace the water with your first calibration standard. Remember that you are trying to determine the absorbance for a given concentration. What does this imply about how you change solutions in the cuvette? Place the cuvette back in the light path and close the lid. Once the absorbance value stabilizes (a second or two) record the absorbance reading. Record a minimum of three measurements, by removing and replacing the same cuvette, in order to be able to average out instrumental fluctuations. Remove the cuvette and replace the solution with your second calibration standard. Repeat data collection for all calibration standards, again taking three readings for each. After analyzing all calibration standards replace the solution in the cuvette with the unknown sample and again record the absorbance. If any solutions spill, clean them up or let your instructor or teaching assistant know.

45 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Table 1. Summary Table for Lab Experiments in Nonmajors Course Experiment

Main Skills

Specialized Equipment Needed

Common Student Difficulties/Mistakes

Crime Scene Processing

Attention to detail

None

Walking through evidence area, missing evidence, not properly labeling evidence

Alcohol Lab

Dilutions, calibration curve

UV/VIS Spectrometer

Incorrect dilution calculations, incorrect use of pipettes, incorrect creation of calibration curve or use of linear fit equation

Ink Lab

Thin layer None chromatography (TLC)

Complete evaporation of solvent, too little sample applied to TLC plate

DNA Fingerprinting Lab

Use of micropipettors, gel electrophoresis

Agarose gel electrophoresis rigs

Not letting gel separation go long enough or letting it go too long, not destaining gel long enough

Drug Lab

pH-driven liquid extraction, analysis of mass spectral results

GC-MS

Not putting enough powder in the narcotics identification kit, extracting the wrong solvent layer

Fingerprinting Lab

Attention to detail, appreciation for skill needed for fingerprint collection

None

Applying too much dusting powder, fuming with superglue to long

Fiber Lab

Spectral matching

FTIR Spectrometer

Not looking at spectra carefully to observe subtle differences

Blood Lab

Spatter analysis

None

Incorrect mathematical (trigonometry) analysis

Poison Lab

Dilutions, calibration curve

AAS

Incorrect dilution calculations, incorrect use of pipettes, incorrect creation of calibration curve or use of linear fit equation

Glass and Paint Lab Density, color analysis, limitations of human perception

Reflectance Spectrometer Incorrect use of reflectance probe

Overdose Lab

UV/VIS Spectrometer

Dilutions, calibration curve

Incorrect dilution calculations, incorrect use of pipettes, incorrect creation of calibration curve or use of linear fit equation

46 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Using Microsoft Excel, the average of each calibration point should be plotted and a calibration trendline should be computed. Your instructor can help with this task if needed. If any data points do not lie near the calibration trendline, the corresponding solution should be remade and retested. Once a satisfactory calibration curve and trendline has been obtained, use the calibration equation to calculate the concentration (if any) of diphenhydramine HCl in the victim. Compare this result with the lethal dose of diphenhydramine HCl of 20 mg/kg. Assume the victim weighed 200 pounds with a total blood volume of 4 L. A summary of the main skills, equipment needed and common student errors for the labs in this course is given in Table 1. Only specialized equipment, defined as costing over $1000, is listed. It should also be noted that only the most common student errors are listed. Since the course is designed for nonmajors, there are a wide range of errors that occur, including simple mathematical calculation errors.

Forensics in Analytical Chemistry While most chemistry majors do not take the Intro to Forensic Science course, they all are required to complete the Analytical Chemistry course and lab. As the vast majority of forensic work is analytical chemistry, this course is a natural place to incorporate forensic science concepts. While we have chosen to incorporate a variety of labs, which are each four hours, stemming from different areas of analytical science, including industrial, environmental, and product development, we have also developed and incorporated a forensic-based experiment that utilizes high-performance liquid chromatography (HPLC) for the analysis of a simulated blood extract from a deceased subject (15). In “The Mysterious Death Dilemma” lab, students are tasked with identifying the identity and quantity of various pharmaceutical agents in the blood sample and then, based on the results, making a determination of whether the death was due to a drug overdose and, if so, whether it was likely accidental or intentional. This is generally one of the experiments that students rate as their favorite in the Analytical lab. They enjoy solving a forensic question while at the same time learning about HPLC separations. The Analytical Chemistry labs at St. Olaf College are run using a unique role-playing format developed by John Walters (16–18). Students work in “Companies” of four students, with each group member assigned a specific role. The role of Manager is in charge of the entire Company. Manager makes decisions as to how to best accomplish the lab objectives and ensures that all other members know their tasks and complete them in a timely fashion. The Chemist role is tasked with making the solutions or dilutions needed as well as performing any reactions. Chemist is also in charge of the proper documentation and disposal of any waste generated. The student assigned as Software is in charge of collecting safety data sheets for any chemicals to be used. They also operate any instrument software required for the experiment and are in charge of the data analysis, including the preparation of any graphs that are needed. The final role, Hardware, must learn how to safely and correctly use the instrumentation pertinent to the experiment being performed. Successful completion of the overall experiment requires teamwork, good communication, and time management. Every Company member must perform their assigned tasks in order for the Company to accomplish the goals. These roles rotate weekly to allow all group members to experience each role multiple times over the course of the semester. In the case of the Mysterious Death lab, the Chemist must make dilutions of the drug standards, pass these to Hardware, who injects them into the HPLC, with data collection and analysis completed by Software. Manager then needs to use this data in order to decide which drug(s) are in the deceased subject’s blood, informing Chemist which drugs need to be further studied to allow for quantitation to be determined. While there is no formal written report 47 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

for labs in the Analytical Chemistry course, the Manager meets with the professor within a week after the lab to provide an oral report on how the Company performed the analysis, what they found, and what conclusions they have made. In this experiment, students are presented with a simulated blood extract obtained from a deceased subject. They are told that the subject was a 48-year-old male, 200 lbs, 6 ft 1 in., with a known heart condition. Four medications, disopyramide, lidocaine, procainamide, and quinidine, were found on his bedside nightstand. There were no signs of foul play, so either natural causes or a drug overdose are suspected as possible causes of death. They are given standard samples of each drug as well as a blood extract from the deceased. They are given the linear working range of the HPLC for each of the drugs and tasked with determining if any of the four drugs were present in the blood and, if so, if any of them were present at a lethal level that could indicate a drug overdose as the cause of death. For any drug present at a lethal level, the Company needs to make a determination of whether the overdose was likely accidental or intentional. In practice, when the simulated blood samples are made, one of the drugs is often present at levels several times the lethal dose, making the determination of intentional overdose fairly straightforward, assuming the analyses are done correctly.

Figure 8. Separation of procainamide, lidocaine, and disopyramide used in the Mysterious Death Dilemma lab. The Companies need to devise the best way of identifying and quantitating the drug(s) in their blood samples. This ultimately generally includes injecting each of the four drugs separately in order to obtain elution times for each and then comparing these to the elution time(s) for the blood sample. Once they have identified the drug(s) present in the blood sample, they can make standards of varying concentrations in order to construct a calibration curve based on peak areas. It is up to Manager to decide how many standards will be used to make the calibration curve. The peak area of the drug(s) in the blood sample can then be used to determine the concentration of drug(s) in the deceased subject’s system. Software can look up the normal and toxic levels for any drug(s) found in the blood sample. These are often given in mass of drug per body mass. The Companies can convert the mass per volume concentrations obtained from the HPLC into mass per body mass by using the

48 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

weight of the deceased subject and an average estimate of the volume of blood for a person of his size and weight. This then allows the Companies to make a determination regarding cause of death. An example chromatogram of procainamide, lidocaine, and disopyramide is given in Figure 8. Additional details can be found in the student lab manual below or from the published manuscript describing this experiment (15).

The Mysterious Death Dilemma Objectives The Company has been asked to analyze blood serum extract from a patient found dead in his bed. Four prescription drug containers were found on the nightstand. Manager has to determine whether the death is due to an overdose and, if so, whether the overdose was likely accidental or intentional. In order to make this determination, Manager needs Chemist to prepare solutions containing known quantities of the suspect drugs, Hardware to inject these solutions into the HPLC, and Software to collect the chromatographic data from the instrument. The Dilemma Your Company has been asked to analyze the blood of a man found dead in his bed last night. There was no sign of foul play, but four prescription medications were found on the nightstand next to his bed. These drugs are disopyramide, lidocaine, procainamide, and quinidine. It is Manager’s job to decide if the death was due to an overdose and, if so, to assess whether the overdose was accidental or intentional. In order to separate what is sure to be a complex mixture of chemicals, Hardware suggests using chromatography, specifically HPLC. Manager remembers reading about HPLC separations and thinks that this sounds like a good idea, so Chemist begins making the required mobile phase needed for the separation. Since Manager does not know much about the four drugs found on the nightstand, Software is sent to find out the drugs identity, their therapeutic uses, and the normal therapeutic and toxic dosage levels. Hardware begins to learn about the instrumentation required so that, once Chemist is ready with the mobile phase and samples, the analyses can be performed. If Manager decides that the blood extract does contain one or more of the drugs found on the nightstand, then Software will have to calculate the concentration of drug found in the clinical sample in order to assess whether the death was accidental or intentional. This will likely require the creation of a calibration curve for each drug found in the blood sample. Your stockroom has standard samples of each individual drug at a concentration of 2.5 µg/µL in water. According to Company records, the linear working range of the four drugs in your HPLC are given in Table 2. Table 2. HPLC Working Ranges for the Four Drugs Being Studieda Drug

Linear Range (µg/µl)

Disopyramide

0.01–0.5

Lidocaine

0.025–2.5

Procainamide

0.00125–0.04

Quinidine

0.00125–0.1

a Reproduced with permission from reference (14). Copyright 2007 American Chemical Society.

49 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

The Experiment Chemist will be given a blood extract sample from the deceased as well as a set of standards containing the four drugs found on the nightstand. Based on Hardware’s analysis, Manager will need to decide how best to analyze the mixture to determine if any of the drugs are present in the blood sample. Chemist will undoubtedly have to prepare standard solutions, solvents, dilutions, mixtures, or all of the above. Software can begin to set up any spreadsheets or calculations needed to solve the dilemma as well as research (online, library, etc.) the identity of the drugs. Upper Management will be available for help, but will not make decisions for the Company. The success of each Company will depend on how well each member is prepared, how well Manager manages the group, and how well everyone in the Company works together. Remember, even though everyone has an individual role, you are a group working on a common task. If one person in the group cannot perform his or her role, you will all be affected. Group dynamics are the key. Manager’s Responsibilities Manager will be in charge of the Company and will be directly responsible for all aspects of the experiment. The Manager will direct the other members of the Company and insure that the experiment runs smoothly. The Manager will handle any problems within the Company. It is the Manager’s responsibility that all members of the Company know and complete their tasks, as well as understand what all other group members are doing. Note that although each member will have a separate task, it is the Manager’s responsibility that all members understand the task of all other members. Manager will report directly to Upper Management during the Management Interview held following completion of the experiment. Manager will have to decide how to best determine if the clinical sample contains any of the drugs found on the nightstand. Manager is in charge of documenting the procedures used in a running log of activities for the afternoon. Manager will have to decide, based on data collected by Hardware and Software, if the deceased died of natural causes or due to an overdose of one or more of the drugs (either accidentally or intentionally). If any of the drugs are found in the sample, Manager can ask Chemist to prepare a standard mixture containing the suspected drugs or several standard samples of known concentration to use in the creation of a calibration curve. Working ranges for the four drugs are provided above. Software can then inform Manager of the concentration of any drugs found in the clinical sample. If the clinical sample contains one or more of the drugs found on the nightstand, Manager must decide if the death was accidental or a suicide. Hardware’s Challenge The Hardware operator will be in charge of learning how to operate the HPLCs correctly and safely. Upper Management will give Hardware a brief tutorial on the instrumentation at the beginning of the lab, but Hardware may also need to read the instrument manuals or information contained in the course textbook that applies to HPLC instrumentation. The Hardware operator will be responsible for performing the analyses. While Manager has the final responsibility of deciding about the cause of death, Hardware has a critical role in making it possible to make a decision. Hardware is responsible for operating the HPLC instrument. This involves injecting microliter quantities of Chemist’s solutions into the instrument. The port into which the syringe is inserted is in the front of the injector valve. This requires some real dexterity to keep an air bubble from getting into the syringe when inserting it.

50 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Other responsibilities that Hardware has are to properly set the pump and detector controls on the signals feeding Software’s computer. The detector wavelength for this experiment is 254 nm, and since we are using the integrator output, there is no need to set the output range. To zero the detector, simply depress the center button labeled “autozero;” the display reads 0.00XX. Alternatively, Software can trigger an autozero signal from the software to do this function; it is a management call. To start the HPLC flow, set the flow rate to 0.200. Now press the RUN/STOP button and the piston will begin moving. Slowly increase the flow rate in 0.2 mL/min increments, making sure to let each new flow rate equilibrate for 1–2 minutes before further increasing until the desired value of 1.000 mL/min is obtained. Once the final flow rate is reached, the instrument should be allowed to equilibrate for about 10 minutes before use. To stop the flow, simply reverse the above process. Slowly decrease the flow rate below 0.20 mL/min, then press the RUN/STOP button to stop the piston. Remember you are performing Reverse-Phase Liquid Chromatography. The stationary phase is C18, and the mobile phase is a buffered mixture of methanol and water. CAUTION: Never leave a buffered or salt containing mobile phase in an instrument. Make sure you flush the system with unbuffered mobile phase after you are finished collecting data! Hardware will need to load 10–20 µL of each sample into the injector sample loop. Once Software is ready, Hardware can inject the sample into the column and then wash the syringe with 4–5 syringe volumes of fresh mobile phase. Software’s Labor The Software expert will be in charge of gathering data on the various drugs that were found on the nightstand including the structures, the therapeutic uses, the normal dosage, toxic levels, and so forth. Software will also be responsible for collecting the data obtained by Hardware using LabVIEW. Software will also have to perform any needed computer data manipulation (such as preparing graphs, or using a spreadsheet to make calculations). If Manager decides that the clinical sample does contain drugs found on the nightstand, Software will have to provide Manager with concentration values for the drugs. Software will have a brief tutorial from Upper Management on how to run the software to collect and process data from the HPLC. Interpretation of the data is a task that Software and Manager will probably want to work on together, especially interpretation of summary tables, overlaying chromatograms, determining retention times and peak areas, and exporting data for additional analysis with Excel. Chemist’s Challenge The Chemist will be responsible for making all solutions needed, including the HPLC mobile phase (63% water, 30% methanol, 6.5% acetic acid, and 0.5% triethylamine) and the standard samples that the Company will need. Chemist will also have to complete Safety Data Sheets for the drugs being analyzed as well as the mobile phase components. The standard mobile phase used by the Company for these types of analyses contains water, methanol, acetic acid, and triethylamine. Chemist will use any equipment needed to prepare all samples, solvents, dilutions, standard curves, and so forth that Manager decides are needed. The Chemist should understand how to use any equipment needed and how to perform any calculations required to make the needed solutions. It is also the Chemist’s responsibility to handle all chemicals safely and dispose of any waste properly.

51 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Senior-Level Analytical Chemistry: Instrumental Analysis Forensics is also used to supply context to a lab taught in the upper-level Instrumental Analysis course. This course is an elective course taken by chemistry students after they have completed the Analytical Chemistry course and lab. The four-hour lab sections are capped at four students with about half of the labs done individually and the other half done in groups of four. The role-playing lab format is relaxed in this course in favor of more open-ended experiential labs. As the pass/ fail lab sections have a maximum of four students, there is no formal lab report required. Rather, students are evaluated during the lab period on their effort, engagement, and results. Students were previously exposed to atomic absorption spectroscopy in the Analytical Chemistry course, allowing the Instrumental Analysis course to build upon their previous experience by introducing an inductively-coupled plasma optical emission spectroscopy (ICP-OES) experiment. This is the experiment with the forensic connection in this course. Kasamatsu et al. described the use of ICP-OES for the forensic analysis of match heads based on their elemental composition (19). This analysis could be useful in the investigation of arson cases. They found six elements, present at trace levels, that could be used to distinguish matches from different sources. We have used their research to develop an analytical teaching lab using the same six elements (20). Briefly, this experiment has students analyze two different types of wooden matches, strike-on-box and strike-anywhere for aluminum, barium, calcium, iron, and zinc. Students compare unburned samples of each type against each other as well as comparing burned samples of each match type. Statistical analysis is performed on the results in order to determine if any of the elements can be used to distinguish the two different types of matches from one another. Aside from the elements to be analyzed, the experimental parameters are kept as open as possible to allow students to make decisions about how to best answer the question of whether the match types can be differentiated. Students are allowed to decide how many replicate samples to create in order to get good statistical data, how many calibration points they want for each element, and whether they will create a standard calibration curve or use the method of standard additions. Students are also tasked with deciding the dilution factor(s) used for the matches. They are given a rough idea of the concentration of each element in the average matchhead so that they have some point of reference. They are also given the general working range for the ICP-OES instrument. One of the critical decisions the student groups must make is how to collect the samples after they have been digested using nitric acid. A significant amount of solid matchhead material, primarily glass powder and fillers, remains after elemental extraction with nitric acid. This solid material must be removed prior to analysis by ICP-OES, otherwise it may plug the sample introduction capillary. Students are given the task of deciding how best to remove the solid material while maintaining a quantitative transfer of the liquid portion of the sample. A number of methods have been utilized by different groups over the past several years. These include passing the samples through filter paper, packing a transfer pipette with glass wool and allowing the sample to pass through, centrifuging the samples and carefully aspirating the liquid portion, and using syringe filters. Students often start with one method only to decide that it is either too slow or too difficult to retain all of the liquid sample before they settle on a different method of extraction. Once the samples have been extracted and diluted and the standards have been prepared, students obtain spectroscopic data using the ICP-OES instrument. These results are then used to perform statistical tests including outlier testing, F-tests, and the appropriate T-tests to reach the final decision as to whether any of the elements can be used to differentiate the match types. For institutions that do not have access to an ICP-OES instrument, this experiment could be performed using a graphite furnace AAS. This might require a smaller number of elements to be investigated, 52 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

since the ability to quickly analyze a sample for multiple elements based on emission will be lost using a graphite furnace. Additionally, if a graphite furnace is not available, the experiment could be completed using a flame AAS, although the working range would likely have to be significantly increased. This might require larger masses of matchhead material to be used in the extraction step in order to provide higher initial concentrations of the elements being studied.

ICP-OES for the Analysis of Matches In 2017, almost 41,000 cases of arson were reported to the FBI (21). This number is likely lower than the actual number, as not all law enforcement agencies reported data to the FBI. While these can range from small backyard brush fires to fires that destroy entire buildings and even result in deaths, all arson cases are serious. There are a number of methods by which fires are started, but a common method involves the use of matches for initial ignition. Indeed, the presence of used matches at a suspected arson scene can help determine both the initial ignition point as well as the cause. For this reason, the analysis of matches is of forensic interest. Can different types of matches be distinguished from one another chemically? Your task is to investigate whether or not two different types of wooden matches can be differentiated based on elemental analysis using the Inductively-Coupled Plasma Optical Emission Spectrometer (ICP-OES). You will look at both burned and unburned “strike-on-box” and “strikeanywhere” wooden matches. In order to make a statistically sound determination, you will need to analyze multiple samples of each type of match (burned strike-anywhere, burned strike-on-box, unburned strike-anywhere, unburned strike-on-box). The number of samples of each is up to you, but remember you will ultimately be doing T-testing to compare results, so you might want to review how this is done before deciding how many samples you will collect.

Figure 9. Items needed for matchhead sample preparation. Match samples will need to be prepared several days prior to lab. Since all we are interested in is the possible chemical differences between the match heads (i.e., not the wooden part of the match), you will need to carefully scrape off the match head(s) for each sample using a clean razor blade, as shown in Figure 9. In order to get enough material for the analysis, you will need 10–20 mg of match head material per sample. It is not crucial that each sample have the exact same mass of material, but it is critical that you accurately know the mass, so that you can later determine the concentration (in ppm) of each element in each sample. For the burned match samples, you will want to let each match burn for the same amount of time before extinguishing and collecting the burned residue. In order 53 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

to extract the elements from the match head material (containing a lot of carbon), you will need to digest the samples in concentrated trace nitric acid for several nights. This can be done by placing each match sample in a separate 1.5-mL centrifuge tube, adding about 1 mL of nitric acid, sealing the tube, and letting it sit. Trace nitric acid has trace amounts of most elemental contaminants, and thus must be used here rather than regular nitric acid. The same is true for water which, can contain significant levels of various elements. You will want to use 18 MΩ water for all parts of this analysis. You should also consider the cleanliness of your glassware. It is standard procedure to wash glassware used for low level elemental analysis with acid prior to using it. Glassware you get from the stockroom will not have been acid washed. Once the samples have been digested, you will need to separate the liquid portion of the sample (which should contain the digested elements of interest) from the solid match residue. It is up to you to decide how to do this. You could, for example, filter through filter paper, filter through glass wool, use a syringe filter, or centrifuge the tubes and collect the liquid portion. However you decide to do this, you must ensure that you get a quantitative transfer. Each sample will ultimately be placed into a separate disposable borosilicate test tube to be loaded into the instrument’s autosampler. You will need a total of 10–25 mL of each solution for the instrument. Again, it is not critical that the volumes be the same, but it is important that you know the volumes. You can use 5% nitric acid (made from trace nitric acid and 18 MΩ water) for all dilutions. In order to determine the amount of each element in your samples, you will need to use some sort of method to compare what you find from your samples to some standard solution. It is up to you to decide if you should use the standard addition method or if a straight calibration curve will work. You will also need to decide how many solutions you will use for each calibration function. The instrumental working range for each element is between 0.1 and 10 ppm, so you will need to make sure that each solution (including both standards and samples) is ultimately in this range. The elements of interest for this lab are Al, Ba, Ca, Fe, and Zn. You will have access to standard reference solutions of each separate element at a concentration of 1000 ppm. These will clearly need to be diluted, again with the 5% nitric acid solution. The approximate range of each element in the match heads (listed per gram of match head) is given in Table 3. Table 3. Rough Concentrations of Each Element in Match Heads Element

Approximate Concentration per Gram of Match Head

Al

200 µg/g

Ba

1000 µg/g

Ca

5000 µg/g

Fe

200 µg/g

Zn

2000 µg/g

You will need to take these values into consideration when you are diluting your samples to make sure the final solutions are within the working range of the instrument for each element. Also, as opposed to the AAS that require a different hollow cathode lamp for each element analyzed and thus usually different solutions for each element, the ICP monitors emission and thus does not use a light source for absorption. This means that multiple elements can be analyzed for in a single solution. Instead of manually switching out the light source and adjusting the monochromator to select the appropriate wavelength, all that is needed here is for the monochromator to be adjusted.

54 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

This instrument can do this in a matter of a second or so. This means that multiple elements can be combined and analyzed in a single solution if desired. Once all of your samples/standards are prepared and in test tubes, they can be loaded into the ICP autosampler. The Standard Operating Procedure document is attached at the end of this experiment and will also be available at the instrument. Once all the data has been collected, you will need to do the statistical analysis (Grubb’s/Q-test, F-test, T-test, etc.) to determine if you can differentiate strike-on-box and strike-anywhere burned matches as well as unburned strike-anywhere from unburned strike-on-box matches.

Conclusions This chapter has presented several examples of the incorporation of forensic science questions or topics into the nonmajor and analytical chemistry lab curriculum. This is clearly not an exhaustive list. In fact, virtually any analysis technique or instrumental method can be applied to a forensic scenario. While there are numerous important nonforensic applications that should also be incorporated into teaching labs, the inclusion of experiments built around forensic scenarios can increase student interest. Students can readily appreciate the application of the technique they are studying and relate it to a real-world scenario. Students also enjoy having to figure out an answer to a mystery, leading to an increased focus. While the nonmajors course was designed for students completing general education requirements and not as a gateway to becoming science majors, there have been a few students who decided to become chemistry or biology majors after taking the Intro to Forensic Science course. In addition, the popularity of this course has led to the department offering it more often and eliminating a previous nonmajors chemistry course that had significantly lower enrollment. The inclusion of forensic-based labs in the majors level courses has also been well received by students. End of semester student reviews generally show that regardless of the class, the most popular experiments and the ones students remember the most are those that are built around a forensic application. Since these labs have been included in the curriculum, several students have pursued forensic science careers, either directly or by selecting graduate school programs with forensic science applications. This interest in forensic science careers is also likely due in part to the popularity of forensics in pop culture and the fact that most of these students have participated in forensic science based independent research projects in the author’s lab. Even with these other influences, the incorporation of forensic-based labs into the curriculum has helped maintain and focus student interest in this area.

Acknowledgments The students who have taken the Intro to Forensic Science, Analytical Chemistry, and Instrumental Analysis labs and the TAs who have helped teach them have provided valuable feedback about the forensic labs that I developed for each course. Their insights and testing of various iterations of the experiments have helped me refine these experiments over the past 15 years. Erin Hoffman did much of the development of the paint lab described from the Intro to Forensic Science course. Nick Parker and Josh Halverson did much of the initial testing and development of the ICP matchhead experiment used in the Instrumental Analysis course. My St. Olaf colleagues, Paul Jackson and Mary Walczak have continued to refine and provide suggestions for the Mysterious Death HPLC lab used in the Analytical Chemistry lab.

55 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12.

13. 14. 15. 16. 17. 18. 19.

Steenberg, L. Forensic Science in Contemporary American Popular Culture; Routledge Publishing: Abingdon, United Kingdom, 2012. Berry, K. O.; Nigh, W. G. Forensic Science—Course for January Interim. J. Chem. Educ. 1973, 50, 208–209. Breedlove, C. H. The Analysis of Ball-Point Inks for Forensic Purposes. J. Chem. Educ. 1989, 66, 170–171. Henck, C.; Nally, L. GC-MS Analysis Gamma-Hydroxybutyric Acid Analogs: A Forensic Chemistry Experiment. J. Chem. Educ. 2007, 84, 1813–1815. Schurter, E. J.; Zook-Gerdau, L. A.; Szalay, P. Analysis of a Suspected Drug Sample. J. Chem. Educ. 2011, 88, 1416–1418. Friesen, J. B. Activities Designed for Fingerprint Dusting and the Chemical Revelation of Latent Fingerprints. J. Chem. Educ. 2015, 92, 505–508. Millard, J. T.; Pilon, A. M. Identification of Forensic Samples via Mitochondrial DNA in the Undergraduate Biochemistry Laboratory. J. Chem. Educ. 2003, 80, 444–446. Houck, M. M.; Siegel, J. A. Fundamentals of Forensic Science, 3rd ed.; Academic Press: Cambridge, MA, 2015. James, S. H.; Nordby, J. J.; Bell, S. Forensic Science: An Introduction to Scientific and Investigative Techniques, 4th ed.; CRC Press: Boca Raton, FL, 2014. Saferstein, R. Criminalistics: An Introduction to Forensic Science, 12th ed.; Pearson: London, 2018. Lillsunde, P.; Korte, T. Comprehensive Drug Screening in Urine Using Solid-Phase Extraction and Combined TLC and GC/MS Identification. J. Anal. Toxicol. 1991, 15, 71–81. Webb, J. L.; Creamer, J. I.; Quickenden, T. I. A Comparison of the Presumptive Luminol Test for Blood with Four Non-Chemiluminescent Forensic Techniques. Luminescence 2006, 21, 214–220. Standard Test Method for Density of Glass by the Sink-Float Comparator; ASTM C729-11(2016); ASTM International: West Conshohocken, PA, 2016. www.astm.org (accessed May 20, 2019). Hoffman, E. M.; Beussman, D. J. Paint Analysis Using Visible Reflectance Spectroscopy: An Undergraduate Forensic Lab. J. Chem. Educ. 2007, 84, 1806–1808. Beussman, D. J. The Mysterious Death: An HPLC Lab Experiment. J. Chem. Educ. 2007, 84, 1809–1812. Walters, J. P. Role-Playing Analytical-Chemistry Laboratories. 1. Structural and Pedagogical Ideas. Anal. Chem. 1991, 63, 977A–985A. Walters, J. P. Role-Playing Analytical-Chemistry Laboratories. 2. Physical Resources. Anal. Chem. 1991, 63, 1077A–1087A. Walters, J. P. Role-Playing Analytical-Chemistry Laboratories. 3. Experimental Objectives and Design. Anal. Chem. 1991, 63, 1179A–1191A. Kasamatsu, M.; Suzuki, Y.; Sugita, R.; Suzuki, S. Forensic Discrimination of Match Heads by Elemental Analysis with Inductively Coupled Plasma-Atomic Emission Spectrometry. J. Forensic Sci. 2005, 50, 883–886.

56 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

20. Parker, N.; Halverson, J. P.; Beussman, D. J. Atomic Analysis of Match Head Compositions Using ICP Spectroscopy: A Forensic Science Experiment. J. Chem. Educ. Submitted for publication. 21. FBI. 2017 Crime in the United States—Arson page. https://ucr.fbi.gov/crime-in-the-u.s/2017/ crime-in-the-u.s.-2017/topic-pages/arson (accessed Jan 26, 2019).

57 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Chapter 4

Using Forensic Science To Engage Nontraditional Learners

Downloaded via MIAMI UNIV on November 5, 2019 at 02:19:33 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Megan L. Pajski* University of Mount Olive, 634 Henderson Street, Mount Olive, North Carolina 28365, United States *E-mail: [email protected].

Nontraditional students present a set of course design challenges different from those presented by traditional students. Although they are often more engaged in their own educational success, their level of preparation can vary much more, especially for courses outside their major. In this chapter, a five-week accelerated forensic science laboratory is described for nontraditional, non-science-major students. Designed to meet a portion of the critical thinking requirements for a general education in the liberal arts, this laboratory is a standalone course with no accompanying lecture. Subject matter is presented via mini lectures and supported with video clips and case studies. Critical thinking skills are assessed according to how well students meet certain problem-solving objectives, including the ability to point to certain pieces of data in support of experimental conclusions and to address the weaknesses of various forensic techniques. Experiments and materials are chosen to minimize chemical hazards, and this course may therefore be held at locations lacking a laboratory space. This course acts as a fun nonmajor science option that could be implemented at any small four-year or two-year undergraduate institution.

Introduction The forensic science laboratory described in this chapter was designed to meet a portion of the critical thinking general education requirements (GERs) for nontraditional students at the author’s institution. This institution is a small, four-year liberal arts university, serving mostly local, firstgeneration students. The course itself is a single, standalone lab that meets one night a week, 4 h a night, for five weeks. It is designed to be accessible for working adults with a wide range of science backgrounds. Nontraditional students are students aged 28 years or older who do not follow an unbroken path from high school to college (1). These students either return to or enter college for the first time for a variety of reasons. Among these are financial investment, career development, and life transitions (2). To these students, a college degree is a necessary component for increasing their career worth, gaining a promotion, or opening doors to new opportunities. Studies comparing nontraditional and © 2019 American Chemical Society Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

traditional students have found a number of differences that present unique challenges with respect to course design. For instance, nontraditional students are more likely to be academically motivated by grades, yet they tend to have a greater number of stressors at home (3). Thus, courses designed specifically for nontraditional students need to be especially clear about the connections between work and grades and be flexible with regard to sudden emergencies. As with traditional students, nontraditional students’ success may be partially predicted by interest in a subject (1), and so there is additional challenge associated with designing engaging courses in subjects not directly related to a student’s major. Many degrees, particularly those at liberal arts institutions, have some selection of required general education courses. Ideally, students appreciate learning for its own sake, but the reality is that motivation correlates strongly with interest in a subject, and GERs are unfortunately treated as hoops for a student to jump through rather than having value on their own. Science courses for nonscience majors can be a particular challenge to design and implement because student success also depends on their history of preparation. Even introductory or nonmajor science courses tend to be more technical than other GERs, and they require a kind of thinking that many nonscience students are not familiar with. Thus, a major component of successful GER science course design is finding a way to engage and motivate students to succeed while providing the tools to do so despite disparate backgrounds and preparation. As a vehicle for teaching applied skills, labs are often linked to a lecture course. In this format, instructors present basic background information in the lecture and later have students apply that information during lab. In the case of the course described here, the laboratory needed to be able to stand independently of a lecture component. At the author’s institution, all students are required to take seven credit hours of science to complete a critical thinking GER. Students often fulfill this requirement by taking two science courses (three credit hours each), one with an accompanying lab (one credit hour). However, it is not unusual for a nontraditional transfer student not to have lab transfer credit. These students therefore need a standalone lab to complete their critical thinking GERs, saving their tuition money for major classes and avoiding “disappearing credits” (4, 5). Some of these students have taken science courses in recent semesters, while others have not, and thus the extent to which students are prepared to produce scientific thought can vary widely. Furthermore, as the institution policy does not allow prerequisite courses for GER classes, some students have taken biology, some chemistry, some astronomy, and some a general science survey course. Additionally, as a science course for nonmajors, there are no math prerequisites for this course, and students come in with a varied level of preparation for logical problem solving. Thus, this lab course had to meet multiple challenges to give nontraditional students an interesting, meaningful experience in science. Currently, it is taught at an undergraduate, four-year liberal arts institution, but the author feels that similar designs would be equally applicable at community colleges, two-year institutions, high schools, or science summer camps intended for high school students.

Engaging Interest with Forensic Science The idea of designing a lab around forensic science was selected for a number of reasons. First, the popularity of crime dramas and true-crime television contributes a certain intrinsic cultural interest in forensic science, which acts as a hook to catch student attention. Second, students who watch these forms of entertainment have some background knowledge of the basic techniques and concepts of forensic science, dramatized though they may be, without having attended a forensic science lecture. Third, many basic forensic science activities, such as fingerprint lifting, taking 60 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

impressions, or comparing the effects of simple chemical reactions, can be done easily with minimal expense, without using hazardous substances, and in locations lacking a formal laboratory space. The overall concept behind this course is hardly new. Using forensic science to pique the interest of nonmajors has been described previously (6–12). Like this design, the courses previously described relied on students’ inherent cultural interest in forensic science to engage them. However, these courses were able to rely on accompanying lecture sections to introduce students to basic forensic science concepts, while, as explained previously, this course needed to be a standalone lab. Additionally, other forensic science courses were offered in the traditional semester-long format, while this course needed to be completed in a much shorter time frame. To design this course, certain key aspects were borrowed from previously published iterations (6–13) and combined with lessons from prior experience (14) with standalone science labs to create a unique accelerated course that has been successful at the author’s institution. There were several advantages to engaging nontraditional students using forensic science. To the frustration of many an experienced forensic scientist, the capabilities of forensic techniques are greatly exaggerated in the average crime drama. By selecting key scenes that relate to the lab activities (these clips are often available online), students can be drawn into discussions about the limitations of certain techniques and how sound any given conclusion may be. In addition, students come to appreciate the scientific process, perhaps more so than is typically observed in the average nonmajors lab. There is an element of magic to the utilization of forensic science in the media, and as students collect, process, and maintain the chain of custody for even small crime scenes, they develop an appreciation for the sheer amount of work and good record keeping necessary for the kinds of 30 min breakthroughs shown in the form of drama. Finally, a number of forensic techniques have connections to the daytime lives of students, acting as good starting points for class interactions. An average class has a number of nurses or individuals who work in the criminal justice system, many of whom have real-world experience with related techniques, such as having fingerprints collected to work in certain fields or performing routine medical bloodwork. By encouraging these students to speak up and share their experiences, instructors are able to get them and their fellow students to engage more with the material. This also encourages students to make hands-on connections with the subject matter.

Engaging Thinking with the Scientific Method To ensure consistent, skill-based improvement among students with a wide range of science backgrounds, this lab focuses on use of the scientific method. Most science courses are assessed according to how well students gain new scientific knowledge (15). However, students only receive one hour of credit for taking this lab, despite spending the same number of hours in class as an accelerated, five-week lecture course for which they earn three hours of credit. To compensate, students expect much less homework in the lower-credit lab course. The combination of limited formal lectures and minimal homework inhibits the use of subject tests as a means of assessment. By focusing on the scientific method, this course takes advantage of the fact that, no matter what their scientific backgrounds, all students have had some exposure to the basic principles of the scientific method. What these students lack is experience applying tit. From a skill standpoint, memorizing the steps of the scientific method is simple, but applying the concepts of having a testable hypothesis and so forth are much more difficult for a nonscientist to master. In essence, this course is designed to strengthen practical understanding and application of the scientific method. Furthermore, as a form of critical thinking, the scientific method process has applications beyond science (16, 17), making it an ideal general education skill. 61 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

To underscore use of the scientific method, most aspects of the course are designed around explicit use of the method’s steps. Students review and produce a reasonably generic list of the basic steps, such as the one shown in Figure 1. This list of steps is then used for the remainder of the course. Real-life examples of forensic crimes are discussed using the scientific method framework, student lab notebooks are filled out using each step as section headings, and students complete one written report analyzing the use of the scientific method in an example selected by the instructor. By repeatedly emphasizing the application of these steps, students are encouraged to think about them in the context of their lab activities.

Figure 1. An example of the list of scientific method steps that students might create and follow over the course of the lab.

Course Overview and Content Nontraditional labs at the author’s institution meet one night a week, 4 h a night, for five weeks. A portion of each lab period (approximately the first hour) is set aside for general explanations of the activities, the scientific principles behind them, and their use in the forensic science field. To capitalize on student interest in forensic television shows, each lecture period often includes short, relevant clips from popular forensic crime dramas and an instructor-led discussion about a case study showing the activities being performed in the field. Following this initial introduction, students spend the remainder of the lab period working in groups to complete two or three activities, recording their progress in lab notebooks, and answering a series of questions before they leave for the day. Lab notebooks with carbon-copy pages are utilized so that students turn in their work each week to receive feedback by the next class period. Table 1 provides a list of topics, case studies, and lab activities for context as each component is discussed later. 62 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Table 1. Example Weekly Schedule for the Forensic Science Lab

Mini Lectures and Video Clips Each lab period incorporates a mini lecture to review relevant concepts of basic science (e.g., lab safety and chemical hygiene, chemical and physical properties, the laws of motion, etc.) and emphasize their relationship to the activities of the day. For instance, during the second week of lab, when students analyze fingerprints and take toolmark impressions, the mini lecture incorporates ideas of chemical and physical properties, how different substances have different profiles of properties, and how forensic technicians can take advantage of these to differentiate between substances. These lectures are accompanied by PowerPoint slides, occasional animations to better explain certain concepts, video clips from forensic dramas to act as discussion starting points, and discussions of case studies as examples of the lecture material in action. Although the mini lectures are necessary for ensuring that students have some baseline understanding of the science behind lab activities, the use of video clips, case studies, and in-class discussions have been integral to encouraging student engagement. Students enjoy using the overdramatized fictional forensic crime plots as a basis for discussing the importance of accuracy and for understanding the limitations of forensic techniques. As an example, the average video clip showing the dramatic use of high-performance liquid chromatography or gas chromatography–mass spectrometry (18) often results in the forensic team determining the exact identity of a compound, down to the place and date of origin. In context with a mini lecture introducing how chemical bonds interact with excitation energies to produce a signal, the class can discuss why information about 63 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

exact locations or dates of manufacture may be unlikely. In another example, the author likes to use an Adam Ruins Everything video (19) detailing the pitfalls of fingerprint analysis to initiate a discussion on precisely the limitations of fingerprinting libraries and the dangers of placing too much importance on them. These discussions are enhanced by encouraging students to share relevant experiences. Students who have worked in medical labs may be encouraged to share their experiences using instruments to determine certain aspects of bloodwork, while someone in the criminal justice field might discuss experiences with breathalyzer instruments. By placing these activities in contrast with other, “realworld” applications, students more fully engage with the material. Using week 1 as a detailed example, a typical lab might run as follows. After an introduction to the lab course, outlining what students can expect, what is expected of them, a breakdown of grades, and where to find course materials, the instructor segues into a general introduction to criminalistics and forensic science. Students are introduced to the idea of considering themselves forensic technicians, who follow basic lab safety rules and handle evidence in a way that prevents contamination and preserves its admissibility to a court of law. The lecture portion ends with a discussion of the importance of proper record keeping. Students might be asked to give examples of times when they personally would have benefitted from better record keeping, for instance when keeping track of documents necessary for filing taxes. A case study in which improper evidence handling and record keeping interfered with a criminal investigation is then presented, and students discuss how adequate chain of custody and other factors might have improved the legal outcome. With these discussions fresh in their minds, they then spend the remainder of the lab period processing their “crime scene.”

Case Studies Case studies are historical forensic examples incorporated into the introductory lecture of each lab. Each case study highlights the use of at least one of the same tests that students will be using. For the sake of time, only one case study is discussed per class. Often, these are famous cases, such as the Enrique Camarena disappearance or the Wayne Williams serial abductions and murders. Richard Saferstein’s textbook, Criminalistics: An Introduction to Forensic Science (20) has been an invaluable resource for case study ideas. Some of these ideas are listed in Table 1. The instructor leads students in an analysis of the case study, specifically following the steps of the scientific method. Particular attention is given to conclusions drawn from the data. One example that has been used in the past is the Brandon Mayfield fingerprints affair. Students are given a description of the events surrounding the 2004 Madrid train bombings (the “make observations” step of the scientific process) and presented with a hypothesis (e.g., “Fingerprints found on detonation devices used in the train bombings likely match the person who used the device”), as well as the data (fingerprints found on detonation devices resembled the fingerprints of Brandon Mayfield). Since the author likes to use the Adam Ruins Everything “Why Fingerprints are Flawed” (18) video for this lecture portion, students are already aware that Mayfield was later cleared of suspicion. Thus, some good points for discussion are the flaws in the conclusions, why fingerprints are not as unique as popularly thought, and the dangers of overreliance on database matching. For the sake of simplicity, the author usually avoids discussing the politics behind the investigation. The initial design of this course did not include case studies. Although the benefits of case studies have long been recognized (8, 15), the author was concerned that the time constraints imposed by keeping all lecture and discussions limited to the first hour of class would forestall adequate case study discussion time. However, case studies were added following student feedback requests for more 64 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

“real-world” examples of techniques. Despite the need for greater attention to the efficiency of time usage, using case studies both increases student interest and improves the kinds of answers students provide to their postlab questions. This improvement may be due to a combination of three reasons. First, student engagement is increased following proof that that what they do in lab is used in real forensic science. Second, how the instructor models the case study discussions acts as a model for students to follow when they explain their answers to questions about their own data. And third, going over a case study immediately following an introduction to the basic science of a technique reinforces the information. Students are thus more interested and better prepared to analyze their own data. During the last two weeks of the lab, students are given a final case study to analyze on their own. Students are expected to complete it outside of lab time, making it one of the course’s few true homework assignments. The instructor writes up a simplified case study summarization, which students are expected to read and write a short report on. In this report, they state the original hypothesis, explain how that hypothesis was tested, summarize the conclusions, and then critique whether or not the data supported those conclusions. This assignment requires that they be able to independently recognize use of the scientific method by other scientists. It can also reveal to the instructor precisely what students do and do not understand about the scientific method.

Lab Experiments A number of considerations were necessary when planning activities. First, activities needed to be feasible for students with limited lab experience. Second, the activities needed to be relatively low cost, because the author’s institution does not collect lab fees and all materials must therefore come from the departmental budget. Third, some of the locations where this course is offered lack a traditional lab space, so the activities needed to involve minimal chemical hazards. Finally, due to limited introductory time and lack of a formal lecture, the activities needed to be easily explained to students. A selection of the more popular activities is described as follows, including some information about how they were designed to meet these needs. A full list of activities and their associated costs and hazards is included at the end of this section in Table 2. Students derive a great deal of satisfaction from processing their original crime scene. For the first lab meeting, the instructor creates a simple, small crime scene with pieces of evidence that the groups will test throughout the remainder of the course. The crime scene rarely consists of more than a handful of items scattered about a relatively small area. During the lecture portion of this class, students have discussed the problems posed by compromised evidence, so they take securing the crime scene very seriously. Students take time to methodically describe and sketch the scene in their notebooks before carefully numbering, collecting, and securing each piece of evidence. In terms of space and cost, the selection of items is left up to the instructor. Items might include pieces of fabric for later fabric analysis, scattered nontoxic powder for later chemical testing, or pieces of plastic or some other item for other tests, depending on what activities the instructor has chosen. Students usually do not dust for fingerprints because, due to the way the fingerprint lab has been set up (as described in next paragraph), this would not be a beneficial use of time. During the second week of lab, students practice collecting and lifting fingerprints. During the lecture portion of this lab period, the instructor explains the anatomy of a fingerprint and what is left behind when an object is touched, with a brief differentiation between latent, patent, and plastic prints. In combination with a review of the basic idea of chemical and physical properties, students are given a simplified explanation of the ninhydrin–primary amine reaction that allows certain latent prints to be visualized. Each student is then given one smooth object, such as a microscope slide, 65 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

and one porous object, such as a 3 × 5 in. card, to touch and turn in. These are labeled ahead of time by the professor, and students make note of which slide or card they touch. After collecting ink impressions of their own fingerprints, students spend the remainder of the lab working in groups to lift prints from the microscope slides and 3 × 5 in. cards using powder or ninhydrin staining and then identify who touched which object. This is a good early semester lab because students find that describing fingerprints in such a way that they can adequately refer to specific features in their conclusion statements is more difficult than they expect. Generally, they enjoy lifting prints, get frustrated trying to find a way to adequately write conclusions that are accepted by the instructor, and then spend the rest of the course being much more detail-oriented and specific in terms of how they describe experimental data. From an instructional standpoint, the ninhydrin staining process is a good, visible example of a chemical reaction to initiate further discussions of chemical and physical properties. From a lab safety standpoint, the individual consumable items are relatively low cost (see Table 2). The main chemical hazard is ninhydrin, but by using it as a 5% solution in water, the fire hazard associated with the typical alcohol-based solution is eliminated. To reduce the associated health hazard, students must handle all items wearing gloves and goggles. (This is the case for the entirety of the lab, and students are graded for wearing appropriate personal protective equipment.) During the fourth week of lab, students perform simple wet chemistry tests. This activity is based on the Forensic Chemistry of Unknown Substances Kit from Carolina Biological (item #699850), although materials may be purchased elsewhere. Students are provided six known compounds and have previously collected samples of an unknown substance from their first week “crime scene.” A portion of each substance is mixed with a series of reagents, and students describe the physical and chemical changes observed. By comparing the reaction of their evidence to that of known substances, they identify their unknown. This activity is interesting in a number of ways, particularly when working with students who have no chemical experience. They often do not know what gelatin is supposed to do, and so the gelling reaction fascinates them. They also may not be aware that talc powder is a component of baby powder, so they describe the mixture as smelling like a baby or pregnancy ward. Following purchase of the original kit (about $100), the materials are replaced by bulk purchase, reducing the subsequent cost of replacement. Each dry chemical is safe to store, and neither the iodine nor biuret reagents are kept in sufficient amounts to need special storage considerations. The design of this course makes it difficult to give students a real sense of instrumental chemical analysis or DNA analysis. Instead, the instructors have created a series of worksheets for students to answer very basic questions in relation to a set selection of infrared (IR), NMR, or mass spectra. Students are provided a simplified summary of where to find peaks on an NMR or IR spectrum, similar to what might be provided to organic chemistry students. They are then given NMR and IR spectra for various compounds and asked to identify certain peaks. As an example, students might be given the IR and NMR spectra for ethanol. The lab worksheet asks them to identify which IR peak and which NMR peak correspond to the presence of the hydroxyl functional group (see Figure 2). To analyze mass spectra, students are given the spectrum of an unknown, along with a library of half a dozen known compounds, and are simply asked to identify which compound the unknown matches. The ultimate activity described by many published descriptions of forensic labs is the creation of a mock crime scene for students to secure and analyze. In many iterations, the crime scene is set up at the beginning of the class, and students progressively collect and analyze evidence for the remainder of the lab. In this course, this model is modified slightly. As previously discussed, students spend the first class period securing a crime scene, with evidence that they test in subsequent weeks. During these weekly activities, key information necessary for understanding the data and conclusions is fresh in students’ minds from the preceding lecture and discussions. 66 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Table 2. Lab Activities and Their Safety Hazards and Associated Costsa

During the last lab meeting, the instructor sets up a new mock crime scene for each group. These scenes are more complex than the original, with more evidence, and include files containing instrumental spectra or suspect interviews from imaginary postscene processing. Now that this course has been run several times, first-time or adjunct instructors have a library of examples and crime scene descriptions set aside to choose from, although they are free to create their own. A selection of examples is shown in Table 3, while an example of the crime scene description that might be shown to students is shown in Figure 3. Note that the description in Figure 3 corresponds to “Group 3” in Table 3. Each group of students spends the lab period securing and processing their crime scene and then testing evidence using appropriate tests. Groups may leave once they can fully 67 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

defend a conclusion to their scene to the satisfaction of the instructor. Although students often feel intimidated by this last week, it is a good way of demonstrating cumulative knowledge and skill gain.

Figure 2. An example infrared spectrum for ethanol, with an accompanying worksheet question to guide student analysis. Students are provided with a table of IR bands. Ethanol spectrum was collected on an Agilent Cary 630 Fourier transform infrared spectrometer at the author’s institution. Due to the accelerated nature of the course, students may only miss one class period before being automatically removed from the course. However, when a student must miss a class, a few alternative assignments have been developed that they may complete at home. These assignments must be recorded in the lab notebook but otherwise may be substituted for grades that would have otherwise been earned during the missed class. In one of these assignments, students may collect fingerprints from three family members and three unrelated individuals and then compare the prints to determine if there are any correlations between fingerprint characteristics and heredity. Like an in-class activity, there are a series of questions to answer, and students must be specific in their explanations. A second option allows students to perform a simple paper chromatography separation on inks from pens that they own using rubbing alcohol and discuss whether or not they could identify which pen was used to write a grocery list based on their observations. Students are asked to treat the final week as an exam, but in the rare event that a student has needed to miss the final crime scene activity, the author has arranged to meet with the student at a convenient time to process a final crime scene individually. 68 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Lab Notebooks

Figure 4. An example lab notebook page using steps of the scientific method as headings under which lab information is organized. Throughout this course, students are expected to maintain a lab notebook, which is divided into sections according to the scientific method. Students fill out each section as they complete the activities, producing something similar to what is shown in Figure 4. Notebook guidelines are otherwise typical for science lab notebooks (i.e., they must be written in ink, in real time, with an updated table of contents, etc.). Students are strongly encouraged to use lab notebooks provided by the institution’s bookstore, which have copy sheets that can be removed and turned in to the professor at the end of each lab period. A standard set of scientific method steps (see Figure 1) includes observations, hypothesis, tests performed, data, and conclusions. Most notebook entries do not have individual observations sections; instead, students are told that their “observations” are their descriptions of the original crime scene(s). Each lab activity starts on a new page, including a title and date, followed by the hypothesis, a brief description of the test or procedures performed, data, and conclusions and postlab questions. We have found two key requirements that improve how students write hypotheses and conclusions. First, we require that hypotheses be written following a specific format. Student 70 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

hypotheses must state which items of evidence are being tested, what tests are being used, and what the student expects to learn by testing the evidence. By requiring this format for a hypothesis, we have found that the specificity of student hypotheses is improved. For instance, without the requirement, students might provide a hypothesis such as, “To identify the unknown.” With the requirement, students are more likely to write the following hypothesis: “Our hypothesis is that by comparing the way Evidence #04-18B reacts with three liquids to the way known chemicals react, we will be able to identify what it is.” This format also helps students organize their approach to thinking about the activity and the exact purpose of each test. In addition to writing an overall concluding statement, students are given specific leading questions to answer at the completion of an activity. The second requirement implemented is that when students answer these questions, they must refer explicitly to which pieces of data support their answers, including the page numbers where the data are found. This, too, has helped ensure that students answer questions in a precise fashion and do more with their data than just record it and ignore it. Students also learn to be very exact when they record data, discovering that vague, imprecise data make it more difficult to answer conclusion questions sufficiently for full credit. Requiring that the lab notebook use scientific method headings, that the hypothesis be written in a certain way, and that answers to postlab questions specifically cite data collected by the students helps avoid a scenario in which students start solving their mysteries before they have collected all their data. A continuing challenge with nontraditional students is that, since many of them are working adults, they are used to solving problems. Their instinct is to solve lab questions exactly the way they solve problems at work, which is often instinctive rather than scientific. As a result of this, students often try to determine a conclusion before all data have been collected and then interpret later data in light of the conclusion they have already drawn. Requiring students to follow a rote series of steps in a set order helps curb this impulse.

Assessment of Student Learning A detailed breakdown of course grades and student learning objectives (SLOs) may be found in Table 4. To summarize, course expectations and assignments are designed around the SLOs, which have been assigned by the department in which this course is housed. Students are assigned points following a set rubric for each graded component of the course, and the final point values for each component are weighted accordingly. Table 4. Breakdown of Grades and SLOs for the Forensic Science Lab

71 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

To ensure that students comply with safety and regulatory guidelines (SLO 6) and to assess students working as a team (SLO 4), the instructor provides a participation grade and a separate evidence handling/safety grade for every class period the student attends. There are no guidelines awarding points; instead, students are informed of the expectations for behavior, safety, and evidence handling during the first lecture period, and instructors assign points at their discretion. The majority of participation points awarded by the instructor come from attendance; however, at the end of the semester, students are asked to peer review members of their group, rating them on a scale of 1–5 for participation and communication. This is incorporated into the overall participation grade and makes up roughly 20% of the total. Points for evidence handling and safety are primarily awarded according to how consistently each group wears safety goggles and gloves and maintains a chain of custody. The remaining course grades (postlab questions, written assignment, final crime scene, and lab notebook) are used to ensure that students meet the remaining learning objectives (SLOs 1–3, 5), and all have rubrics that are made available to students at the start of the course. Answers to postlab questions (partial assessment of SLO 1) are due by the end of each lab period and are graded according to both correctness and strength of data supporting the answer. This rubric is shown in Table 5. The written assignment, which is more fully described in the “Engaging Thinking with the Scientific Method” section, is graded primarily according to how well students identify each aspect of the scientific method (SLOs 1 and 2). A partial rubric for this assignment is shown in Table 6. The sections shown have been selected as providing the clearest snapshot of how points are awarded for this assignment. Additional points are awarded for following formatting guidelines for the assignment. Half of a student’s grade comes collectively from the final crime scene and the lab notebook, which assesses the students’ abilities to synthesize all learning objectives. The lab notebook acts as a record of students using the steps of the scientific method and their progress in learning how to explicitly use collected data to support conclusions. This rubric is filled out by the instructor throughout the semester. An example section, with points from weeks 1 and 2, is shown in Table 7. The full rubric includes sections for the hypothesis, procedures, and data from weeks 3 and 4. Processing of the final crime scene acts as a measure of students’ abilities to synthesize what they have learned the previous weeks in a larger, more complex problem. A portion of the rubric for this assignment is shown in Table 8.

72 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Table 5. Grading Rubric for Postlab Questionsa

Table 6. Scientific Method Assignment Rubric (Partial)a

73 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Table 7. Lab Notebook Rubric (Partial)a

74 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Table 8. Week 5 Crime Scene Rubric (Partial)a

Course Assessment, Success, and Challenges Since its inception, this forensic science lab has become the most popular nontraditional standalone lab at the author’s institution. Students report having a lot of fun with it while learning a lot about forensic science and the roles that general science and chemistry have in their daily lives. Seventy percent of students who take the course pass with an A, while another 25% receive a B or C. Most students receive a score greater than 93% on participation and evidence handling, with the number of points received for evidence handling increasing between weeks 1 and 4 and decreasing during the final week. The average student receives an 85% on the scientific method assignment, with most points being lost due to misstatement of what techniques were used to test the hypothesis. The average grade on the final crime scene is an 83%, and the average notebook grade is an 86%. Rarely do students fail the course; those who do typically failed to follow directions or did not complete multiple assignments. There are a few administrative challenges associated with running this course. Because this course is sometimes taught at locations off campus, this creates challenges with instructor training and inventory maintenance. Off main campus, the lab is most often taught by adjuncts, some of whom have no formal training in chemistry or forensic science. An effort has been made to make course materials clear and user friendly for adjuncts. This includes documents written specifically for adjuncts listing common troubleshooting issues with each lab and providing general background information in an attempt to make them feel prepared for the lesson. The course designers have also tried to make themselves easily available for questions from adjuncts or in case issues arise. Inventory maintenance is a continuing problem. Instructors are provided a list of materials and are asked to inform course designers when materials need to be ordered; however, it still happens that materials

75 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

will run out during a course and need to be emergency ordered or taken from main campus supplies and dropped off at the location. Beyond the usual instructional challenges, students often want to carry out lab work on scrap paper and transfer the information to their lab notebook at home. Students see this as a way to avoid making a mistake when putting information in the lab notebook, not realizing that the notebook is supposed to be where mistakes occur and how difficult it is to remember details accurately after the fact. This is primarily corrected during the first couple of weeks of the course by extra attention from the instructor. Finally, during the final lab session, when students process an entire crime scene from start to finish, they often neglect the use of proper safety procedures and fail to maintain an adequate chain of custody for evidence. To combat this, we recommend that instructors take special care to remind students that these processes are still relevant and important.

Conclusions This concludes an overview of a five-week course designed to meet one semester hour of general education critical thinking credit for nontraditional, nonscience majors. The course was specifically designed to be low cost and easy to implement at locations lacking a formal lab setup. It was also designed to allow some flexibility on the part of the instructor as to which activities are used and in what order, depending on student abilities and interest or available resources. Therefore, while this course is taught at a small four-year institution, it could be successful at community colleges, high schools, or science summer camps. Critical thinking is practiced using repeated references to the steps of the scientific method, and students are expected to be able to make explicit connections between their conclusions based on forensic evidence and the data collected. Overall, this has been a very successful course for the author’s institution.

Acknowledgments The author would like to acknowledge the Chemistry Collaborations, Workshops and Communities of Scholars program for holding the Forensic Science workshop and Dr. Lawrence Kaplan for sharing his knowledge and resources for the creation of forensic science labs.

References 1. 2. 3. 4. 5.

6.

Bye, D.; Pushkar, D.; Conway, M. Motivation, Interest, and Positive Affect in Traditional and Nontraditional Undergraduate Students. Adult Education Quarterly 2007, 57, 141–158. Chao, R.; Good, G. E. Nontraditional Students’ Perspectives on College Education: A Qualitative Study. Journal of College Counseling 2011, 7, 5–12. Dill, P. L.; Henley, T. B. Stressors of College: A Comparison of Traditional and Nontraditional Students. J. Psychol. 1998, 132, 25–32. U.S. Government Accountability Office. Higher Education: Students Need More Information to Help Reduce Challenges in Transferring College Credits; GAO-17-574; August 2017. Simone, S. A. Transferability of Postsecondary Credit Following Student Transfer or Coenrollment: Statistical Analysis Report; NCES 2014-163; IES National Center for Educational Statistics: Washington, DC, August 2014. Kaplan, L. J. Chemistry and Crime: From Sherlock Holmes to Modern Forensic Science; A Science Course for Non-Science Majors. Crime Laboratory Digest 1992, 19, 107–132.

76 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

7. 8. 9. 10. 11. 12. 13. 14.

15. 16.

17.

18. 19. 20.

Arwood, L. Teaching Cell Biology to Nonscience Majors Through Forensics, or How to Design a Killer Course. Cell Biology Education 2004, 3, 69–140. Kaplan, L. J. Forensic Science: Crime in the Chemistry Curriculum. J. Chem. Educ. 1993, 70, 574. Mozayani, A.; Noziglia, C. The Forensic Laboratory Handbook: Procedures and Practice, 2nd ed. Humana Press: New York, NY, 2011. Clark, M. J.; Keegel, J. F. Chemistry in the Crime Lab. A Forensic Science Course. J. Chem. Educ. 1977, 54, 38. Nienhouse, E. J. Chemistry and Crime: A Laboratory-Based Forensic Science Techniques Course as an Alternative to a Natural Science Requirement. J. Chem. Educ. 1985, 62, 1047. Nabais, J. M. V.; Costa, S. D. A Forensic Experiment: The Case of the Crime at the Cinema. J. Chem. Educ. 2017, 94, 1111–1117. Kanu, A. B.; Kaplan, L. J. The Quest for Confirmatory Data in Crime Scene Investigations. Chem. Educ. 2016, 21, 231–239. Kanu, A. B.; Pajski, M.; Hartman, M.; Kimaru, I.; Marine, S.; Kaplan, J. K. Exploring Perspectives and Identifying Potential Challenges Encountered with Crime Scene Investigations when Developing Chemistry Curricula. J. Chem. Educ. 2015, 92, 1353–1358. Willingham, D. T. Critical Thinking: Why Is It So Hard To Teach? Arts Education Policy Review 2008, 109, 21–32. Abrahams, I.; Millar, R. Does Practical Work Really Work? A Study of the Effectiveness of Practical Work as a Teaching and Learning Method in School Science. International Journal of Science Education 2008, 14, 1945–1969. van Ede, J. Scientific Method Core of All Improvement Methods. Business-improvement.eu, 2017. http://www.business-improvement.eu/worldclass/scientific_method_process_ improvement.php (accessed Dec 14, 2018). Galindo, A. NCIS 7x16 id marry mass spec 40 sec. YouTube, March 15, 2017. https://www. youtube.com/watch?v=b9f Iwpvn8GQ (accessed March 6, 2019). truTV. Adam Ruins Everything – Why Fingerprinting is Flawed. YouTube, Oct 16, 2015. https://www.youtube.com/watch?v=vM1QgwaKv4s (accessed March 6, 2019). Saferstein, R. Criminalistics, An Introduction to Forensic Science, 10th ed.; Prentice Hall: Upper Saddle River, NJ, 2009.

77 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Chapter 5

Teaching Introductory Forensic Chemistry Using Open Educational and Digital Resources Downloaded via MIAMI UNIV on November 5, 2019 at 02:19:07 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Brooke Taylor* Science Division, Lane Community College, Eugene, Oregon 97405, United States *E-mail: [email protected].

To help reduce costs for students and to maintain the currency of the course content, open educational and digital resources, along with an instructordeveloped lab packet, are now used exclusively as course materials for Introductory Forensic Chemistry. In this chapter, major themes, course learning outcomes, topics, digital and open educational resources, labs, and assignments will be detailed.

Introduction In an effort to reduce textbook costs for students and improve the quality and currency of course materials, Introductory Forensic Chemistry, CH 114, at Lane Community College now uses a variety of digital and open educational resources to replace the costly textbook. All sources are available online for free under fair use permission if they are not openly licensed. Sources include the National Institute of Justice, Oregon State Police, and National Forensic Science Technology Center (NFSTC), among others. Current forensic cases reported in national newspapers and by news organizations are also used. With the exception of one video edited for use in the classroom, no modifications have been made to the resources. This chapter will also detail major themes of the course, outcomes, and types of evidence, as well as labs and assignments.

Digital and Open Educational Resources Fair use of copyrighted material is permitted by nonprofit educational institutions for teaching, research, and scholarship, but the material may not be used for public distribution (1, 2). Many digital resources on government and educational Web sites are available under fair use copyright. For example, “information on Department of Justice Web sites is in the public domain and may be copied and distributed without permission. Citation of the Department of Justice as source of the information is appreciated, as appropriate. The use of any Department of Justice seals, however, is protected and requires advance authorization, as described below” (3). Open educational resources (OER) “are educational materials and resources offered freely and openly for anyone to use and © 2019 American Chemical Society Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

under various licenses to re-mix, improve and redistribute” (4). Creative Commons is the most widely used licenser of OER. Unlike fair use of copyrighted materials, resources with Creative Commons copyright may be retained, reused, revised, remixed, and redistributed (4). The ability to make and own copies of resources, modify and edit resources, and share resources provides educators more open and flexible use of content. The challenge for some topics, such as introductory forensic chemistry, is the lack of resources licensed by Creative Commons and the lack of a single open educational textbook, such as those found on OpenStax for introductory, general, or organic chemistry. OpenStax publishes peer-reviewed, openly licensed, college textbooks in many areas of science, including chemistry, physics, and biology, but not forensics. Online access to OpenStax material is free, and low-cost printed copies are available (5). A large number of forensic-related photographs and videos have a Creative Commons copyright. Few resources appropriate for academic use to replace a textbook have an open copyright and would therefore not be considered an OER. Given the lack of a single digital OER for introductory forensics, a variety of free digital resources are currently used in the course. A link to each resource is included on the Introduction to Forensic Chemistry, CH 114, course Moodle site (6). Access to the course Moodle site is available by logging in as a guest. The course Web page, along with other open educational courses taught in the state of Oregon, may also be accessed from the Open Oregon Educational Resources Web site (7). Table 1. Article Review Assignment. Reproduced with permission from reference (10). Copyright Creative Commons Attribution 4.0 International. General Theme

Specific Questions

Introduction

Briefly summarize the article in your own words. What topic(s) or questions does the article discuss? What background information is provided? Briefly summarize one or two of the most interesting points of the article for you.

Forensics and Science How does the article relate the topic of forensics with science? Briefly summarize the supporting data and/or key evidence presented in the article. What topics mentioned in the article have been (or will be) discussed in CH 114? Please include the digital resource or a related lab or a class date for discussion of the topic. Evaluating the Source

Briefly evaluate the article using the CRRAPP criteria. Is the article popular or scholarly?

Conclusions

What conclusions or points do the article authors make? What questions did the article raise but not answer? How might you be able to answer these questions?

Given the wealth of information on the Internet, resource selection is important and must be evaluated critically by both the instructor and students. In an effort to further build scientific literacy, students first view a tutorial about evaluating online sources (8). Students are then introduced to the CRRAPP rubric, a tool designed to critically evaluate Web sites or other resources: for currency, relevance, reliability, authority, purpose, and point of view (CRRAPP) (9). The CRRAPP rubric is a component of each article review assignment completed in the course. See Table 1 for the specific components of the article review assignment used throughout the course (10). Students also view a short tutorial about popular and scholarly articles (11). Focusing on scholarly resources but also including popular sources provides instructors the opportunity to maintain a high level of currency and relevance for students. 80 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

For each specific type of evidence, students are responsible for determining how and from what locations the evidence is collected. The proper collection and storage of evidence is based on the Oregon State Police Evidence Manual (12). Evidence collection is certainly an area where details matter. Initially student responses were so broad and brief any evidence collected in the manner described by students would not have been preserved properly for analysis and would therefore not be admissible in court. To help focus student responses, specific evidence collection assignments were developed such as open-note quizzes or open-source quizzes. Table 2 provides an example where students complete the assignment using the online Oregon State Police Evidence Manual. Table 2. Arson Evidence Collection Open-Source Quiza. Reproduced with permission from reference (13). Copyright Creative Commons Attribution 4.0 International. What type of container is used for suspected arson evidence? Briefly explain why this type of packaging is required. Should arson evidence, like blood and DNA evidence, be dried? YES or NO? What observations at the scene may indicate arson? List them all. Is one more likely to indicate arson than any others? If so, which one? What types of evidence should be collected? Which type of evidence is most likely to indicate arson? What is a comparison sample? What is a control sample? In addition to your answer for question #1, what other types of materials may be used to collect evidence? What materials used to collect suspected arson evidence should not be submitted to the lab? a Students would complete this assignment online using the Oregon State Police Evidence Manual.

The lab work in Introductory Forensic Chemistry focuses on how the evidence is analyzed in a crime lab using proper scientific technique, and whether each test is presumptive or confirmatory. A presumptive test screens for the presence of a substance based on general characteristics. A presumptive test is not conclusive; multiple substances may yield the same positive test results. A confirmatory test positively identifies a substance and generally uses analytical instruments. Students prepare for labs by reviewing the NFSTC Web site (14) as well as other sources if the type of evidence is not discussed on the NFSTC site. The focus of the NFSTC Web site is to simplify forensic science. Topics relevant to this course include bloodstains, crime scene investigation (CSI), DNA, drugs, explosives, toxicology, and trace evidence. Other topics on the NFSTC site related to forensics but not discussed in this class include digital evidence, firearms, evidence and witnesses, footwear and tires, and photography. For each topic, NFSTC introduces the topic, explains principles of the analysis, describes the analysis, has frequently asked questions, and provides a list of references and resources. Permission was requested and granted to use the NFSTC Web site.

Course Development, Structure, and Learning Outcomes Introductory Forensic Chemistry was developed as a four-credit lecture lab course at Lane Community College in 2006 to expand the application-based course offerings of the chemistry discipline and to meet the general education lab science requirement for community college students to earn a degree to transfer to a four-year school in Oregon (Oregon transfer degrees). Course learning outcomes are listed in Table 3. The course meets for 6 h each week during a 10-week quarter. Final exams are scheduled for the 11th week of the quarter.

81 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Course grades are assigned based on two in-class exams worth 20–30% of the course grade, labs worth 20–30%, and activities, including article reviews and in-class assignments, worth 15–25% of the course grade. Students determine the specific percentage each category is worth on a grade distribution form. Table 3. Course Learning Outcomes and Assessment Methods Course Learning Outcomes

Assessment Methods

Understand basic chemical concepts used in forensics such as density, measurement, chromatography

Class discussions, reading assignments, article review, labs, exams

Apply concepts to solve a problem

Evidence analysis, reading assignments, labs, exams

Be familiar with methods of scientific thought and Class discussions, reading assignments, article review, labs, exams inquiry Critically analyze data

Class discussions, reading assignments, article review, labs, exams

Build scientific literacy

Class discussions and reading assignments related to forensics in the news, and scientific journal articles

Work safely in a lab

Weekly labs

The schedule used for the most recent course is shown in Table 4. Each class period was 3 h. A shorter lab is often completed in addition to the classroom lecture or discussion during the first class meeting of the week. The second class period of the week is primarily spent in lab. Web site links, the course packet containing most class assignments, and all labs and additional assignments not part of the course packet are posted on the course Web site (6). Table 4. Weekly Schedule for Introductory Forensic Chemistry from Spring Term, 2018 Class

Topic

1

Course Intro, CSI Effect, Physical Evidence, Scientific Method, Lab Safety

2

Measurement Lab

3

Evidence, Metric System, Glass Lab: Refractive Index of Glass

4

Determination of Glass Density Lab

5

Organic Analysis, Lab: Observing Chemical Reactions

6

Identification of Drugs and Poisons Lab

7

Drug Analysis

8

Thin Layer Chromatography Lab

9

Inorganic Analysis, Lab: Serial # Restoration

10

Spectroscopy Lab

11

Forensic Serology, Exam #1

12

Visible Spectroscopy Lab

82 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Table 4. (Continued). Weekly Schedule for Introductory Forensic Chemistry from Spring Term, 2018 Class

Topic

13

Forensic Serology (Blood), Lab: Blood Typing

14

Blood Identification Lab

15

DNA Gel Electrophoresis Practice

16

DNA Extraction and Polymerase Chain Reaction (PCR) Lab

17

No Class, Memorial Day

18

DNA Lab, Gel Electrophoresis

19

Arson

20

Identification of Unknowns by IR Lab

21

Exam #2

General Course Themes, Topics, and Types of Evidence The course introduces chemistry in a forensic context by emphasizing the chemical analysis work completed in a forensic lab. Major themes include an introduction to forensics, the application and assessment of the scientific method to forensics, evidence collection, evidence analysis, the limitations of the evidence analysis, and case studies. Types of evidence include glass, drugs, trace elements, accelerants, blood, and DNA. Past offerings of the course have also included the collection and analysis of soil. Selected chemistry topics complement the specific forensic themes and types of evidence. Chemistry topics include the metric system, physical properties including density and refractive index, atoms, elements and compounds, analytical techniques, chromatography, spectrophotometry, organic analysis, and inorganic analysis. Based on the types of evidence and forensic themes, typical introductory chemistry topics such as balancing equations, stoichiometry, gas laws, enthalpy, and electron configurations are absent from Introductory Forensic Chemistry. Forensic chemistry students are not introduced to the mole anywhere in the course, unless a mole happens to leave trace evidence at the scene of a crime. Lane Community College also offers an Introductory Forensic Science course through the criminal justice discipline in the social science division. This course is intended for criminal justice majors, including peace officers and public safety officers. Topics include modern techniques employed in investigating crimes, the value and significance of physical evidence, the collection and preservation of evidence, understanding criminal lab methods, procedures for analyzing evidence, and the effective preparation of physical evidence for trial (15). Many of the general themes in the criminal justice class are the same in the Introductory Forensic Chemistry class. The two courses complement each other with each focusing on different types of evidence and processes in the collection and analysis of the evidence. The criminal justice class focuses more on documenting the crime scene and the collection of the evidence prior to submitting the evidence to the lab for analysis, whereas the chemistry course focuses more on the analysis of the evidence in the lab. Students are able to take both courses to meet their general education lab science requirement.

83 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Introductory Forensic Chemistry Labs Introductory Forensic Chemistry is a lab-focused course with a minimum of 13 labs completed during the term with some labs only requiring approximately 1 h and others requiring lab work during multiple class periods. For some labs, the college does not own the necessary instrument for analysis. In these cases, we improvise for the lab activities but always include a discussion of how the evidence would actually be analyzed in the crime lab. Students also analyze the same piece of evidence in multiple ways to improve the evidential value of the physical evidence, and whenever possible, students test both known and unknown samples for comparison. Table 5. Introductory Forensic Chemistry Labs Lab Title

Green Materials or Techniques

Measurement and Density

Metal slug and ethanol

Refractive Index of Glass

Clove oil, olive oil, 50:50 mix of clove and olive oils

Glass Density

Archimedes principle and/or water displacement

Observing Chemical Reactions

Neutralization, precipitation, and two oxidation–reduction reactions

Thin-Layer Chromatography of Ink

Water-soluble ink, ethanol solvent to extract ink and develop plate

Spectroscopy

Gas lamps and flame tests

Identification of Drugs and Poisons

Water solubility and color tests for aspirin, sucrose, fructose, ascorbic acid, and caffeine

Restoring Serial Numbers

Iron (III) chloride and acid solutions

Identification of Unknowns by IR Ethanol, methanol, or hexane, and electronic library Spectroscopy Blood Type and Identification

Purchased kit and sheep’s blood and red food dye for identification

Visible Spectroscopy

Red food dye

Isolation of DNA

Wheat germ and strawberries

Forensic PCR Investigation of DNA

PCR followed by gel electrophoresis, purchased kit

Lab curricula at Lane Community College, including forensic labs, are green. Green chemistry “is an effort to change or reinvent processes used in chemical manufacturing to ensure a safer, cleaner environment” (16). Of the six actions recommended by the American Chemical Society, labs in Introductory Forensic Chemistry focus on getting off to a safe start by using nonhazardous materials, using safer solvents to eliminate the use of toxic solvents, and avoiding waste (17). For example, instead of using toxic solvents to test actual controlled substances (drugs), students complete color tests on more benign substances using safer solvents in the Identifications of Drugs and Poisons lab. Table 5 lists other green materials used in the labs and a note about technique if the lab title isn’t technique-specific. When soil is analyzed in the lab with a density column, instead of using carcinogenic bromoform and xylene, ethanol and concentrated sugar solutions are used instead. More details about each lab are included in the Course Description by Topic section. All labs are

84 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

available in the course packet posted on the course Web site (6) and can also be accessed from the Open Oregon Educational Resources Web site (7). Introductory Forensic Chemistry was originally developed in coordination with the criminal justice course. One of those instructors, an Oregon State Police Crime Lab Forensic Scientist, made an important point that forensic scientists do not solve crimes; they analyze the evidence. In order to maintain that consistency and foster a realistic image of what forensic chemistry is and is not, all labs focus on analyzing the evidence without much context or a story about the purported crime. While many forensic courses may weave the lab work into a narrative about the crime, this, in my opinion, perpetuates the unrealistic portrayal of forensics on past and current television dramas, known as the CSI Effect. To support a more realistic image, labs focus on testing both known and unknown samples and, when possible, identifying the unknowns. Students have already determined if the test is presumptive or confirmatory, and this focus is maintained in the lab. Students are not penalized if they are unable to identify the unknown. Sometimes narrowing down the possibilities is the best we can do, and it is an important demonstration of the limits of a particular test. To provide students with a realistic view of daily activities in a crime lab, a local forensic scientist is invited to speak to the class. Past guests have included latent print analysts, as well as a controlled substance analyst.

Specific Course Details by Topic and Type of Evidence The course begins with a general introduction to forensic science (18), a summary of crime lab operations (19, 20), the definition of Locard’s exchange principle (21), an introduction to types of evidence (22), and an introduction to the scientific method and the impact of research and development (23). During the first week of lab, students complete a basic measurement lab to practice using a variety of typical laboratory glassware. Students experimentally determine the density of an oddly shaped object using water displacement to measure the volume. The density of a known liquid is also experimentally determined and the percent error calculated. To address the application and assessment of the scientific method to forensics, a variety of popular and scholarly articles have been used. During the spring of 2018, students completed an initial article review of Balko and Carrington’s article, “Bad Science Puts Innocent People in Jail—and Keeps Them There” (24). Students then viewed the “Forensic Science” episode of Last Week Tonight with John Oliver, edited for the classroom (25). Students completed an initial draft of the article review assignment described in Table 2 for the newspaper article, and then submitted a final draft incorporating the video after a class discussion. Both Balko and Carrington and John Oliver reference the reports from the National Academy of Sciences (26) and the President’s Council of Advisors on Science and Technology (27). Each report is extensive and beyond the introductory scope of the class. During the first term the course was taught using digital resources, neither the President’s Council nor the National Academy of Sciences reports were assigned to students. Future classes will likely include assigned sections from each report. Throughout the course, the class returns to the theme of applying and assessing the scientific method to the specific types of evidence and analysis techniques. After the introduction, the class begins to discuss specific types of physical evidence. Students complete an assignment about the proper collection and storage of each type of evidence using the Oregon State Police Evidence Manual (12). The first type of evidence students analyze is glass. During lecture, the class examines how glass fractures, how to determine the direction and order of impact, and how to properly collect glass evidence from a crime scene. In class, students examine a diagram of both sides of a window with glass fractures to determine the order and direction of 85 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

each impact. An example of this assignment is in the course packet available on the course Web site (6). Students also examine a sample of a small window with the fracture side clearly marked. In the lab, students determine the density of known and unknown glass samples, including types typically found at crime scenes such as window, bottle, and eye glass. Pyrex lab glass is also included. Forensic analysis of glass involves the determination of indices of refraction using a refractometer. Lane Community College does not own such a device, so instead known and unknown glass samples are submerged in three different oils with different indices of refraction. Students observe the presence or absence of a bright halo near the edge of the piece of glass immersed in a liquid of different refractive index, known as the Becke line (28). Organic analysis is the next major topic discussed in class. Prior to beginning work in the lab, students are introduced to functional groups using molecular model kits. Spectrophotometry, the wave nature of light, and the electromagnetic spectrum are also introduced prior to the lab. The presumptive and confirmatory aspects of different types of spectrophotometry are also discussed with specific examples of spectra collected in forensic labs used whenever possible. Different types of chromatography are introduced, including thin-layer and gas chromatography. Drugs and ink are typical examples of evidence analyzed in a crime lab using organic analysis techniques. Prior to the Identification of Drugs and Poisons lab, students begin with observing chemical reactions as an introduction to chemical changes. In the Identification of Drugs and Poisons lab, instead of testing controlled substances typical in a forensic lab, students test more common substances using reagents following the principles of green chemistry. Students conduct color tests on each known substance and an unknown to presumptively identify their unknown. The identity of the unknown is confirmed by comparison to provided IR spectra. Lane Community College purchased a Thermo Scientific Nicolet iS5 IR spectrometer during the summer of 2018. Future students will collect their own IR spectra to analyze through an electronic library. Thin-layer chromatography may also be used to presumptively identify substances used in the color test lab. Ink is also analyzed in a separate thin-layer chromatography lab. For either chromatography lab, solvents are selected that meet the criteria of green chemistry. Testing controlled substances requires hazardous solvents for both the presumptive color tests and chromatography. A license may also be required to work with controlled substances. Lane Community College has elected to not use hazardous substances in academic labs if other options are available. To complete the introduction to organic analysis section of the course, students are also introduced to the use of organic analysis in a toxicology context. Inorganic analysis follows. Students are introduced to the structure of atoms and the periodic table, and they complete an assignment from the course packet asking about particular elements and isotopes in a forensic context. The importance of trace elements in forensics is discussed, and students observe the emission spectrum of a variety of gas lamps as well as flame tests in the lab, and use their observations of known substances to presumptively identify unknowns. The class then discusses inductively coupled plasma-optical emission spectroscopy and how it compares to the emission spectroscopy lab. As a demonstration of the use of inductively coupled plasma-optical emission spectroscopy in forensics, the class discusses the Tri-State Crematory case. The final lab in this section has students channel their inner criminal. Lane Community College has a large number of old lab drawer keys with numbers engraved by hand. Forensic chemistry students file away the number, and then another student attempts to chemically restore the “serial” number. Biological evidence is the focus of the next section of the course, beginning with serology and finishing with DNA. As a lead-in to serology, students use visible spectroscopy to determine the concentration of a substance in a collected “blood sample” as a mock toxicology lab. Both serology and DNA have a significant variety of online and open educational resources available. Introductory forensic chemistry characterizes blood but does not analyze blood spatter; that topic is part of the 86 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

criminal justice course. To forensically characterize blood, the analyst must first determine if the sample is blood, then determine from which species the blood sample originated, and, lastly, associate the sample with a particular person. In the lab, students presumptively determine if samples and unknowns are blood using luminal and reduced phenolphthalein, called the Kastle–Meyer test. Associating blood with a particular person based on blood type is completed using a purchased blood typing kit from Carolina Biological (29). Prior to the lab, students complete an online tutorial on blood typing (30). The class also discusses the limitations of blood typing to associate a blood sample with a particular person and the value of DNA evidence, the next topic in the course. Students are introduced to the basics of DNA in class with an online tutorial from the DNA Learning Center to supplement the lecture discussion (31). In class, students use model kits to build an informational strand of DNA consisting of about 12 base pairs. Student groups then work to build the template or complementary strand. The sequence of mRNA is then determined as are the amino acids that small strand of DNA codes. DNA replication is introduced and connected to the forensic use of the PCR to replicate DNA for lab analysis. The importance of analyzing junk DNA for forensics is discussed, and the Druid Dracula case study is introduced (32). This case includes the analysis of the amelogin gene to determine if the DNA sample is from a male or female source. In the lab, students isolate DNA. Using a kit purchased from Carolina Biological (33), students replicate DNA using PCR and analyze the DNA using gel electrophoresis. To supplement the lab work, students complete a virtual lab extracting DNA (34) and complete an online tutorial about gel electrophoresis (35). The goal for future classes is for students to isolate a sample of DNA and then run PCR and gel electrophoresis on that same sample. Arson is the final topic of the course and returns to organic analysis previously discussed. Students complete a final case study titled Burning Down the House (36) while analyzing accelerant samples from a suspected arson on the IR spectrometer. The case study was written for analysis using gas chromatography; however, at the current time, Lane Community College does not have the required equipment. In addition to a basic introduction of combustion reactions, the class discusses the different requirements to collect arson evidence compared to other types of physical evidence addressed in the Arson Evidence Collection assignment in Table 2. As the course wraps up, typical signs of arson are discussed. Returning to the initial theme of the scientific method in forensics and the limitations of evidence analysis, students either read an excerpt from a New Yorker (37) article or view an excerpt of a Frontline episode, “Death by Fire” (38). Both resources describe cases where a controlled test fire (nonarson origin) left evidence typically and historically thought to be associated only with arson fires started with accelerants.

Case Studies and Research and Development Reports Forensic science is a constantly evolving topic. New cases are reported in newspapers and by news outlets on a regular basis. Case studies from the current media generally involve the article review assignment described in Table 1. Students complete a rough draft article review and then resubmit a final draft after the class discussion of the case. A large number of educational and government sources also publish online resources. These sources focus on research and development studies in forensics. Additionally, the National Center for Case Study Teaching in Science, in Buffalo, New York (39), publishes case studies in a variety of topics for use in the classroom. These case studies are available for free, however, access to answer keys and teaching notes requires a small annual subscription. Burning Down the House (36) and Druid Dracula (32) are both cases currently used in the class published by the National Center for Case Study Teaching in Science. 87 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Examples of case studies from popular sources include the Smell of Death in the Casey Anthony case in 2008 as an example of the validity of new scientific research and its admissibility in court based on the Frye standard. The Tri-State Crematory case of Ray Marsh in 2002 is an example of using an analytical technique to identify a trace element in cremains, the remains after cremation. The most recent case is that of the Golden State killer, Joseph DeAngelo, in 2018. In this case, DNA collected from the crime scene (unknown suspect) was submitted to a commercial genealogy Web site by law enforcement under an assumed name. The suspect DNA was related to a sample submitted voluntarily by a distant relative of the suspect, and Joseph DeAngelo was identified as the suspect in murders going back 30 or 40 years. Articles and reports about research and development studies have included the CSI Effect (40), an improved method for removing lipstick samples from surfaces (41), the effect of luminol on other presumptive tests to identify blood and DNA analysis (42), and the use of microcrystalline and infrared microscopy as developing analytical techniques to analyze emerging drugs (43). In an attempt to keep the course up-to-date, a different research and development article is used each year. ScienceDaily is an excellent source to find summaries of new developments in forensic research (44). Students are encouraged to check this page frequently and may be assigned to review an article of their choice featured on ScienceDaily.

Future Work and Next Steps In an effort to continue to improve Introductory Forensic Chemistry, assignments will continue to be developed to assist students in focusing on the detailed work of properly collecting and storing evidence. The application and assessment of the scientific method to forensics with a focus on research and development in evidence analysis and the limitations of the analysis will be improved by including specific sections of the reports from the National Academy of Sciences (26) and the President’s Council of Advisors on Science and Technology (27). Current cases will continue to be added to the course, and new developments in research in forensic science will be used for article review assignments. As lab equipment is upgraded and added, labs will be revised to replicate as closely as possible the evidence analysis techniques used in a crime lab, while also maintaining the commitment to green chemistry. A reading list with a brief description of each resource and current and active links for all of the resources included in this chapter is in the works. The intent is for the reading list to have a Creative Commons copyright so other instructors may remix it, improve it, retain it, reuse it, revise it, and redistribute it. As new resources are found, the reading list will be updated. Once complete, it will be available on the Open Oregon Educational Resources Web site (7).

Conclusion Using free digital and open educational resources from the Internet provides instructors up-todate course materials at a significant reduction in costs for students. Additionally, the use of digital resources provides instructors the opportunity to tailor the course to specific topics and outcomes, without being obligated by topics in the textbook. The flexibility created by using free digital and open educational resources also provides instructors the opportunity to use a select set of resources in a chemistry class not specific to forensics. For example, inductively coupled plasma-optical emission spectroscopy and the case of the Tri-State Crematory could be discussed in a general or introductory chemistry class during the atomic structure section. 88 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Staying current with the resources is challenging, however, because no single source exists, and Web site links can change, move, or disappear. How many times have you clicked a link to a site used in a class the previous year to find it no longer there? The loss of resources creates a challenge, especially when an entire course is based on digital resources. Following a series of links from an initial resource may also result in an overwhelming amount of information for both the instructor and the students. However, the development of new and improved resources for inclusion in a digital resource course and the built-in relevance of those resources are a great benefit. The National Institute of Justice Forensic Science Web site (18) is an excellent starting point for general aspects of forensics. The research and development page (45) lists forensic research and development projects, and makes an excellent place to start for focusing on scientific inquiry in the field of forensics. ScienceDaily (44) is also an excellent source for up-to-date summaries of research in forensic sciences. The resources listed in the reading guide and referenced in this paper have been used successfully with students and will continue to be the foundation of the Introductory Forensic Chemistry taught at Lane Community College. As new resources are discovered, the reading guide will be updated, one of the significant benefits of using online and open educational resources. This a very exciting and educational time to study forensics. The use of digital and open educational resources keeps forensic topics and courses relevant and interesting for both the student and the instructor.

References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10.

11. 12. 13.

Lane Community College Library Guide. https://libraryguides.lanecc.edu/fair (accessed Nov 29, 2018). Fair Use Checklist, University System of Georgia. https://www.usg.edu/assets/usg/docs/copyright_docs/ Fair_Use_Checklist_(1-10-17_final).pdf (accessed Nov 29, 2018). Legal Policies and Disclaimers; U.S. Department of Justice: Washington, DC, 2016. https://www.justice. gov/legalpolicies (accessed Jan 7, 2019). Lane Community College Research Guide, Open Educational Resources. https://libraryguides.lanecc.edu/c. php?g=391465&p=2658775 (accessed Nov 29, 2018). OpenStax, About. https://openstax.org/about (accessed Nov 29, 2018). CH 114 Forensic Chemistry OER Course Moodle Site, Lane Community College. https://classes.lanecc.edu/ course/view.php?id=86872 (login as a guest; accessed Dec 6, 2018). Open Oregon Educational Resources, Resources. https://openoregon.org/resources/ (accessed Jan 5, 2019). Evaluating Online Sources, Lane Community College Library Research How-Tos. https://library.lanecc.edu/ howtos/video/evaluating_internet_sources (accessed March 28, 2018). CRRAPP Criteria Rubric, Lane Community College Library Research How-Tos. https://library.lanecc.edu/ sites/default/files/handouts/CRRAPPcriteria.pdf (accessed March 28, 2018). Taylor, B. CH 114 Article Review Assignment. In CH 114 Introduction to Forensic Chemistry Packet, Spring Term; Lane Community College: Eugene, OR, 2018; p 7. https://classes.lanecc.edu/pluginfile. php/3074100/mod_resource/content/4/CH%20114%20Lab%20ManualSP18edits.pdf (accessed Jan 25, 2019). Popular and Scholarly Sources, Lane Community College Library Research How-Tos. https://library.lanecc. edu/howtos/video/popular_and_scholarly_sources (accessed March 28, 2018). Physical Evidence Manual; Oregon State Police: Salem, OR, 2017. https://www.oregon.gov/osp/Docs/ PhysicalEvidenceManual.pdf (accessed Feb 25, 2019). Taylor, B. Arson Evidence Collection Assignment; Lane Community College: Eugene, OR, 2018. https://classes.lanecc.edu/pluginfile.php/3074141/mod_resource/content/3/Arson%20evidence. pdf (accessed Feb 25, 2019).

89 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

14. A Simplified Guide to Crime Scene Investigation; National Forensic Science Technology Center: Largo, FL, September 2013. http://www.forensicsciencesimplified.org (accessed Dec 18, 2018). 15. CJA 214, Intro to Forensic Science, Course Description, Spring 2018, Lane Community College Catalog. https://crater.lanecc.edu/banp/bwckctlg.p_disp_dyn_ctlg (accessed Jan 5, 2019). 16. Anastas, P.; Warner, J. Green Chemistry: Theory and Practice; Oxford University Press: New York, 1998; p 30. 17. Ryan, M.; Tinnesand, M. Introduction to Green Chemistry: Instructional Activities for Introductory Chemistry; American Chemical Society: Washington, DC, 2002. 18. Forensic Sciences; National Institute of Justice: Washington, DC, April 2018. https://www.nij.gov/ topics/forensics/Pages/welcome.aspx (accessed Jan 7, 2019). 19. Crime Lab Operations; National Institute of Justice: Washington, DC, April 2018. https://nij.gov/topics/ forensics/lab-operations/Pages/welcome.aspx (accessed Jan 7, 2019). 20. Inside the Crime Lab: Forensic Chemistry Unit; Denver Police Crime Lab, March 26, 2017. https://www. youtube.com/watch?v=8AnRtHHwxYc (accessed Jan 7, 2019). 21. A Simplified Guide to Trace Evidence; National Forensic Science Technology Center: Largo, FL, September 2013. http://www.forensicsciencesimplified.org/trace/principles.html (accessed Jan 7, 2019). 22. Forensic Science: Types of Evidence; National Institute of Justice: Washington, DC, April 2018. https://nij. gov/topics/forensics/evidence/pages/welcome.aspx (accessed Jan 7, 2019). 23. Impact of Research and Development on Lab Efficiency; National Institute of Justice: Washington, DC, April 2018. https://nij.gov/multimedia/Pages/video-impact-of-research-on-crime-labs.aspx (accessed Jan 7, 2019). 24. Balko, R.; Carrington, T. Bad Science Puts Innocent People in Jail—and Keeps Them There. The Washington Post, March 21, 2018. https://www.washingtonpost.com/outlook/bad-science-putsinnocent-people-in-jail--and-keeps-them-there/2018/03/20/f1fffd08-263e-11e8-b79df3d931db7f68_story.html?noredirect=on&utm_term=.748dc926f78b (accessed Jan 9, 2019). 25. Forensic Science; Last Week Tonight with John Oliver (HBO), October 1, 2017. https://www.youtube. com/watch?v=ScmJvmzDcG0 (accessed April 4, 2018). 26. Strengthening Forensic Science in the United States: A Path Forward; Committee on Identifying the Needs of the Forensic Sciences Community, National Research Council, National Academy of Sciences: Washington, DC, August 2009. https://www.ncjrs.gov/pdffiles1/nij/grants/228091.pdf (accessed April 9, 2018). 27. Report to the President, Forensic Science in Criminal Courts: Ensuring Scientific Validity of Feature-Comparison Methods; President’s Council of Advisors on Science and Technology: Washington, DC, September 2016. https://obamawhitehouse.archives.gov/sites/default/files/microsites/ostp/PCAST/pcast_ forensic_science_report_final.pdf (accessed Jan 5, 2019). 28. Glass Refractive Index Determination. Forensic Science Communications; Federal Bureau of Investigation: Washington, DC, January 2005. https://archives.fbi.gov/archives/about-us/lab/forensic-sciencecommunications/fsc/jan2005/standards/2005standards9.htm (accessed Feb 22, 2019). 29. ABO-Rh Blood Typing with Synthetic Blood Value Kit; Carolina Biological: Burlington, NC, 2017. https://www.carolina.com/blood-typing/carolina-abo-rh-blood-typing-with-synthetic-blood-valuekit/700160.pr?question=ABO-Rh+blood+typing+kit (accessed Jan 8, 2019). 30. Blood Types Tutorial; The Biology Project, Human Biology, University of Arizona, Oct 23, 1997. http://www.biology.arizona.edu/human_bio/problem_sets/blood_types/genotypes.html (accessed May 7, 2018). 31. DNA Applications; DNA Learning Center, Cold Spring Harbor Laboratory: Cold Spring Harbor, NY, 2003. http://www.dnai.org/d/index.html (accessed Jan 8, 2019). 32. Brickman, P. The Case of the Druid Dracula; National Center for Case Study Teaching in Science, University at Buffalo, State University of New York: Buffalo, NY, 2006. http://sciencecases.lib.buffalo. edu/cs/collection/detail.asp?case_id=492&id=492 (accessed May 23, 2018).

90 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

33. Forensic PCR Investigation Amplification and Electrophoreses Kit with CarolinaBLU; Carolina Biological: Burlington, NC, 2017. https://www.carolina.com/pcr-kits/forensic-pcr-investigation-amplificationand-electrophoresis-kit-with-carolinablu-with-prepaid-coupon/211402.pr?question=PCR+kit (accessed Jan 8, 2019). 34. DNA Extraction; Genetic Science Learning Center, University of Utah: Salt Lake City, UT, 2006. https://learn.genetics.utah.edu/content/labs/extraction/ (accessed Jan 8, 2019). 35. Gel Electrophoresis; Genetic Science Learning Center, University of Utah: Salt Lake City, UT, 2006. https://learn.genetics.utah.edu/content/labs/gel/ (accessed Jan 8, 2019). 36. Boyd, A.; Larson, R. Burning Down the House: A Case Study in Forensic Instrumental Analysis; National Center for Case Study Teaching in Science, University at Buffalo, State University of New York: Buffalo, NY, 2005. http://sciencecases.lib.buffalo.edu/cs/collection/detail.asp?case_id=456&id=456 (accessed June 4, 2018). 37. Grann, D. Trial by Fire. The New Yorker, Sept 7, 2009. https://www.newyorker.com/magazine/2009/ 09/07/trial-by-fire (accessed Jan 8, 2019). 38. Death by Fire. Frontline (Public Broadcasting System). https://www.pbs.org/wgbh/frontline/film/deathby-fire/ (accessed April 19, 2018). 39. National Center for Case Study Teaching in Science. http://sciencecases.lib.buffalo.edu/cs/ (accessed Jan 5, 2019). 40. Shelton, D. The ‘CSI Effect’: Does It Really Exist? NIJ Journal 2008, 259, 221501NCJ. 41. Bellott, B. Typing Lipstick Smears from Crime Scene to Specific Brands. Presented at the 251st National Meeting and Exposition of the American Chemical Society, San Diego, CA, March 14, 2016. https://www. youtube.com/watch?v=lHF_UKiiFsk&list=PLLG7h7f PoH8L8o4Um_LZTS2lHxorDgHAH& index=6 (accessed Jan 8, 2019). 42. Gross, A.; Harris, K.; Kalden, G. The Effect of Luminol on Presumptive Tests and DNA Analysis Using the Polymerase Chain Reaction. J. Forensic Sci. 1999, 44, 837–840. 43. Joshi, M. A Systematic Evaluation of the Analysis of Drug Microcrystals Using Infrared Microspectroscopy; National Institute of Justice Award, Office of Justice Programs’ National Criminal Justice Reference Service: Washington, DC, August 2017. https://www.ncjrs.gov/pdffiles1/nij/grants/250547.pdf (accessed Jan 8, 2019). 44. Forensic Research News. ScienceDaily. https://www.sciencedaily.com/news/matter_energy/forensics (accessed March 26, 2018). 45. Forensic Science Research and Development Projects; National Institute of Justice. https://www.nij.gov/ topics/forensics/Pages/research-development-projects.aspx (accessed Jan 4, 2019).

91 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Chapter 6

On Utilizing Forensic Science To Motivate Students in a FirstSemester General Chemistry Laboratory Downloaded via COLUMBIA UNIV on October 22, 2019 at 08:15:37 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Stephen M. Testa* 27 Chemistry-Physics Building, 505 Rose Street, Department of Chemistry, University of Kentucky, Lexington, Kentucky 40506-0055, United States *E-mail: [email protected].

First-semester general chemistry students often have difficulty relating to the course material in a meaningful way. A consequence of this is that student interest in and enthusiasm for learning chemistry content can be low. One approach to increase student motivation is to present course content using a context that is more student-friendly, either by making the content more interesting or more relevant to their lives. To this end, teaching chemistry within the broader context of the forensic sciences has wide appeal. Students routinely encounter aspects of forensic science via news outlets and TV crime dramas. They are somewhat familiar with the topic and its importance. Chemistry instructors, for their part, are aware of the interdependence of forensic science techniques with chemistry and biochemistry. Using a forensic science context to teach chemistry, therefore, seems fitting. This report describes how we developed a general chemistry laboratory experiment that puts oxidation-reduction (redox) reactions within the framework of a fictional murder mystery. In this experiment, the perpetrator is identified by color-matching redox reactions and using permanganate as the oxidizing agent and DNA nucleotides as reducing agents. The four different nucleotides mimic the DNA of four different suspects. I will then summarize previous reports that utilized a forensic science framework to teach chemistry content, including for introductory college chemistry courses.

© 2019 American Chemical Society Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Introduction I am not a forensic scientist. I have not taught a forensic science course. I have not taken a forensic science course. I have not attended a forensic science–focused workshop or even read a forensic science textbook. I have certainly never participated in forensic science research. One could easily conclude, and I would not disagree with them, that I am wholly unqualified to be an educator in any forensic science field. Yet, once a semester, I (or a surrogate teaching assistant) don the requisite personal protective devices, march into the first-semester general chemistry laboratory, and, with the acting skills of a fourth-rate character on an episode of CSI, proceed to brief the students on a fictional unsolved murder. The victim, by all accounts, was rather unethical, intentionally harming the careers of his associates. The suspects all had justifiable reasons to be angry with the victim, and one of them, presumably, took things too far and committed… cold-blooded murder. I put forth the central question to the class: Who is the perpetrator? Everyone has their suspicions, but we, as scientists, do not deal in suspicions. We will use the tools of chemistry to analyze evidence, and we will take a scientific approach to identify the most likely perpetrator. The students are itching to get started, timidly conferring with their lab partners about the details of the very first step of the upcoming experiment. They look with anticipation at the chemicals, glassware, and DNA samples that they will use to experimentally confirm the identity of the killer. They eagerly participate in the experiment, are motivated, and are learning chemistry. They are not, however, learning specifically about forensic science. They are instead learning about oxidation-reduction (redox) reactions. In this experiment, the teaching of redox reactions has been framed within the context of a murder mystery. For the typical first-year college student (most particularly the vast majority of whom are not chemistry majors), the concepts behind redox reactions seem complex and complicated. To make matters worse, even when students can correctly identify a redox reaction, and distinguish it from a precipitation reaction, they often fail to see how this information can possibly be useful in a real-world context. Laboratory experiments that simply demonstrate the properties of redox reactions, which can be very interesting and exciting to those of us in the field (including chemistry majors), can be uninteresting to the bulk of the class, leaving many of the students underwhelmed. Therefore, to help motivate the students to learn about redox reactions, these reactions have been framed within the context of an entertaining mystery: a CSIthemed murder mystery. Teaching chemistry within the context of forensic science is a fairly common approach to motivating students. This approach can be used to teach virtually any chemistry concept at virtually any cognitive level, from elementary school up to graduate school. The details of our redox experiment have recently been published in the Journal of Chemical Education (1). Around the same time, I presented the development and application of this laboratory experiment at the 25th Biennial Conference on Chemical Education, in a symposium on “Teaching Chemistry in the Context of Forensic Science” (2). This chapter will briefly summarize the laboratory experiment and its relevance to chemical education. I will then summarize previous publications that have utilized a forensic science framework to teach chemistry content, with a specific emphasis on introductory college chemistry.

The Murder Mystery Redox Experiment The decision to develop this experiment was born out of the convergence of many factors. First, the academic coordinator of the first-year chemistry laboratories was “modernizing” the laboratory curriculum. Outdated experiments were retired, and experiments that introduced the students to 94 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

newer technologies, more advanced practices in teaching and learning chemistry, and experiments to which the students could more easily relate were added. Second, my research on colorimetrically distinguishing between different DNA sequences and structures (3) resulted in the development of redox reactions that were readily adaptable to an educational laboratory environment. Third, these redox reactions involve monomers of DNA (nucleotides), an important biomolecule with which students are already familiar. Fourth, the experiment is simple to conduct, relatively inexpensive, and the results are easy to detect and measure (for example, by eye or by camera phone). Fifth, in the firstsemester general chemistry curriculum at my university, redox reactions were not being conducted in the lab session, which was a missed opportunity for students to interact with this challenging topic. In the murder mystery, there are four suspects, each of whom provided witness statements that are summarized in the student laboratory manual. They also provided (simulated) DNA samples, one of which will match the DNA found on the handle of the murder weapon. The experiment was purposely written so that each of the four suspects could be the perpetrator. This allows the laboratory instructor (or director) to easily alter the identity of the perpetrator, either from semester to semester, class to class, or student to student. The experimental methods and results, the identification of the oxidation states of the manganese reaction products, and the limitations of the experiment as a tool for teaching redox reactions can be found in the relevant publication (1) and its accompanying supporting information. To summarize, students mix together two types of previously prepared solutions: one solution containing a DNA nucleotide in water and the other solution containing a buffer. Four different nucleotide solutions were prepared, using either adenosine monophosphate (AMP), thymidine monophosphate (TMP), cytidine monophosphate (CMP), or guanosine monophosphate (GMP). The four different DNA nucleotides (AMP, CMP, GMP, and TMP) were used to mimic DNA samples from four different suspects (Arnold, Cindy, Gary, and Tina, respectively). Each nucleotide solution was paired with a particular nucleotide-specific buffer solution. Once mixed, all of these solutions are colorless. The redox reactions are then initiated by adding the strong oxidizing agent potassium permanganate (KMnO4, Mn7+), which is a bold violet solution. The students then observe each redox reaction solution change color as the permanganate is differentially reduced by the different DNA nucleotides. This approach allows for a single oxidizing agent (permanganate, MnO4-, Mn7+) to be reduced to different extents (Mn6+, Mn5+, Mn4+, Mn3+, and Mn2+); each of these changes results in an easily identifiable solution color (green, yellow or blue, brown, pink, and colorless, respectively). A non-DNA sample, which stays violet, is also run as a negative control. When each of the six redox reactions are simultaneously conducted (using the four DNA suspect controls, one noDNA control, and one unknown DNA sample collected from the crime scene), the solution with the unknown sample will change color in parallel with only one of the suspects. The lab manual includes a table that correlates the solution colors with manganese oxidation states (for these particular nucleotide reactions in these particular buffers). Armed with this information, students assign manganese oxidation states to each reaction solution. By further consulting the laboratory manual, the students then identify which nucleotide was present in each reaction solution. Because the four nucleotides mimic the DNA from the four suspects, the perpetrator can now be identified. Each student records their data with a single camera phone image, writes a lab report, and answers redox reaction questions from the laboratory manual (1). By framing this experiment as an inquiry-based forensic science mystery, the students were intellectually engaged. They were interested in conducting the experiment, in analyzing the results, and in developing their conclusions. On a personal note, the intellectual energy that the students 95 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

displayed (for conducting redox reactions no less) was worth the time and effort put into developing and implementing the experiment.

Teaching Chemistry with Forensic Science A fundamental feature of forensic science is that, as a discipline, it often takes advantage of instrumentation and techniques that are traditionally taught in chemistry programs (including general chemistry courses). To state the obvious, one fundamental goal in general chemistry is for students to learn about the chemical and physical properties of matter, instrumentation, analytical techniques, and the identification of compounds, all of which are also important in forensic science. This overlap lends itself well to aligning chemistry content with multiple aspects of forensic science. Forensic Science in the Chemistry Curriculum An important benefit of using forensic science framing is that students, at nearly all grade and experience levels, can readily understand the context within which forensic scientists work. Students have grown up in a world where fictional and nonfictional crime scenarios are constantly in the media, either as news stories or as film and TV dramas. Students understand these scenarios: the general gist of what the problem is (Who Shot J.R.?) (4), how to solve it (gather and analyze evidence), and the conclusions (Kristin Shepard shot J.R.–and we can prove it). While their understanding of the technical details behind these scenarios can be limited, and sometimes even wrong, students still understand the overall issues at play. Who doesn’t like to solve a good mystery? Teaching chemistry within the framework of forensic science can occur at any grade level, from elementary school up through graduate school. This experience can be from within (for example, as part of a school’s normal curriculum) or from outside (for example, as part of outreach and professional development programs). There are many reports of these types of outreach programs in the literature (5–16), often developed and implemented by high school or college instructors and faculty, along with their associated students. Generally, activities targeting elementary (7, 14, 17), middle (6, 9, 11, 13, 15), and high school students (15, 18–27) focus on stimulating interest in STEM disciplines, while also increasing scientific literacy. College students, usually those who are chemistry, biochemistry, or forensic science majors (but not exclusively so), can be exposed to the forensic sciences at any point in their degree programs, from science electives for non-science majors to general chemistry to analytical and instrumental analysis courses. At the lower course levels, this framework is frequently used to stimulate student interest, increase scientific literacy, and introduce students to inquiry-based, “authentic” scientific experiences. Presumably, the content in introductory college courses is at a higher cognitive level than what is targeted at precollege students. Forensic science activities could be developed to teach any chemistry topic, including those in inorganic, organic, analytical, instrumental, physical, and biochemistry. Quite a few activities have already been implemented. Examples at the general chemistry level include (but are not limited to) lessons on chemical properties (28–30), physical properties (29, 31, 32), density (5, 33), osmolarity (34), equilibrium (30), acid-base reactions (30), redox reactions (1, 35), catalysis (35), and intermolecular forces (36, 37). These activities have been used in a wide variety of educational settings, including seminar-based courses, lecture-only courses, laboratory courses, professional development opportunities, and workshop environments. While it seems clear that students can be taught and can learn chemistry using forensic science as a framework, it is not clear what material is appropriate for any given course and any given student, 96 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

including for students in college-level general chemistry courses. Vygotsky’s social theory of learning (38) suggests that effective teaching occurs in a student’s zone of proximal development, which is the cognitive level where the material is not too easy (which induces boredom and disinterest) or too difficult (which induces hopelessness and disinterest). Making it all the more problematic is that this zone is student-specific, sometimes not even overlapping between different students taking the same course. This can be particularly frustrating for instructors and for students in lower-level science courses like general chemistry. There is no definitive, agreed-upon solution to this problem, and potential remedies will not be addressed in this chapter. However, there are encouraging reports from the literature that describe situations where advanced chemistry content was delivered effectively to students in lower-level college courses (35, 39–43) and, perhaps surprisingly, to high school students (19, 20, 22, 23, 26, 27). In fact, bringing more advanced topics and methods into more basic courses (and vice versa) can have considerable educational value. Therefore, it does not seem to be constructive to try to make cognitive-level, course-level, or degree-level distinctions or recommendations for what chemistry content, or what chemistry or forensic science approaches, are appropriate for general chemistry students. It should be left up to the instructor to make decisions regarding the appropriateness of the content for their students. Educational Approaches to Delivering Chemistry Content In-class active learning methods can be more effective than in-class traditional methods, such as simply listening to an instructor speak. Reference (44) provides more information regarding active learning as it relates to chemical education. Because of the very nature of forensic science as a handson, open-ended, problem-solving discipline, teaching chemistry using a forensic science framework, particularly with regard to the laboratory environment, usually falls squarely in the active learning spectrum. These active learning methods, presumably, can help students develop an enhanced understanding of chemistry, which can increase their higher-order thinking skills (in both STEM and non-STEM disciplines). This approach can lead to students who are more independent thinkers and who are more effective problem-solvers. There are a number of educational approaches that have been used to deliver chemistry content within a forensic science framework, most of which encompass some degree of active learning. Notwithstanding the fact that it is often hard to distinguish between these different approaches, they seem to be adequately represented in the chemical education literature. Examples include problembased learning (which applies to most of the activities referenced), case-based learning (42, 45–56), game-based learning (57), role-playing (58–62), storytelling (10, 14, 28, 63–68), and guided inquiry (31, 39, 46, 69, 70), among others. The decision on which approach is most appropriate is dependent on a number of factors: the cognitive level of the students, the teaching environment, the available time and resources, the desired learning outcomes, and the skills and desires of the instructor. Forensic Science Scenarios There is no shortage of forensic science scenarios created for the express purpose of teaching chemistry, most often presented in a laboratory setting. For those interested in using previously published scenarios or creating their own unique scenarios, it is instructive to know what has already been developed and implemented in this field. Scenarios typically revolve around solving a crime. For simplicity, scenarios will be listed here only in the most basic terms. Scenarios involving primarily living or dead materials include the identification of a perpetrator in a murder mystery (1, 40, 50, 57, 60, 71, 72), cause of death and 97 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

foul play (45, 49, 73), babies switched at birth in a hospital (27), stolen parakeets (27), canine genders (74), race horse pedigrees (56), paternity results (75), missing persons (75), and a suspicious drowning (9). Scenarios involving primarily the identification of non-living materials include arson (28, 36, 39, 76–82), chemical spills (11), authenticity of artwork (9, 42), industrial poisoning (83), written forgeries (84, 85), counterfeiting (59), sabotage (11, 23, 86), a pharmacy break-in (69), crime location (60), gun discharge (87), theft (7, 11, 86), spying (35), kidnapping (88), a car accident (53), laboratory misconduct (47), and an escape-room scenario (57). It is quite common for educators to utilize more than one scenario in their published reports (7, 50, 82, 89). It is important to note that these simple scenario descriptions do not do justice to the clever and unique ways that many of these scenarios were developed and delivered to the students. The reader is strongly encouraged to review the individual publications. One of many notable examples is the Chocolate Science Investigation (CSI) scenario (23). This industrial sabotage scenario takes place in the context of two rival chocolate companies. Students conduct a variety of experimental analyses in order to detect and identify counterfeit chocolate bars, poison, the ink on a ransom note, blood type (simulated), and the DNA (simulated) from different cocoa beans. This scenario takes advantage of the analytical properties of high performance liquid chromatography (HPLC), thin-layer chromatography, gel electrophoresis, polymerase chain reaction (simulated), and color analysis. During the scenario, students are distributed into three role-playing groups of investigators: one for the defense, one for the plaintiff, and one neutral party. Each group has to gather evidence from a mock crime scene, interview suspects, analyze evidence, make a legal case based on the evidence, and then present their case to a judge. While this scenario was developed for a STEMbased summer camp for high school students, its components (in whole or in part) are readily transferable to the general chemistry college curriculum. Finally, some educators take advantage of these scenarios to teach students about conducting science as part of a team (26, 70, 90), which is often necessary when solving real criminal cases. The ability to be an effective team player is an important, but often overlooked skill, one that employers universally value. Forensic Science Materials Forensic science scenarios in chemistry often involve identifying and comparing materials and using a known material as a standard or reference for identifying unknown materials. Scenarios that focus primarily on living materials (real or simulated) have been used to identify biopsy material (59), the presence of blood (40, 69, 91–93), blood typing (23, 27, 72, 94, 95), blood samples from suspects (7, 21, 25, 49, 52, 57, 73, 96), bone (60, 70), brain matter (96), DNA (1, 56, 69, 74, 75, 86, 88, 97–104), food (50, 70), hair (52, 72, 74, 88), saliva (50, 74), serum samples (93, 105), teeth (88), and urine (52, 58, 96, 104, 106–108). Scenarios that focus primarily on nonliving materials (real or simulated) have been used to detect or identify alcohol (105, 109, 110), drugs (10, 29, 34, 43, 73, 102, 107, 111–114), dyes (32), fibers (26, 115–117), fire accelerants (28, 36, 39, 76–82), glass (33, 118), gun residue (4, 87, 119), inks (6, 39, 84, 85, 120), lipstick (13, 116), paints (9, 42, 53, 121, 122), papers (24, 32), plastics (50), poisons and herbs (10, 23, 45, 48, 52, 57, 61, 62, 67, 71, 123), rubber (33, 118), metal (6, 42, 52, 108, 122, 124), soil (31, 60, 125), and water (126). Materials that do not readily fit into one of the above categories, but have been used in forensic scenarios, are bite marks (21), blood spatter (21, 25), fingerprints patterns and chemical composition 98 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

(37, 55, 97, 98, 127–131), footprints (21, 127), handwriting comparisons (21, 25, 72), and lip prints (13, 21). In fact, most educators utilize more than one type of material in their activities (7, 20, 21, 40, 50, 72, 82, 89, 96). Analytical Methods and Instrumentation Many of the tools of the forensic science trade are first introduced to students in chemistry courses. Chemistry is a discipline that relies heavily on methods and instruments than can be used for analyzing, detecting, and identifying materials, often the key to solving certain types of crime. Therefore, it is fairly intuitive to develop crime-based forensic science scenarios to teach students about analytical methods and instrumentation. In addition, this approach can accurately simulate an authentic crime lab experience (27, 47). Many of the methods presented below were developed for use in upper-level analytical chemistry, instrumental methods of analysis, and biochemistry college courses. It is worth mentioning that some upper-level chemistry content can be presented to students in lower-level courses, provided the content is presented at the appropriate cognitive level. One example is teaching luminescence to middle school students. These students will not have the background to understand how luminescence occurs or how its brightness and its color are scientifically measured. However, they can be introduced to the topic using fireflies, jellyfish, or a forensic science scenario that involves the discovery of a crime scene using luminol (91). Observing visible color changes is a powerful and memorable tool that can be effectively used at all educational levels. Therefore, the instructor should carefully choose which methods to use and at what cognitive level the methods should be taught. Teaching methods for separating chemical components is the focal point of many forensic science scenarios. The most prevalent of these methods is arguably chromatography, which can take many forms. Chromatography methods include gas chromatography (36, 77, 127), HPLC (23, 73, 96, 111, 127), paper chromatography (24, 32, 39), and thin-layer chromatography (19, 23, 29, 43, 49, 85, 107, 116, 120, 132). More biologically oriented scenarios exploit the separation properties of agarose gel (23, 56, 69, 74, 75, 86), polyacrylamide gel (98), and capillary gel (96) electrophoresis. Methods for detecting and identifying chemical compounds are also commonly used in forensic science scenarios. The most prevalent of these is arguably spectroscopy, for which there are a large number of different types. These include atomic absorption (4, 50, 58, 96), gas chromatography and mass spectrometry (39, 43, 50, 58, 76, 106, 125, 129, 130, 133–136), fluorescence (58, 96, 124, 131, 137), inductively coupled plasma mass spectrometry (62, 125, 135), infrared (19, 26, 50, 58, 96, 104, 109, 116, 117, 136, 138), laser-induced breakdown atomic emission (87, 122), HPLC mass spectrometry (114), mass (106, 139), nuclear magnetic resonance (26, 108, 139, 140), Raman (50), reflectance (50, 121), ultraviolet–visible (19, 50, 58, 73, 75), and X-ray (40, 42, 62, 81, 96, 137). Several scenarios take advantage of low and high power microscopy techniques. These include atomic force (62), optical (50, 116, 141), and scanning microscopy (58, 62, 142). Other useful methods of analysis include differential scanning calorimetry (115), isothermal titration calorimetry (62), osmometry (34), potentiometry (50), and thermogravimetric analysis (115). Specific to more biochemically-oriented courses are scenarios that utilize immunological assays (50, 81, 94, 95) and polymerase chain reactions (23, 27, 45, 56, 74, 75, 86, 98, 99). Importantly, many forensic science scenarios do not require complex instrumentation. They instead tend to exploit the differences between chemical properties of compounds (e.g. flammability, reactivity) or the differences between physical properties of compounds (e.g. solubility, density, 99 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

osmotic strength) for sample identification (21, 50). One example is a drug test, perhaps conducted by a law enforcement officer, in which a chemical reaction turns a white powder to a blue liquid if the white powder is a particular illicit drug.

Summary Teaching general chemistry, or any level of chemistry, within a forensic science context can help lessen some of the negative feelings that many students have toward the discipline. Like other natural sciences (unfortunately referred to as hard sciences), chemistry has a relatively high degree of rigor associated with its content. To facilitate meaningful understanding (143, 144) of chemistry subjects (44), one needs to be somewhat skilled in mathematics, lower-ordered thinking skills (i.e., remember, understand), higher-ordered thinking skills (i.e., apply, analyze, evaluate, and perhaps create), and laboratory protocol. For many students, the confluence of these necessary skills can lead to cognitive overload and disinterest. Chemistry educators have long pondered how to best teach these skills, a discussion that will continue as long as our understanding of the nature of learning increases. To make matters worse, the conceptual nature of chemistry can act like a smokescreen, preventing students from finding personal relevance in the content. For this reason, students are rarely excited when they go into a chemistry lab, where they will mix a few things, observe a few things, and then write an imminently forgettable report. This is especially true for many students in lower-level college chemistry courses, as relatively few of them are interested in a career in chemistry. Using a forensic science framework to teach chemistry can help ameliorate some of these issues. This chapter has presented many examples, in many contexts, that demonstrate how our peers are successfully using forensic science to teach chemistry in ways that instill a sense of relevance, excitement, and motivation in their students. I have had the good fortune to teach students in junior- and senior-level biochemistry courses that have previously conducted the murder mystery redox experiment (1) as part of their general chemistry lab course. When asked if they remember the lab, the students usually do. Many even remember the scenario, who the killer was in their individual scenario, the changing colors during the chemical reaction, that colors were different because of the different (oxidation) states of a transition metal, and that they were conducting redox reactions. The scenario seems to act as a cognitive cue, allowing students to access information about redox reactions that had long ago been stored in their brains. I attribute their ability to remember much of this information, at least in part, to the memorable forensic science scenario that was at the core of the experiment.

References 1.

2.

3.

Testa, S. M.; Selegue, J. P.; French, A.; Criswell, B. Permanganate Oxidation of DNA Nucleotides: An Introductory Redox Laboratory Framed as a Murder Mystery. J. Chem. Educ. 2018, 95, 1840–1847. Testa, S. M.; Selegue, J. P.; French, A.; Criswell, B. Permanganate Oxidation of DNA Nucleotides: An Introductory Redox Laboratory Framed as a Forensic Investigation. In Proceedings of the 25th Biennial Conference on Chemical Education, South Bend, Indiana, July 29–Aug 2, 2018. Thomas, E. M.; Testa, S. M. The colorimetric determination of selectively cleaved adenosines and guanosines in DNA oligomers using bicinchoninic acid and copper. J. Biol. Inorg. Chem. 2017, 22, 31–46. 100 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

4. 5.

6. 7.

8.

9. 10. 11.

12.

13.

14.

15.

16. 17. 18. 19. 20.

Dahl, D. B.; Lott, P. F. A Forensic Laboratory Experiment: Who Shot J.R.? J. Chem. Educ. 1991, 68, 1025–1026. Heinze, K.; Allen, J.; Jacobsen, E. Encouraging Tomorrows Chemists - University Outreach Program Bringing Hands-On Experiments to Local Students. J. Chem. Educ. 1995, 72, 167–169. Vandoren, J.; Nestor, L.; Knighton, W. Engaging Students in the Action of Chemistry - An Effective, Fun, and Inexpensive Outreach Program. J. Chem. Educ. 1997, 74, 1178–1179. Koehler, B. G.; Park, L. Y.; Kaplan, L. J. Science for Kids Outreach Programs: College Students Teaching Science to Elementary School Students and Their Parents. J. Chem. Educ. 1999, 76, 1505–1509. Charkoudian, L.; Heymann, J.; Adler, M.; Haas, K.; Mies, K.; Bonk, J. Forensics as a Gateway: Promoting Undergraduate Interest in Science and Graduate Student Professional Development through a First-Year Seminar Course. J. Chem. Educ. 2008, 85, 807–812. Harmon, K.; Miller, L.; Millard, J. Crime Scene Investigation in the Art World: The Case of the Missing Masterpiece. J. Chem. Educ. 2009, 86, 817–819. André, J. P. Opera and Poison: A Secret and Enjoyable Approach to Teaching and Learning Chemistry. J. Chem. Educ. 2013, 90, 352–357. Meyer, A. F.; Knutson, C. M.; Finkenstaedt-Quinn, S. A.; Gruba, S. M.; Meyer, B. M.; Thompson, J. W.; Maurer-Jones, M. A.; Halderman, S.; Tillman, A. S.; Destefano, L.; Haynes, C. L. Activities for Middle School Students to Sleuth a Chemistry “Whodunit” and Investigate the Scientific Method. J. Chem. Educ. 2014, 91, 410–413. Kanu, A. B.; Pajski, M.; Hartman, M.; Kimaru, I.; Marine, S. Exploring Perspectives and Identifying Potential Challenges Encountered with Crime Scene Investigations When Developing Chemistry Curricula. J. Chem. Educ. 2015, 92, 1353–1358. Levine, M.; Serio, N.; Radaram, B.; Chaudhuri, S.; Talbert, W. Addressing the STEM Gender Gap by Designing and Implementing an Educational Outreach Chemistry Camp for Middle School Girls. J. Chem. Educ. 2015, 92, 1639–1644. Morais, C. Storytelling with Chemistry and Related Hands-On Activities: Informal Learning Experiences to Prevent “Chemophobia” and Promote Young Children’s Scientific Literacy. J. Chem. Educ. 2015, 92, 58–65. Griep, M. A.; Mikasen, M. L. Close Encounters with Creative Chemical Thinking: An Outreach Presentation Using Movie Clips About the Elemental Composition of Aliens and Extraterrestrial Minerals. Educación Química 2016, 27, 154–162. Thompson, R. Q. Forensic Chemistry and Its Flip Side. J. Chem. Educ. 2016, 93, 1677–1678. Frey, C.; Mikasen, M.; Griep, M. A. Put Some Movie Wow! in Your Chemistry Teaching. J. Chem. Educ. 2012, 89, 1138–1143. Horan, J.; Rothchild, R. Introduction to Forensic Science: An advanced elective for high schools. J. Chem. Educ. 1980, 57, 599–600. Siggins, B. A.; Hendricks, B. W. Forensic Drug Chemistry: A Cooperative Program. J. Chem. Educ. 1993, 70, 312–314. Koppang, M.; Webb, K.; Srinivasan, R. Not So Late Night Chemistry with USD. J. Chem. Educ. 1994, 71, 929–931.

101 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

21. Guzzetti, B. Thinking like a Forensic Scientist: Learning with Academic and Everyday Texts. J. Adolesc. Adult Lit. 2009, 53, 192–203. 22. Ahrenkiel, L.; Worm-Leonhard, M. Offering a Forensic Science Camp to Introduce and Engage High School Students in Interdisciplinary Science Topics. J. Chem. Educ. 2014, 91, 340–344. 23. Marle, P. D.; Decker, L.; Taylor, V.; Fitzpatrick, K.; Khaliqi, D.; Owens, J. E.; Henry, R. M. CSI-Chocolate Science Investigation and the Case of the Recipe Rip-Off: Using an Extended Problem-Based Scenario to Enhance High School Students’ Science Engagement. J. Chem. Educ. 2014, 91, 345–350. 24. Ravgiala, R.; Weisburd, S.; Sleeper, R.; Martinez, A.; Rozkiewicz, D.; Whitesides, G.; Hollar, K. Using Paper-Based Diagnostics with High School Students To Model Forensic Investigation and Colorimetric Analysis. J. Chem. Educ. 2014, 91, 107–111. 25. Williams, K. Engaging High School Students with Forensic Science: Bringing CSI into the Classroom to Teach Multiple Disciplines [Pipelining: Attractive Programs for Women]. IEEE Women in Engineering Magazine 2018, 12, 43–45. 26. Long, G. A. Simulation of a Forensic Chemistry Problem: A Multidisciplinary Project for Secondary School Chemistry Students. J. Chem. Educ. 1995, 72, 803–804. 27. Duncan, K.; Daly-Engel, T. Using Forensic Science Problems as Teaching Tools. The Science Teacher 2006, 73, 38–43. 28. Waddell, T. G.; Rybolt, T. R. The Chemical Adventures of Sherlock Holmes: The Baker Street Burning. J. Chem. Educ. 1998, 75, 484–486. 29. Hasan, S.; Bromfield-Lee, D.; Oliver-Hoyo, M. T.; Cintron-Maldonado, J. A. Using Laboratory Chemicals to Imitate Illicit Drugs in a Forensic Chemistry Activity. J. Chem. Educ. 2008, 85, 813–816. 30. Eilks, I.; Gulacar, O.; Sandoval, J. Exploring the Mysterious Substances, X and Y: Challenging Students’ Thinking on Acid–Base Chemistry and Chemical Equilibrium. J. Chem. Educ. 2018, 95, 601–604. 31. Shmaefsky, B. R. Forensic Analysis Demonstration via Hawaii Five-O. J. Coll. Sci. Teach. 2006, 36, 14–16. 32. Smith, M. J.; Vale, I. C.; Gray, F. M. Identification of Paper by Stationary Phase Performance. J. Chem. Educ. 2014, 9, 1679–1683. 33. Saccocio, L. A.; Carroll, M. K. Density Determination by Water Displacement and Flotation: An Introductory Experiment in Forensic Chemistry. J. Chem. Educ. 2006, 83, 1187–1189. 34. Rothchild, R. A Forensic Science Laboratory Experiment: Molecular Weights of Drugs and Diluents by Osmometry. J. Chem. Educ. 1982, 59, 672–673. 35. Macrakis, K.; Bell, E. K.; Perry, D. L.; Sweeder, R. D. Invisible Ink Revealed: Concept, Context, and Chemical Principles of “Cold War” Writing. J. Chem. Educ. 2012, 89, 529–532. 36. Wedvik, J.; McManaman, C.; Anderson, J.; Carroll, M. Intermolecular Forces in Introductory Chemistry Studied by Gas Chromatography, Computer Models, and Viscometry. J. Chem. Educ. 1998, 75, 885–888. 37. Friesen, J. B. Forensic Chemistry: The Revelation of Latent Fingerprints. J. Chem. Educ. 2015, 92, 497–504. 38. Vygotskiĭ, L. S.; Cole, M. Mind in Society: The Development of Higher Psychological Processes; Harvard University Press: Cambridge, MA, 1978. 102 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

39. Maurer, M.; Bukowski, M.; Menachery, M.; Zatorsky, A. Inquiry-Based Arson Investigation for General Chemistry Using GC-MS. J. Chem. Educ. 2010, 87, 311–313. 40. Basso, A.; Chiorri, C.; Bracco, F.; Carnasciali, M. M.; Alloisio, M.; Grotti, M. Improving the Interest of High-School Students Toward Chemistry by Crime Scene Investigation. Chem. Educ. Res. Pract. 2018, 19, 558–566. 41. Kaplan, L. J. Chemistry and Crime: From Sherlock Holmes to Modern Forensic Science: A Science Course for Non-Science Majors. Crime Laboratory Digest 1992, 4, 107–132. 42. Nivens, D. A.; Padgett, C. W.; Chase, J. M.; Verges, K. J.; Jamieson, D. S. Art, Meet Chemistry; Chemistry, Meet Art: Case Studies, Current Literature, and Instrumental Methods Combined to Create a Hands-On Experience for Nonmajors and Instrumental Analysis Students. J. Chem. Educ. 2010, 87, 1089–1093. 43. Szalay, P. S.; Zook-Gerdau, L. A.; Schurter, E. J. A Multi-Technique Forensic Experiment for a Nonscience-Major Chemistry Course. J. Chem. Educ. 2011, 88, 1419–1421. 44. Devetak, I.; Glazar, S. A. Learning with Understanding in the Chemistry Classroom; Devetak, I., Glažar, S. A., Eds.; Springer Netherlands: Dordrecht, 2014. 45. Childs-Disney, J. L.; Kauffmann, A. D.; Poplawski, S. G.; Lysiak, D. R.; Stewart, R. J.; Arcadi, J. K.; Dinan, F. J. A Metabolic Murder Mystery: A Case-Based Experiment for the Undergraduate Biochemistry Laboratory. J. Chem. Educ. 2010, 87, 1110–1112. 46. Cresswell, S. L.; Loughlin, W. A. A Case-Based Scenario with Interdisciplinary Guided-Inquiry in Chemistry and Biology: Experiences of First Year Forensic Science Students. J. Chem. Educ. 2017, 94, 1074–1082. 47. Contakes, S. M. Misconduct at the Lab? A Performance Task Case Study for Teaching Data Analysis and Critical Thinking. J. Chem. Educ. 2016, 93, 314–317. 48. Labianca, D. A. Science and the Nonscience Major: Addressing the Fear Factor in the Chemical Arena Using Forensic Science. J. Coll. Sci. Teach. 2007, 37, 28–33. 49. Konaklieva, M. Accidental Drowning or Foul Play?: A Case Study in Organic Chemistry. J. Coll. Sci. Teach. 2004, 34, 10–13. 50. Thompson, R. Q.; Edmiston, P. L. Ohio Crime Solvers. Anal. Chem. 2001, 73, 678A–684A. 51. Summerfield, S.; Overton, T.; Belt, S. Problem-Solving Case Studies. Anal. Chem. 2003, 75, 181A–182A. 52. Belt, S.; Evans, E.; McCreedy, T.; Overton, T.; Summerfield, S. A Problem Based Learning Approach to Analytical and Applied Chemistry. Univ. Chem. Educ. 2002, 6, 65–72. 53. Eyring, M. B. Paint and Criminal Forensic Science. Paint & Coatings Industry 2008, 24, 72–74. 54. Ferguson, A.; Dronsfield, A. Forensic Science and the Case of Dr. Mario Jascalevich. Educ. Chem. 2011, 48, 84–86. 55. Kimbrough, D.; Delorenzo, R. Solving the Mystery of Fading Fingerprints with London Dispersion Forces. J. Chem. Educ. 1998, 75, 1300–1301. 56. Millard, J.; Chuang, E.; Lucas, J.; Nagy, E.; Davis, G. Case-Study Investigation of Equine Maternity via PCR-RFLP: A Biochemistry Laboratory Experiment. J. Chem. Educ. 2013, 90, 1518–1521. 57. Ferreiro-González, M.; Amores-Arrocha, A.; Espada-Bellido, E.; Aliaño-Gonzalez, M. J.; Vázquez-Espinosa, M.; González-de-Peredo, A. V.; Sancho-Galán, P.; Álvarez-Saura, J. Á.;

103 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

58.

59.

60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 70.

71. 72. 73. 74.

75.

Barbero, G. F.; Cejudo-Bastante, C. Escape ClassRoom: Can You Solve a Crime Using the Analytical Process? J. Chem. Educ. 2019, 96, 267–273. Grannas, A.; Lagalante, A. So These Numbers Really Mean Something? A Role Playing Scenario-Based Approach to the Undergraduate Instrumental Analysis Laboratory. J. Chem. Educ. 2010, 87, 416–418. Maciejowska, I.; Wietecha-Posłuszny, R.; Woźniakiewicz, M.; Kościelniak, P. Application of Case Study and Role-Playing in Forensic Chemistry and Analytical Chemistry Education: Students’, Graduates’ and Teachers’ Points of View. In Learning with Understanding in the Chemistry Classroom; Devetak, I., Glažar, S. A., Eds.; Springer Netherlands: Dordrecht, 2014; pp 287–301. Dinan, F.; Szczepankiewic, S.; Carnahan, M.; Colvin, M. The Analysis of a Murder, a Case Study. J. Chem. Educ. 2007, 84, 617–618. Tarr, M. Solving a Mock Arsenic-Poisoning Case Using Atomic Spectroscopy. J. Chem. Educ. 2001, 78, 61–62. Kloepper, K. D. Bringing in the Bard: Shakespearean Plays as Context for Instrumental Analysis Projects. J. Chem. Educ. 2015, 92, 79–85. Folino, D. A. Stories and Anecdotes in the Chemistry Classroom. J. Chem. Educ. 2001, 78, 1615–1618. Griep, M. A.; Mikasen, M. L. Based on a True Story: Using Movies as Source Material for General Chemistry Reports. J. Chem. Educ. 2005, 82, 1501–1503. Kitson, T. Telling the Stories of Chemistry. J. Chem. Educ. 2005, 82, 1504. Shaw, K. The Chemical Adventures of Sherlock Holmes: The Serpentine Remains. J. Chem. Educ. 2008, 85, 507–513. Harper-Leatherman, A. S.; Miecznikowski, J. R. O True Apothecary: How Forensic Science Helps Solve a Classic Crime. J. Chem. Educ. 2012, 89, 629–635. Waddell, T.; Rybolt, T. Prologue to The Chemical Adventures of Sherlock Holmes. J. Chem. Educ. 2011, 88, 370–371. Cresswell, S. L.; Loughlin, W. A. An Interdisciplinary Guided Inquiry Laboratory for First Year Undergraduate Forensic Science Students. J. Chem. Educ. 2015, 92, 1730–1735. Fakayode, S. O.; Mayes, J. P.; Kanipes, M. I.; Johnson, D.; Cuthbertson, E. L. Promoting Student Learning in Criminal Justice, STEM, and Forensic Science: Aggie Sleuth Initiative (AggieSI)-Guided Inquiry Learning Experience. J. Crim. Just. Educ. 2017, 28, 192–206. Grove, N.; Bretz, S. L. Sherlock Holmes and the Case of the Raven and the Ambassador’s Wife: An Inquiry-Based Murder Mystery. J. Chem. Educ. 2005, 82, 1532–1533. Sandage, B. J. Who Killed Myra Mains? Sci. Teach. 2002, 69, 28–32. Beussman, D. J. The Mysterious Death: An HPLC Lab Experiment. An Undergraduate Forensic Lab. J. Chem. Educ. 2007, 84, 1809–1812. Carson, T. M.; Bradley, S. Q.; Fekete, B. L.; Millard, J. T.; LaRiviere, F. J. Forensic Analysis of Canine DNA Samples in the Undergraduate Biochemistry Laboratory. J. Chem. Educ. 2009, 86, 376–378. Elkins, K. M.; Kadunc, R. E. An Undergraduate Laboratory Experiment for Upper-Level Forensic Science, Biochemistry, or Molecular Biology Courses: Human DNA Amplification

104 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

76.

77.

78. 79. 80.

81. 82.

83. 84.

85. 86. 87.

88. 89. 90. 91. 92.

Using STR Single Locus Primers by Real-Time PCR with SYBR Green Detection. J. Chem. Educ. 2012, 89, 784–790. Anzivino, B.; Tilley, L. J.; Ingalls, L. R.; Hall, A. B.; Drugan, J. E. Got a Match? Ion Extraction GC-MS Characterization of Accelerants Adsorbed in Charcoal Using Negative Pressure Dynamic Headspace Concentration. J. Chem. Educ. 2009, 86, 55–59. Elderd, D.; Kildahl, N.; Berka, L. Experiments for Modern Introductory Chemistry: Identification of Arson Accelerants by Gas Chromatography. J. Chem. Educ. 1996, 73, 675–677. Olesen, B. An Economical Method for Static Headspace Enrichment for Arson Analysis. J. Chem. Educ. 2010, 87, 314–315. Sodeman, D.; Lillard, S. Who Set the Fire? Determination of Arson Accelerants by GC-MS in an Instrumental Methods Course. J. Chem. Educ. 2001, 78, 1228–1230. Warnke, M. M.; Erickson, A. E.; Smith, E. T. Simplex Optimization of Headspace-Enrichment Conditions of Residual Petroleum Distillates Used by Arsonists. J. Chem. Educ. 2005, 82, 1082–1085. Briner, R. C. Forensic Chemistry. J. Chem. Educ. 1982, 59, 41–44. Nienhouse, E. J. Chemistry and Crime: A Laboratory-Based Forensic Science Techniques Course as an Alternative to a Natural Science Requirement. J. Chem. Educ. 1985, 62, 1047–1049. Zdravkovich, V.; Cunniff, P. Spectroscopic Sleuthing: An Introduction to Forensic Science. The Science Teacher 1991, 58, 21–23. Wietecha, R.; Mania, J.; Wozniakiewicz, M.; Madej, K.; Maciejowska, I. Is This Last Will Fraudulent? – An Experiment for the Forensic Chemistry Laboratory Course for Undergraduate Students. Forensic Sci. Int. 2003, 136, 181–182. Quigley, M. N.; Qi, H. A Chemistry Whodunit: Forensic Examination of Pen Inks. J. Chem. Educ. 1991, 68, 596–597. Jackson, D. D.; Abbey, C. S.; Nugent, D. DNA Profiling of the D1S80 Locus: A Forensic Analysis for the Undergraduate Biochemistry Laboratory. J. Chem. Educ. 2006, 83, 774–776. Dockery, C. R.; Turner, J.; Rosenberg, M. B.; Kammerdiener, K.; Mungai, S. W. Gunshot Residue Analysis in the Undergraduate Laboratory Using Toy Cap Guns. Spectrosc. Lett. 2010, 43, 534–538. Millard, J. T.; Pilon, A. M. Identification of Forensic Samples via Mitochondrial DNA in the Undergraduate Biochemistry Laboratory. J. Chem. Educ. 2003, 80, 444–445. Clark, M. J.; Keegel, J. F. Chemistry in the Crime Lab - A Forensic Science Course. J. Chem. Educ. 1977, 54, 38–40. Koscielniak, P. An Example of Educational Collaboration Between a Forensic Institute and a University in the Field of Forensic Chemistry. Forensic Sci. Int. 2003, 136, 185. Burke, B. A.; Golestaneh, K.; Samson, H. Luminosity, My Dear Watson, Luminosity! – Or, Are Those Bloodstains? J. Chem. Educ. 1999, 76, 65–67. Pinkus, J. L.; Goldman, L. S. A Benzidine Rearrangement and the Detection of Trace Quantities of Blood. A Laboratory Experiment in Criminalistics. J. Chem. Educ. 1977, 54, 380–381.

105 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

93. Sheehan, F. X.; Kobilinsky, L. Human Blood Identification: A Forensic Science Approach. J. Chem. Educ. 1984, 61, 542–546. 94. Kobilinsky, L.; Sheehan, F. X. Human Blood Typing: A Forensic Science Approach – Part I: Background. J. Chem. Educ. 1988, 65, 531–533. 95. Kobilinsky, L.; Sheehan, F. X. Human Blood Typing: A Forensio Science Approach – Part II: Experiments. J. Chem. Educ. 1988, 65, 624–626. 96. Frederick, K. Using Forensic Science to Teach Method Development in the Undergraduate Analytical Laboratory. Anal. Bioanal. Chem. 2013, 405, 5623–5626. 97. Palladino, M. A. A DNA Fingerprinting Simulation Laboratory for Biology Students: HandsOn Experimentation to Solve a Mock Forensic Problem. Am. Biol. Teach. 2001, 63, 596–605. 98. McNamara-Schroeder, K.; Olonan, C.; Chu, S.; Montoya, M. C.; Alviri, M.; Ginty, S.; Love, J. J. DNA Fingerprint Analysis of Three Short Tandem Repeat (STR) Loci for Biochemistry and Forensic Science Laboratory Courses. Biochem. Mol. Biol. Educ. 2006, 34, 378–383. 99. Elkins, K. M. Designing Polymerase Chain Reaction (PCR) Primer Multiplexes in the Forensic Laboratory. J. Chem. Educ. 2011, 88, 1422–1427. 100. Millard, J. T. DNA Profiling of Convicted Offender Samples for the Combined DNA Index System. J. Chem. Educ. 2011, 88, 1385–1388. 101. Stacey, L. R. Development of a Bi-Disciplinary Course in Forensic Science. J. Micro. Bio. Educ. 2013, 14, 250–251. 102. Coticone, S.; Van Houten, L. DNA, Drugs, and Detectives: An Interdisciplinary Special Topics Course for Undergraduate Students in Forensic Science. J. Coll. Sci. Teach. 2015, 45, 24–29. 103. Arnaud, C. Thirty Years of DNA Forensics Analysis Techniques Have Become Faster, More Sensitive, and More User-Friendly. Chem. Eng. News 2017, 95, 16–20. 104. Mistek, E.; Lednev, I. FT-IR Spectroscopy for Identification of Biological Stains for Forensic Purposes. Spectroscopy 2018, 33, 8–28. 105. Zabzdyr, J. L.; Lillard, S. J. Was the Driver Drunk? An Instrumental Methods Experiment for the Determination of Blood Alcohol Content. J. Chem. Educ. 2001, 78, 1225–1227. 106. Beck, R. C.; Erdmann, M. A. Reinforcing Mass Spectrometry Concepts through an Undergraduate Laboratory Exercise Utilizing a Direct Analysis in Real Time Enabled Mass Spectrometer. J. Chem. Educ. 2018, 95, 652–661. 107. Macherone, A.; Siek, T. Identification of Pharmaceuticals via Computer-Aided Thin Layer Chromatography. J. Chem. Educ. 2000, 77, 366–367. 108. Wilson, A.; Myers, C.; Crull, G.; Curtis, M. Analysis of Soft Drinks Using Nuclear Magnetic Resonance Spectroscopy: A Mentorship. J. Chem. Educ. 1999, 76, 1414–1416. 109. Kniesel, A.; Bellamy, M. Measuring Breath Alcohol Concentrations with an FTIR Spectrometer. J. Chem. Educ. 2003, 80, 1448–1450. 110. Timmer, W. C. An Experiment in Forensic Chemistry – The Breathalyzer. J. Chem. Educ. 1986, 63, 897–898. 111. Bowen, R. S.; Picard, D. R.; Verberne-Sutton, S.; Brame, C. J. Incorporating Student Design in an HPLC Lab Activity Promotes Student Metacognition and Argumentation. J. Chem. Educ. 2018, 95, 108–115. 112. Coleman, W. F. Molecular Models of Real and Mock Illicit Drugs from a Forensic Chemistry Activity. J. Chem. Educ. 2008, 85, 880. 106 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

113. Griep, M.; Mikasen, M. Lights, Camera… ReAction! Chemistry in the Movies. Nat. Chem. 2010, 2, 71. 114. Parker, P. D.; Beers, B.; Vergne, M. J. What Is in Your Wallet? Quantitation of Drugs of Abuse on Paper Currency with a Rapid LC-MS/MS Method. J. Chem. Educ. 2017, 94, 1522–1526. 115. Gray, F. M.; Smith, M. J.; Silva, M. B. Identification and Characterization of Textile Fibers by Thermal Analysis. J. Chem. Educ. 2011, 88, 476–479. 116. Valente Nabais, J. M.; Costa, S. D. A Forensic Experiment: The Case of the Crime at the Cinema. J. Chem. Educ. 2017, 94, 1111–1117. 117. Bender, S.; Lillard, S. Was the Suspect in Contact with the Victim? An Instrumental Methods Experiment for the Analysis of Single Fibers Using FT-IR Microscopy. J. Chem. Educ. 2003, 80, 437–440. 118. Berry, K. O. A Tool for Forensic Science. J. Chem. Educ. 1986, 63, 701–702. 119. Hern, J. A. The Great Detective Caper: An Experiment in Sampling and Forensic Analysis. J. Chem. Educ. 1988, 65, 1096. 120. Breedlove, C. The Analysis of Ball-Point Inks for Forensic Purposes. J. Chem. Educ. 1989, 66, 170–171. 121. Hoffman, E. M.; Beussman, D. J. Paint Analysis Using Visible Reflectance Spectroscopy: An Undergraduate Forensic Lab. J. Chem. Educ. 2007, 84, 1806–1808. 122. Randall, D. W.; Hayes, R. T.; Wong, P. A. A Simple Laser Induced Breakdown Spectroscopy (LIBS) System for Use at Multiple Levels in the Undergraduate Chemistry Curriculum. J. Chem. Educ. 2013, 90, 456–462. 123. Zuidema, D. R.; Herndon, L. B. Using “The Poisoner’s Handbook” in Conjunction with Teaching a First-Term General/Organic/Biochemistry Course. J. Chem. Educ. 2016, 93, 98–102. 124. Porter, S. E. G.; Barber, A. E.; Colella, O. K.; Roach, T. D. Using Biological Organisms as Chemical Sensors: The MicRoboCop Project. J. Chem. Educ. 2018, 95, 1392–1397. 125. Reidy, L.; Bu, K.; Godfrey, M.; Cizdziel, J. V. Elemental Fingerprinting of Soils Using ICPMS and Multivariate Statistics: A Study for and by Forensic Chemistry Majors. Forensic Sci. Int. 2013, 233, 37–44. 126. Boucher, M. A.; Specht, K. M. A Forensic-Themed Case Study for the Organic Lab. J. Chem. Educ. 2009, 86, 847–848. 127. Rose, R. E.; Fitzgerald, N. Incorporating a Forensic Science Problem-Solving Activity into an Analytical Chemistry Laboratory Course. Chem. Educ. 2011, 16, 1–4. 128. Friesen, J. B. Activities Designed for Fingerprint Dusting and the Chemical Revelation of Latent Fingerprints. J. Chem. Educ. 2015, 92, 505–508. 129. Hartzell-Baguley, B.; Morgan, S. Chemical Composition of Latent Fingerprints by Gas Chromatography-Mass Spectrometry. An Experiment for an Instrumental Analysis Course. J. Chem. Educ. 2007, 84, 689–691. 130. Ifa, D. R.; Manicke, N. E.; Dill, A. L.; Cooks, R. G. Latent Fingerprint Chemical Imaging by Mass Spectrometry. Science 2008, 321, 805. 131. LaFratta, C. N.; Huh, S. P.; Mallillin, A. C.; Riviello, P. J.; Walt, D. R. Visualizing Fluorescence: Using a Homemade Fluorescence “Microscope” to View Latent Fingerprints on Paper. J. Chem. Educ. 2010, 87, 1105–1107. 107 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

132. Dickson, H.; Kittredge, K. W.; Sarquis, A. M. Thin-Layer Chromatography: The “Eyes” of the Organic Chemist. J. Chem. Educ. 2004, 81, 1023–1025. 133. Henck, C.; Nally, L. GC-MS Analysis of γ-Hydroxybutyric Acid Analogs: A Forensic Chemistry Experiment. J. Chem. Educ. 2007, 84, 1813–1815. 134. Allison, J. Mass Spectrometry Theatre: A Model for Big-Screen Instrumental Analysis. J. Chem. Educ. 2008, 85, 1582–1583. 135. Houk, S. ICP-MS for Forensic Applications. Spectroscopy 2011, 26, 57–58. 136. Schurter, E.; Zook-Gerdau, L.; Szalay, P. Analysis of a Suspected Drug Sample. J. Chem. Educ. 2011, 88, 1416–1418. 137. Palmer, P. T. Energy-Dispersive X-ray Fluorescence Spectrometry: A Long Overdue Addition to the Chemistry Curriculum. J. Chem. Educ. 2011, 88, 868–872. 138. Smith, G. D.; Nunan, E.; Walker, C.; Kushel, D. Inexpensive, Near-Infrared Imaging of Artwork Using a Night-Vision Webcam for Chemistry-of-Art Courses. J. Chem. Educ. 2009, 86, 1382–1388. 139. Marino, M. A.; Voyer, B.; Cody, R. B.; Dane, A. J.; Veltri, M.; Huang, L. Rapid Identification of Synthetic Cannabinoids in Herbal Incenses with DART‐MS and NMR. J. Forensic Sci. 2016, 61, S82–S91. 140. Callahan, R.; Kobilinsky, L.; Rothchild, R. Organic Chemistry Course Development in a Forensic Science Program: Use of FT-NMR. J. Chem. Educ. 1999, 76, 1332–1333. 141. Babcock, S.; Warny, S. Forensic Palynology as Classroom Inquiry. Science Activities 2014, 51, 116–128. 142. Berka, K. M.; Berka, L. H. Developing Student Speaking Skills: A Project/Independent Study in Forensic Science. J. Chem. Educ. 1996, 73, 931–933. 143. Bloom, B. S.; Krathwohl, D. R. Taxonomy of Educational Objectives: The Classification of Educational Goals, 1st ed.; Longmans, Green: New York, 1956. 144. Anderson, L. W.; Krathwohl, D. R.; Bloom, B. S. A Taxonomy for Learning, Teaching, and Assessing: A Revision of Bloom’s Taxonomy of Educational Objectives, Complete ed.; Longmans, Green: New York, 2001.

108 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Chapter 7

Downloaded via CHALMERS UNIV OF TECHNOLOGY on October 6, 2019 at 15:01:36 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Interdisciplinary Learning Communities: Bridging the Gap between the Sciences and the Humanities through Forensic Science Maria Danielle Garrett* Department of Chemistry and Physics, Belmont University, 1900 Belmont Boulevard, Nashville, Tennessee 37212, United States *E-mail: [email protected].

Interdisciplinary Learning Communities are an integral part of the Belmont Experience: Learning for Life (BELL) Core for first- or second-year students at Belmont University. In an Interdisciplinary Learning Community, a single group of undergraduate students takes two courses in different subjects that are linked together through a common theme. CEM 1620 (general chemistry II), the second course in a two-semester sequence of chemistry for science majors, has been successfully linked with ENL 2015 (forensics in literature) over the course of four semesters. The common theme for these two courses is forensic science. Through this link, students begin to develop both deeper insight into the importance of understanding scientific processes and the skills necessary to accurately and effectively analyze and communicate scientific information. This chapter addresses the unique challenges faced in linking a science-major course with a humanities course, including introducing nontraditional chemistry texts into the curriculum and engaging students in an exploration of scientific analysis techniques not common to general chemistry. It discusses how to incorporate meaningful interdisciplinary experiences as well as how to analyze outcomes in the students’ final projects—writing an original fictional mystery, including scientifically accurate chemical evidence.

Introduction Although there are many variations of learning communities in the undergraduate setting, in general, learning communities consist of a group of students taking a series of academic courses together. Zhao and Kuh report that student participation in learning communities has a positive correlation with many favorable outcomes, including but not limited to academic performance, active and collaborative learning, and satisfaction with the college experience (1). Although learning communities cannot be attributed as the sole cause, emphasis on learning communities has been © 2019 American Chemical Society Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

seen to have a positive impact on student retention (2). Utilizing science courses as part of a learning community is not uncommon (2–5). Often, the science courses chosen for learning communities are non-science majors’ courses. L’Heureux highlights learning communities taught at Holyoke Community College, pairing a laboratory topics in science course and a writing course—providing students with the opportunity to engage in in-depth discussions, writing analysis, and relating science content to bigger ideas, looking at human and environmental impact (6). It was reported that students participating in the learning communities at Holyoke Community College typically showed an increase in academic performance, earning grades of C or better in these courses. At Wofford College, first-year students take a learning community that pairs a 4-hour laboratory science course for non-science majors with a humanities course, typically a seminar that first-year students must complete (7). Some science–humanities course pairings include biology, physics, or psychology–English; biology or physics–philosophy; biology–sociology; and psychology–history. Wofford’s model reaches beyond the typical learning community models—engaging not only college faculty and first-year students but also upper-level students, who serve as part of the teaching teams in the learning communities. Additionally, the learning community courses at Wofford reach outside the institutional walls, developing and implementing programming for the outside community—typically for elementary school children. Some of the feedback from first-year students participating in the learning communities at Wofford included an increase in appreciation for science and an increase in accountability for their academic success and their role in the community. Although it is more challenging to incorporate science courses for science majors into learning communities because of the breadth and depth of content that must be covered in these courses, several institutions have chosen to implement the learning community concept with majors’ courses, such as those in engineering (4, 5), life sciences (5), and physical sciences (2, 5). In some cases, such pairings have been shown to possibly relate to an increase in the number of majors in a given program (2). Richardson, Tooker, and Eshleman outline a three-course Learning Community model at Wagner College for first-year students (8). This model includes a core science course, a course in a different content area—such as a social science, philosophy, or microeconomics—and a reflective tutorial. In this study, students taking a Learning Community general chemistry I course were found to score as well or better on exams in general chemistry II than those students who completed a traditional version of general chemistry I. It was also reported that a high percentage of students completing their chemistry major at Wagner College identified the general chemistry I Learning Community as one of the reasons they choose to major in chemistry.

Interdisciplinary Learning Communities General education at Belmont University is comprised of the Belmont Experience: Learning for Life (BELL) Core Experience. Through the BELL Core curriculum, undergraduate students build and strengthen their ability to make connections between different disciplines, communicate, collaborate, think critically, and engage in citizenship (9). Over the course of 4 years, the BELL Core requires that students take a series of signature courses (Belmont-specific liberal arts), foundation courses (traditional liberal arts), and degree cognates (degree specific) (Table 1). The signature courses were designed with a vertical structure as students move through four stages of learning (launching, intersecting, broadening, and reflecting) (10). Interdisciplinary Learning Communities are one part of the signature courses and are designed to correspond to the “intersecting” stage of learning. In an Interdisciplinary Learning Community, a single group of students take two courses 110 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

in different subjects linked through a common theme. Additionally, faculty may apply for a Belmont Quality Enhancement Plan (QEP) grant to fund a common experience that will serve to enhance the students’ learning experience. Faculty teaching the two courses are encouraged to attend each other’s courses, if possible. The two courses may share one or more common assignments that are designed to help the students see the interconnectedness between the two disciplines. To emphasize the interdisciplinary experience for the students, the two linked courses are from different Colleges. For example, students would not find an Interdisciplinary Learning Community offering for a chemistry course and a mathematics course because both these departments are housed in the College of Sciences and Mathematics. There are a wide variety of course pairings from which students can choose (Table 2). Table 1. BELL Core Program Requirementsa Signature courses (16 hrs)

Foundations courses (22 hrs)

Degree cognatesc for Bachelor of Science (15 hrs)

First-Year Seminarb (3 hrs)

Quantitative Reasoning (3 hrs)

Social Science (6 additional hrs)

Interdisciplinary Learning Communitiesb,d

Oral Communication (3 hrs)

Humanities (3 additional hrs)

Junior Cornerstoneb,d

Lab Science (4 hrs)

Math (3 additional hrs)

Senior Capstoneb (1 hr)

Social Science (3 hrs)

Science (3 additional hrs)

First-Year Writing (3 hrs)

Humanities (3 hrs)

Third-Year Writing (3 hrs)

Fine Art (3 hrs)

First-Year Religion (3 hrs)

Wellness (3 hrs)

Third-Year Religion (3 hrs) a Descriptions

of BELL Core program requirements are detailed on the BELL Core website (11). b These signature courses correspond to the four stages of learning: launching (First Year Seminar, intersection (Interdisciplinary Learning Communities), broadening (Junior Cornerstone), and reflecting (Senior Capstone) (10). c Degree Cognates range from 0 to 15 hours and are specific to the degree the student is pursuing (Bachelor of Science [15 hrs], Bachelor of Arts [15 hrs], Bachelor of Business Administration [12 hrs], Bachelor of Science in Nursing [10 hrs], Bachelor of Science in Public Health [13 hrs], Bachelor of Social Work [12 hrs], Bachelor of Music [0 hrs], or Bachelor of Fine Arts [0 hrs]). d Hours for the Interdisciplinary Learning Community and Junior Cornerstone count elsewhere in the program. For example, students might take a Junior Cornerstone that would also give them credit for their third-year religion signature course requirement.

111 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Table 2. Examples of Recent Interdisciplinary Learning Community Course Offerings Course 1

Course 2

Urban Wildlife and Nature (biology, 3 hrs)

Environmental Advocacy (communications, 3 hrs)

A Process of Inquiry (physics, 4 hrs)

Science Fiction; Science Fact (English literature, 3 hrs)

Principles of Biology II (biology, 4 hrs)

Creationism and Evolution (philosophy, 3 hrs)

Principles of Biology II (biology, 4 hrs)

Wisdom or Warning (English writing, 3 hrs)

General Chemistry II (chemistry, 4 hrs)

Forensics in Literature (English literature, 3 hrs)

Principles of Macroeconomics (economics, 3 hrs)

Elementary Statistics (mathematics, 3 hrs)

The Medieval World (history, 3 hrs)

Culture and Arts Medieval World (humanities, 3 hrs)

Literature and the Stage (English literature, 3 hrs)

The Theatre Experience (theatre and drama, 3 hrs)

Critical Listening (audio engineering technology, 1 hr)

Writing with Sound (English writing, 3 hrs)

History of Recording Business (music business, 3 hrs)

The Rhetoric of Country Music (English writing, 3 hrs)

Asian Humanities (humanities, 3 hrs)

Survey of East Asian History (history, 3 hrs)

Most students enroll in a set of Interdisciplinary Learning Community courses the spring of their first year or the fall of their second year. The learning outcomes for the Interdisciplinary Learning Communities were developed to align with the “intersecting” stage of learning, in which students develop their understanding of “ways of knowing.” By the end of a set of Interdisciplinary Learning Community courses, students should exhibit the ability to: • Distinguish between the kinds of knowledge and the types of thinking and learning processes that are represented in the two disciplines; • Recognize the interconnectedness of knowledge through examination of an area of overlap between two courses; • Integrate learning from each of the disciplines into the other and provide a specific example of how something learned in one class contributed to your understanding of the other; and • Evaluate various information and experiences from the perspective of each of the disciplines.

112 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Linking General Chemistry II and Literature Since spring 2015, an Interdisciplinary Learning Community linking general chemistry II (CEM 1620) and a literature course (ENL 2015) has been offered four times. The common theme for this link was forensic science. Students’ grades were based on both traditional assessments typical to the nature and content of each course and a final common project (Table 3). Table 3. Assessment Items for General Chemistry II and Forensics in Literature and Science Course CEM 1620a

ENL 2015

Activity

Grade distribution

Exams (3) and final

60%

Online Online homework

5%

Online Online reading quizzes, participation, and attendance

5%

Laboratory

15%

Interdisciplinary Learning Community final projectb

15%

Sociocultural analysis of a mystery story

30%

Participation and daily activities

15%

Quizzes and tests

20%

Reading responses

20%

Interdisciplinary Learning Community final project

15%

a The

grade distribution for the Interdisciplinary Learning Community version of CEM 1620 has a different grade distribution than does the traditional version of the course (60% exams and final; 8% online homework; 7% online reading quizzes, participation, and attendance; and 25% laboratory) because the traditional course does not include a final project. For the linked course, students were assigned some additional assignments, including readings, writing assignments, and small-group work. Points for these assignments fell under the online homework grade category. Several additional forensic-themed laboratory activities were also included. Points for these activities fell under the laboratory grade category. b A minimum of 15% of the grade for an Interdisciplinary Learning Community course is required to be based on a common assessment.

CEM 1620 is the second course in a two-semester sequence of general chemistry for science majors. It is a 4-hour course, with both a lecture and a lab component. Topics covered in this course typically include chemical equilibrium, acid–base chemistry, thermochemistry and thermodynamics, electrochemistry, kinetics, intermolecular forces, and the physical properties of liquids, solids, and solutions (Tables 4–6). Course-specific goals include: • Students will be able to read, understand, and solve problems in the context of chemical principles. • Students will be able to read and process information from each chapter, then apply that knowledge to new situations. • Students will be able to learn how to carry out a laboratory experiment and relate concepts from lecture—interpreting data in a meaningful way. • Students will be able to become responsible for their own academic success. 113 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Table 4. Traditional CEM 1620 Lecture Schedule for Fall 2018 Semestera,b Class introduction and chemical equilibrium Chemical equilibrium Chemical equilibrium Acids and bases Acids and bases and acid–base equilibrium Acid–base equilibrium Acid–base equilibrium Solubility Solubility Exam 1 Energy Energy Spontaneity, entropy, free energy Spontaneity, entropy, free energy Spontaneity, entropy, free energy Electrochemistry Electrochemistry Review Exam 2 Kinetics Kinetics Kinetics Kinetics and nuclear chemistryc Liquids and solids Liquids and solids Solutions Solutions Exam 3 Final exam a The lecture schedule for fall 2018 is based on a TR class time. Students meet for lecture for 75-minute blocks

two times a week. b Slight modifications to the length of time spent on a given topic vary from semester to semester. c Chemistry instructors have the option to include a chapter of their choice in general chemistry II. As such, nuclear chemistry is not covered in every section of general chemistry II. However, it is covered in both the traditional and Interdisciplinary Learning Community versions compared in this chapter.

114 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Table 5. Interdisciplinary Learning Community CEM 1620 Lecture Schedule for Spring 2018 Semestera, b Class introduction and chemical equilibrium Chemical equilibrium Chemical equilibrium Chemical equilibrium Acids and bases Acids and bases Acids and bases Acids and bases Nuclear chemistry lecture (during lab time)c,d Acid–base equilibrium Acid–base equilibrium Acid–base equilibrium Solubility Solubility Exam 1 Energy Energy Energy Forensic science presentations (1–5) Spontaneity, entropy, free energy Spontaneity, entropy, free energy Spontaneity, entropy, free energy Spontaneity, entropy, free energy Forensic science presentations (6–10) Electrochemistry Electrochemistry Electrochemistry Guest speaker: during lab time (in-depth discussion with class) and afternoon talk (general science talk open to the university) Electrochemistry Exam 2 Kinetics Guest speaker reflection due Kinetics

115 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Table 5. (Continued). Interdisciplinary Learning Community CEM 1620 Lecture Schedule for Spring 2018 Semestera, b Kinetics Kinetics Liquids and solids Liquids and solids Liquids and solids Solutions Solutions Solutions Solutions Exam 3 Student presentations (an original fictional mystery) These presentations may run through both CEM and ENL Final exam a The lecture schedule for spring 2018 is based on a MWF class time. Students meet for lecture for 50-minute

blocks three times a week. b Slight modifications to the length of time spent on a given topic vary from semester to semester. c Chemistry instructors have the option to include a chapter of their choice in general chemistry II. As such, nuclear chemistry is not covered in every section of general chemistry II. However, it is covered in both the traditional and Interdisciplinary Learning Community versions compared in this chapter. d To account for the additional time constraints caused by the inclusion of the Interdisciplinary Learning Community content, some of the traditional course content crosses over into laboratory time. For example, nuclear chemistry is specifically covered during a scheduled laboratory time. Occasionally, the last part of a lecture may need to be covered at the beginning of a laboratory time in order to stay on schedule.

ENL 2015 is a topical literature course that focuses on the development of forensic science in literature. Course-specific goals include: • Students will be able to explain, in writing or orally, the role of inductive and deductive reasoning in the analysis of crime data. • Students will be able to write a sociocultural analysis of a mystery story. • Students will be able to explain, in writing or orally, how knowledge of the scientific method demonstrates a particular “way of knowing.” • Students will know how to locate critical analyses of literature in the library and will demonstrate that they know how to cite sources using appropriate documentation.

116 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Table 6. A Comparative Laboratory Schedule for CEM 1620 Week

CEM 1620 fall 2018 lab schedule

CEM 1620 spring 2018 lab schedule

1

Check-in

Fictional case study project – Art Under Scrutiny

2

Equilibrium and lab safety

Check-in, lab safety, and equilibrium

3

No lab

Nuclear chemistry lecture

4

Titrationsa

Titrationsa

5

No lab

Bloodstain pattern analysis

6

Calorimetry Ia

Calorimetry Ia

7

Calorimetry IIa

Calorimetry IIa

8

Enthalpy/entropy/dilutions

Enthalpy/entropy/dilutions

9

No lab

No lab

10

Electrochemistry

Guest speaker during lab and afternoon talk

11

Kinetics

Electrochemistry

12

No lab

No lab

13

Intermolecular forces

Kinetics

14

No lab (holiday/break)

Intermolecular forces and Fiber Analysis

15

Colligative properties

Colligative properties

a Students were required to write a formal lab report for both the titration lab and the two-week calorimetry lab.

Challenges Although using forensic science as a common theme is a natural fit for both the general chemistry and literature courses, finding a way to help students clearly see the overlap between the content of the two courses presented some unique challenges: • Introducing nontraditional chemistry texts into the curriculum (CEM 1620); • Making clear to the students what links the two courses (ENL 2015); and • Balancing the emphasis on literature and science (ENL 2015). CEM 1620 has the additional challenge of time constraints. In addition to covering the normal content for general chemistry II, the Interdisciplinary Learning Community version of the course must also incorporate engaging ways to let the students explore scientific analysis techniques not commonly addressed in general chemistry. Addressing these challenges is a continuous process. Each time this course pairing is offered, forensic-themed content is added or changed to try to enhance the overall learning experience for the students (Table 7). Student Population Although CEM 1620 can be used by both science and non-science majors to complete the 4hour BELL Core lab science requirement (foundation courses), it is designed for science majors. Most students do not take this course only to fulfill the Interdisciplinary Learning Community 117 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

because the content is challenging. Additionally, in order to receive credit for the Interdisciplinary Learning Community, students must pass both courses. If they do not pass one of the courses, they must either retake both of the linked courses the next time they are offered or take another set of linked courses to fulfill the Interdisciplinary Learning Community requirement. As such, most students taking the course need general chemistry II to complete their major or minor. Although students are not currently polled regarding their choice to take the CEM 1620–ENL 2015 link, students taking the Interdisciplinary Learning Community version of general chemistry II enroll for a variety reasons, including enjoying the content, completing degree requirements, and scheduling. Some students are interested in forensic science. These students typically fall into one of two categories—those who want to pursue a career in forensic science or those who enjoy forensicthemed television shows, films, and books. Some of the students who need to take general chemistry II for their major or minor take the course to complete the BELL Core learning community requirement. Finally, there are those students who take the course because the linked general chemistry II section fits into their overall course schedule. Because these enrollment reasons are not fundamentally different from those in a traditional general chemistry II course, student population does not present a unique challenge to teaching the Interdisciplinary Learning Community version of general chemistry II. Table 7. Current Course Content Used To Link General Chemistry II and Forensics in Literature and Science Course

Content

CEM 1620

ENL 2015

Description

Molecules of Murder: Criminal Molecules and Classic Cases (12, 13)

Text: Explores several toxic molecules that have been used in infamous murders

As Light Meets Matter: Art Under Scrutiny (14)

Fictional case study: Explores the use of scientific instrumentation and analysis to determine the authenticity of a painting

Forensic presentations

Small group project: Analysis of a forensic science or investigation topic and how it relates to both a real-world case and a fictional work

Forensic-themed lab activities

Bloodstain pattern analysis and fiber analysis

The Scientific Sherlock Holmes: Cracking the Case with Science and Forensics (15)

Text: Explores Sherlock Holmes’ use of science, the scientific method, and investigation techniques

How Sherlock Changed the World (16) PBS film: Explores the influence Sherlock Holmes had on actual forensic science and investigations CEM 1620 and ENL 2015

Interdisciplinary experiences

On- and off-campus activities, including a tour of the Tennessee Bureau of Investigation, an interactive mystery performance, and guest speakers (a local detective fiction writer and a local forensic odontologist)

CEM 1620 and ENL 2015

Interdisciplinary Learning Community final project

Small group project: Original work of detective fiction with Appendix on scientific evidence

118 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Although this chapter does not focus on the challenges faced by students taking the CEM 1620–ENL 2015 link, it is worth noting a few informal qualitative observations. The additional assignments and projects added to enhance learning in the general chemistry II course can be a challenge for some students. Time management is key because students are still responsible for the same types of assignments, laboratory work, and exams that the traditional general chemistry II course also requires. Additionally, because the majority of the students are science majors, some students may find that the literature course presents them with some unexpected challenges. ENL courses are just one option for the required Humanities courses for students. Not all of the students enrolled in the CEM 1620–ENL 2015 link would necessarily have chosen an ENL option for one of their Humanities requirements had it not been linked to general chemistry II. Not only do literature courses usually require a great deal of reading, but the writing style and critical analysis of text is different from what the students typically experience in their science classes. Depending on students’ background in writing, the very nature of a literature course may also pose a challenge to some students.

Introducing Nontraditional Chemistry Texts into the Curriculum Introducing Nuclear Chemistry through a Criminal Case Chapter 10 in Emsley’s Molecules of Murder: Criminal Molecules and Classic Cases (12, 13), focuses on the use of polonium-210 in the 2006 murder of a former officer for the Committee for Safety Security (KGB), Alexander Litvinenko. Litvinenko was poisoned on November 1, 2006. Within 22 days (Table 8), he experienced all the symptoms of polonium poisoning (12): vomiting, decrease in white blood cell count, liver damage, peritonitis (17), toxic shock syndrome, and irreparable heart damage. Table 8. Timeline Overview of the Poisoning of Alexander Litvinenko Date

Description

November 1, 2006

Litvinenko was poisoned with a cup of tea. Later that evening, he began vomiting.

November 2–November 5, 2006

Litvinenko went to the hospital several times, receiving painkillers and antibiotics. Later, it was found that his white blood cell counted dropped.

November 6, 2006

Litvinenko had to be fed by a tube and could hardly talk.

November 17, 2006

Litvinenko had lost all of his hair and was moved to a top hospital in London.

November 20, 2006

Litvinenko was transferred to intensive care.

November 23, 2006

Litvinenko’s heart stopped. He was revived, and polonium-210 was finally detected. Litvinenko passed away later that evening.

In CEM 1620, this case is used as a starting point for an introduction to nuclear chemistry: nuclear reactions, radiation and the human body, and radioactive decay. After discussing common types of nuclear reactions, the focus is shifted toward the effect that radiation has on the human body. The biological effect of a range of values of single dose exposure (in roentgen equivalent man [rem]) to radiation is discussed. Dosage data can be found in the Nuclear Chemistry chapter of many general chemistry textbooks (18–21). This data is also used to introduce students to several common 119 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

units of radiation (gray [Gy], radiation absorbed dose [rad], and rem). Looking at the approximate absorbed radiation dose resulting in death within a certain time (30 Gy, 2 days and 2 Gy, 6 weeks) (22), students are asked to convert those values to rem, using eq 1 and eq 2 (18):

where the factor depends on the type of radiation and tissue being subjected to the radiation. For both the traditional and Interdisciplinary Learning Community versions of general chemistry II, student understanding of nuclear chemistry is assessed though online homework assignments and exams. In the traditional version, nuclear chemistry is introduced after kinetics and is assessed on Exam 3. In the Interdisciplinary Learning Community version, nuclear chemistry is introduced much earlier in the semester and is assessed on Exam 1. Although students in the linked CEM 1620 course are not exposed to first-order integrated rate laws prior to learning about radioactive decay, there has not been a significant difference in student assessment performance concerning this content. Additionally, the decision was made to move this content forward for the Interdisciplinary Learning Community version of general chemistry II because the difficulty of the material is lower compared with other topics covered during the semester, which makes it ideal for being used as the second piece of forensic-themed content to which the students are exposed. Exploring Artistic Forgery through a Fictional Case Study As Light Meets Matter: Art Under Scrutiny (14), a fictional case study that tasks students with determining the authenticity of a Cézanne painting, was used to introduce students to techniques and instrumentation they might not normally encounter in general chemistry (Table 9). This case study was introduced in CEM 1620 early in the semester and was the first forensic-themed content to which the students were exposed. Although this activity did require student work outside of normal scheduled course meeting times, students were given time to begin working on this case study during the first lab meeting of the semester. The points for this assignment counted toward the online homework component of their grade. Table 9. Analysis Techniques for Authenticating Paintings Technique

How the technique was used in the fictional case study

UV absorption

Used to detect levels of polyenes, which can be related to the age of the painting

IR spectroscopy

Used to detect the presence of paints or binders

Underdrawings

The presence or absence of underdrawings could be used to help verify authenticity, based on whether a painter was known to use underdrawings

X-ray fluorescence

Used to identify pigments

UV-vis fluorescence

Used to identify whether there were retouches to the painting

Students broke into four groups (four to six students per group). The students were tasked with creating a poster that described and explained the chemical evidence provided in the case study. Using this evidence, two of the groups had to support the idea that the Cézanne painting was

120 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

authentic, and the other two groups had to support the idea that the painting was fake. Students were required to: • Describe and explain the chemical evidence (including but not limited to UV absorption, IR spectroscopy, underdrawings, X-ray fluorescence, and UV-vis fluorescence); • Address their conclusion on whether the painting is authentic, based on analysis of the evidence provided in the case study; and • Answer the broader question—can a painting truly be authenticated? Students were told to use the questions provided in the case study to guide their scientific arguments and conclusions but not to simply list answers to the questions on their poster. In the most recent version of this assignment, students were also provided with an electronic template for an academic poster. The addition of this template allowed students to create a conference-style academic poster, a skill that is especially useful to Belmont science majors because formal presentations are part of their research requirement. Students were graded on the following: • Completion of all the requirements, including the use of at least four relevant visuals, proper use of figure captions, grammar, and use of the ACS citation style; • How well they explained the science associated with their conclusions about the case study; • Creativity and visual interest; and • Coherence and flow. Students were required to submit their posters electronically to the instructor. All four presentations were made available to the class. Each student was then required to submit a twoparagraph reflection on whether he or she believed the artwork was authentic or fake—which may or may not agree with the group to which they were assigned. They were required to support their answers using scientific evidence.

Engaging Students in an Exploration of Scientific Analysis Techniques In order to provide students with more cohesive exposure to how forensic science ties into chemistry, additional projects and activities were incorporated into the course curriculum—forensic science presentations and forensic-themed lab activities. These assignments pushed the students beyond the confines of traditional general chemistry content, having them research and explore forensic science topics and conduct hands-on activities, making connections between subject matter and real-world applications. Forensic Science Presentations Small group projects were added to the CEM 1620 Interdisciplinary Learning Community Course. Students worked in pairs to research one of several possible forensic science or forensic investigation topics: fingerprints; bloodstains; DNA evidence; impressions (including shoes, tires, and tools); handwriting (including forgery and the validity of documents); analyzing firearms evidence; and arsenic, cyanide, strychnine, antimony, and lead poisoning and poisoning via living things (plants and animals). They were tasked with creating a presentation on how their topics related to criminal investigations. Students were required to: 121 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

• Describe scientific evidence related to the topic (in terms of facts, concepts, principles, and theories), explaining any new terminology required to understand the topic; • Discuss how the evidence is analyzed (in terms of methods, instrumentation, and procedures), showing examples to support the discussion—for example, images of fingerprint patterns, sketches of shapes or patterns of blood spatter, and sample gas chromatography, mass spectrometry, or infrared spectra; • Discuss at least one real example of the topic being used in a forensic case, giving an overview of the case but focusing on how the evidence was gathered and analyzed; • Discuss at least one fictional work (literature or film) of the topic being an integral part of the plot, focusing on how the crime was implemented and how the evidence was gathered and analyzed (if covered in the plot); and • Utilize and formally cite at least three references, with at least one of the references being a book (23). Along with their presentation, student groups were required to create a bullet point overview of their presentations. These overviews were compiled, copied, and handed out to the class toward the end of the semester. These overviews served as an informational resource for their Interdisciplinary Learning Community final project. Students were also assessed on the equality of the contribution of each group member and the length of the presentation (10 to 11 minutes). While many of the forensic science or forensic investigation topics have direct tie-ins to the field of chemistry—using gas chromatography–mass spectrometry to detect or confirm the presence of a poison in a victim or using ninhydrin to detect fingerprints—some of them do not. For example, bloodstain investigations focus on data analysis of evidence including patterns, size, shape, and angle of impact. Beyond being a reference tool for the students’ final projects, ultimately, these presentations are used to enhance student understanding of the scientific method in practice through exploration of these applied real-world analytical investigation techniques. Additionally, these presentations serve as a surface-level introduction to various instrumentation that some of these students will be exposed to if they continue on to take chemistry courses such as organic, inorganic, and instrumental analysis. Forensic-Themed Lab Activities Two forensic-themed activities were added to laboratory experience for students in CEM 1620. These activities were popular among the students and provided a real-world connection for them. Students engaged in a simulated bloodstain pattern analysis activity (Flinn Scientific) (24). In this activity, students explored the impact of distance and angle on simulated bloodstains and patterns (Figure 1). Simulated blood was dropped onto paper from a series of different heights and angles. Students created graphs of drop height versus diameter of bloodstain, drop angle versus length of bloodstain, and drop angle versus width of bloodstain. Students used their results to try to interpret and recreate a simulated bloodstain pattern from a fictional crime scene. Students also engaged in a second activity—finding evidence in fibers (Flinn Scientific) (25). In this activity, students spent time investigating and analyzing different synthetic and natural fibers (wool, acrylic, polyester, nylon, cotton, and acetate) through primary and secondary analysis. Through burn analysis (primary), students observed burn time, coating, curling, melting, color, smell, and smoke. They also conducted microscope analysis (secondary), in which they observed winding and color. Students used their results to identify a fiber sample obtained from a fictional crime scene. 122 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

These two activities were supplemental content for the Interdisciplinary Learning Community general chemistry II laboratory coursework and were used to give students hands-on experience with some basic concepts and techniques concerning bloodstain pattern and fiber analysis. Instrumental analysis techniques are not part of the current CEM 1620 curriculum. Students are not introduced to concepts such as gas chromatography–mass spectrometry, high-performance liquid chromatography, and fluorescence spectroscopy until subsequent chemistry courses. However, depending on the sequencing of curriculum and student access to laboratory instrumentation, these activities could be modified to have a more direct link to specific chemistry content: • Principles of biochemistry and molecular biology: Another important part of bloodstain pattern analysis is determining which stains are blood so that they may be collected for DNA analysis (26). Additionally, polymerase chain reaction analysis plays an important role in DNA profiling. • Principles of organic chemistry: Dyes on fibers can be analyzed using techniques such as ultraviolet-visible spectroscopy, chromatography, and mass spectrometry (27). Additionally, Fourier transform infrared spectroscopy can be used to determine structure and composition of natural fibers (28).

Figure 1. Samples of bloodstain patterns, with simulated blood dropped from different heights and angles. Photo courtesy of Maria Danielle Garrett.

Interdisciplinary Experiences Each fall, the instructors of CEM 1620 and ENL 2015 applied for a QEP grant to fund a common experience for the linked Interdisciplinary Learning Community courses in the spring. These interdisciplinary experiences were chosen to pique student interest in the shared forensic 123 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

science theme, help stimulate the creative process for their final project, and help students see the real-world connections between forensic science in both chemistry and literature. In spring 2015, the CEM 1620–ENL 2015 class took a tour of the Tennessee Bureau of Investigation (TBI). The tour started with a brief video introduction concerning the history of the TBI—specifically, the 1949 murder that led to the development of what is now known as the TBI. Students were also given an overview about the educational and training requirements needed for one to become involved with the TBI, as well as internship opportunities for college students. A TBI officer led a walking tour of many of the secured access areas of the facility. Visitors were not allowed to enter the labs themselves; however, the labs, equipment, and scientists were able to be viewed from the hallway. Outside each lab, there was a display of sample evidence and data analysis. Not only did the officer go into great detail concerning the collection, processing, and analysis of evidence for the various labs, but he also shared many interesting criminal cases with the students. Students learned about evidence and data analysis concerning fingerprints, shoe prints, DNA, toxicology, firearms, phone taps, background checks, and data file recovery. In spring 2016, the CEM 1620–ENL 2015 class engaged in an activity that aligned more closely with goals of the literature course. An interactive performance was presented on campus by a local murder mystery troupe. The troupe performed specifically for the CEM 1620–ENL 2015 Interdisciplinary Learning Community course. Three actors came to the classroom, bringing props and costumes. Some students were asked to read the parts of characters. Toward the end of the performance, all the students engaged in using the qualitative evidence provided and deductive reasoning to try to solve the crime. In spring 2017, instructors sought to find a more balanced common experience that addressed ideas and topics related to both chemistry and literature. Clay Stafford, an author, screenwriter, producer, director, actor, composer, entrepreneur, and environmentalist, was brought to campus in a two-part event. First, a general talk, required for CEM 1620–ENL 2015 students to attend, was offered to students across Belmont University. Stafford’s diverse background and experiences generated a wide appeal to students across the Belmont community. However, it was his work as a local writer of detective fiction and founder of Killer Nashville, an international writers’ organization, that was the focus of a more in-depth question-and-answer session that took place with just the CEM 1620–ENL 2015 students after the general talk. In this small group setting, Stafford provided details on the process of writing mystery fiction, including advancing the plot and methods of researching scientific content. Based on the positive outcome from the spring 2017 event, in spring 2018, the CEM 1620–ENL 2015 class interacted with another guest speaker, Mike Tabor, DDS. Not only has Tabor served as the Chief Forensic Odontologist for the State Medical Examiner’s office in Nashville, TN, since 1983, but he also has his own private clinical practice, has served as president of the Tennessee Board of Dental Examiners and president of the American Board of Forensic Odontology, and is a senior consultant for the Criminal Investigations Division of TBI, under its Medicaid Fraud Unit. Tabor has lectured both nationally and internationally, coauthored two scientific texts, and pursued his own creative writing. As before, the speaker participated in a two-part event—an in-depth talk with the CEM 1620–ENL 2015 students and a general science talk open to the Belmont community. This experience helped the students with the following goals, which were developed by the instructors as part of the QEP grant application process:

124 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

• Students gained an understanding of the importance of a strong foundation of science necessary in writing creative, compelling, and scientifically accurate forensic mystery stories. • Students identified and reflected upon the important roles that science knowledge and laboratory methodology play in criminal investigations, from both a scientific and legal perspective. • Students gained real-world perspective on the importance of development of character behavior and plot in the creation of a mystery work of fiction. • Students learned more detail about the differences in the way that forensic investigations have been portrayed in literature and popular culture compared with real life. • Students learned more about multidisciplinary career opportunities. • Students saw how to incorporate creative, relevant, and accurate scientific information into their end-of-semester linked class project (creating an original fictional mystery, including scientifically accurate chemical evidence).

Culminating Final Project: An Original Fictional Mystery Students worked in groups of three to five to write an original work of detective fiction. Their story had to clearly highlight all the literary components they had learned about in ENL 2015 over the course of the semester, including development of a crime, characters, motive, and evidence. Their story was required to completely introduce, explain, and solve the mystery. The English faculty instructor graded students on their overall plot, pieces of forensic evidence, plausible motive, character development, description of setting, use of literary devices, and proper grammar and mechanics (Table 10). Each group was also required to write a detailed Appendix explaining the evidence used to solve their mystery (Figures 2–4). Students chose to use a variety of criminal evidence and analysis techniques, including but not limited to bruising, imprints, DNA, poisoning, bloodstain pattern analysis, blunt force trauma, latent fingerprints, and gas chromatography. The chemistry faculty instructor graded students on scientific knowledge and content, application of scientific content, significance of scientific evidence, and communication of scientific evidence (Table 11). Students were required to: • • • •

Include a detailed explanation of at least three pieces of scientific evidence; Explain how their applications of scientific evidence were relevant to the story; Analyze what the investigators learned from the evidence; and Submit written and visual communication that was well-organized, coherent, and supported their explanations.

After submitting their story and Appendix, each group presented their short story to the class. Presentations were graded by both instructors (Table 12). The presentations are listed on the same date on the syllabi for both courses—typically the last class meeting before finals. Presentations start in whichever course occurs first and continue, if needed, into the second course. Any remaining time can be used for final exam review. During their presentations, students were required to set the scene for their story and introduce the characters, background, and motive for the crime. They had to describe how the investigators discovered the crime and clearly present the evidence to the rest of the class. Students were given creative freedom in their choice of presentation. Over the course of time, this creativity manifested 125 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

itself in many forms, including interactive PowerPoint, dramatic readings, live performances, sock puppet theater, and video reenactments. Toward the end of each presentation, the class was tasked with attempting to solve the crime, using deductive reasoning skills. After the class members were encouraged to share their deductions and reasoning, the student presenters revealed the actual criminal.

Figure 2. Sample of student Appendix on scientific evidence—mercury poisoning. Students were required to format their Appendix as a scientific paper, including proper use of table headings, figure captions, and citations. Reproduced with permission from reference (29). The references cited in this sample of student work are listed in reference (30).

126 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Figure 3. Sample of student Appendix on scientific evidence—bloodstain pattern analysis. Bloodstain pattern analysis is covered in both the student forensic science presentations and in the Flinn Scientific (24) bloodstain pattern analysis activity. Beyond any information they receive in class, students are required to do more in-depth research on the evidence and analysis techniques they include in their Appendix. Reproduced with permission from reference (31). The references cited in this sample of student work are listed in reference (32).

127 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

128 Figure 4. Sample of student Appendix on scientific evidence—(a) strychnine poisoning and (b) gas chromatography. Some student Appendices were very comprehensive, clearly indicating how their scientific evidence was integrated into their story and providing detailed background information not only about the evidence itself but also about the theory behind the forensic and analytical analysis. Reproduced with permission from reference (31). The references cited in this sample of student work are listed in reference (32).

Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Rubrics and Assessment Students were provided rubrics for the crime story (Table 10), Appendix on scientific evidence (Table 11), and presentation (Table 12). The crime story and Appendix were scored by the instructors for ENL 2015 and CEM 1620, respectively. Both instructors scored the final presentations. At the end of the semester, instructors were also required to evaluate the final projects using an Interdisciplinary Learning Community assessment rubric (Table 13). This rubric aligns with the BELL Core learning outcomes for Interdisciplinary Learning Communities and is used as part of the program assessment. In CEM 1620 and ENL 2015, these rubric scores are not factored into the students’ grades. The majority of student groups score in the developing or mastering range for their abilities to recognize and apply content from both disciplines. There have been no student groups to reach mastery of integration of content. The qualifying statement that students must provide “unique insight” to their final project has proved challenging for the students to achieve (33). Table 10. Crime Story Rubrica 100–90

89–80

79–70

69–60

Plot

Clear beginning, middle, and end.

Two of three components are clear.

One of three None of components components is clear. is clear.

Evidence

Three or more pieces of forensic evidence.

Three pieces of forensic evidence.

Two or three pieces One or two pieces of of forensic evidence. forensic evidence.

Motive

Motive is plausible Motive is plausible Motive is unclear and/or does not fit and connects to and ties tightly to well with character. character. character.

Character

Story has at least three welldeveloped characters.

Story has two to three welldeveloped characters.

Story has two welldeveloped characters.

Description/ Detail

Setting and characters are described with detail.

Setting and characters are described.

Setting and Setting and characters could use characters are more development. minimally described.

Literary Devices

Plot contains at least three devices common to mystery fiction.

Plot contains at least two devices common to mystery fiction.

Plot contains at least Plot contains none of one device common the devices common to mystery fiction. to mystery fiction.

Grammar/ Mechanics

Story contains no more than two minor errors and no major errors.

Story contains no more than two minor errors and one major error.

Story contains no more than two minor and two major errors.

Motive is not obvious or too obvious. Story has only one character or characters lack development.

Story contains more than three minor and/or three major errors.

a Reproduced with permission from reference (34). Copyright 2015 Dr. Linda Holt, Ph.D.

129 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Table 11. Appendix of Scientific Evidence Rubrica,b 14.0

12.4

11.0

9.6

0

Scientific knowledge and content Description of at least three pieces of scientific evidence are thorough and accurate (in terms of facts, concepts, principles, theories, and methods).

Description of at least three pieces of scientific evidence are lacking some detail but are accurate (in terms of facts, concepts, principles, theories, and methods).

Descriptions of scientific evidence are lacking some detail but are accurate (in terms of facts, concepts, principles, theories, and methods). Only two pieces of scientific evidence are provided.

Descriptions of scientific evidence are lacking detail but are accurate (in terms of facts, concepts, principles, theories, and methods). Only two pieces of scientific evidence are provided.

Less than two pieces of scientific evidence are provided or descriptions of scientific evidence are not present.

Application of scientific content Applications of scientific evidence are extremely relevant to the story and explained thoroughly in terms of both why and how each test or method was used.

Applications of scientific evidence are extremely relevant to the story and explained somewhat thoroughly in terms of both why and how each test or method was used.

Applications of scientific evidence are somewhat relevant to the story and explained somewhat thoroughly in terms of both why and how each test or method was used.

Applications of scientific evidence are somewhat relevant to the story. An explanation of either why or how a test or method was used is missing.

Applications of scientific evidence are missing. There is no explanation of why or how a certain test or method was used.

Significance of scientific evidence A thorough analysis of what the investigators learned from the evidence (what the results proved or disproved) is written directly based on the evidence.

An analysis of what the investigators learned from the evidence (what the results proved or disproved) is written directly based on the evidence.

An analysis of what the investigators learned from the evidence (what the results proved or disproved) is written but not thoroughly tied to the evidence.

An analysis of what the investigators learned from the evidence (what the results proved or disproved) is written but not tied to the evidence.

An analysis of what the investigators learned (what the results proved or disproved) is not present.

Communication of scientific evidence Written and visual communication (e.g., tables, graphs, figures) are wellorganized and flow coherently. All listed

Most of the written and visual communication (e.g., tables, graphs, figures) are wellorganized and flow coherently. All listed

Most of the written and visual communication (e.g., tables, graphs, figures) are well-organized and flow coherently.

Some of the written and visual communication (e.g., tables, graphs, figures) are well-organized and flow coherently.

Written and visual communication (e.g., tables, graphs, figures) lack organization and coherency.

130 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Table 11. (Continued). Appendix of Scientific Evidence Rubrica,b 14.0

12.4

requirements are followed.

11.0

requirements are followed.

9.6

Most listed requirements are followed.

0

Most listed requirements are followed.

References ----At least four references (two being physical books) are cited ACS style.

-----

At least four references (two being physical books) are cited without proper ACS style.

a The final grade for the Appendix was scaled from 70 points to 100 points.

Fewer than four references are used or at least two physical books are not used.

b Adapted with permission from

reference (35). Copyright 2014 Illinois State Board of Education.

Table 12. Presentation Rubrica 100–90

89–80

79–70

Group Participation?

All three group members have visible roles and no one monopolizes.

All have visible roles but One person has no one person dominates. visible role.

Creativity of Presentation?

Highly creative.

Has some creative components.

Solvability?

The crime is solvable with the Crime is easily solvable. Crime does not evidence presented, but not too correspond well to easily. evidence.

Forensic Evidence?

Forensic evidence is explained well and integrated effectively into the story and the presentation.

Forensic evidence is explained well, but integration needs attention.

Little creativity; predictable.

Forensic evidence does not integrate well into the story and/or the presentation.

a Reproduced with permission from reference (34). Copyright 2015 Dr. Linda Holt, Ph.D.

131 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Table 13. Belmont University Interdisciplinary Learning Community Assessment Rubrica Emerging (1)

Developing (2)

Recognizes interconnectedness of both disciplines as applied to an area of overlap

Primarily acknowledges only one discipline relevant to the issue/topic

Acknowledges components of each discipline relevant to the issue/topic separately

Applies knowledge from both disciplines

Shows a lack of Appropriately uses understanding of knowledge or concepts knowledge or concepts from each discipline from one or both disciplines

Demonstrates sophisticated knowledge or concepts from each discipline

Integrates course content/ knowledge from both disciplines

Uses disciplinary content from both courses without 'weaving'

Weaves disciplinary content/knowledge from both courses together to provide unique insight into the issue/topic

Applies content/ knowledge from both courses but provides no unique insight into the issue/topic

Mastering (3) Skillfully acknowledges that the issue/topic has components of both disciplines

a Reproduced with permission from reference (36). Copyright 2015 Rachel Rigsby, Ph.D.

Conclusions While having the opportunity to incorporate forensic science into general chemistry II content has presented many unique challenges, the addition of new assignments, activities, and common experiences through the Interdisciplinary Learning Community course experience has served to enrich and deepen the learning experience of students involved in these courses, supporting the course-specific goals as they relate to the Interdisciplinary Learning Community course outcomes (Figure 5). As the interdisciplinary experiences and projects are developed to help meet the Interdisciplinary Learning Community course outcomes, they also serve to reinforce the CEM 1620 and ENL 2015 course-specific goals. The main benchmark, indicating the relative success of these additions to the course, is the Interdisciplinary Learning Community final project. Qualitative and quantitative improvement motivates the instructors to keep modifying the courses, finding new and exciting ways to connect the students to the content. The addition of the topical presentations gave the students a richer perspective on some of the connections between forensic science and methodologies and instrumental analysis techniques used in chemistry and provided a larger knowledge base to draw upon for the scientific evidence presented in their final projects. Similarly, the bloodstain pattern and fiber analysis activities not only gave the students hands-on experience with simulated real-world forensics analysis, but the activities also helped develop their applied critical thinking skills. In these activities, students were required to test a series of known samples and interpret their results, using those results to analyze and identify an unknown sample. Students also encounter this type of applied critical thinking in several of the traditional labs in general chemistry II, including enthalpy/entropy/dilutions, in which students generate a standard Beer–Lambert calibration curve based on the red-colored product of an equilibrium reaction and then utilize that calibration curve to interpret the temperature dependence on thermodynamic properties, and electrochemistry, in which students generate a reduction potential table based on a standard lead electrode and then use the reduction potential table to interpret a series of galvanic cells. 132 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

The 2015 experience at the TBI helped students recognize and analyze the important role that foundational knowledge of chemistry and laboratory techniques play in forensic science investigations, and it enabled them to compare and contrast the ways that forensic investigations have been portrayed in literature and popular culture with real life. Based on the creativity of the final presentations and stories in 2016, bringing the murder mystery performance troupe to campus improved the quality of all aspects of the final project. The stories were more detailed and the presentations included clever uses of props and role-playing. Perhaps the most impactful of all was the incorporation of an expert guest speaker. In 2017, the ENL 2015 instructor noted that, “In the final reflective writing that students composed, many mentioned advice they received from Clay [Stafford] about the necessary components of detective fiction—in particular, that setting is of paramount importance and is key to how the story progresses and how the crime is solved.” After the 2018 event, in CEM 1620, students were asked to write a reflection about the event. The responses were extremely positive: • “Overall I found Dr. Taber’s visit to be very informative and entertaining, as well as a valuable part of the LCC courses. I hope this continues on next year for those taking this LCC.” • “…I found the whole seminar very interesting and I loved looking at the pictures he showed us because I’ve always been super interested in crime shows and it was cool to see how the process works in real life. It was a really neat experience!” • “The part that was the most interesting was how Walk of Death [Tabor’s book] was a true story with just characters changed and how it took them over 10 years just to solve the case. It showed how dedicated he was in solving the case.” • “All the cases he describes gave a sense of mortality to the lecture and was very interesting.” Not only did these events positively impact the outcome of the final project, especially in terms of the Appendix of scientific evidence, but they also enriched the overall class experience.

Figure 5. Course-specific goals for both CEM 1620 and ENL 2015 as they relate to the Interdisciplinary Learning Community course outcomes.

133 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Future Assessment Because this Interdisciplinary Learning Community was not designed to be part of a research study, the Institutional Review Board approval necessary to provide quantitative data concerning student outcomes was not obtained. Generalizations about student performance are also difficult to make, as the overall grade distribution for the Interdisciplinary Learning Community version and traditional version of CEM 1620 is different. Additionally, motivation for academic success may be slightly different for some students taking the Interdisciplinary Learning Community version. As mentioned in the section on Student Populations, students must pass both courses in order to receive BELL Core credit for the Interdisciplinary Learning Community. After obtaining the proper Institutional Review Board approval, future assessment plans include: • Developing a more comprehensive multipart question on nuclear chemistry; • Comparing and analyzing Interdisciplinary Learning Community and traditional CEM 1620 student performance on the nuclear chemistry question to see if learning the material in the context of forensic science has any significant impact on student understanding; • Developing and implementing a student questionnaire that explores student perceptions of general chemistry II, the perceived impact that participating in an Interdisciplinary Learning Community with a forensic science theme has on classroom learning, and the value added for each interdisciplinary experience included in the course; and • Comparing and analyzing Interdisciplinary Learning Community and traditional CEM 1620 student performance on the general chemistry II final exam.

Acknowledgments The author thanks Linda Holt, Ph.D., associate dean of the College of Liberal Arts and Social Sciences and professor of English, for her collaboration in this Interdisciplinary Learning Community. The author would also like to thank Taylor Gerson, Caleb Holdener, Arielle Manabat, Corinn Marakovits, Michael Ottavio, Phoenix Slone, Jacob Utley, and Bethany West for the use of their Appendices. Funding for various resources and events were provided by the Belmont Office of General Education (QEP Grant) and the Department of Chemistry and Physics, College of Sciences and Mathematics, and the College of Liberal Arts and Social Sciences at Belmont University.

References 1. 2.

3. 4.

5.

Zhao, C-M.; Kuh, G. D. Adding Value: Learning Communities and Student Engagement. Res. High. Educ. 2004, 45 (2), 115–138. deProphetis Driscoll, W.; Gelabert, M.; Richardson, N. Efficacy of Using Learning Communities To Improve Core Chemistry Education and Increase Student Interest and Retention in Chemistry. J. Chem. Edu. 2010, 87 (1), 49–53. McBride, K. K. Linking Science Fiction and Physics Courses. The Physics Teacher 2016, 54 (5), 280–284. Calculus and Physics for Engineers in Themed Learning Communities Fall Courses: Fall 2018. https://niu.edu/learning-communities/themed-learning/fall-courses.shtml (accessed Jan 13, 2019). Iowa State University Learning Communities. https://www.lc.iastate.edu/communities (accessed Jan 13, 2019). 134 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

6. 7.

8. 9. 10.

11. 12. 13. 14.

15. 16. 17. 18. 19. 20. 21. 22.

23.

L’Heureux, G. A. A Learning Community Retrospective Integrating Science and Literature. Learning Communities Research and Practice 2018, 6 (2)Article 3. Goldey, E. Disciplinary Integration: The Science and Humanities in Learning Communities. The Power of Interdisciplinary/Multidisciplinary Courses and Curricula. https://www.aaas.org/sites/ default/files/07_IMC_Goldey.pdf (accessed April 1, 2019). Richardson, N.; Tooker, P. A.; Eshleman, A. Core Sciences in First-Year Learning Communities. Learning Communities Research and Practice. 2014, 2 (1)Article 4. BELL Core (Belmont’s General Education). http://www.belmont.edu/bellcore/ (accessed Jan 2, 2019). Daus, K.; Rigsby, R. In The Promise of Chemical Education: Addressing Our Students’ Needs; Daus, K., Rigsby, R., Eds.; ACS Symposium Series 1193; American Chemical Society: Washington, DC, 2015; pp 101–113. Description of Program Requirements. http://www.belmont.edu/bellcore/courses.html (accessed April 1, 2019). Emsley, J. Molecules of Murder: Criminal Molecules and Classic Cases; RSC Publishing: Cambridge, 2008; pp 197–213. Royal Society of Chemistry. https://pubs.rsc.org/en/content/ebook/978-1-78262-474-5 (accessed Jan 2, 2019). Del Federico, E.; Diver, S. T.; Konaklieva, M. I..; Ludescher, R. As Light Meets Matter: Art Under Scrutiny; National Center for Case Study Teaching in Science. http://sciencecases.lib. buffalo.edu/cs/collection/detail.asp?case_id=466&id=466 (accessed Jan 2, 2019). O’Brien, J. The Scientific Sherlock Holmes: Cracking the Case with Science and Forensics; Oxford University Press: New York, NY, 2017. Bernays, P., Director. How Sherlock Changed the World; LOVE Productions, 2013. Peritonitis is inflammation of the tissue that lines the abdominal wall and covers and supports many abdominal organs. Ball, D. W. Introductory Chemistry; Flat World Knowledge, Inc.: Irvington, NY, 2011; pp 395–396. Zumdahl, S. S.; Zumdahl, S. A. Chemistry: An Atoms First Approach, 2nd ed.; Cengage Learning: Boston, MA, 2014; p 747. Gilbert, T. R.; Kirss, R. V.; Foster, N.; Bretz, S. L. Chemistry: An Atoms-Focused Approach, 2nd ed.; W. W. Norton & Company, Inc.: New York, NY, 2018; p 999. More comprehensive tables of the effects of single-dose exposure to radiation can also be found through online searches. Bushberg, J. T. Radiation Exposure and Contamination. https://www.merckmanuals.com/ professional/injuries-poisoning/radiation-exposure-and-contamination/radiation-exposureand-contamination (accessed Jan 12, 2019). Students had access to, but were not required to use, the following books through the CEM 1620 Interdisciplinary Learning Community course instructor: (1) Emsley, J. The Elements of Murder: A History of Poison; Oxford University Press Inc.: New York, NY, 2006. (2) Lofland, L. HowDunit – Police Procedure & Investigation: A Guide for Writers; Writer’s Digest Books: Cincinnati, OH, 2007. (3) Lyle, D. P. Forensics for Dummies; Wiley Publishing, Inc.: Indianapolis, IN, 2004. (4) Lyle, D. P. HowDunit – Forensics: A Guide for Writers; Writer’s

135 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

24. 25. 26. 27. 28. 29. 30.

31. 32.

33.

34.

35. 36.

Digest Books: Cincinnati, OH, 2008. (5) Stevens, S. HowDunit – The Book of Poisons: A Guide for Writers; Writer’s Digest Books: Cincinnati, OH, 2007. Bloodstain Pattern Analysis – Forensic Laboratory Kit. https://www.flinnsci.com/bloodstainpattern-analysis---forensic-laboratory-kit/fb1643/ (accessed Jan 12, 2019). Flinn Forensic Files – Finding Evidence in Fibers. https://www.flinnsci.com/flinn-forensic-files--finding-evidence-in-fibers/fb2096/ (accessed Jan 12, 2019). Schiro, G. Collection and Preservation of Blood Evidence from Crime Scenes. https://www.crimescene-investigator.net/blood.html (accessed April 1, 2019). Goodpaster, J.; Liszewski, E. A. Forensic Analysis of Dyed Textile Fibers. Anal. Bioanal. Chem. 2009, 394 (8), 2009–2018. Fan, M.; Dai, D.; Huang, B. In Fourier Transform - Materials Analysis; Salih, S. M., Ed.; IntechOpen, 2012; pp 45–68. Holdener, C.; Marakovits, C.; Utley, J.; West, B. Appendix of Scientific Evidence for A Janus-Faced Crime; Belmont University: Nashville, TN, 2018. Unpublished work. The references shown in this sample of student work: (1) Hryhorczuk, D.; Persky, V.; Piorkowski, J.; Davis, J.; Moomey, C. M.; Krantz, A.; Runkle, K. D.; Saxer, T.; Baughman, T.; McCann, K. Residential Mercury Spills from Gas Regulators. Environ. Health Perspect. 2006, 114 (6), 848–852. (2) Haddad, L. M.; Winchester, J. F. Haddad and Winchester’s Clinical Management of Poisoning and Drug Overdose; Saunders: Philadelphia, PA, 1983; pp 637–643. Manabat, A.; Ottavio, M.; Slone, P.; Gerson, T. Appendix of Scientific Evidence for Accidents Happen; Belmont University: Nashville, TN, 2017. Unpublished work. The references shown in this sample of student work: (7) Brodbeck, S. Introduction to Bloodstain Pattern Analysis. SIAK-Journal – Journal for Police Science and Practice. 2012, 2, 51–57. (9) Clingolani, M.; Froldi, R.; Mencarelli, R.; Rodriguz, D. Analytical Detection and Quantitation of Strychnine in Chemically Fixed Organ Tissues. Journal of Analytical Toxicology 1999, 23 (3), 219–221. (10) Kodikara, S. Strychnine in Amoxicillin Capsules: A Means of Homicide. Journal of Forensic and Legal Medicine 2012, 19 (1), 40–41. (11) Draper, W. M. Optimizing Nitrogen-Phosphorus Detector Gas Chromatography for Pesticide Analysis. J. Agr. Food Chem. 1995, 43 (8), 2077–2082. The instructors teaching the CEM 1620–ENL 2015 Interdisciplinary Learning Community courses have had multiple discussions both about defining the term “unique insight” and defining what such insight would look like in an original fictional mystery with a supporting Appendix on accurate scientific evidence. While many strong projects have been submitted, none of them have taken the integration of science and literature to the next plane—detailing a comprehensive and clear interpretation of a complicated problem. The rubrics for the crime story and presentation were provided by Linda Holt, Ph.D., Associate Dean of the College of Liberal Arts and Social Sciences and Professor of English at Belmont University. Illinois State Board of Education. Science Rubric. https://www.isbe.net/Documents/rubric.pdf (accessed April 1, 2019). The rubric for the Interdisciplinary Learning Community assessment was provided by Rachel Rigsby, Ph.D., Professor of Chemistry at Belmont University.

136 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Chapter 8

Interdisciplinary Learning Activity Incorporating Forensic Science and Forensic Nursing Downloaded via MIAMI UNIV on November 5, 2019 at 02:21:33 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Amanda S. Harper-Leatherman1,* and Linda N. Roney2 1Department of Chemistry and Biochemistry, Fairfield University,

1073 North Benson Road, Fairfield, Connecticut 06824, United States 2Egan School of Nursing and Health Studies, Fairfield University, 1073 North Benson Road, Fairfield, Connecticut 06824, United States *E-mail: [email protected].

Incorporating forensic science into chemistry or other science laboratories can help motivate students to learn and can build connections to an interdisciplinary career field. However, a complete picture of the interdisciplinary, collaborative aspects of forensic science is not always possible through individual experiments. Nursing students have a need to learn about forensic science as well. Many times, patients coming into emergency rooms may be victims of crimes, and health care professionals can be the first to have the opportunity to collect and preserve evidence. Despite their importance, it is not always possible to fit forensic science learning objectives into the undergraduate nursing curriculum. This chapter outlines an interdisciplinary extracurricular learning activity on forensic science in the health care setting that was attended by 14 nursing major and science major students. Students learned background information, cared for a simulated mock crime victim in a simulated hospital room, collected fiber evidence, and analyzed the fiber evidence in the laboratory using three chemistry techniques over the course of three hours. Both nursing and science students gained experience in chemical analysis and patient care that would not normally be possible. The surveyed participant students self-reported an increased ability to recognize and document evidence and/or potentially crime-related injury on a person in a health care setting and an increased ability to collect and preserve the evidence. This type of interdisciplinary activity can be used as a model for incorporation of similar activities at other academic institutions.

© 2019 American Chemical Society Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Introduction and Background Using forensic science to teach chemistry concepts and content is a popular method with many literature references documenting the various approaches used (1–3). Active learning and laboratory experiments utilizing forensic science concepts are especially common to help students work in the lab while solving a “mystery (4–9).” This approach helps build student motivation for learning about science and helps connect scientific material with the real world and with other disciplines very easily. Students learn the scientific content and also gain some knowledge about an interdisciplinary career field. However, science experiments alone cannot demonstrate the truly interdisciplinary and collaborative nature of the forensic science field. In the nursing field, there are a wide variety of topics that must be covered in the curriculum to meet accreditation requirements as well as to prepare students to take the National Council on Licensure Exam for Registered Nurses. Forensic nursing is not a required topic, but it is often included in the nursing curriculum in some way. Typical examples of forensic nursing in university curricula include situating it in either graduate program education (10) or continuing education programs for hospital-based nurses (11). Education about forensic nursing in undergraduate programs has received little attention (12). However, regardless of practice setting, many graduates of undergraduate nursing programs will take an active role working with patients, families, and communities as part of potential patient-related forensic investigations. Whether working with clients in cases of suspected child abuse, caring for gunshot victims in trauma rooms, or identifying signs of abuse in the elderly, graduates of bachelor’s of science in nursing programs play a significant role in providing holistic nursing care to vulnerable patients. Part of this care may involve collecting forensic evidence from a patient, documenting assessment findings, and collaborating with social workers, police officers, and state agencies. As important as these skills are to develop, clinical partners often have policies that limit or potentially restrict students from working with the type of patients who may be victims of crimes, limiting a student’s ability to develop necessary clinical forensic skills as part of his or her undergraduate nursing student experience. When considering whether to include forensic nursing in the undergraduate curriculum, the essential learning objectives are to help the students recognize, document, and collect evidence properly from a patient in a health care setting. Nurses are not involved in the scientific analysis of the evidence they collect. However, gaining basic knowledge about the scientific techniques used to analyze the evidence can help nurses realize the importance of their role in the investigative process. Understanding their role will improve their performance in forensic nursing–related tasks. Although nursing students take both chemistry and biology laboratory classes in the undergraduate curriculum, the authors cannot find any published laboratory experiments specifically related to forensic nursing scenarios that may help students appreciate the significance of proper evidence treatment. By developing a simulated crime scenario that includes nursing care of a simulated victim and subsequent corresponding forensic chemistry laboratory experiments in which both undergraduate nursing students and science students participated as an extracurricular event, we helped to enhance the interdisciplinary learning of both populations of students. The nursing students not only learned to recognize, document, and collect evidence on patients but also learned how to analyze the evidence scientifically. The goal was to help the nursing students feel more grounded in the importance of evidence collection by seeing the full spectrum of implications and how this evidence is later used to make a determination about a particular case. The science students not only learned to analyze forensic evidence, but also learned about necessary patient care procedures and the medical 138 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

realities related to crime, thus revealing the true collaboration between a variety of professions in forensic science. Teaching chemistry experiments with the added motivation of assisting in a mock crime investigation is always motivating for students. In addition, working with a robotic patient simulator added more realism to the mock crime. All of the students involved in this interdisciplinary learning experience benefitted by becoming more informed and aware with regard to health care, law enforcement, the scientific method, and the interdisciplinary and collaborative nature of forensic science. In addition, the two populations of students offered a unique opportunity for interdisciplinary education as the students taught each other about their disciplines.

Interdisciplinary Learning Activity Description Goals In designing this activity, we had specific learning objectives in mind for the students. We wanted each group of students from either nursing or science to gain experience in the other discipline represented. In other words, we wanted the nursing students to learn to analyze evidence scientifically, and we wanted the science students to learn necessary patient care procedures so that each group would leave the event with a clearer picture of the collaborative and interdisciplinary nature of forensic science. We wanted all of the students to build confidence in the proper steps to take when encountering a potential victim of a crime. In addition, we wanted them to recognize evidence and/or potentially crime-related injury on a person in a health care setting, to document evidence and/or potentially crime-related injury on a person in a health care setting, and to collect and preserve evidence on a person in a health care setting. We hoped the activity would help students understand the interdisciplinary work involved in the field of forensic science. In addition to content knowledge, we also thought the event would help students gain some skills in terms of learning to interact with and teach students from another discipline about topics of their expertise. Student Participants The goal was to have nine science major students (at the sophomore, junior, or senior level in order to have enough lab experience) and nine nursing major students (at the junior or senior level in order to have enough patient care experience) take part in the event. This number of students was chosen to be manageable in terms of working in the simulated hospital rooms and chemistry laboratory, with three groups of six each (three nursing students and three science students per group) in three different simulated hospital rooms and split up into three different stations within one chemistry laboratory. Due to some last-minute cancellations, 14 students ended up participating in the event. Nine of the students were nursing students, with seven of the nine being seniors and two of the nine being juniors. Five science students participated, three of whom were biology majors, one of whom was a physics major, and one of whom was a biochemistry major. One of these science students was a senior, two were juniors, and two were sophomores. The three groups of these 14 participants consisted of three nursing students per group and one to two science students per group. Facilitators The two coauthors of this chapter facilitated the event, and gave a mini-lecture on forensic science and forensic nursing, respectively. In addition, guest speakers from the university’s Department of Public Safety and the local police department also spoke. Members of the 139 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Department of Public Safety acted as police officers during the mock hospital work and laboratory work, taking collected evidence from the simulated hospital and then submitting the collected evidence to the simulated crime laboratory along with reference samples from the mock suspects’ blankets and the mock victim’s clothing. Graduate nursing students assisted in the simulated hospital rooms. Two graduate nursing student volunteers worked in each of the three simulated hospital rooms, with one student acting as the mother of the mock victim within the room and one graduate student giving voice to the robotic patient simulated child using a walkie-talkie while sitting in an adjacent room. However, it should be noted that a high-fidelity patient simulator is not necessary as a trained actor, also known as a standardized patient, could also be used. The graduate students served the dual purpose of also advising the undergraduates as needed if the nursing professor facilitator was not available. Upper-level chemistry or biochemistry majors assisted in the chemistry laboratory with one student assigned to each experimental station to assist with the lab work. Pre-Event Information for Students and Surveys Table 1. Select Pre-Event and Post-Event Survey Questions Used for Student Learning Assessment

Prior to the event, the participating students were assigned to a “learning community” within an online learning management system to have access to material to review and surveys to complete. 140 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

The posted material consisted of introductory videos of each of the coauthors as the leaders of the event within the different event venues, a list of terms to be familiar with, and the pre-event survey link for students to view and complete before the event (Table 1). There was also a folder with information disseminated during the event such as the video report on the simulated mock crime and victim (outlined in a section below), the list of assigned student groups, a forensic nursing tip sheet, instructions for the students during the hospital simulation, PowerPoint presentations (with background information presented at the beginning of the activity as described below), and laboratory experiment instructions. After the event, the post-event survey link was posted for students to complete (Table 1). Overall Structure of Activity The interdisciplinary learning activity was an extracurricular voluntary event occurring one evening for three hours in both the Egan School of Nursing and Health Studies and the Bannow Science Center at Fairfield University. As shown in Table 2, the first hour began in a classroom with an introduction and background discussion. Table 2. Schedule of Activities within the Three-Hour Interdisciplinary Learning Activity Event Order of Activities

Activity

Length of Activity (hr)

Location

1

Introduction and background discussion

1

Classroom

2

Care of mock victim and collection of evidence

0.5

Simulated hospital room

3

Chemical analysis of collected evidence

1

Chemistry laboratory

4

Activity debrief

0.5

Classroom

In order to provide the necessary background before beginning the mock crime simulation, the coauthors held an introductory discussion about forensic science and forensic science in the health care setting with help from guest speakers from the Fairfield University Department of Public Safety and the Fairfield Police Department. PowerPoint slides were used and were also shared with the students within the online learning management system for future reference. Forensic science was broadly defined and then the field of criminalistics or lab sciences was defined within the broad field of forensic science. Students learned the general steps for processing a crime scene with an emphasis on the importance of proper collection and preservation of evidence, maintaining a chain of custody, and obtaining reference samples for comparison with evidence samples. To give students some background on different types of evidence, the concepts of class evidence and individual evidence were also introduced and defined. Finally, fiber evidence was discussed in detail, and the experiments that students would conduct to analyze fiber evidence collected from the mock crime victim were introduced. The director of the Fairfield University Public Safety Department and a Fairfield Police Department detective then followed with specifics on how they are involved at crime scenes, how a mock crime such as the one used in this event would be handled, and how they would interface with health care professionals. The introductory presentations were then completed by explaining the role 141 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

of nurses in recognizing, documenting, collecting, and preserving forensic evidence or signs of injury on potential crime victims through discussions of real-life examples. The mock crime for the evening was then introduced through a video that depicted an emergency room nurse relaying information about the scenario and the admitted mock victim. During the next half hour, students went to the hospital simulation area within the Egan School of Nursing and Health Studies and assumed roles as health care providers as described below. Each group of students cared for their mock victim in a separate simulated hospital room and collected evidence. After evidence collection was complete, the students went into a chemistry laboratory for about an hour and switched roles from health care providers to forensic scientists in a simulated crime laboratory. The students continued working within the three small groups, and the groups rotated between three 20 minute experimental stations to analyze the samples with three different methods over the hour as will be described in more detail in a subsequent section. For the last half hour of the event, the students proceeded back to the classroom to debrief on the activity. Each coauthor led a discussion on how things went in the simulated hospital rooms and the laboratory, respectively, and students shared their thoughts and ideas. Students also reported the conclusions they came to from the fiber analysis, and it was emphasized that the fibers were just one piece of evidence that did not individualize a single suspect due to the mass production of fibers in society. More evidence would be needed in real life to link a suspect to this mock crime. Students were asked to complete a post-event survey afterwards, the results of which are outlined later in this chapter. Mock Crime Scene Scenario The mock crime scenario began with a witness observing someone sitting with a young toddler wrapped in a blanket in the campus dining area. The next time the witness looked over, the child was alone, crying, and sitting under the table without the blanket. Fairfield’s Department of Public Safety was called, and the child was transported to the hospital’s emergency department. The patient matched the description from a recent America’s Missing: Broadcast Emergency Response (AMBER) alert. Later that evening, someone else called the Department of Public Safety saying that there was a suspicious-looking van driving around campus. Two vans were then located that matched the description, and they each had a blanket inside. The blankets were collected for submission to the forensic lab so that reference fibers from the blankets could be compared with loose evidence fibers that would be collected off the child’s clothing in the hospital. The scenario was provided to the students as a video of an emergency room nurse reporting on the situation to the students. She reported, “The patient’s mother left her unattended in the shopping cart at the grocery store at 1500 today. When she returned to the cart, Sophia was not there. Her mother screamed for help; the store manager immediately locked all of the doors to the store and called 911. Sophia was not found in the store and the Fairfield police activated an AMBER ALERT. At approximately, 1800, a Fairfield University Public Safety Officer found Sophia hiding underneath a table at the dining hall and brought her to you for an evaluation in the Pediatric Emergency Department of the Egan School of Nursing and Health Studies.” The nurse then made a recommendation on care to the students within the video stating, “I would recommend that you proceed with your care of this child, recognizing the typical growth and development parameters and responses for a 20-month-old child and anticipating those for a child who has been separated from her parent. Be prepared to collect forensic evidence from the patient as needed.”

142 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Assigned Student Roles Within the simulated hospital room, the nursing students were assigned to assess the patient’s health (vitals, head-to-toe assessment, and pain assessment), provide wound care or medication as needed, and document the assessment. The nursing students were also asked to provide therapeutic communication to the mock patient’s mother and explain what they were doing to the science students to help bring them into the process. The science students were to observe and learn about the medical procedures and were also given the responsibility to search for loose evidence fibers on the patient, collect and preserve the evidence fibers in an available box properly, and give this collected evidence to Public Safety officers. The officers stood outside the rooms acting as police officers might act in a real hospital to obtain and deliver evidence to the crime laboratory. Within the simulated crime laboratory, the science students took more of a lead role with the nursing students learning from them. At each station, all of the group members were taught the technique by the volunteer chemistry student teaching assistant, and then each student had a chance to test a piece of evidence or a standard reference sample. Students recorded all of their observations for comparison in an electronic lab notebook document using the university-supported app, Quip, which students could easily access from their smartphones or other electronic devices.

Hospital/Clinical Simulation with Evidence Collection Although it is an essential skill in many clinical practice areas, nursing students are often not permitted to participate in the care of patients who are undergoing a forensic investigation. Once they graduate, pass their registered nurse licensure exam (the National Council on Licensure Exam for Registered Nurses), and are hired to work at a facility, the expectation is that they will learn how to collect evidence on the job. Given the potential implications of such a practice, the coauthors included the process of evidence collection in a hospital setting as part of the simulation. During the introductory portion of the event, the students were oriented to documentation in the electronic health records used in health care facilities and on required state evidence forms (e.g., the state sexual assault evidence collection kit). Chain of custody and examples of this application in the clinical setting were provided. For example, in a hospital emergency department, if a nurse collects evidence from a patient, the ideal situation is to hand off the evidence to a police officer as the kit can never be left unattended. There are times when this is not possible, and the nurse must follow institutionspecific protocols for securing the evidence in a locked evidence refrigerator without breaking the chain of custody. Typically in the hospital setting, as mentioned previously, the state sexual assault evidence collection box is the only formal kit that nurses regularly use to collect evidence in the clinical setting. Jurisdictions and clinical agencies may have their forms to complete for evidence collection and documenting the chain of custody, but the actual receptacle for transporting the evidence may only be a paper bag labeled with the patient’s identifying information. To standardize the process and make sure that evidence collection was transpired in a uniform and systematic way within this student simulation, white, hinged paperboard boxes (8.5 × 5 × 2.25 in.) were purchased for the participants to use for evidence collection. In addition, hospital-grade forceps, disposable bandage scissors, and an envelope were provided inside the box to aid in evidence collection. An identification label (Figure 1) was also included and affixed to the top of the box for the students to complete before providing the simulated evidence to the Public Safety officer in the same type of hand off that takes place in the clinical setting.

143 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Forensic Laboratory Analysis of Fiber Evidence The introductory portion of the event included background on fiber evidence (13). It was discussed that while ripped fabric samples may have some individualizing characteristics or characteristics that are unique to only one source, single fibers are considered class evidence in that their characteristics are common to all fibers of the same type. This makes it possible to associate each fiber only with a group (fabric type, e.g., cotton, polyester, etc.) and not a single source. It was discussed that class evidence is still valuable in an investigation as it can corroborate other evidence or lead to other more individualizing evidence that can be associated with a single source. Fibers were also introduced as organic polymers, either synthetic or natural, each type with unique chemical structures as well as physical and chemical properties. Example structures of different polymeric fibers were shown. The students were told that three experiments would be completed to help compare the types of fibers of the reference samples from the suspects’ blankets and the mock victim’s clothes to the evidence samples of the fibers collected off of the mock victim’s clothes that may link to one of the suspects’ blankets. The experiments were a burn analysis, a differential staining analysis, and an infrared spectroscopic analysis. Each of the 20 min experimental stations gave students different laboratory experience and helped clarify for students the difference between presumptive (screening) and confirmatory tests. The infrared spectroscopic analysis is the most confirmatory of the three tests and the most likely to be used in a forensic laboratory for fiber identity confirmation. In the laboratory, all of the students wore gloves and safety goggles, long pants, lab coats, and closed-toe shoes for safety. Any long hair was tied back, and no eating or drinking was allowed. All students participated in the experiments equally by being in charge of testing different fibers using the different experiments, but the science students took the lead, having more familiarity with lab work, and helped the nursing students within their groups as necessary. Burn Analysis Reagents and Supplies Standard reference fiber samples of cotton, wool, nylon, and polyester were pulled from different sections of the Multifiber Fabric, Style #43 from Testfabrics Inc. (14) A Bunsen burner flame and metal forceps were used. Experimental Procedure The procedure was adapted from Flinn Scientific Forensics of Fibers Student Laboratory Kit, FB2022. Students were directed with the assistance of the senior student lab assistant to use forceps to hold a standard reference cotton fiber over a Bunsen burner flame, within the flame, and then outside of the flame, making observations at each step. A beaker of water was kept nearby in case a flame needed to be extinguished. The students looked for charring (indication of natural fiber) or melting (indication of synthetic fiber). Students also observed rate of burning, any smoke produced, and other unique characteristics. Standard reference sample fibers of wool, polyester, and nylon were tested in the same way. Then an evidence fiber collected from the mock victim was tested. Finally, reference fiber samples from the known suspects’ blankets and the mock victim’s clothes were tested. Students recorded all of their observations for comparison in a table within an electronic lab notebook using the university-supported app, Quip, which students could easily access from their own smart phones or other electronic devices. Example data is shown in Table 3.

145 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Table 3. Example Data from Burn Fiber Analysis Experiment Observations When Heated

Sample

Observations When Observations When Removed from Ignited (Odor? Curling? Ash? Smoke? Flame (Melt? Rate?) Drip? Residue?)

Reference Cotton

Nothing

White smoke, fast rate Nothing

Reference Wool

Begins to burn and turn brown

White smoke, slow rate

Black residue

Reference Nylon

Nothing

Curling, fast rate, disappears quickly

Residue on tweezers

Reference Polyester

Nothing

Flame grew and then disappeared quickly, smoke

Smoke (black)

Possible Fiber Identification

Evidence Fiber from Mock Victim

Black smoke

Polyester

Reference Fiber from Nothing Mock Victim’s Clothing

Ignites and burns quickly and entirely

Cotton

Reference Fiber from Curling Suspect #1’s Blanket

Curls away, quick rate

No residue

Nylon

Reference Fiber from Nothing Suspect #2’s Blanket

Black smoke

Black smoke

Polyester

Differential Staining Analysis Reagents and Supplies Students utilized the two stains (#1 and #3A) available from Testfabrics Inc. (14) that differentially stain different fiber types different colors based on the different functional groups and charges on the different polymeric fibers interacting with the dyes differently. One inch wide strips of 4 inch long Multifiber Fabric, Style #43 from Testfabrics Inc. (14) were cut and used as reference standards. The Multifiber Fabric contains 13 different commercial fibers woven together. Experimental Procedure The procedure was adapted from the Laboratory Manual for Forensic Science from the National Science Foundation–Sponsored Center for Workshops in the Chemical Sciences 2007 Forensic Science Workshop (15). Approximately 200 mL of each stain mixture was prepared according to Testfabrics Inc. instructions in two separate beakers on hotplates in a fume hood by the senior chemistry lab assistants prior to the students arriving in the lab. Students were then directed to use forceps and a stirring rod to dip a standard Multifiber Fabric strip into the beaker of the warm #1 stain mixture and to dip a separate standard Multifiber Fabric strip into the beaker of the warm #3A stain mixture. The strips remained in the stain mixtures for approximately 2–5 minutes. The strips were then removed, rinsed in a beaker of water, rinsed under running hot water, and then left to dry 146 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

on a paper towel. Students then repeated the process with an evidence fiber from the mock victim, a reference fiber from the mock victim’s clothes, a reference fiber from suspect #1’s blanket, and a reference fiber from suspect #2’s blanket, making sure to hold the fibers with the forceps while immersing so as not to lose the fibers. The colors of the dyed fibers were then compared to the colors on the dyed Multifiber Fabric strips to identify the type of fabric. While one stain may have dyed two different fibers the same color, using the two stains helped differentiate the different fiber types by comparison with the colors on the standard dyed Multifiber Fabric strips. Students took photos of the fibers and standards and uploaded these photos into their electronic lab notebook documents on Quip for documentation. Examples of the colors formed on the standard Multifiber Fabric strips when dyed with stain #1 and stain #3A are shown in Figure 2.

Figure 2. Examples of the colors formed on the standard Multifiber Fabric strips, Style #43 from Testfabrics Inc., when dyed with stain #1 and stain #3A compared to an undyed strip. Infrared Spectroscopic Analysis Reagents and Supplies A Bruker alpha platinum-attenuated total reflectance (ATR) Fourier transform infrared (FTIR) instrument was used with OPUS Version 7.2 software loaded with a FDM FTIR Polymers and Polymer Additives Spectral Database. Standard reference fiber samples of cotton, wool, nylon, and polyester were pulled from different sections of the Multifiber Fabric, Style #43, from Testfabrics Inc. (14) 147 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Experimental Procedure

Figure 3. Example fiber ATR-FTIR spectra of nylon and polyester fibers. Taking advantage of the fact that different polymeric fibers absorb different wavelengths of infrared radiation depending on the exact chemical structure of the fiber, students were directed to take spectra of each reference suspect blanket fiber, the reference fiber from the mock victim’s clothes, and the evidence fiber collected off of the mock victim’s clothes. The students took spectra of standard reference cotton, wool, nylon, and polyester fibers as well. Fibers were simply laid on the ATR crystal and held down with the lever before taking spectra. The spectra were searched against a commercial library of standard polymer spectra to make a determination of fiber type and identity. Images of the spectra were copied and pasted into the electronic lab notebook for comparison. Example spectra of two fibers are shown in Figure 3. The synthetic fibers of nylon and polyester had

148 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

matches within the commercial library of standard polymer spectra. The polyester fiber specifically matched a poly(ethylene terephthalate) spectrum in the library, which is another name for polyester.

Student Learning and Assessment

Figure 4. Pre-survey and post-survey results of students participating in the interdisciplinary learning activity. Survey questions shown in detail in Table 1. Ten students completed each survey.

Figure 5. Pre-survey and Post-survey results of control group of students who did not participate in the interdisciplinary learning activity. Survey questions shown in detail in Table 1. Seventeen control students took the pre-survey, and 12 control students took the post-survey. To assess how well the event helped students progress in relation to the learning objectives, the authors administered anonymous online surveys following approval of the Fairfield University Institutional Review Board (Protocol #0540-2018) for this project (select questions shown in Table 1). The questions were essentially the same for both the pre-event and post-event surveys, with 149 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

slight wording changes as shown in Table 1. Other survey questions were included to measure nursing students’ ability to analyze evidence scientifically and science students’ ability to follow necessary patient care procedures, but the results were not clear due to a survey error in identifying the responses from each type of student. In addition to administering surveys to event participants, surveys were also administered to a control group of students who did not attend the event, but who either came from an Introduction to Forensic Science course or a Pediatric Nursing course. Ten participating students completed the pre-event survey and 10 participating students completed the post-event survey, and these results were compared graphically (Figure 4). Seventeen control students took a pre-survey and 12 control students took a post-survey, and these results were compared graphically (Figure 5). When asked to rate their confidence in the proper steps to take when encountering a potential victim of a crime in a health care setting, half the students attending the event reported no or low confidence and half reported medium confidence before the event. After the event, 90% of the students reported medium or high confidence. About 70% of the control group of students reported no or low confidence both before and after the event. When asked to rate their ability to recognize forensic evidence and/or potentially crime-related injury on a person in a health care setting, 40% of the students attending the event reported low confidence and 60% reported medium confidence before the event. After the event, 90% of the students reported medium and 10% reported high confidence. About half of the control group of students reported no or low confidence both before and after the event. When asked to rate their ability to document forensic evidence and/or potentially crime-related injury on a person in a health care setting, 60% of the students attending the event reported no or low confidence and 40% reported medium confidence before the event. After the event, 90% of the students reported medium confidence. About 60% of the control group of students reported no or low confidence before and 75% reported no or low confidence after the event. When asked to rate their ability to collect and preserve forensic evidence on a person in a health care setting, 70% of the students attending the event reported no or low confidence and 30% reported medium confidence before the event. After the event, 90% of the students reported medium confidence. About 65% of the control group of students reported no or low confidence before and 50% reported no or low confidence after the event. When asked to rate their opinion about how much interdisciplinary work forensics involves, 10% of the students attending the event reported a moderate amount of interdisciplinary work and 90% reported a lot of interdisciplinary work before the event. After the event, 100% of the students reported a lot of interdisciplinary work. About 18% of the control group of students reported a moderate amount of interdisciplinary work and 82% reported a lot of interdisciplinary work before and 100% reported a lot of interdisciplinary work after the event. When asked to freely respond about the most significant thing learned from the event, some themes came through in the participant responses. Four different students mentioned learning the proper methods for identifying, collecting, or preserving evidence in responses such as “how sensitive collecting evidence from patients is” or “how to properly identify evidence.” Three different students mentioned ideas related to the interdisciplinary, collaborative nature of forensics in responses such as “interdisciplinary team skills,” “juggling the responsibilities and respecting each member of the team while collecting evidence and giving care,” and “how to work with those not in the health care setting such as police.” When asked to freely respond to what was the most confusing part of the event or what questions still remain unanswered, one nursing student mentioned the lab work, and some of the responses from the science students related to the nursing work, such as “a lot of the nursing protocols before 150 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

the evidence was collected” and “I didn’t really understand the nursing aspect, granted I’m not a nursing major.” However, some of the responses just related to knowing general roles to play and where to find things within the simulated hospital room, such as “knowing when to jump in during the simulation (what my role was)” and “the beginning of the simulation because I really had no clue what I was doing.” The students were gaining exposure to the other discipline and were learning as they were going, which was natural in this type of one-time short event. When asked if the students would suggest any changes to the event that would have helped them learn more, some students did not have suggestions, but some students brought up the timing of the event as the event did end up going over time due to the initial introductory presentations taking a bit longer than originally planned. A few students also asked for clearer instructions and roles that can be clarified more the next time the event is run. However, some things being a little unknown is not always a bad thing as it forces the students to work together and figure things out. In addition, having the assistants available in both the simulated hospital rooms and the lab was definitely key in keeping things running smoothly. One student also offered a good suggestion to record a small group in the simulated hospital and the lab and then watch key parts of the recording for the purposes of discussion and learning. When asked for additional comments related to the event, many positive responses were given, such as how much the event was enjoyed, how much was learned, and how much the students enjoyed the participation of the Department of Public Safety and police officers as well. When asked if students would participate in another extracurricular event related to forensic health care, all agreed that they would. In addition, 70% of the students attending the event said they would take a course related to forensic health care if they had the opportunity. This question was also asked of the control group of students, and 75% of these students said they would take a course related to forensic health care if they had the opportunity.

Alternative Forensic Laboratory Analyses The time limitations of the event made it prudent to choose one type of evidence to analyze so that students could focus on one collection in the simulated hospital rooms and analysis of one type of evidence in the laboratory for simplicity and to avoid confusion. Analyzing one type of evidence with one longer experiment would be another alternative to setting up shorter experiments to be rotated. Working on short experiments and rotating through them seemed ideal for this type of onetime evening event to keep student attention and interest. Fibers as a type of evidence to analyze worked well with three short experiments. Solubility tests of the fibers would be another possibility for an alternate station to those described. In addition, an optical microscopy station or a scanning electron microscopy station could also be set up if instrumentation was available (16). In addition to fibers, another piece of physical evidence that could lend itself well to short experimental stations would be unknown powder or drug analysis. Simple wet chemical tests in well plates for pH, solubility, or reactivity could be set up as one station and thin-layer chromatography would be another station possibility. Optical microscopy and ATR-FTIR would also be possible short experimental stations, in addition to gas chromatography if a quick chromatography method were designed. Biological samples could also be focused on and could be analyzed with short stations easily if simulated samples were purchased and used. Simulated blood typing kits are available for microscopic analysis, as well as simulated urine analysis using thin-layer chromatography for heavy metal poisoning such as those from Ward’s Science or Flinn Scientific. Blood spatter on a victim or suspect might also be documented in a hospital and analyzed. A DNA analysis is a longer-term 151 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

project, but could still be possible to incorporate if not all steps in the process were focused on and perhaps only the last electrophoretic results were completed. Otherwise, three short DNA stations could be set up with simulated samples if students prepared samples for polymerase chain reaction (PCR) at a first station, then were given samples that had already gone through PCR to set up in gel electrophoresis chambers in a second station, and then were given completed gel electrophoresis results to analyze in a third station. BioRad sells a simulated DNA profiling kit known as the Crime Scene Investigator PCR Basics Kit.

Conclusion Designing an interdisciplinary extracurricular learning activity on forensic science in the health care setting made it possible to expose science students and nursing students to forensic science as it relates to care of a patient and as it relates to the scientific analysis of collected evidence. In one evening, through the collaboration of a School of Nursing and Health Studies, a Chemistry and Biochemistry Department, and a Department of Public Safety, students experienced the full picture of how different professions work together to answer crime-related questions. Due to heavy course loads and other extracurricular activities, science and nursing students do not often have the chance to learn together what may be relevant to both types of students. This type of extracurricular event gives students a chance to enhance their learning with a relatively small time commitment. Despite the brevity of the event, student participants reported increased confidence in processing a potential crime victim in a health care setting, increased ability to recognize and document evidence and/or potentially crime-related injury on a person in a health care setting, and increased ability to collect and preserve evidence. This type of event could easily be implemented at other universities as many universities will have chemistry departments, schools of nursing, and public safety departments.

Acknowledgments The authors wish to thank the Marion Peckham Egan School of Nursing and Health Studies Simulation Grant, the Department of Chemistry and Biochemistry, and Fairfield’s Honors Program for the funding and support of the extracurricular learning activity. We would also like to thank the many colleagues who specifically helped with and supported our event, including Professor Susan Reynolds, Dr. Michael Andreychik, Dr. John Miecznikowski, Dr. Eileen O’Shea, Dr. L. Kraig Steffen, Dr. Dorothy Sobczynski, Dr. Matthew Kubasik, and Dr. Meredith Kazer. We would like to thank Director Todd Pelazza and the many Department of Public Safety officers at Fairfield University for their support and involvement in our event. We thank Fairfield Police Detective Fred Caruso for speaking at our event. We are indebted to our undergraduate senior chemistry major lab assistants and graduate nursing assistants. Finally, we would like to thank all of the students who took part in the event for their time and helpful feedback.

References 1.

2.

Kanu, A. B.; Pajski, M.; Hartman, M.; Kimaru, I.; Marine, S.; Kaplan, L. J. Exploring Perspectives and Identifying Potential Challenges Encountered with Crime Scene Investigations When Developing Chemistry Curricula. J. Chem. Educ. 2015, 92, 1353–1358. Ahrenkiel, L.; Worm-Leonhard, M. Offering a Forensic Science Camp to Introduce and Engage High School Students in Interdisciplinary Science Topics. J. Chem. Educ. 2014, 91, 340–344.

152 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

3.

4.

5. 6. 7. 8.

9. 10. 11. 12. 13. 14. 15. 16.

Charkoudian, L. K.; Heymann, J. J.; Adler, M. J.; Haas, K. L.; Mies, K. A.; Bonk, J. F. Forensics as a Gateway: Promoting Undergraduate Interest in Science, and Graduate Student Professional Development Through a First-Year Seminar Course. J. Chem. Educ. 2008, 85, 807–812. Testa, S. M.; Selegue, J. P.; French, A.; Criswell, B. Permanganate Oxidation of DNA Nucleotides: An Introductory Redox Laboratory Framed as a Murder Mystery. J. Chem. Educ. 2018, 95, 1840–1847. Parker, P. D.; Beers, B.; Vergne, M. J. What Is in Your Wallet? Quantitation of Drugs of Abuse on Paper Currency with a Rapid LC-MS/MS Method. J. Chem. Educ. 2017, 94, 1522–1526. Valente Nabais, J. M.; Costa, S. D. A Forensic Experiment: The Case of the Crime at the Cinema. J. Chem. Educ. 2017, 94, 1111–1117. Frederick, K. A. Using Forensic Science to Teach Method Development in the Undergraduate Analytical Laboratory. Anal. Bioanal. Chem. 2013, 405, 5623–5626. Elkins, K. M.; Kadunc, R. E. An Undergraduate Laboratory Experiment for Upper-Level Forensic Science, Biochemistry, or Molecular Biology Courses: Human DNA Amplification Using STR Single Locus Primers by Real-Time PCR with SYBR Green Detection. J. Chem. Educ. 2012, 89, 784–790. Specht, K. M.; Boucher, M. A. A Forensic-Themed Case Study for the Organic Lab. J. Chem. Educ. 2009, 286, 847–848. Simmons, B. Graduate Forensic Nursing Education: How to Better Educate Nurses to Care for This Patient Population. Nurse Educ. 2014, 39, 184–187. Yoo, Y.; Cha, K.; Cho, O.; Lee, S. Emergency Department Nurses’ Recognition of and Educational Needs for Forensics Nursing Education. Korean J. Adult Nurs. 2012, 24, 499–508. Freedberg, P. Integrating Forensic Nursing into the Undergraduate Nursing Curriculum: A Solution for a Disconnect. J. Nurs. Educ. 2008, 47, 201–208. Saferstein, R. Criminalistics: An Introduction to Forensic Science, 12th ed.; Pearson: Boston, MA, 2018. TESTFABRICS. http://www.testfabrics.com/ (accessed March 31, 2019). Kaplan, L. J. A Laboratory Manual for Forensic Science; Williams College: Williamstown, MA, 2007. Meloan, C. E.; James, R. E.; Brettel, T.; Saferstein, R. Lab Manual for Criminalistics: An Introduction to Forensic Science, 10th ed.; Prentice Hall: Upper Saddle River, NJ, 2011.

153 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Chapter 9

Drugs and DNA: Forensic Topics Ideal for the Analytical Chemistry Curriculum Downloaded via MIAMI UNIV on November 5, 2019 at 02:23:22 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Ling Huang* Department of Chemistry, Hofstra University, Hempstead, New York 11549, United States *E-mail: [email protected].

Because of the popularity of CSI and other similar TV shows, high school and college students naturally enter chemistry classrooms with a strong interest in the application of instrumental techniques in the context of forensic science. As chemical educators, we can channel that interest into the active learning of analytical chemistry. Drugs and DNA are the two main areas that provide the most employment opportunities in forensic science. They also happen to be two highly relevant topics in the teaching of analytical chemistry. This chapter reveals how real-world drug analysis can be used as a teaching tool in quantitative analysis and instrumental analysis courses. Mimics of controlled substances can be used in analytical lab courses. Many concepts and practices in spectroscopy, spectrometry, and separations can be covered with existing classroom resources on a manageable budget. DNA genotyping involves many separation and spectroscopic techniques that can be used to introduce analytical chemistry applications. These forensic topics can be used to train and inspire students to learn analytical chemistry, expand their practical skill sets, and potentially earn job opportunities in the increasing field of forensic labwork.

Introduction Because of the popularity of CSI and other similar TV shows in the past two decades, many students have become interested in the chemistry of forensic science, specifically the analytical chemistry involved in the collection, analysis, and presentation of crime scene evidence such as blood, saliva, gunshot residue, drug powder, and fibers. Subsequently, many high school science teachers started to use forensic science as a platform to introduce students to STEM fields (1–3). Naturally, many of those students gained further interests that took them to chemistry or forensic science programs in college. This created a tremendous need for tailored curricula and teaching materials for high school and college instructors to properly teach chemistry in the context of forensic science and to seamlessly integrate the teaching needs with existing teaching guidelines for subjects such as general, organic, analytical, or biochemistry. © 2019 American Chemical Society Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

This chapter will focus on using two of the most popular and relevant forensic science topics, drugs and DNA, to teach analytical chemistry in quantitative analysis and instrumental analysis courses. These courses are typically taught in the second, third, or fourth year of college. The examples and covered topics will hopefully benefit both the lecture and lab instructors. Many of the resources provided in this chapter are freely available online or for a very affordable price so even a high school teacher could potentially use these toolkits to teach analytical concepts related to drugs and DNA. All the abbreviations used in this chapter are spelled out in Table 1. The illicit drugs or controlled substances section is usually the largest or core unit in many state, county, or city forensic labs where many chemistry-trained students become employed (4). The current opioid pandemic along with traditionally strong drug law enforcement makes drug analysis a hot field for employment opportunities. Forensic DNA typing is the other area where many current job opportunities exist as jurors and judges nowadays demand DNA analytical results for criminal or civil cases. These two areas have seen rapid growth in the last twenty years with the dramatic evolution of analytical technologies such as mass spectrometry, NMR, miniaturized (sometimes handheld) infrared spectrometers, Raman spectroscopy, and fast separations. All of this makes drug and DNA topics two ideal vehicles to carry out the teaching mission of analytical chemistry. Before we lay out the actual teaching materials on drugs and DNA, let us talk about the common misconceptions among our students regarding forensic drug or DNA analysis. A lot of these misconceptions originate from TV shows or movies in which 45 minutes or less were devoted to a whole episode. The analyst is only given a few hours to finish the analytical project and reach meaningful answers. Real practitioners in forensic science, however, take many hours, days, or sometimes even weeks to find out the identity of a drug or a whole DNA profile that is of rigorous scientific quality so that the results can stand up in court. The painstaking science training to achieve those scientific standards should be the core teaching point in our analytical chemistry courses. Our goals are to correct the misconceptions from some of our forensic-minded chemistry students and to turn them into professional chemists or biochemists through the teaching of authentic analytical techniques used in actual forensic labs. Today in the American legal system, the Frye standard is increasingly being replaced by the Daubert standard in evaluating the admissibility of scientific evidence (4). Consequently, the evidence has to be analyzed with an instrumental analytical method that has been peer-reviewed, validated, and demonstrated to present repeatability, robustness, reliability, and ruggedness. This also means that the data has to be presented with enough statistical analysis such as error rate, precision (often as standard deviation), probability of a random match, sources of errors, and other important contribution factors that can be either inclusionary or exclusionary. In order to meet this stringent demand, many subfields in forensic science have been revamped or questioned. Multiple scandals involving hair analysis, fingerprint analysis, and toolmark analysis have increased the public distrust in the analytical tools of forensic analysis (5, 6). On the other hand, the drug section and DNA section in forensic labs have stood up to scientific scrutiny and have been challenged and improved to be the two fields with the most rigor. These fields also involve the most cases, funding, and job opportunities for students studying analytical chemistry or forensic science. As previously mentioned, the popularity of CSI combined with the strong market demand for drug chemists and DNA technicians naturally bring many students to analytical chemistry or related courses such as quantitative analysis (and corresponding lab classes), instrumental analysis (and corresponding lab classes), forensic biology, and DNA analysis (where a lot of instrumental analytical tools are used). How do you channel these strong interests and enthusiasm into the learning of

156 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

analytical chemistry? Hopefully, the following sections will provide the readers with some insight and inspiration in this field.

Analysis of Illicit Drugs The analysis of illicit drugs usually happens in the biggest section of crime labs (4). The recent national pandemic of opioid addiction and overdoses exacerbates the workload of forensic analytical chemists (7). The illicit drugs that are analyzed are typically Schedule I or Schedule II controlled substances such as heroin, cocaine, opioids, cannabinoids, or methamphetamines, as classified in the Controlled Substance Act (8) of 1970 with constant updates by emergency scheduling by the Drug Enforcement Administration (DEA). Another large area of drug analysis involves analyzing drug metabolites in toxicology labs that support the investigations related to driving with influence of drugs (DWI), driving under the influence of drugs or alcohol (DUI), or cause of death (COD). This second group of investigative activities usually are found in medical examiner’s offices or private or public toxicology labs. The analytical techniques involved in drug evidence analysis include TLC, GC-MS, FTIR, LC-MS, HPLC-DAD, and NMR, almost all of which are covered in the instrumental analysis curriculum. In addition, drug samples or metabolites can be positively identified with colorimetrybased immunoassay. There are additional spectroscopic techniques such as fluorescence, Raman spectroscopy, X-ray crystallography, and ICP-MS that can be used to characterize illicit drug compounds. One difficulty for the teaching labs is that educators require a license to possess and handle controlled substances commonly investigated in real crime labs. This problem can be circumvented with two different strategies. One is to use structurally close or functionally similar compounds, many of which are over-the-counter (OTC) drugs. There have been several published studies (9–11) on using the OTC drugs as substitutes to teach students about spectroscopic analysis of illicit drugs. The second route is to use noncontrolled formulations of controlled substances, which are legal and available for sale through major chemical vendors as substitutes. For example, 1 mg/ mL of methanol solutions of methamphetamine can be found for sale through Sigma-Aldrich legally without the requirement of a license. This type of standard can be used for GC-MS, GC-FID, or LCMS analysis. The solvent can also be evaporated under a vacuum to produce the solid powder for other separation techniques or spectroscopic investigations. According to recommendation IIIB2 from the Scientific Working Group for the Analysis of Seized Drugs (SWGDRUG) (12), there needs to be a combination of two or more instrumental methods in order to confirm the identification of a drug. There are three categories: Category A includes FTIR, MS, NMR, Raman, and XRD; Category B includes CE, GC, IMS, LC, and microcrystalline tests; Category C includes color tests, fluorescence spectroscopy, immunoassay, melting point, TLC, and UV spectroscopy. They recommend that “when a validated Category A technique is incorporated into an analytical scheme, at least one other technique (from either Category A, B, or C) shall be used,” and “when a Category A technique is not used, at least three different validated techniques shall be employed. Two of the three techniques shall be based on uncorrelated techniques from Category B (12).” This created a greater need for students of analytical chemistry, especially advanced students of instrumental analysis, to study thoroughly the previously mentioned instrumental techniques in order to be qualified for a drug analyst job. At the minimum, students should have the opportunity to grasp the concepts of high-use instruments such as FTIR, GC-MS, HPLC, and UV-Vis. In the next section, several examples of instrumental applications are demonstrated in three case studies to indicate the pedagogical value as well as the proximity to realworld drug crime lab scenarios. 157 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Case Study 1: Analysis of OTC Analgesics Many OTC analgesics can be analyzed with TLC, FTIR, GC-MS, and NMR. The students can run standards of acetaminophen, aspirin (or acetylsalicylic acid), phenacetin, ibuprofen, and salicylamide (structures shown in Figure 1) on TLC plates with 0.5% v/v acetic acid in ethyl acetate to learn chromatographic separation basics, particularly normal phase chromatography as the stationary phase is a thin layer of polar silica particles that attracts the more polar analytes. The concepts of “like dissolves like”, competition for analyte between stationary and mobile phases, relative polarity, intermolecular forces, UV-absorbance detection, and retardation factors could be strengthened under the context of this separation science application. After completing the analysis, the students can be asked to analyze an unknown OTC drug mixture to use retardation factors and the color of the spots to identify each separated compound.

Figure 1. Structures of OTC drugs. After TLC experiments, the students can use those standard powder samples to generate ATRFTIR spectra and conduct library searches to find the matching reference spectra. The students can use the peak-picking function in their IR software to identify the specific type of bond vibration (e.g. C-H stretch vs. C=O stretch), and then comment on the width and relative height of the peaks and how the information related back to the structure of the compounds. In addition, the students can test an unknown OTC drug mixture to see if some of these vibrational signals can be isolated for partial identification. If the mixture spectrum is difficult to interpret, the student can discover the shortfall of FTIR in the identification of a drug in the mixture profile. Furthermore, in some of the advanced analytical courses, instructors can introduce simple chemometric techniques, such as principal component analysis (PCA), to help students deconvolute mixture profiles (13). Advanced instrumental analysis lab students can be directed to use GC-MS to run a methanol (or acetyl nitrile) solution of OTC drug extract or unknown mixture for GC separation. Most of the OTC drugs depicted in Figure 1 are compatible with GC analysis. After separation, the students can learn how to use the MS library search to find the drug and potentially use the internal standard calibration method to quantify it. Typically the GC-compatible OTC drugs are present in the NISTFDA or NISTDRUG library. The SWGDRUG library is free and available for download for further investigating an illicit drug if needed. Students can learn the makeup of these libraries. Any internal standard such as the deuterated version of the found compound can be spiked in the drug extract solution and injected into the GC for separation. The students will then learn how to process total ion chromatograms to get extracted ion chromatograms (Figure 2) and how to use peak response factors to calculate concentrations based on integrated peak areas. 158 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Figure 2. Sample toxicological LC-MS TIC and EICs for amphetamines, commonly found in a toxicology lab where urine drug tests are conducted. For an institution with an NMR, the students can be assigned an unknown, yet pure OTC drug compound. The students can then use proton and carbon NMR spectroscopic techniques (1-D or 2D correlation spectroscopy) to generate multiple spectra for the identification of the compound. For these spectrometric techniques, it is invaluable for students to learn library search or different unknown and reference spectral comparisons for identification purposes; although they must also understand the limitation of these approaches through the understanding of hit scores and match percentages. The student can also learn that chromatographic peaks that are not fully resolved and overlapping signals in FTIR and NMR will diminish the quality of these methods. The students can also learn the difference between a destructive method (such as MS) and a nondestructive method (such as FTIR or NMR). During the spectroscopic investigation, the students can practice the determination of a signal to noise ratio, determine the noise level, and learn how to improve the signal to noise ratio by increasing the scan numbers (there is also a numeric relationship). If quantitative NMR is used, the student has the opportunity to study the internal standard method and potentially standard addition method by using the integrated peak areas in proton NMR spectra.

159 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Besides these experiments, there have already been practices in analytical chemistry labs to identify and quantify other OTC drugs such as dextromethorphan using HPLC-DAD and UVVis (14). Besides OTC drugs, caffeine identification and quantification with the same instruments are also very popular choices in lab studies (15). At our institution, we have also identified and quantified nicotine in tobacco products, nicotine gums, and e-cigarette fluids using CE, GC-MS, and HPLC (16). The students can use external standard calibration, internal standard calibration and standard addition methods to generate a linear calibration curve to quantify these drug analytes in the presence of a sample matrix. Besides deuterated internal standards, cheaper alternative such as phenol can also be used in the case of quinine quantification (in the presence of tonic water) (17). In the calibration graph processing, the students can review the concepts of limit of detection (LOD), limit of quantitation (LOQ), limit of linearity (LOL), and dynamic range (18). If multiple separation and spectroscopic methods are used, the students have the opportunity to synthesize the mixed outcomes to arrive at a more solid conclusion, thus mimicking the process of the combined instrumental analytical procedures recommended by SWGDRUG (12). By doing so, the learning experience in an analytical lab can truly reflect the day-to-day operations in a forensic drug analytical lab, which gives students a better sense of utility and authenticity. Case Study 2: Analysis of Controlled Substances

Figure 3. Structural examples of controlled substances. Controlled substances (examples shown in Figure 3) often cannot be acquired without a teaching license from the state or federal government (DEA). Even in labs with the licenses, there are restrictions on using these substances for teaching purposes in most undergraduate institutions. The standard drugs can also be very expensive. There are three solutions to this practical challenge: (1) buy the noncontrolled version; (2) use spectra to substitute; or (3) use something similar. As previously mentioned, although most pure powders of controlled substances require a federal license from the DEA and a state license to legally purchase often at a heavy price tag, teaching labs can still obtain noncontrolled versions of those drugs for instructional purposes. Typically, they come in the form of a 1-mg/mL solution in methanol (with 1-mg total quantity). The deuterated standard of these drugs can also be bought in the same format. These forms are more than adequate for GC-MS or LC-MS investigations thanks to the high sensitivity of MS detectors that can easily detect smaller than ppm levels of analyte (in forensic terms, often referred to as trace analysis). Occasionally for some of these noncontrolled standards, methanol can be evaporated to generate dry solute powder for nondestructive spectroscopic analyses such as NMR. Safety protocols should be strictly followed to minimize the hazard of aerosolization of the powder and inhalation. The students 160 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

should wear facial masks, gloves, lab coats, and goggles. For NMR analysis, the student should dissolve the dry powder in deuterated NMR solvents immediately to minimize risk. For some highly toxic and addictive drugs such as opioids, the student should handle the dried standard in a glove box with the supervision of trained instructors. In the absence of the actual chemicals, spectral analysis can still be performed with existing spectral databases in literature and online sources (18, 19). This could be an excellent resource for practice in an instrumental analysis lecture. For example, the SWGDRUG spectral database contains the ATR-FTIR spectrum of fentanyl, an opioid used in hospital that is also widely abused in the ongoing opioid pandemic in the United States. The spectrum contains peaks linking to C=O stretch, aromatic C=C stretch, and out-of-plane bending signals around aromatic rings. Figure 4 demonstrates an ATR-FTIR spectrum above a Raman spectrum of a synthetic cannabinoid AM2201 from our research lab. Similar signals can be analyzed and used to identify this compound sprayed on a herbal incense that was sold as a designer street drug. Many of these FTIR, NMR, and MS spectra can be downloaded free of charge from SWGDRUG (19) and NIST websites (20) or can be found in published journal papers and books such as the Clarke’s manual (21). Although we cannot replicate an actual lab experience, the students can still learn how to perform peak assignment as part of their spectroscopy curriculum. Many analytical chemistry instructors use safer and noncontrolled compounds as analogues of controlled substances for illustrative purpose. These chemicals typically contain the same major functional groups as presented in illicit drugs. These chemicals often behave similarly in a chromatographic separation or a spectroscopic investigation. Consequently, the associated analytical chemistry concepts can be conveyed adequately so the students can evaluate terminology such as retention time, adjusted retention time, resolution, selectivity factor, theoretical plate height, and plate number. Students can also learn the rationale behind the selection of columns and mobile phase conditions as well as the criteria for MS scanning parameters. Particularly, through GC-MS, the students will learn the different heating programs needed for different kinds of drugs. And sometimes, because it is thermally labile, the analyte may react to form another compound and demonstrate a second peak in the chromatogram.

Figure 4. ATR-FTIR and Raman spectra of a synthetic cannabinoid AM-2201.

161 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Case Study 3: Toxicological Study of Drug Metabolites Analytical chemistry is routinely used in toxicological investigations of COD due to poisoning or drug overdose. The sample matrix in a biological fluid such as urine, blood, or saliva is often very complex. Typically, the toxicology lab starts their direct immunoassay testing of the biological fluid as a presumptive test method, which can produce false positives or false negatives. Subsequently, drug metabolites in the urine, blood, or saliva samples are extracted through organic solvent extraction before GC-MS or LC-MS analysis. Although it might be difficult to use these biological fluids in analytical chemistry labs due to the biosafety concerns, actual toxicology data like the LC-MS EIC chromatograms in Figure 2 along with MS spectra could be used to illustrate the identification procedure. For GC-MS methods, a full scan is usually run to screen for a list of major metabolites in controlled substances. After preliminary identification, individual separation is carried out with the addition of a deuterated internal standard to quantify each identified compound through external and internal standard calibration. For all instrumental analysis of illicit drugs or metabolites, the students should be taught the importance of running positive and negative controls and how to calibrate the instrument with a standard. The appropriate “blanks” should be run to reflect the true background signal. This is especially important when the external standard calibration method is selected. These procedural requirements can be introduced through the presentation of forensic accreditation protocols (22). Using the previous three case studies, readers can hopefully be inspired to utilize the cited resources including lab experiments, instrumental techniques, databases, and other literature to create analytical chemistry curricular modules for lectures and labs. Many of the discussed examples can be expanded or modified for other analytes investigated in a forensic lab. DNA is one of the prime examples and is discussed in the following section.

Forensic DNA Typing and Associated Analytical Chemistry Concepts As previously mentioned, DNA evidence is often requested from both the prosecution and the defense for the inclusion or exclusion of the presence of an individual in most legal proceedings today. The forensic typing of DNA was standardized in late 1990s and matured in the early 2000s. There have been continuous technological breakthroughs in the second decade of this century, and many of the advancements are within the realm of analytical chemistry and specifically bioanalytical chemistry. The whole process from DNA sample collection to identification (Figure 5) can be used to introduce important instrumental analysis concepts and demonstrate their value in real-world applications.

Figure 5. Forensic DNA typing and associated analytical techniques. 162 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Forensic human DNA typing is the most used human identification technology in crime labs. It usually starts with either a swab of the inside cheek of a suspect or a swab of evidence collected at the crime scene. The DNA- containing swab then goes through organic extraction or solid phase extraction (SPE) to recover maximal amount of purified DNA. Subsequently, the recovered DNA molecules are quantified using quantitative polymerase chain reaction (qPCR) utilizing fluorescence quenching as a spectrometric detection technique. Once amplifiable DNA is quantified, a multiplex PCR amplification is performed to produce millions of copies of short tandem repeat (STR) fragments with fluorescent tags that can be used for detection in the final capillary gel electrophoresis (22). As displayed in Figure 5, DNA extraction incorporates SPE, liquid-liquid, or organic solvent extraction techniques. The students can be exposed to the discussion of extraction efficiency, the binding mechanism, the distribution constant, and intermolecular forces used to attract and elute DNA. The solid phase could be silica beads, resins, or magnetic beads. Currently, many forensic labs have adopted robotic DNA extraction equipment to increase efficiency and minimize contamination. The students of instrumental analysis can learn the latest automation technology by watching videos of these robotic platforms (23). In forensic science, there is an increasing demand for “closed” analytical solutions to isolate the evidence from potential contamination sources such as the environmental DNA and the DNA of the analysts. The extraction part usually is not covered extensively in traditional instrumental analysis textbooks. Because of its frontal position in a streamlined analytical process, the importance of extraction shouldn’t be overlooked. DNA extraction is a unique application that can be used to compare and contrast multiple techniques and to give students this opportunity to learn about extraction on a deeper level. The DNA quantification part involves a fluorescence dye-labelled DNA primer attached to a quencher. As the PCR amplification occurs, the quencher tag is cut, and the fluorescence signal is displayed and used to quantify the amplified DNA. In order to quantify the template DNA, threshold cycle number Ct is plotted against log[DNA] to produce a linear external standard calibration curve. The quantification step is necessary to determine if amplifiable DNA is present and the right amount of DNA to be introduced in the next multiplex amplification step. Real-time qPCR and later multiplex PCR both involve positive and negative controls. The positive controls are run with known amounts of starting copies of DNA that are predicted to amplify successfully. This is used to check the quality of the matrix including other important components such as Taq polymerase, bases, salt, and water purity. Negative controls must be included to ensure the absence of contaminant DNA and is often composed of everything but the template DNA. In analytical chemistry curriculum, +/controls are usually not stressed or discussed so this is a good place to bring it up (24). The critical DNA separation step is carried out using capillary gel electrophoresis. Since 2017, the FBI requires forensic DNA labs to amplify and separate 20 loci of STR pieces for human DNA typing (25). This creates a lot of challenges for the effective separation and sizing of the STR fragments and for the multichannel laser-induced fluorescence detection and optical signal separation. Students can learn about the state-of-the-art multicapillary separation instrument as well as the design of the fluorescence tags for multiple optical channel detection. Last, but not least, after the electropherograms are produced, the migration time of each fluorescently labelled STR product is compared to a “sizing ladder,” or a mixture of external standards to determine the number of repeats at each locus before the entire genetic barcode for human identification can be generated. Typically, an internal standard mixture with a different fluorescence tag is also used to calibrate the length of each STR product (24). Students can reinforce

163 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

the concepts of internal versus external standard calibration for quantification purposes (in this case, the number of repeats present in an STR). Most of this case study is intended to provide examples or application notes for analytical chemistry lecture courses. If there is hardware and available time in the teaching lab, some of the DNA typing experiments can be done to demonstrate the actual implementation of the analytical techniques. Some of the experiments can be modified to simplify the procedures and lower the instrumental costs. For instance, the CGE can be replaced by slab gel electrophoresis, and robotic extraction can be replaced with manual extraction using the same chemistry carried out with a similar SPE kit. Detailed experimental designs are included in most published lab manuals (26). In addition, there are many free online resources (many are governmental or non-profit) that provide teaching materials on forensic DNA typing techniques and associated data interpretation online tools (27). Even if teaching lab budgets cannot always afford the actual chemicals and supplies, the students can look at the data and learn how to analyze and interpret it to reach meaningful conclusions that can make sense. As one of the two top job areas in forensic science, DNA analysts are in high demand throughout the world. Understanding the analytical chemistry behind each DNA typing step will greatly benefit our students who are interested in this career. Even for non-forensic students, learning the application of instrumental analysis in the context of human DNA typing provides insights and motivation to animate the sometimes dry and difficult learning process. In this section, the brief overview of the analytical chemistry techniques used in DNA fingerprinting can be used by readers to visualize the potential teaching practices in the chemistry curriculum and can hopefully further explore the linked references to create more interesting teaching materials. In summary, by targeting the two “hottest” areas in forensic science, we can fully utilize the teaching potential of drugs and DNA along with their associated analytical concepts, instrumental theory, and practices to engage our students in analytical chemistry. The added benefits include opening up more job prospects for students and enhancing and motivating students to participate in active learning and experiential learning. The possibilities of using drugs and DNA to effectively teach analytical chemistry is undoubtedly not limited to the content of this chapter. Hopefully, the examples and tips provided in this chapter can inspire our colleagues to try new and creative teaching methods involving the broader forensic sciences. After exposure to drugs and DNA analytical methods, the students can enhance their understanding of the relevance of these sometimes dry analytical chemistry concepts to actual real-world forensic analysis and hopefully be ready for a career in a forensics-related field. Table 1. Abbreviations ATR

attenuated total reflection

CE

capillary electrophoresis

CGE

capillary gel electrophoresis

COD

cause of death

CSI

crime scene investigation

DAD

diode array detector

DEA

Drug Enforcement Administration

DNA

deoxyribonucleic acid

164 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Table 1. (Continued). Abbreviations DUI

driving under the influence (of alcohol or drugs)

DWI

driving with influence (of drugs or alcohol)

EIC

extracted ion chromatogram

FBI

Federal Bureau of Investigation

FDA

Food and Drug Administration

FID

flame ionization detection

FTIR

Fourier-transform infrared spectroscopy

GC

gas chromatography

HPLC

high-performance liquid chromatography

ICP

inductively-coupled plasma

IMS

ion mobility spectrometry

LC

liquid chromatography

LOD

limit of detection

LOL

limit of linearity

LOQ

limit of quantification

MS

mass spectrometry

NIST

National Institute of Standards and Technology

NMR

nuclear magnetic resonance (spectroscopy)

OTC

Over the counter (medication)

PCA

Principal component analysis

PCR

polymerase chain reaction

SPE

solid phase extraction

STEM

Science, Technology, Engineering and Math

STR

short tandem repeat

SWGDRUG

Scientific Working Group for the Analysis of Seized Drugs

TIC

total ion chromatogram

TLC

thin layer chromatography

UV-Vis

Ultra-violet visible (spectrometry)

References 1.

Ravgiala, R. R.; Weisburd, S.; Sleeper, R.; Martinez, A.; Rozkiewicz, D.; Whitesides, G. M.; Hollar, K. Using Paper-Based Diagnostics with High School Students To Model Forensic Investigation and Colorimetric Analysis. J. Chem. Educ. 2014, 91, 107–111.

165 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

2.

3. 4. 5. 6. 7.

8. 9.

10. 11.

12. 13. 14.

15. 16. 17.

18. 19.

Ahrenkiel, L.; Worm-Leonhard, M. Offering a Forensic Science Camp To Introduce and Engage High School Students in Interdisciplinary Science Topics. J. Chem. Educ. 2014, 91, 340–344. Cleaves, A. The Formation of Science Choices in Secondary School. Int. J. Sci. Educ. 2005, 27, 471–486. Bell, S. Quality Assurance and Quality Control. In Forensic Chemistry, 2nd ed.; Pearson: Essex, England, 2014; pp 43–76. Giannelli, P. C. Forensic Science: Under the Microscope. Ohio N.U. L. Rev. 2008, 34, 315. Cole, S. A. Scandal, Fraud, and the Reform of Forensic Science: The Case of Fingerprint Analysis. W. Va. L. Rev. 2016, 119, 523. Morrow, J.; Ropero-Miller, J.; Catlin, M. L.; Winokur, A. D.; Cadwallader, A. B.; Staymates, J. L.; Williams, S. R.; McGrath, J. G.; Logan, B. K.; McCormick, M. M.; Nolte, K. B.; Gilson, T. P.; Menendez, M. J.; Goldberger, B. A. The Opioid Epidemic: Moving Toward an Integrated, Holistic Analytical Response. J. Anal. Toxicol. 2019, 43, 1–9. The Controlled Substances Act. https://www.dea.gov/controlled-substances-act (accessed May 30, 2019). Hasan, S.; Bromfield-Lee, D.; Oliver-Hoyo, M. T.; Cintron-Maldonado, J. A. Using Laboratory Chemicals To Imitate Illicit Drugs in a Forensic Chemistry Activity. J. Chem. Educ. 2008, 85, 813. Schurter, E. J.; Zook-Gerdau, L. A.; Szalay, P. Analysis of a Suspected Drug Sample. J. Chem. Educ. 2011, 88, 1416–1418. Mojica, E. R.; Zapata, J.; Vedad, J.; Desamero, R. Z. B.; Dai, Z. Analysis of Over-the-Counter Drugs Using Raman Spectroscopy. In Raman Spectroscopy in the Undergraduate Curriculum; Sonntag, M. D. Ed.; ACS Symposium Series 1305; American Chemical Society: Washington, DC, 2018; pp 69–91. SWGDRUG Recommendations Version 7.1; 2016-June-09. http://swgdrug.org/Documents/ SWGDRUG%20Recommendations%20Version%207-1.pdf (accessed May 30, 2019). Massart, D. L.; Vandeginste, B. G. M.; Deming, S. M.; Michotte, Y.; Kaufman, L. Chemometrics: A Textbook, 5th ed.; Elsevier Science B.V.: Amsterdam, 2003. Patel, B. H.; Patel, M. M.; Patel, J. R.; Suhagia, B. N. HPLC Analysis for Simultaneous Determination of Rabeprazole and Domperidone in Pharmaceutical Formulation. J. Liq. Chromatogr. Relat. Technol. 2007, 30, 439–445. Ferguson, G. K. Quantitative HPLC Analysis of an Analgesic/Caffeine Formulation: Determination of Caffeine. J. Chem. Educ. 1998, 75, 467. Huang, L. Instrumental Analysis Lab Manual; Hofstra University: Hempstead, NY, 2018; Unpublished. Herman, H. B.; Jezorek, J. R.; Tang, Z. Analysis of Diet Tonic Water Using Capillary Electrophoresis: An Undergraduate Instrumental Analysis Experiment. J. Chem. Educ. 2000, 77, 743–744. Skoog, D. A.; Holler, F. J.; Crouch, S. R. Principles of Instrumental Analysis; Cengage Learning: Boston, MA, 2007. SWGDRUG Mass Spectral Library. http://www.swgdrug.org/ms.htm (accessed May 30, 2019). 166 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

20. Forensic Database Trace Evidence Table. https://www.nist.gov/oles/forensic-database-traceevidence-table (accessed May 30, 2019). 21. Clarke’s Analysis of Drugs and Poisons, 4th ed.; Moffat, A. C., Osselton, M. D., Widdop, B. , Eds.; Pharmaceutical Press: Chicago, IL, 2011. 22. The Ascld Policy Library. https://www.ascld.org/resource-library/ascld-policy-library/ (accessed May 30, 2019). 23. QIAGEN QIAsymphony SP Sample Preparation Fully-Automated DNA RNA Purification. https://youtu.be/7TnUNmuJkXY (accessed May 30, 2019). 24. Butler, J. Fundamentals of Forensic DNA Typing, 1st ed.; Academic Press (Elsevier): Burlington, MA, 2010. 25. Hares, D. R. Selection and Implementation of Expanded CODIS Core Loci in the United States. Forens. Sci. Int-Gen. 2015, 17, 33–34. 26. Elkins, K. M. Forensic DNA Biology: A Laboratory Manual, 1st ed.; Academic Press (Elsevier): Burlington, MA, 2012. 27. Short Tandem Repeat DNA Internet Database. https://strbase.nist.gov//srm_tab.htm (accessed May 30, 2019).

167 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Chapter 10

Downloaded via COLUMBIA UNIV on October 22, 2019 at 19:13:07 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

From DUIs to Stolen Treasure: Using Real-World Sample Analysis To Increase Engagement and Critical Thinking in Analytical Chemistry Courses Sarah E. C. Gray1,2,* 1Chemistry Program, School of Natural Science and Mathematics, Stockton University,

101 Vera King Farris Drive, Galloway, New Jersey 08205, United States 2Department of Chemistry and Biochemistry, Georgia Southern University, 11935 Abercorn Street, Savannah, Georgia 31419, United States *E-mail: [email protected].

With the popularity of CSI and other criminal investigation shows, forensic topics have become a prevalent way to engage students in nonmajor science classes. These topics can also successfully engage chemistry and biochemistry majors, particularly in analytical courses where students often find the level of detail tedious. Laboratory activities designed around a forensic theme demonstrate why the details matter and provide critical-thinking opportunities for students. This chapter will cover two ways to integrate forensics into analytical chemistry courses: in a multiple-week laboratory rotation in an instrumental analysis course and in a hybrid lecture–lab elective course in forensics. The methods described here build on skills learned during the Chemistry Collaborations, Workshops, and Community of Scholars Forensic Science workshop sponsored by the National Science Foundation. In the instrumental analysis course, students spend five to six laboratory periods evaluating the guilt of a teaching assistant (TA) on a variety of charges, after she was pulled over for erratic driving. Techniques incorporated include fluorescence spectroscopy, atomic absorption (AA), gas chromatography–flame ionization detection (GC–FID), gas chromatography–mass spectroscopy (GC–MS), and high-performance liquid chromatography (HPLC). In the forensics elective course, a new fictional case was developed around the theft of potentially rare coins from the university chemistry laboratory, which was asked to help determine their authenticity. Students learn techniques for evaluating different types of evidence and then apply these skills to solve the case over the course of the semester.

© 2019 American Chemical Society Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Introduction Forensic topics have become increasingly popular for engaging students in science curricula and teaching critical thinking at both the K–12 (1–3) and undergraduate levels (4–6). In this context, engagement is considered the degree of attention, curiosity, and interest the students exhibit toward what they are learning. Several factors make forensic themes a good fit. First, criminal cases, real or imaginary, are relatable to students. Elementary students can crack the case of who stole the teacher’s pen, and undergraduates can investigate the arson at the dormitory. Students are more interested in critically thinking about an experiment, rather than simply trying to get the right answer for a good grade, when it is tied to a compelling problem with real-world samples (7, 8). Critical thinking here is considered the process of actively and skillfully applying, analyzing, and evaluating information gathered by observation (9). Second, forensic cases often produce limited amounts of evidence, with someone’s freedom potentially riding on the outcome of the analyses. These factors enable instructors to highlight the importance of best practices in the laboratory, including data management (10), and the effects on people’s lives when high accuracy and precision are not achieved. This chapter covers two different ways to integrate forensic topics into analytical chemistry classes in the undergraduate curriculum. The first section details a forensic rotation used in the laboratory of an instrumental analysis course. The second section presents a new crime scene scenario for use in an upper-division elective course on forensics.

Instrumental Analysis Traditionally, the instrumental analysis laboratory has been used to expose students to as many different scientific instruments as possible using samples that produce good results, but that may not be relevant for students. Current pedagogy indicates that students develop scientific analysis skills and critical thinking more effectively through problem- and inquiry-based learning (11–14). These changes are reflected in the 2015 ACS guidelines for certified degree programs, which encourage the use of these pedagogies to promote the application of appropriate laboratory skills and instrumentation to solve problems (15). A topic that lends itself to these types of projects, as well as being interesting to undergraduate students, is forensic science. The examination of evidence in a forensic case was incorporated into a five- to six-week laboratory rotation in which students were tasked with determining if a teaching assistant (TA) should be charged with any crimes. Due to the limited quantity of each instrument, five to six groups of two to four students took turns moving through each laboratory over the course of the rotation. Since each group or pair of groups was carrying out a different laboratory, typically course-wide prelaboratory lectures were not given. Groups were allowed to get started after a brief safety discussion, and the instructor circled around to each group to check in. Instrumentation used included fluorescence spectroscopy, atomic absorption (AA) spectroscopy, gas chromatography–flame ionization detection (GC–FID), gas chromatography–mass spectroscopy (GC–MS), and high-performance liquid chromatography (HPLC). A week after the completion of each laboratory, students submitted the analyzed data and results for that experiment. The data and results were graded and returned before the final report was due, so that students could correct any calculation errors before discussing what their data meant in terms of the guilt or innocence of the TA. At the end of the rotation, each group presented a chalk talk of the data analysis for one laboratory to the class to be sure all students were on the same page. This put the explanations in the hands of the students after they had a chance to think about the data, rather than the instructor carrying out a lengthy prelaboratory lecture. Students then wrote a combined abstract, introduction, and discussion encompassing all five experiments. As part 170 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

of the discussion, students also evaluated whether the methods they used were the best suited for each sample, using knowledge they gained in both the lecture and the laboratory. Students had to consider figures of merit, instrument and sample analysis costs, analysis time, waste produced, and the destructiveness of the technique. In place of a conclusion section, students summarized their results in a letter to the district attorney’s (DA’s) office, which the DA could use to determine whether charges should be filed against the TA. Student feedback regarding this format has been very positive, with students commenting that the themed rotation gave them more insight into how the instruments can be applied to real-world problems.

Scenario The scenario for the case was adapted from two cases by Grannas and Lagalante (7), as follows. A graduate TA was arrested. After spending the day at a friend’s party, at 2:30 a.m. she was pulled over on the Atlantic City Expressway by a police officer for speeding. She was clocked by radar as traveling 85 mph in a 65 mph zone. The officer reported that her car was swerving on the road and that, upon inspection, empty tonic water bottles (evidence A1 and A2) and, more seriously, an empty gin and vodka bottle were found in the car (evidence A3 and A4, respectively). Naturally, the officer suspected that the TA was driving under the influence of alcohol. The TA claimed she had not been drinking alcohol and explained that, since her friends were not going to recycle the empty bottles from the party, she was going to recycle the bottles with her trash. The TA told the officer that she drinks tonic water all the time without alcohol and that her car is often filled with empty tonic water bottles. The officer’s suspicions were aroused when she noted the TA was speaking quickly and showed signs of paranoia and nervousness. In the TA’s defense, she was very tired and grew nervous when the officer began to insinuate she had been drinking. The officer had probable cause to search the car and, upon examination, found several Hamilton syringes (evidence B), a bag containing several solid tablets (evidence C), and several wax paper packets containing a white powder (evidence D). The TA tried to explain to the officer that the tablets were antacids she got from her friend at the party for an upset stomach (which is why they are in a bag and not a labeled bottle) and that, as a graduate student in chemistry, she uses the needles for chromatographic research testing the white powders (which she claimed were analgesics for headache relief) for a consulting case. However, the officer was highly suspicious that illegal drugs could be involved. The TA passed the roadside sobriety test but was taken to the police station under suspicion of drug and alcohol use. Her car was towed to the police impound lot. Once at the station, the police were able to verify that the TA does indeed work in the chemistry department at the local university. A blood sample was collected from the TA 30 min after arriving at the police station and was sent to the lab for blood alcohol testing by GC–FID (evidence E). The syringes generated enough circumstantial evidence of drug use to warrant testing for substances in the urine that could be used as evidence in a trial. Urine samples (evidence F1 and F2) were collected for testing by fluorescence spectroscopy and GC–MS. Quinine is a common adulterant in illicit drug preparations such as methamphetamine and cocaine. Finding quinine in urine (using fluorescence) can be an indication that drugs were used; however, quinine can also be found in over-the-counter medications and in tonic water. The urine sample could also be analyzed by GC–MS and compared to a mass spectral library to search for any illegal drugs present. The white powders could be analyzed by thin-layer chromatography (TLC) and HPLC to see if they contain the analgesics the TA claimed they do, and the white tablets could be analyzed for calcium content by AA to test the TA’s claim that they are antacids. 171 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Laboratories Each laboratory was completed within one 4 h laboratory period. For two of the chromatography labs (GC–MS and HPLC), autosamplers were used, and the students came back on a separate day to process and save their data. For shorter laboratory blocks, each experiment could span two periods, which would also allow students to process their chromatography data during normal class time. A summary of the labs included in the rotation is shown in Table 1, and each of them is discussed in detail. The instructor can determine the amount or type of analyte to spike each sample with, depending on the desired guilt or innocence of the TA. Learning objectives for each exercise centered on (1) demonstrating proficiency in making calibration curves and operating instrumentation and (2) analyzing and evaluating the data produced. The first objective was assessed through observation by the instructor during the laboratory and grading of the laboratory notebooks. The second objective was assessed by grading the laboratory reports and the report for the DA. Table 1. Summary of the Instruments, Samples, Analytes, and Calibration Methods Used for the Forensic Rotation in an Instrumental Analysis Laboratory Instrument

Sample

Evidence Label

Analyte

Calibration Technique

Fluorimeter

TA’s urine

F1

quinine

external standards

AA

white tablets

C

calcium

standard addition

GC–FID

TA’s blood

E

ethanol

external + internal standard

GC–MS

TA’s urine

F2

quinine, caffeine, codeine

----

HPLC

white powders

D

caffeine, acetaminophen

external + internal standard

Figure 1. Evidence as it was provided to the students. (Top) Evidence C and D. (Bottom) Evidence E, F1, and F2. Photographs courtesy of Sarah Gray. 172 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Figure 1 shows the evidence as it was provided to the students. The urine and extracted blood samples were put into urine sample cups (Lynn Peavey, catalog # 01706) with a temperature strip. The instructor briefly explained to the students that the temperature strip is used to check that the sample is fresh and that the suspect is not trying to beat the test by using someone else’s clean urine. Fluorescence For the fluorescence laboratory, students determined the quinine concentration in the urine sample (evidence F1) collected from the TA at the police station. Quinine has therapeutic uses for the treatment of malaria in oral doses of 500–650 mg every 8 h, and it is used in tonic water at concentrations of approximately 60 µg/mL to give it its bitter taste (16). Quinine has also been used as an adulterant for cocaine and heroin (17, 18), and its presence can indicate drug use, since excreted values would be much higher than expected from drinking tonic water alone. The TA claimed that she does not consume illegal drugs but does drink tonic water regularly, so her urine would be expected to have a low quinine concentration. The percentage of consumed quinine that is excreted in urine after 2.5 h is approximately 7% (19). The instructor can choose the concentration of quinine with which to spike a clean urine sample. Here, samples were spiked with 30 ppm quinine, which required students to carry out several dilutions of the unknown sample to produce a fluorescence in the linear range of their calibration curve, which covered 0.1–1.0 ppm quinine. The learning objectives for the laboratory were as follows: Students will be able to: 1. Demonstrate the correct procedures for carrying out a liquid–liquid extraction, creating an external calibration curve, and operating a fluorimeter; 2. Justify why an extraction was carried out on the urine sample before analysis; 3. Analyze raw data to determine the concentration of quinine in the TA’s urine sample (evidence F1); and 4. Evaluate processed data to determine if the TA should be charged with the consumption of illegal drugs. Students were instructed to answer the following questions as technicians working on the TA’s case: 1. Does the urine contain quinine? If so, at what concentration, in units of parts per million? 2. If quinine is present, does the concentration indicate illegal drug use, malaria treatment, or the consumption of tonic water? Note that high quinine levels by themselves do not prove illegal drug use. A person could be taking quinine for treatment of malaria with a valid prescription. You can ask the instructor additional clarifying questions once you have collected your data. The laboratory procedure involved extracting quinine from a synthetic urine sample (evidence F1), which can be made (20) or purchased from Flinn Scientific (catalog #FB1444), following the procedure of O’Reilly (21). Briefly, the quinine and urine sample was adjusted to a pH between 9 and 10 with a few drops of 3.7 M NH4OH (checked with pH paper) and was shaken or mixed on a vortexer with 3:1 v/v chloroform–isopropanol to extract the quinine from the urine matrix. The organic layer was then acidified with 0.05 M H2SO4 and shaken or vortexed to extract the quinine 173 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

back into the aqueous layer, which was separated from the organic layer for measurement. An external calibration curve was created from five quinine standards made in 0.05 M H2SO4 in the 0.1–1.0 ppm range. Students carried a “blank” urine sample, consisting only of the synthetic urine, through the extraction procedure to determine if any background fluorescence was present. An “extraction check” sample was created by making a 0.60 ppm quinine standard in clean urine (rather than in 0.05 M H2SO4), which was then also carried through the extraction procedure and measured to determine the extraction efficiency. Using a benchtop fluorescence spectrometer, students used a midrange concentration quinine standard to determine the excitation and emission wavelengths. The students were instructed to use a cuvette specifically for fluorescence measurements (i.e., clear on all four sides). All standards for the external calibration curve, blank, extraction check, and unknown samples were then measured using the determined wavelengths. To establish the extraction efficiency, students compared the fluorescence intensity of the 0.60 ppm extraction check sample to the fluorescence intensity predicted by their calibration curves for that concentration. Students typically found their extraction efficiency to be around 60%. Students corrected the TA’s quinine concentration for any blank fluorescence and for extraction efficiencies less than 100%. A sample student external calibration curve is shown in Figure 2, along with the 0.60 ppm quinine extraction check. This group found the extraction efficiency to be 62.15% and the final concentration of quinine in the TA’s urine to be 42.95 ppm, after correcting for the extraction check.

Figure 2. Sample student external calibration curve for fluorescence measurements of quinine extracted from a synthetic urine sample. If fluorescence was not covered in the lecture before the laboratory was carried out, students struggled slightly with the need to determine two separate maximum wavelength values (as opposed to one for ultraviolet–visible (UV–vis) absorption measurements, which they were more familiar with from lower-division courses). In addition, students had difficulty determining how to calculate their extraction efficiency from their measured values, and several students calculated percentage error rather than the simple percentage calculation required. Extraction efficiency numbers varied depending on how effectively students mixed their samples during the liquid–liquid extraction. More consistent results were obtained when students mixed samples with a vortexer rather than by hand. The same urine sample could potentially be used for both the fluorescence and the GC–MS analysis, but the instructor may want to check that any drugs added for the GC–MS analysis do not interfere with the quinine extraction process.

174 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

AA For the AA laboratory, students analyzed the solid tablets (evidence C) found in the TA’s car. The TA asserted that they are regular strength antacids she got from her friend for her upset stomach. The unknown was in fact a regular-strength antacid tablet with 500 mg CaCO3 as the active ingredient. Students tested the tablets for calcium content and had to research what the expected calcium level would be for a regular-strength antacid to determine if their measured numbers supported the TA’s claim or not. A standard addition calibration curve was used to correct for any matrix effects from the tablet binders and dyes and to give students practical experience with different types of calibration curves. Students often struggled with how to actually create standard addition calibrations and carry out the calculations, since they may have only encountered these principles in a lecture setting. The hands-on practice in the laboratory helped to cement their knowledge of the operation and use of the technique. The learning objectives for the laboratory were as follows: Students will be able to: 1. Demonstrate the correct procedures for creating a standard addition calibration curve and operating an AA spectrometer; 2. Justify why a standard addition calibration curve was needed for the analysis; 3. Analyze raw data to determine the concentration of calcium in the white tablets (evidence C); and 4. Evaluate processed data to determine if the TA should be charged with the possession of illegal drugs. Students need to answer the following questions as technicians working on the TA’s case: 1. Does the tablet contain Ca? If so, at what concentration, in units of milligrams of calcium per gram of tablet? 2. Is this concentration of calcium consistent with a regular-strength antacid, supporting the TA’s claim that the tablets do not contain illegal drugs? The laboratory procedure involved crushing the white tablet (evidence C), dissolving it in 1.0 M HCl with gentle heating to remove CO2, filtering it through a medium porosity filter paper, and diluting it in a 250 mL volumetric flask with 0.1 M HCl. Students were told that the total concentration of calcium in their calibration curve solutions should fall between 1 and 12 ppm calcium to stay within the linear range of the method. To estimate the concentration of calcium in the unknown, students prepared a one-point external calibration standard at 4 ppm and a solution of their unknown with 500 uL of the tablet solution diluted to 25 mL in 0.1 M HCl. This produced a concentration of calcium close to 4 ppm when using a regular-strength antacid. If multiple weeks are available for the laboratory, the instructions for diluting the unknown can be omitted, and students can be tasked with determining a volume that produces an absorbance near the 4 ppm external standard. Once the needed volume of the unknown was approximated, students selected concentrations of standard to add to their unknown, keeping the total calcium concentration within the linear range, and prepared three replicates with four standard concentrations and one sample with just their unknown (15 solutions total). The standard addition solutions were then measured on a flame AA spectrometer with a calcium hollow cathode lamp installed. If the instrument does 175 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

not automatically set the maximum wavelength from the lamp that is installed, 422.7 nm should be used for calcium. Students generated a calibration curve of average absorbance versus concentration of standard added, with error bars on each point accounting for the precision among their three replicates. The linear regression line was extrapolated to the x intercept to determine the concentration of calcium in the diluted sample. After undoing the dilution, students determined calcium concentration in units of milligrams of calcium per gram of tablet. Figure 3 shows a sample student data set. Using the 4 ppm external standard check, it was determined that 250 µL of the unknown solution should be added to the 25 mL volumetric flasks for the standard curve. The unknown was found to have a concentration of 214.6 mg Ca/g tablet.

Figure 3. Sample student standard addition calibration curve. Twenty-five µL of the unknown solution were added to the 25 mL volumetric flasks. Students were initially confused by using two different types of calibration techniques, external and standard addition, in the same laboratory and were unsure how to apply the data from the onepoint external standard to making their standard addition curve. Calculations on these techniques were subsequently added to the pre-laboratory activity so that students felt more confident with the techniques during the laboratory period. Despite making the standard addition curve during the laboratory, when students submitted their data and results, many treated their data like it was produced from an external calibration curve and completed the wrong calculations. It may be worthwhile to remind students to think carefully about how to calculate the concentration of their unknown. Once calculation errors were corrected, all groups were able to produce a calibration curve with an R2 of at least 0.9. Some groups had each group member prepare one of the replicates, and these groups typically saw lower precision in their data due to varying accuracy in pipetting between students. GC–FID For the GC–FID laboratory, students analyzed the blood sample (evidence E) collected from the TA at the police station for percentage blood alcohol content (% BAC). The TA asserted that she had not been drinking the night of her arrest. The unknown sample consisted of ethanol in deionized water, and the students were told the ethanol had already been extracted from the blood matrix. Bovine serum can be used instead of deionized water for a more realistic approach. Zabzdyr and Lillard (22) provide a conversion factor to account for the differences in ethanol partitioning between 176 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

red blood cells and blood serum. The unknown was made at a blood alcohol concentration of 0.05%, a level that contradicts the TA’s claim that she did not consume any alcohol but is still below the legal limit of 0.08% in most states. The learning objectives for the laboratory are as follows: Students will be able to: 1. Demonstrate the correct procedures for creating an external calibration curve with an internal standard and operating a GC–FID instrument using headspace analysis; 2. Justify why an internal standard was needed for the analysis; 3. Analyze raw data to determine the presence and concentration of % BAC in an extracted blood sample (evidence E); and 4. Evaluate processed data to determine if the TA should be charged with driving under the influence of alcohol. Students need to answer the following questions as technicians working on the TA’s case: 1. Does the TA’s blood contain ethanol? If so, at what concentration, in units of % BAC? 2. Is this concentration of ethanol above the legal limit while driving a car? 3. If there is ethanol present, should the TA be charged with driving under the influence? The laboratory procedure involved headspace GC–FID analysis of ethanol from simulated blood samples, following a modified version of the procedure described by Zabzdyr and Lillard (22). If specific instrumentation for headspace analysis is not available, as was the case here, students can instead add 0.5 mL of each solution to an autosampler vial and let the solution sit for at least 15 min to equilibrate with the headspace. The minimum depth in the vials for the autosampler needle was set to above 1.0 mL, so only gas was injected on the instrument with the 0.5 mL sample volume. A 500 µL injection volume was used. Students were provided with stock solutions of ethanol and n-propanol, the latter of which served as an internal standard. Dilutions of the ethanol and n-propanol stock solutions were prepared and measured so that students could determine the retention time for each standard and practice their method to ensure the sample volatilization timing was consistent with each sample. The ethanol stock solution was then diluted with deionized water to make the ethanol standards, which were spiked with the n-propanol. Exact concentrations and instrumental details are provided in the supplemental material from Zabzdyr and Lillard (22). The standards were made on a staggered schedule so that each one only sat for the required amount of equilibration time before measurement. The TA’s extracted blood sample was diluted, spiked with the internal standard, and measured following the same procedure as the standards. For the data analysis, students were required to make both an external calibration curve and an external calibration curve plus their internal standard. This allowed students to see how an internal standard can improve their data without requiring any additional measurements. Students typically suggest running more replicates to improve poor calibration curves rather than changing their method, so this laboratory helps reinforce the latter technique. If students have not previously carried out a laboratory in which reaction time of the samples matters, they may need a reminder to prepare their samples on a staggered schedule rather than making all of them at once. In this course, the GC–FID laboratory was used as a method-development exercise, and the students were not given an explicit procedure. Instead, they were given information about the 177 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

chemicals available and asked a series of questions to help guide them in developing a valid procedure. An abbreviated version of this information is given here: • Provided chemicals include a % BAC standard, an n-propanol internal standard, and evidence E. • You will need to prepare an external calibration curve with an internal standard. The concentrations of ethanol for your curve should bracket the legal limit for percent blood alcohol in your state. • To sample the headspace in the autosampler vials, you need to have your analytes move from the liquid phase into the vapor phase. How can you facilitate this? How long should you allow your samples to sit before measurement? • If you have both ethanol and n-propanol in your solution, how will you know which peak is which on your chromatogram? • Once you have developed a valid calibration curve, how can you test that your method correctly determines the % BAC in an unknown sample before you test evidence E? Students worked toward creating a procedure by measuring the two standards individually to determine retention times, testing several equilibration times for forming the headspace, determining whether to equilibrate their vials on the countertop or in a water bath [Zabzdyr and Lillard (22) recommend on the countertop], measuring their calibration curve, creating a new % BAC standard at a known concentration and running it as an unknown to validate their calibration curve, and then measuring evidence using their new procedure. Due to time constraints in the semester when this laboratory was run, each group only worked on the project for one week, summarized their findings, and made suggestions for the next group to try. While this type of exercise can help train students how to conduct research within a larger group, they found the format very difficult and had a hard time following the data from the previous groups. In the future, the schedule will be reworked to allow each group several laboratory periods to create a valid procedure and measure evidence E.

GC–MS For the GC–MS laboratory, students analyzed the urine sample (evidence F2) collected from the TA at the police station. The TA asserted that she does not consume any illegal drugs. The unknown was 40 mL of synthetic urine (20) spiked with two 1 mL ampoules of a 1 mg/mL codeine certified reference material, purchased from Sigma-Aldrich (catalog #C-006) through Cerilliant. Many other drug standards are available from Cerilliant, and the instructor can substitute their legal or illegal drug of choice. Codeine was chosen as a drug that could be either legal or illegal depending on whether the TA had a prescription. The urine sample was also spiked with quinine, to align with the fluorescence laboratory, and with caffeine to simulate another drug the TA might have consumed at the party.

178 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

The learning objectives for the laboratory are as follows: Students will be able to: 1. Demonstrate the correct procedures for carrying out a solid phase extraction (SPE) and operating a GC–MS; 2. Justify why an extraction was carried out on the urine sample before analysis; 3. Analyze gas chromatography chromatograms and mass spectra to determine the chemicals present in the TA’s urine (evidence F2); and 4. Evaluate the data to determine whether the TA should be charged with the consumption of illegal drugs. Students need to answer the following questions as technicians working on the TA’s case: 1. What chemicals are present in the TA’s extracted urine? 2. Are any of the chemicals present restricted drugs? If so, what class and schedule of drug are they? 3. If there are restricted drugs present, is the TA breaking any laws by having these drugs in her system? Remember, some drugs are legal if you have a valid prescription but illegal if bought on the street (e.g., oxycodone). You can ask the instructor additional clarifying questions once you have collected your data. The laboratory procedure was taken from the forensic science laboratory manual developed by Kaplan (23) as part of the National Science Foundation’s Chemistry Collaborations, Workshops, and Communities of Scholars program. Acidic, neutral, and basic drugs were extracted from the urine sample (evidence F2) using an SPE tube from United Chemical Technologies (Clean Screen DAU, catalog #CSDAU206). The extraction procedure (23) was a modified version of the “Narcotics/Metabolites Panel in Blood, Plasma/Serum, Urine, Tissue by LC-MS/MS or GC-MS” technical procedure from United Chemical (24). The urine was added to 0.1 M phosphate buffer in a centrifuge tube and shaken by hand or mixed with a vortexer. Since a vacuum manifold was not available, students used the air lines in the ventilation hood to push the sample and reagents through the SPE tube, at a rate of 1–2 mL/min, into test tubes below. The SPE tube was preconditioned with 3 mL of methanol, 3 mL of deionized water, and 1 mL of 0.1 M phosphate buffer, and the urine and buffer sample was applied to the tube (all in fraction A). The column was then washed with 3 mL of deionized water and 1 mL of 0.1 M acetic acid. The column was dried by drawing air through it for 5 min, which was followed by the addition of 1 mL of hexane (fraction B). Acidic and neutral drugs were eluted with 3 mL of a 1:1 mixture of hexane and ethyl acetate (fraction C). The column was washed with another 3 mL of methanol and again dried for 5 min (fraction D). Basic drugs were eluted with 3 mL of a 78:20:2 mixture of dichloromethane/isopropyl alcohol/ammonium hydroxide (fraction E). The two test tubes with the potential acid or neutral and basic drugs (fractions C and E) were evaporated to dryness using the hood air line, reconstituted in 1.5 mL of acetonitrile, and transferred to GC–MS autosampler vials. An Agilent GC–MS system was used with an HP-5ms (5%-phenyl-methyl siloxane) column, 30.0 m × 0.25m × 0.25 µm. A 10 µL injection was used with the inlet at 275 °C, a 10.0:1 split ratio, pressure of 13.68 psi, and total flow of 8.4 mL/min. The oven was set to an initial temperature of 50.0 °C for 3.00 min, with a 25 °C/min ramp to 175 °C and a hold time of 3.00 min, followed by a 25 °C/min ramp to 270 °C with a hold time of 10.00 min. For the 179 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

mass spectrometry, the heater was set to 280 °C, with a solvent delay of 5.0 min and scan parameters of 50–550 amu. Figures 4 and 5 show sample student total ion chromatograms for fractions C and E. Caffeine was eluted in fraction C (Figure 4), codeine was eluted in fraction E (Figure 5), and the quinine was not identified. Students did not encounter many problems with this laboratory. They were instructed to test their flow rate with methanol before starting the column wash procedure so that they knew how much pressure must be applied from the air line. They were cautioned that if their flow rate was too fast, the analytes would not have enough time to react with the stationary phase in the extraction tube and separate.

Figure 4. Sample student total ion chromatogram for fraction C. The peak at 9.069 min has a 74.8% library match to methylparaben, and the peak at 12.996 min has a 98.8% match for caffeine.

180 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Figure 5. Sample student total ion chromatogram for fraction E. The peak at 5.110 min has an 89.1% match to hexylene glycol, and the peak at 16.171 min has a 66.6% library match to codeine. HPLC For the HPLC laboratory, students analyzed the white powder (evidence D) found in the TA’s car. The TA asserted that the powder is a sample of Pain B Gone, an analgesic mixture that she is testing in a lawsuit against the manufacturer, PharmaCo. Pain B Gone is supposed to contain caffeine, acetaminophen, and acetylsalicylic acid at concentrations typically found in migraine pain relievers (i.e., 250 mg acetaminophen, 250 mg acetylsalicylic acid, 65 mg caffeine) (25). However, PharmaCo is currently being sued based on reports that Pain B Gone either does not contain all three drugs or that the drugs are not present at the stated concentrations. A tension headache medicine containing 500 mg acetaminophen and 65 mg caffeine was chosen as the unknown here, but any analgesic containing one or more of the three drugs given would be a suitable unknown. Students carried out a TLC analysis of the unknown before the HPLC analysis. An internal standard was used to correct for any random or systematic errors in sample injection and to give students practical experience with different types of calibration curves, as with the AA laboratory.

181 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

The learning objectives for the laboratory are as follows: Students will be able to: 1. Demonstrate the correct procedures for creating an external calibration curve with an internal standard and operating a HPLC instrument; 2. Justify why TLC was used before HPLC and why an internal standard was needed; 3. Analyze raw data to determine the presence and concentration of analgesics in the white powder (evidence D); and 4. Evaluate processed data to determine whether the TA should be charged with the possession of illegal drugs and whether PharmaCo’s Pain B Gone sample contains the correct number and concentration of drugs. Students need to answer the following questions as technicians working on the TA’s case: 1. Based on the TLC plate comparison with the standards, are any of the drugs that PharmaCo states are in Pain B Gone present? 2. Based on the HPLC data, what is the concentration (in grams of drug per gram of Pain B Gone) of each major drug in the tablet? 3. Do the composition and concentrations of the drugs present match the claims made by PharmaCo? Should the TA be charged with possession of illegal drugs? The idea for the Pain B Gone case was adapted from Grannas and Lagalante (7) and follows the typical procedures for analyzing analgesics by HPLC (26–28). The white powder was first dissolved in the HPLC mobile phase (60% deionized water with 2% acetic acid, 40% methanol). Since the lawsuit states that some drugs may be missing from the Pain B Gone samples, students ran a TLC plate of the Pain B Gone against standards of the three listed drugs at a concentration of 1.00 × 10−3 g/mL. In the case of the tension headache medicine chosen as the unknown, the TLC results showed that acetylsalicylic acid was missing from the sample. This noninstrumental approach demonstrates to students the benefits and time and cost savings associated with using simpler techniques before a scientific instrument is used. Based on the TLC plate results, students only prepared HPLC standards for two, instead of three, drugs, at the concentration ranges shown here. Note that acetylsalicylic acid concentration ranges have not been tested for actual headache medicines containing all three drugs; concentrations were chosen in the range of the caffeine and acetaminophen standards:

• Caffeine: 5.00 × 10−5, 3.00 × 10−5, 2.00 × 10−5, 1.00 × 10−5, 5.00 × 10−6 g/mL; • Acetaminophen: 1.25 × 10−4, 1.0 × 10−4, 7.00 × 10−5, 5.00 × 10−5, 3.00 × 10−5 g/mL; and • Acetylsalicylic acid: 1.00 × 10−4, 5.00 × 10−5, 3.0 × 10−5, 1.00 × 10−5, 5.00 × 10−6 g/ mL.

182 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Since the concentration of the acetaminophen is an order of magnitude higher than the caffeine concentration, the standard range was adjusted accordingly. The standards were prepared in 10.0 mL volumetric flasks and spiked with 1.0 mL of 1.0 ×10−3 g/mL salicylic acid prepared in the mobile phase as an internal standard. Two dilutions of the Pain B Gone sample were made at dilution ratios of 1:20 and 1:40, to account for the different concentrations of the analytes, and were spiked with the internal standard in the same ratio as the standards. All samples were transferred to autosampler vials and measured on an HPLC instrument with a C18 column, 20 µL injection volume, 1 mL/min flow rate, isocratic elution, and diode array detection at 254 nm. A sample student chromatogram for a 1:20 dilution of the unknown solution and the calibration curves for acetaminophen and caffeine using the salicylic acid internal standard are shown in Figures 6 and 7, respectively. For acetaminophen, the R2 value increased from 0.9988 to 0.9991 with the addition of the internal standard. For caffeine, the R2 value increased from 0.9655 to 0.9925. Using the calibration curve and unknown runs from the student group shown, the acetaminophen concentration was found to be 0.61 g acetaminophen/g powder, while the tension headache medicine was labeled to contain approximately 0.74 g acetaminophen/g tablet. Since not every tablet weighs the same, a subset of tablets were weighed to determine an average tablet weight to calculate the grams of drug per gram of tablet. The caffeine concentration was found to be 0.080 g caffeine/g tablet compared to the labeled amount of 0.097 g caffeine/g tablet. Both values were low and may indicate that the drugs were not fully solubilized from the powder with the mobile phase.

Figure 6. Sample student chromatogram of a 1:20 dilution of their unknown solution. The peak at 4.34 min is acetaminophen, the peak at 5.64 min is caffeine, and the peak at 19.35 min is salicylic acid. Students did not encounter many problems with either the sample preparation or measurement with this laboratory. All groups were able produce a calibration curve with an R2 value of at least 0.9 using their internal standard. As part of the TLC analysis, students first pulled their own spotters using Pasteur pipettes and a Bunsen burner. Most students enjoyed pulling the spotters, since they do not typically learn glassblowing as part of the chemistry curriculum, with the possible exception of pulling spotters in the organic laboratory. Some students had to repeat their TLC analyses with commercial spotters, since the spotters they pulled were not fine enough and their spots ran together on the plate.

183 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Figure 7. Sample student calibration curve data for acetaminophen (R2=0.991) and caffeine (R2=0.9925), using the salicylic acid internal standard. Postlab Data and Method Evaluations Once all five laboratories had all been carried out, students completed a formal laboratory report and wrote a letter to the DA summarizing their findings. They had to state the final result for each sample (Table 1) and advise whether or not they thought the TA should be charged with any crimes other than speeding. In addition, students also wrote a summary for the instructor evaluating whether each instrumental technique used was the best method for that sample in a real forensic case, supporting their answers with knowledge gained in both the lecture and laboratory portions of the course. This forced students to apply knowledge from lecture to a specific laboratory problem. One example would be the use of AA to analyze the white tablets. If all possible instruments were readily available to the students, inductively coupled plasma–atomic emission spectroscopy could have been used to simultaneously determine multiple elements in the tablet rather than using AA to evaluate only calcium. As another example, analyzing the fluorescence of quinine in urine would likely be skipped in a true forensic laboratory in favor of the GC–MS technique, which can both identify multiple drug components in the sample directly and quantify components once they are identified. Finally, in the HPLC analysis, an actual forensic laboratory would likely prepare all three standards of drugs that are supposed to be in Pain B Gone, since TLC may not detect the presence of drugs at low concentrations. The question regarding the best methods is open ended, and students do not need to give these specific answers. As long as their answers are plausible and well supported by material from the lecture and laboratory or literature references, they receive credit. Assessment of Student Engagement and Critical-Thinking Skills The forensic theme was very successful at engaging students in the material. Students across numerous semesters have asked for a recap of each experiment at the end of the forensic rotation to find out how the case was intended to play out. This interest went beyond simply desiring a good grade, since students had already received their graded assignments when they asked for the recap. After the summary of each case, covering the intended value for each piece of evidence and what that value means for the guilt or innocence of the TA, the student group whose summary to the DA was the most complete was declared the winner, earning bragging rights. Initially, students tended 184 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

to believe the TA was guilty and would make up scenarios for which crimes she should be charged with. This provided the opportunity to talk about ethical considerations and how difficult it can be in real life for criminalists to be impartial when working on evidence from a highly publicized case. Importantly, it also shows that the forensic theme got the students talking and thinking about the laboratory experiments outside of class. The following are representative student quotes from the end-of-semester evaluations: “The design of the lab experiments were great in helping me connect the subject matter of the course to real life applications (i.e. forensics, pharmaceuticals).” “The real-world approach helped bring home the importance of the instruments and clarify any uncertainty such as why a procedure is the way it is.” “The use of the forensics and pharmaceutical rotation lab gave more insight into how the instruments can be applied into real world problems.” These comments demonstrate that the students were engaged in and appreciative of laboratory exercises that tie the science into real-world applications and demonstrate how scientific instruments can be used outside of teaching labs. The overall learning objectives for the laboratory exercises were that students be able to (1) demonstrate proficiency in making calibration curves and operating instrumentation and (2) analyze and evaluate the data produced. All student groups across several semesters were able to operate the instrumentation successfully. As the instructor improved at facilitation, more students correctly made calibration curves for standard addition and internal standard measurements. Facilitation tips were noted with each of the previously described laboratories. Students were also successful at critically analyzing the data to determine whether the TA should be charged with any crimes. The real-world samples gave the students a context in which to analyze their results. Historically, the point of analytical labs has been to get the “right answer” (7), with students submitting a number and waiting for their graded report to find out if it was correct. Here, students were able to critically evaluate the significance of their results in the context of the laws restricting drug and alcohol use, even if their absolute accuracy was off. Students also had to synthesize all five laboratories into a cohesive story to write their summaries to the DA. In addition to the data, students critically analyzed the instrumental methods themselves. This required them to compare and contrast different instrumental methods to determine which would be best for each of the five pieces of evidence, justifying their answers with evidence.

Forensics Upper-Division Elective Course This course has been taught as a two-credit hybrid lecture and laboratory upper-division elective for junior and senior chemistry and biochemistry majors during both the regular semester and a 2.5-week May term. The short, intensive May term format worked well for the course and allowed more flexibility to switch back and forth between lecture and laboratory time during a given day. The course could easily be formatted to fit a three-credit lecture or a lecture plus separate laboratory format. The extra time would allow for each of the topics shown in Table 2 to be covered in more depth and for students to analyze more types of samples in the laboratory. There was considerable overlap between the student populations for the two courses, and most of the students in the 185 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

forensics upper-division course had previously taken the instrumental analysis course with the forensic rotation. If this was not the case, many of the laboratory activities used in the instrumental analysis course could be used in an upper-division elective course on forensics as well. Course objectives were as follows: Students will be able to: 1. Define and describe the major topics in forensic science, including processing a crime scene, physical evidence, firearms and impressions, hair and fibers, drugs, serology, and questioned documents; 2. Analyze and critique the use of forensic evidence in historical crimes; and 3. Perform analyses of evidence in a mock crime scene, evaluate the data to solve a crime scenario, and present key findings in written form. The course was an implementation of the staged crime scene format learned during the NSFsponsored Chemistry Collaborations, Workshops, and Community of Scholars Forensic Science workshop taught by Kaplan (29), with a new crime scene scenario designed by Gray. The lecture portion of the course covered the topics shown in Table 2, generally following the format presented in the textbook Criminalistics: An Introduction to Forensic Science by Saferstein (30). Practitioner documents were also incorporated, including a firearms identification document produced for the National District Attorneys Association (31) and crime scene investigation guides for law enforcement (32) and defense attorneys (33). Table 2. Forensic Topics Covered in the Lecture and Laboratory Portions of the Coursea Topic

Subtopics

Introduction to forensic science

History and developments, functions of a forensic scientist

The crime scene

Proper processing, legal considerations

Fingerprinting

Fundamental principles, classification, methods of detection

Trace evidence

Fibers, glass

Drugs

Presumptive and evidentiary tests

Questioned documents

Ink and handwriting analysis

Forensic serology

Presumptive tests, blood typing

Firearms and impressions

Bullet comparisons, firearms, tool marks, other impressions

a Students

carried out laboratory experiments related to the subtopics to solve the mock crime committed during the course.

After the topics were introduced in lecture, students learned how to analyze the different types of evidence in the laboratory. Students knew that they would be applying the techniques they learned in the lecture and laboratory to a forensic case, but they did not know that the crime would be staged in real time during the class. In addition to investigating the crime scene scenario, students also applied their knowledge from lecture to a famous criminal case of their choosing, such as those of Caylee Anthony, the Unabomber, or the Zodiac Killer. They researched the forensic evidence in a 186 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

case and evaluated whether the evidence at the scene was likely compromised, whether the forensic techniques used would satisfy today’s standards, and whether the conclusions drawn by investigators (or juries, for cases that went to trial) were logical extensions of the forensic evidence. Local resources allowed for two field trips during the course for students to see applications of forensic investigation techniques. Georgia Southern University–Armstrong Campus is recognized as a National Center of Digital Forensics Academic Excellence and has a Cyber Forensics Division (34) Students were able to tour the facility and learn how police officers analyze cyber forensic evidence when not responding to on-campus calls. Students were also able to tour the U.S. Customs and Border Protection Savannah Customs Laboratory to learn the types of samples they investigate and tour their mobile laboratory, which is well stocked with miniaturized versions of scientific instruments. Students were assessed through homework, two lecture exams, participation in lecture and in the laboratory, results from the laboratory case work including a written report, and laboratory notebooks. Crime Scene Scenario Setting the Scene During the first several class periods, students were introduced to the field of forensic science and covered how to process and collect evidence from a crime scene. During the next class period, after a brief introduction in the classroom, the instructor stated that the class would move to the laboratory to look at evidence collection techniques. Attempting to keep everyone together in a group, the instructor walked the class to the laboratory. When they approached the laboratory, the class heard shouting and a crash and saw a person wearing a hooded sweatshirt run out of the lab and down the hallway in the opposite direction. Upon entering the laboratory, they found another faculty member “bleeding.” Students were instructed to follow the procedures for arriving on a crime scene, which include acquiring medical assistance for injured victims, detaining any suspects or witnesses, securing the scene, and calling for additional personnel. Students broke into groups to help and then interview the injured faculty member, search for and interview witnesses, secure the crime scene, and start collecting potential evidence. Given today’s climate, the department was notified in advance that the staging of the crime scene would be occurring, and other faculty members stood at the intersecting hallways in case students or faculty not involved in the case observed the “crime” and were worried it was a real incident. These faculty members were also able to serve as witnesses for the students to interview. Campus police graciously agreed to assist and were standing by to “respond” to a call from the instructor. They helped the students collect witness statements and gave them pointers for collecting evidence from the crime scene. Victim Statement The injured faculty member, Dr. Smith, stated that she was walking down the hallway when she heard a noise coming from the lab and noticed that the rear door of the lab was open. Since the door is usually locked, this was unusual. She walked in and saw a person in a hooded sweatshirt at the back of the laboratory holding something. Since she did not recognize the individual, she asked who they were and what they were doing in the laboratory. The individual turned, threw a piece of glassware at Dr. Smith, and rushed past her toward the door. The glassware shattered on the floor, spraying both Dr. Smith and the suspect. Dr. Smith was able to grab a piece of the suspect’s shirt as the individual ran by but was knocked into the corner of a flammables cabinet and cut her arm. She provided a vague description of the suspect. 187 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Both the faculty member and the suspect were recruited as volunteers. A student from another department was preferred for the suspect so that the class did not instantly recognize the individual as they fled down the hallway. Dr. Smith was interviewed by several different students in the class, and when the transcripts were compared, the statements did not match exactly. The class was able to see that witness statements can be unreliable due to the chaos of the moment, even if the witness is not intending to deceive law enforcement. Case Background Once students had helped the assault victim and talked to witnesses, the instructor explained what she thought the assailant may have been doing in the lab: trying to steal rare coins. Recently, Geoffrey Raleigh, a noted treasure hunter, sent his friend, Dr. Gray, several items he found off the coast of Savannah. Raleigh believed these items may in fact be a part of Captain Thorpe’s treasure. The treasure hunter needed Dr. Gray to verify the providence of the objects to lay claim to the site as well as raise funds to completely excavate it. If the coins were truly from the wreck, the find could be worth millions of dollars, as Captain Thorpe was a well-known pirate hunter, privateer, and explorer. He and his crew sailed the waters from South Carolina to the Caribbean. By the end of his lucrative career, Thorpe was said to have accrued a significant treasure trove of gold and silver. The legend says that as the captain made his final voyage, carrying all his plunder on the way to giving up his adventures on the seven seas, his ship was caught in a bad storm and sank. Explorers and treasure hunters have been searching for the site where his ship went down with all of his wealth ever since. Evidence A summary of the evidence collected from the laboratory after the crime occurred is shown in Table 3. Students were inclined to collect every single item found on the bench or floor in the laboratory, and occasionally the instructor stepped in to let them know an item was not related to the case. All evidence was collected in plastic or paper bags, as appropriate for the evidence type, entered into an evidence log created in Excel, and stored in an evidence locker (i.e., a spare filing cabinet). To maintain the chain of custody, students had to check each piece of evidence out of and back into the evidence log each time they wanted to carry out an analysis. Students had to plan the order of analyses carefully for each piece of evidence, since one test could damage the sample for another test. For example, blood samples were collected from the bloody footprint before ninhydrin was used to detect fingerprints to avoid compromising the blood sample. Multiple types of synthetic blood were necessary to carry out all of the serology tests. The instructor often “extracted” blood samples from the swabs the students collected into centrifuge tubes, so that the correct sample could be substituted before analysis by the students. Stage blood was used on the victim’s arm for a more realistic look. Samples for blood typing were from a kit by Ward’s Science (catalog #470213-350). Synthetic blood from Arrowhead Forensics (catalog #A-1219B) was used for the phenolphthalein and luminol presumptive blood tests. Diphenhydramine HCl, the main ingredient in Benadryl, can be used as a substitute for cocaine powder (35) and will show a positive result on presumptive drug tests (QuickCheck Narcotic Identification, Lynn Peavey, catalog #10130). Synthetic urine (20) and the Cerilliant standards from Sigma-Aldrich described in the instrumental analysis GC–MS laboratory were used for the urine sample.

188 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Table 3. Evidence Collected Immediately after the Crime Occurred and Analyses Carried Out on Each Sample Evidence

Analysis Methods

Blood sample from victim’s wound

• Presumptive blood test • Blood typing

Bloody shirt the victim was wearing

• Presumptive blood test • Blood typing • Dye fiber analysis on shirt • Attenuated total reflectance–Fourier transform infrared fiber analysis on shirt

Scrap of the suspect’s shirt, with a red stain

• Presumptive blood test • Blood typing • Dye fiber analysis on shirt piece • Attenuated total reflectance–Fourier transform infrared fiber analysis on shirt

Red stain on broken glassware

• Presumptive blood test • Blood typing

• Presumptive blood test • Blood typing English 101 syllabus found on the floor with a • Comparison of tread impression to any future suspect’s bloody footprint on the back of it shoes • Ninhydrin fingerprint detection Victim’s shoeprint

• Shoe tread pattern foam impression

Victim’s fingerprints

• Ink fingerprinting of victim

Bloody fingerprints left on the frame of the door to the laboratory

• Fingerprints photographed and collected with fluorescent fingerprint powder and tape

English 101 handwritten class notes

• Ink analysis • Handwriting analysis • Ninhydrin fingerprint detection

White powder found on laboratory floor

• Presumptive drug test

Coins

• Fingerprint analysis, powder, and superglue fuming

Lockbox containing the coins

• Fingerprint analysis, powder • Tool mark impressions collected with Mikrosil putty

Students analyzed the evidence collected from the crime scene over the rest of the semester, as they learned the appropriate techniques. Since this was a forensic investigation, students were cautioned that there was no way to obtain more sample if an analysis did not go well. The class worked on learning each technique in groups of two or three and collaborated between groups. Typically, the person or group the class deemed most successful at a technique collected and processed that particular piece of evidence. For example, some students were more successful than others at collecting fingerprints, particularly off vertical surfaces, so one student was elected to collect the bloody fingerprints from the door frame. As each piece of evidence was evaluated, the results were pooled in the evidence log, since not every group analyzed each piece of evidence.

189 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Once all of the initial evidence had been analyzed, students discovered that there were two different blood types at the scene, one belonging to Dr. Smith and one to the suspect. Fingerprints could not be lifted from the coins but were collected from the benchtop next to the coins, the handle of the lockbox, and the doorway to the laboratory. These prints did not match Dr. Smith’s reference sample. The fiber makeup from the piece of the suspect’s shirt was found to be a cotton/polyester mix, likely too common to be of use. The bloody shoeprint left on the back of the syllabus did not match the pattern on Dr. Smith’s shoes and was thought to belong to the suspect. The white powder on the floor was not drugs, but sugar that was being analyzed by the physical chemistry class. From these results, police officers and criminalists working on the case were able to collect more evidence to narrow down the suspect pool. Using the English 101 syllabus and the class notes, police collected handwriting samples from the course instructor for the five students in the course matching the description of the suspect. Students analyzed the handwriting and ink on each sample and checked each page for fingerprints. The handwriting from the class notes was a potential match to two of the five potential suspects (student D and student E). Police searched both students’ oncampus dorm rooms and collected T-shirts, shoes, and white powders from each room, as well as a set of fingerprints from each student. In addition, a police officer doing a routine sweep of the campus saw two figures, one in a hooded sweatshirt matching the description of the suspect’s, talking closely near the chemistry building. The figure in a hoodie started to hand something to the second person before they spotted the police car. Both parties immediately fled the scene. The police officer, having no evidence yet to detain them, did not give chase. Thinking the two people may have been involved in a drug deal, the officer investigated the nearby area. No drugs were found, but a gold coin, similar to those found in the laboratory after the theft and assault, was found on the ground near a footprint in a flower bed. A summary of the evidence collected from the secondary locations is shown in Table 4. Table 4. Evidence Collected during the Investigation of Secondary Sites and Analyses Carried Out on Each Sample Evidence

Analysis Methods

Handwriting samples from students in the English 101 course whose syllabus was found at the crime scene

• Handwriting analysis • Ink analysis • Ninhydrin fingerprint detection

Shoeprint left in the flower bed outside the chemistry building after campus police spotted the suspect

• Impression, plaster cast of footprint

White T-shirts collected from the two suspects’ dorm rooms, both with brown stains

• Dye fiber analysis • Attenuated total reflectance–Fourier transform infrared fiber analysis • Presumptive blood test • Blood typing

Pairs of shoes from the two suspects’ dorm rooms

• Impressions, foam mold of shoe tread pattern

White powder, collected from the two suspects’ dorm rooms

• Presumptive drug analysis

Fingerprints from the top two suspects

• Ink fingerprinting

Urine sample from student E

• Presumptive drug analysis • Evidentiary drug analysis by GC–MS

190 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

After analyzing all of the new evidence (results shown in Table 5), students found that the blood type, shoeprint, and fingerprints from student D were all a match to the evidence collected at the crime scene. No evidence was found to directly link student E to the crime in the laboratory, but illegal drugs were found in student E’s dorm room and, upon submitting a urine sample, in student E’s system. For their final written reports about the case, students wrote a letter to the DA summarizing their findings and advising whether they thought student D or student E should be charged in any crimes. Table 5. Results of Evidence Analysis Related to Student D and Student E Evidence

Student D

Student E

White T-shirt

Fibers match blend found on the scrap of the suspect’s shirt

Fibers match blend found on the scrap of the suspect’s shirt

Brown stain on T-shirt

Positive for presumptive blood, blood type matches that found at the Negative for presumptive blood scene

Shoe tread pattern

Matches bloody print left at the crime scene and impression left in the flower bed

Does not match any of the impressions in evidence

White powder

Negative for presumptive drugs

Positive for presumptive drugs

Fingerprints

Match fingerprints left at the crime scene

Do not match crime scene

Blood type

Matches suspect’s from the crime scene

Matches suspect’s from the crime scene

Urine sample

Not collected

Positive for amphetamines and methamphetamines

Laboratory Activities Each activity can be completed within a 2 h laboratory period with the exception of the ink analysis, since concentrating the ink sample can take some time. Table 6 lists the sources for chemicals and supplies needed that are not typically already found in a chemistry stockroom. A summary of the laboratory exercises carried out as part of the scenario is shown in Tables 3 and 4, and each of the activities is discussed further in the following sections. The instructor can determine the amount or type of analyte to spike each piece of evidence with. Learning objectives for the laboratory portion of the course center on (1) accurately carrying out forensic analyses, (2) choosing the appropriate technique for each piece of evidence in the scenario and analyzing the evidence, and (3) evaluating the data produced to determine the significance of the findings on the crime scenario. Objective 1 was assessed through observation by the instructor during the laboratory and grading of the laboratory notebooks. Objectives 2 and 3 were assessed by grading the written report about the crime scenario.

191 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Table 6. Manufacturer Sources for Chemicals and Supplies Used for the Tests in Tables 3 and 4 Technique

Chemical/Supplies

Manufacturer

Catalog #

Crime Scene Processing

Strip-n-grip pouches, 9" × 12"

Lynn Peavey

07710

Crime Scene Processing

Strip-n-grip pouches, 4" × 6"

Lynn Peavey

07714

Crime Scene Processing

Yellow markers numbered 1–15, 3"

Lynn Peavey

09902

Crime Scene Processing

Yellow markers numbered 16–50, 3"

Lynn Peavey

09908

Crime Scene Processing

Kraft bag evidence kit

Lynn Peavey

10112

Crime Scene Processing

Evidence labels, 2" × 3 ⅛", roll of 100

Lynn Peavey

05357

Crime Scene Processing

Flexible barrier tape, "crime scene"

Lynn Peavey

56273

Fingerprint Analysis Dual-use magnetic fingerprint powder

Lynn Peavey

55574

Fingerprint Analysis Orange fluorescent magnetic powder

Lynn Peavey

61021

Lynn Peavey

35555

Fingerprint Analysis Fiber duster

Lynn Peavey

05539

Fingerprint Analysis Lifting tape, 4" × 360" roll

Lynn Peavey

05567

Fingerprint Analysis ScaleLift backing cards, 4" × 5", 100 pk Lynn Peavey

05591

Fingerprint Analysis Criminal record cards, 100 pk

Lynn Peavey

05996

Lynn Peavey

09621

Fingerprint Analysis BlackMagic ceramic ink pad, 1.5" round Lynn Peavey

09622

Fingerprint Analysis

Fingerprint Analysis

Fingerprint Analysis

Spring-loaded magnetic fingerprint powder wand

BlackMagic ceramic ink pad, small rectangle

Latent print reference pad (sebaceous oil and amino acid based)

Lynn Peavey

95389

Fingerprint Analysis Ninhydrin solution

Lynn Peavey

05060

Fiber Analysis

Textile identification stain 1, 250 g

Testfabrics, Inc.

505001

Fiber Analysis

Textile identification stain 3A, 250 g

Testfabrics, Inc.

505003

Fiber Analysis

Multifiber test fabric. 43 precut 5 × 10 cm cold cut edges

Testfabrics, Inc.

408001

Presumptive Drug Analysis

QuickCheck Narcotic Identification, cocaine tube

Lynn Peavey

10130

192 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Table 6. (Continued). Manufacturer Sources for Chemicals and Supplies Used for the Tests in Tables 3 and 4 Technique

Chemical/Supplies

Manufacturer

Catalog #

Presumptive Drug Analysis

STATDIP, 5 drugs (THC, opiates, cocaine, amphetamines, methamphetamines)

Micro-Distributing

SDP254

Presumptive Drug Analysis

Adulterant strips

Micro-Distributing

Adulterant strips

Ink Analysis

Silica G TLC Plates with UV254

Fisher Scientific

50-113-8565

Blood Typing

Simulated ABO and Rh blood typing

Ward's Science (VWR)

470213-350

Staged Assault

Stage blood

Amazon

Mehron Makeup Stage Blood

Presumptive Blood Analysis

Synthetic blood, Arrowhead Forensics

Lynn Peavey

01606

Presumptive Blood Analysis

QuickCheck Bloodstain Green

Lynn Peavey

10131

Presumptive Blood Analysis

Luminol powder

Lynn Peavey

PF15031-5g

Presumptive Blood Analysis

Phenolphthalein presumptive test

Lynn Peavey

PF30040-15mL

Tool Mark Analysis

Mikrosil Toolmark Putty, gray

Lynn Peavey

59853

Footprint Analysis

Copy Cast casting powder, 6.25 lb pouch

Lynn Peavey

06822

Footprint Analysis

Casting frame, 12”

Lynn Peavey

05751

Footprint Analysis

Biofoam impression, 2-sided

Lynn Peavey

95019

Crime Scene Processing Following the scenario described in the “Setting the Scene” section, students sealed off the crime scene, searched for evidence, placed evidence marker numbers at each piece of evidence (Figure 8), sketched out the crime scene, photographed all evidence, and collected numerous different types of evidence, as shown in Tables 3 and 4 (30, 32, 33). Biological samples can grow mold if stored in airtight containers and should be collected in paper bags or wrapping paper. Glass, fibers, and other trace evidence is best collected in small plastic or metal pillboxes. Powders should be folded into paper before being placed in an evidence bag to prevent the loss of evidence (30).

193 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Figure 8. Identified and tagged pieces of evidence in the crime scene scenario. (Top) Bloody footprint left on the back of an English 101 syllabus with a ruler included for scale. (Bottom) Case containing suspected Captain Thorpe treasure with tool marks on the lock, fingerprints on the case handle and benchtop, and coins that were in the case before the theft. Photographs courtesy of Sarah Gray. Fingerprinting Fingerprint residue is made up of 98.5% water, which can evaporate, and 1.5% solid components. The solid components are made up of 1% organic materials (lactic acid, fatty acids, riboflavin, pyroxidin, glucose, sugars, ammonia, urea, creatin, albumin, peptides, proteins, isoagglutinogens, amino acids, and lipids) and 0.5% inorganic materials (Na+, K+, Ca2+, Cl−, PO43−, CO32−, and SO42−) (23, 36). Several methods can be employed to lift fingerprints from porous surfaces. One method is iodine fuming, which involves placing the item to test in a fuming chamber with a few heated crystals of solid iodine. The iodine is thought to react with either the fatty acids and oils on the print or the water from the perspiration (23). The iodine prints will fade over time and need to be fixed with a 1% starch solution. This fixing does prevent the use of any other fingerprint development techniques (23). In this course, a beaker with a watch glass on top was used as the fuming chamber. The beaker was placed on top of a hot plate in a fume hood (37). Another detection method used ninhydrin (triketohydrindene hydrate), which reacts with the amino acids present in the fingerprint residue to form Ruhemann’s purple. The ninhydrin solution was sprayed on the porous surface and either left to react for 24 h or subjected to a heat gun to speed up the reaction. This method will not work if the print is too old or if insufficient amino acids were excreted in the perspiration. To ensure prints show up for the crime scenario, an amino acid fingerprint pad can be used to leave the prints. 194 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

For lifting prints from nonporous surfaces, cyanoacrylate (super glue) fuming can be used. The cyanoacrylate reacts with the moisture left in a fingerprint and forms a few long chains of the super glue polymer. This method does not work well for old prints, since there is not enough moisture left in the print for the reaction to take place. Here, a few drops of super glue were deposited into a metal weighing dish on top of a hot plate. The weighing dish, hot plate, and item to be fingerprinted were placed in a fuming chamber (a medium-sized plastic tote). Dusting powders can also be used to visualize prints on nonporous surfaces. Magnetic powders make it easier to remove the excess powder without disturbing the print. Fluorescent powders are also available and fluoresce under UV light, which can help visualize a print on a background with a pattern that might otherwise obscure it. Trace Evidence: Fibers Fiber analysis is used to match fibers found at the crime scene to suspects. In this scenario, the Tshirt scrap torn from the perpetrator’s shirt by the victim was compared to T-shirts in the suspects’ rooms. When initially learning the analysis procedures, students were asked to bring in a fiber sample from home for examination. Many students had very little knowledge of where natural fibers come from (e.g., that cashmere comes from goats) or the chemical structures for synthetic fibers. Both natural and synthetic fibers are made up of polymers, and including them in the discussion brings polymer chemistry into the curriculum in a way that is relatable to students. Nylon is formed from the condensation of a diamine and a dicarboxylic acid, while a polyester such as Dacron is formed by the condensation reaction of an ester with an alcohol (23). The polymers in wool and silk are polyamides composed of monomer units, while cotton is composed of cellulose (23). Fibers can be examined under a microscope and compared to standards. Differential staining can also be used to determine what types of fibers a garment is made of. Since the chemical structures of different fibers are different, each fiber will stain a different color. Two stains from Testfabrics, Inc., test fabric identification stain 1 and 3A, can be used to stain unidentified fabrics. Each stain consists of three different dyes. A test fabric consisting of 13 fibers woven together was also stained and used as a reference. While some fabrics may stain a similar color in one stain, they can be differentiated by using the second stain (23). Students can analyze the dye solutions using UV–vis spectrophotometry to determine how many dyes are present in the mixture by identifying the number of maximum wavelength values. Finally, fibers can be differentiated using attenuated total reflectance–Fourier transform infrared (ATR–FTIR) spectroscopy. With the ATR attachment, single fibers can be analyzed, which would be difficult with the staining technique. In the forensic scenario, students used both differential dying and ATR–FTIR to analyze fiber evidence. Results were compared between the two analysis techniques. Drug Analysis Presumptive analysis of solid drug samples can be completed using quick test kits that involve a characteristic color change when the suspected drug sample is mixed with various reagents. These tests are often carried out before more definitive analysis methods to save time and money on the analysis. To test for cocaine, a cobalt thiocyanate solution (2g/100 mL of water) is reacted with the suspected powder. If cocaine is present, a flaky blue precipitate is formed (23). The Lynn Peavey QuickCheck Narcotic Identification tubes for cocaine were used in the forensic scenario. As mentioned earlier, diphenhydramine HCl, the main ingredient in Benadryl, can be used as a substitute for cocaine powder (35) and will yield a positive result on presumptive drug tests. QuickCheck kits are available from Lynn Peavey for numerous other narcotics, and other types of 195 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

test kits are available for drugs in other categories. Confirmatory analysis of solid drug samples can be carried out using FTIR spectroscopy. An ATR attachment makes the measurement easier, since the salt plate sample holders are not needed. Powdered drug samples may need to be separated from impurities (e.g., binders, stabilizers, or coatings) before measurement. Sugars and starches used for coatings are water soluble and can be eliminated by extracting the drug into an organic solvent. Drugs in the salt form should be made basic before extraction so that they are more soluble in organic solvents (23). Presumptive analyses can also be carried out for drugs in urine samples. In this case, the STATDIP cassette was used to test for THC, opiates, cocaine, amphetamines, and methamphetamines in urine. Once the urine sample has been collected (or prepared, as was the case for the forensic scenario), the cassette lid is removed, revealing five test strips that are dipped into the urine sample. The cassette uses an immunoassay to selectively detect multiple drugs. Colored microbeads are treated with antibodies to a specific drug. In a negative sample, the microbeads interact with some of the antigen (i.e., drug) that has been embedded in the viewing region of the test strip, and two colored lines form. If the drug is present, it binds to the antibody, and the antibody does not have any sites left to react with the antigen in the test strip. A second reference line changes color with or without drugs present to show the test is working (23, 38). Since suspects may try to fool chemical tests by altering their urine samples or trying to flush drugs out of their system ahead of the test, test strips for adulterants are used on the urine sample in addition to the drug analysis cassette. The adulterant test strips test for creatinine, nitrite, glutaraldehyde, pH, specific gravity, bleach, and pyridinium chlorochromate (39). The GCMS method described for the instrumental analysis course can be used for definitive identification of drugs in urine samples. Questioned Documents The ink from questioned documents was extracted from paper evidence and compared to ink extracted from marks on paper made with reference pens and markers. Note that this method is destructive to the piece of evidence. Fingerprint analysis can be carried out on the paper first, but chemical tests may alter the results of the ink analysis. The ink analysis method was taken from Kaplan’s forensic science laboratory manual (23). A portion of the document with ink on it was cut out and placed in a test tube with 2–3 mL of methanol. The test tube was heated in a beaker of water on a hot plate to increase the solubility of the ink. After the ink was extracted from the paper, the liquid was transferred to an Eppendorf tube and evaporated with a stream of air. If all the methanol evaporated, it was reconstituted with a few more drops of methanol. The ink sample and the reference samples were then spotted onto a TLC plate, which was developed in 60:10:20:0.5 n-butyl alcohol, ethyl alcohol, water, and acetic acid. The retention factor of each spot was calculated, and the ink unknowns were compared to the standards for identification. Forensic Serology Several different tests were carried out for presumptive identification of blood. Synthetic blood from Lynn Peavey was used to produce positive results on the presumptive tests. For the phenolphthalein test, phenolphthalein and hydrogen peroxide were reacted with a suspected blood sample using the Lynn Peavey phenolphthalein test kit. Using this test, if blood is present, the pink phenolphthalein anion and water are formed (23). The luminol test was also used to test for blood. A luminol solution was sprayed where it was suspected blood may have been cleaned up and was not visible to the naked eye. The room was darkened, and a UV light source was used to search for 196 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

the presence of blood, as the luminol/blood mixture glows blue under the UV light. QuickCheck Bloodstain Green test kits were also used. The suspected blood sample was rubbed with the cotton swab from the kit, and the reagent was added to the swab. A green color indicates that blood is present. For samples that were determined to be blood, blood typing was carried out using a Ward’s Science blood typing kit. The day after students analyzed the blood swabs by the presumptive techniques, they were provided with an Eppendorf tube with blood for the blood typing analyses (from the kit). Students were told that a technician had extracted some of the blood from the swab for them to facilitate switching out the type of synthetic blood that was needed for each test to work. Firearms and Impressions No firearm evidence was used in the forensic scenario, but students did analyze bullet (Figure 9) and shell casing samples by measuring the diameter to determine bullet caliber and placing them under a microscope to look for land, groove, and firing pin impressions. Bullet samples were obtained through a colleague whose father is a gun collector.

Figure 9. Bullet samples for students to analyze for caliber and land and groove marks. Photograph courtesy of Sarah Gray.

Figure 10. Tool mark impressions from the coin case cast in Mikrosil putty. Marks A and C were consistent with the use of a chiseling tool to open the case. Mark B was wider and more consistent with a flat-head screwdriver. Photograph courtesy of Sarah Gray. 197 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Impressions were collected of the tool marks left on the coin case from the suspect breaking the lock using Mikrosil putty. The Lynn Peavey kit contains two tubes of putty, and a sample of each type was mixed together to form the final putty. The putty was applied to the tool marks on the case shortly after it was mixed. Over the course of a few minutes, the putty set to the consistency of soft rubber. The impression was then peeled from the case (Figure 10). Tools found in the lab were used to make impressions in biofoam. These impressions were compared to the tool mark cast to determine what tool was used to make the marks. A shoeprint (Figure 11, top) left in the soil outside by a suspect was cast using plaster of Paris, which hardens and forms gypsum when reacted with water. Once suspects were identified, shoes were collected from their homes and the biofoam was used to make an impression of the tread markings (Figure 11, bottom). These markings were compared to those on the plaster cast of the shoeprint.

Figure 11. (Top) Shoeprint impression left by the suspect in the dirt outside the chemistry building as they fled the crime scene. (Bottom) Foam impressions of the treads on shoes collected from two suspects. Photographs courtesy of Sarah Gray.

198 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Assessment of Student Engagement and Critical-Thinking Skills The forensic theme successfully grabbed students’ attention before the course even started, convincing them to sign up. In both semesters the course was run, it had 12–15 students enrolled, compared to 3 students who signed up for a polymer upper-division course in the same semester. The students were curious enough about solving the case that they were willing to go along with it despite the crime scenario being a bit contrived. During one semester, the class included two students who were so invested in the scenario that when they found Dr. Smith “bleeding” in the lab they proceeded to use their emergency medical technician training to bandage the wound with gauze wrap they were carrying. Student feedback was generally very positive for the course. Informal evaluations conducted by the instructor at the end of the semester showed each student had a different favorite topic, with almost all covered topics included at least once. These results suggest that the entire class could find at least one forensic analysis area that particularly appealed to them. Students also commented that the laboratory portion of the class contributed to their learning and allowed them to better understand how the lecture material is applied to evidence in real crimes. The learning goals for the laboratory portion of the course were that students would be able to • Perform analyses of evidence in a mock crime scene; • Evaluate the data to solve a crime scenario; and • Present key findings in written form. The chemistry behind the laboratory techniques was covered during lecture time, and students practiced each technique on test samples. Students then had to apply their knowledge of each technique to the crime scenario evidence. They had to decide for themselves what tests to run on each piece of evidence, as this information was not provided to them. As new evidence came in, the class would discuss what techniques they thought should be used. Since this was a crime scene scenario, no more evidence was provided to the students if they used all of it in incorrect analyses or contaminated a sample. Students also had to determine what order to run the analyses in to avoid contaminating the evidence for future tests. Except for blood typing, which often had competing results between students, the class was able to carry out each analysis successfully, and the blood type of samples could usually be corrected if there was enough sample left for repeat analyses. By analyzing the evidence in Table 3, the class was able draw initial conclusions about the data and request that campus police search and interview suspects. This led to the dorm room search for students D and E, and the analysis and evaluation of the evidence from this search allowed the class to the draw the correct conclusions about the guilt of the students on various charges.

Conclusion Overall, the use of forensics in both courses increased student engagement in analytical analyses. Students completed the instrumental analyses laboratories with much more enthusiasm than was shown in previous semesters when none of the laboratory exercises related to one another. The forensic theme of the upper-division elective course attracted students to the course before it even started, and the crime scenario kept them engaged throughout the semester. The crime scenarios provided real-world problems that required critical-thinking skills to solve. Students analyzed data for both cases, evaluated what that data meant for the crime scenarios, and defended conclusions based on their evaluations. This put the focus of both classes on higher-order thinking skills (apply, analyze, and evaluate in Bloom’s taxonomy) rather than on rote memorization of instrumentation parts or forensics facts. 199 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

References 1.

2.

3.

4. 5. 6. 7.

8.

9. 10. 11.

12. 13. 14. 15. 16.

17.

Meyer, A. F.; Knutson, C. M.; Finkenstaedt-Quinn, S. A.; Gruba, S. M.; Meyer, B. M.; Thompson, J. W.; Maurer-Jones, M. A.; Halderman, S.; Tillman, A. S.; DeStefano, L.; Haynes, C. L. Activities for Middle School Students To Sleuth a Chemistry “Whodunit” and Investigate the Scientific Method. J. Chem. Educ. 2014, 91, 410–413. Ahrenkiel, L.; Worm-Leonhard, M. Offering a Forensic Science Camp To Introduce and Engage High School Students in Interdisciplinary Science Topics. J. Chem. Educ. 2014, 91, 340–344. Ravgiala, R. R.; Weisburd, S.; Sleeper, R.; Martinez, A.; Rozkiewicz, D.; Whitesides, G. M.; Hollar, K. A. Using Paper-Based Diagnostics with High School Students to Model Forensic Investigation and Colorimetric Analysis. J. Chem. Educ. 2014, 91, 107–111. Cresswell, S. L.; Loughlin, W. A. An Interdisciplinary Guided Inquiry Laboratory for First Year Undergraduate Forensic Science Students. J. Chem. Educ. 2015, 92, 1730–1735. Specht, K. M.; Boucher, M. A. A Forensic-Themed Case Study for the Organic Lab. J. Chem. Educ. 2009, 86, 847. Szalay, P. S.; Zook-Gerdau, L. A.; Schurter, E. J. A Multi-Technique Forensic Experiment for a Nonscience-Major Chemistry Course. J. Chem. Educ. 2011, 88, 1419–1421. Grannas, A. M.; Lagalante, A. F. So These Numbers Really Mean Something? A Role Playing Scenario-Based Approach to the Undergraduate Instrumental Analysis Laboratory. J. Chem. Educ. 2010, 87, 416–418. Hall, A. B.; Drugan, J. E.; Anzivino, B.; Tilley, L. J.; Ingalls, L. R. Got a Match? Ion Extraction GC–MS Characterization of Accelerants Adsorbed in Charcoal Using Negative Pressure Dynamic Headspace Concentration. J. Chem. Educ. 2009, 86, 55. Critical Thinking: Where to Begin. http://www.criticalthinking.org/pages/critical-thinkingwhere-to-begin/796 (accessed April 4, 2019). ACS Approval Program for Bachelor’s Degree Programs. https://www.acs.org/content/acs/en/ about/governance/committees/training/acsapproved.html (accessed Feb 5, 2019). Seng Tan, O. Students’ Experiences in Problem‐based Learning: Three Blind Mice Episode or Educational Innovation? Innovations in Education and Teaching International 2004, 41, 169–184. Wenzel, T. J. AC Educator: Problem-based Learning: In Need of Supporting Materials. Anal. Chem. 2001, 73, 501 A–502 A. Larive, C. K. Educational Approaches for Analytical Science. Anal. Bioanal. Chem. 2004, 378, 1399–1400. Quattrucci, J. G. Problem-based Approach to Teaching Advanced Chemistry Laboratories and Developing Students’ Critical Thinking Skills. J. Chem. Educ. 2018, 95, 259–266. ACS Guidelines & Supplements. https://www.acs.org/content/acs/en/about/governance/ committees/training/acs-guidelines-supplements.html (accessed Dec 19, 2018). McCloskey, K. L.; Garriott, J. C.; Roberts, S. M. Quinine Concentrations in Blood Following the Consumption of Gin and Tonic Preparations in a Social Setting. J. Anal. Toxicol. 1978, 2, 110–112. Perry, D. C.; Ratcliffe, B. E. Heroin and Cocaine Adulteration. Clin. Toxicol. 1975, 8, 239–243. 200 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

18. Fiorentin, T. R.; Krotulski, A. J.; Martin, D. M.; Browne, T.; Triplett, J.; Conti, T.; Logan, B. K. Detection of Cutting Agents in Drug-Positive Seized Exhibits within the United States. J. Forensic Sci. 2019, 64, 888–896. https://onlinelibrary.wiley.com/doi/abs/10.1111/15564029.13968 (accessed May 2, 2019). 19. Mikuš, P.; Maráková, K.; Veizerová, L.; Piešťanský, J.; Galba, J.; Havránek, E. 2D Capillary Electrophoresis Hyphenated with Spectral Detection for the Determination of Quinine in Human Urine. J. Chromatogr. Sci. 2012, 50, 849–854. 20. Khan, L. B.; Read, H. M.; Ritchie, S. R.; Proft, T. Artificial Urine for Teaching Urinalysis Concepts and Diagnosis of Urinary Tract Infection in the Medical Microbiology Laboratory. Journal of Microbiology & Biology Education 2017, 18, 1–6. 21. O’Reilly, J. E. Fluorescence Experiments with Quinine. J. Chem. Educ. 1975, 52, 610. 22. Zabzdyr, J. L.; Lillard, S. J. Was the Driver Drunk? An Instrumental Methods Experiment for the Determination of Blood Alcohol Content. J. Chem. Educ. 2001, 78, 1225. 23. Kaplan, L. National Science Foundation Sponsored Chemistry Collaborations, Workshops & Communities of Scholars: Successor to the Center for Workshops in the Chemical Sciences: Forensic Science Workshop; Williams College: Williamstown, MA, June 2014. 24. United Chemical Technologies. CLEAN SCREEN DAU. https://sampleprep.unitedchem. com/clean-screen-dau (accessed Jan 7, 2019). 25. Excedrin Migraine. https://www.excedrin.com/products/migraine/ (accessed Jan 15, 2019). 26. Ferguson, G. K. Quantitative HPLC Analysis of an Analgesic/Caffeine Formulation: Determination of Caffeine. J. Chem. Educ. 1998, 75, 467. 27. Sawyer, M.; Kumar, V. A Rapid High-Performance Liquid Chromatographic Method for the Simultaneous Quantitation of Aspirin, Salicylic Acid, and Caffeine in Effervescent Tablets. J. Chromatogr. Sci. 2003, 41, 393–397. 28. Kagel, R. A.; Farwell, S. O. Analysis of Currently Available Analgesic Tablets by Modern Liquid Chromatography: An Undergraduate Laboratory Introduction to HPLC. J. Chem. Educ. 1983, 60, 163. 29. Kanu, A. B.; Pajski, M.; Hartman, M.; Kimaru, I.; Marine, S.; Kaplan, L. J. Exploring Perspectives and Identifying Potential Challenges Encountered with Crime Scene Investigations When Developing Chemistry Curricula. J. Chem. Educ. 2015, 92, 1353–1358. 30. Saferstein, R. Criminalistics: An Introduction to Forensic Science, 11th ed.; Pearson: Boston, MA, 2015. 31. Thompson, R. M. Firearm Identification in the Forensic Science Laboratory; National District Attorneys Association: Alexandria, VA, 2010. 32. Technical Working Group on Crime Scene Investigation. Crime Scene Investigation: A Guide for Law Enforcement; National Forensic Science Technology Center: Largo, FL, September 2013. 33. Larsen, J.; Harris, D. Crime Scene Forensic Evidence Collection Guidelines for Defense Attorneys. The Champion 2011October, 28–35. 34. Georgia Southern University. Cyber Forensics Division, Center for Applied Cyber Education. https://academics.georgiasouthern.edu/cace/services/cyber-forensics-division/ (accessed Jan 12, 2019).

201 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

35. Hasan, S.; Bromfield-Lee, D.; Oliver-Hoyo, M. T.; Cintron-Maldonado, J. A. Using Laboratory Chemicals To Imitate Illicit Drugs in a Forensic Chemistry Activity. J. Chem. Educ. 2008, 85, 813. 36. Brettell, T.; Meloan, C. E.; Saferstein, R. Lab Manual for Criminalistics: An Introduction to Forensic Science; Pearson: Boston, MA, 2015. 37. Flinn Scientific. Iodine Fingerprint. https://www.flinnsci.com/iodine-fingerprint/dc10189/ (accessed April 3, 2019). 38. Micro-Distributing. STATDIP. http://micro-distributing.com/stat_dip.cfm (accessed April 4, 2019). 39. Micro-Distributing. Intect 7 Adulteration Testing. http://micro-distributing.com/prod_intect7. cfm (accessed April 4, 2019).

202 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Chapter 11

Downloaded via MIAMI UNIV on November 5, 2019 at 02:24:24 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Integration of Forensic Themes in Teaching Instrumental Analysis at Pace University Elmer-Rico E. Mojica, Robert Marvin, Normisha Evans, Lauren Reilly, David Mendoza, Styliani Karpadakis, Charles Cusumano, Demosthenes Athanasopoulos, and Zhaohua Dai* Forensic Science Program, Department of Chemistry and Physical Sciences, Pace University, 1 Pace Plaza, New York, New York 10038, United States *E-mail: [email protected].

Pace University’s forensic science program, in the Department of Chemistry and Physical Sciences, originated and was developed as an expansion of the scope of the hard sciences (chemistry, physics, and biology) in the forensic field. This is in line with the findings and recommendations of the National Research Council on forensic science. At Pace, we continue to integrate teaching and research to train high-caliber forensic scientists. We also introduce forensic themes in teaching chemistry. We are in the final stages of completing the revision of Instrumental Analysis, Analytical Spectroscopy, and Forensic Separations chemistry lecture and laboratory curricula. One of the newly developed goals is to take interesting and exciting topics from recent or current research and development in the field of forensic chemistry and adapt them to undergraduate and graduate (master’s level) teaching settings. Another goal is to stimulate students to appreciate the variety of modern experimental methodologies within the subdisciplines of chemistry, and to encourage students to pursue careers in the discipline of chemistry, including forensic chemistry. In this chapter, we will describe our effort in preparing online interactive eTextbooks with online assignments using examples adapted from leading forensic science journals such as the Journal of Forensic Sciences and Microgram Journal and the implementation of such eTextbooks in our teaching. We will also describe our efforts in adapting protocols published in leading journals, such as chiral drug analysis and hydrocarbon analysis, in our teaching labs using gas chromatography–mass spectrometry, capillary electrophoresis, Fourier-transform infrared spectroscopy, Raman spectroscopy, and nuclear magnetic resonance spectroscopy. These topics are related to forensic drug analysis, arson accelerant analysis, and environmental forensic analysis. The impact on student learning will be analyzed and discussed.

© 2019 American Chemical Society Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Introduction and Forensic Science Curricula Forensic science has become an increasingly prominent area of science within the last 15 years. The increase in students studying forensic science comes with the potential to produce high caliber science graduates with sought-after skills in critical thinking, analysis, interpretation, and communication. The forensic science program at Pace University is in the middle of its second decade in operation and it is stronger than ever. We are dedicated to educating future highly qualified professionals in forensics at both the graduate and undergraduate levels. The successful employment of our graduates in forensic facilities of major law enforcement agencies has already built a tradition for the program. The major strengths of the forensic science program at Pace are its scientific rigor and orientation, the dedication and qualifications of its faculty, its flexible organization, and its strategic location. Our strategic location in downtown New York City (NYC) provides endless opportunities for collaborations and affiliations with other forensic science facilities, such as the forensic biology labs of the NYC Office of Chief Medical Examiner, the Drug Enforcement Administration (DEA) labs, the NYC police laboratory, among others. Due to very strong relationships with such prominent and highly recognized (both nationally and internationally) forensic institutions, our program was able to support internships and thesis projects for many students. The vast majority of students were able to find jobs in the field as a direct continuation of the internships and projects that they had been working on. Most of the early forensic science programs in the country emerged as an evolution of traditional criminalistics programs that gradually incorporated scientific methodologies, maintaining in their core philosophy the emphasis on expertise assessment rather than quantitative analysis of forensic evidence. Pace's forensic science program, however, originated and was developed as an expansion of the scope of the hard sciences (chemistry, physics, and biology) in the forensic field. The component of criminalistics was added as a considerable but not dominant element in the curriculum. The forensic methodology of our program aims to express the "beyond reasonable doubt" convincing evidence in mathematical terms of statistics and probability, using established scientific criteria. This is in line with the findings and recommendations of the report of the National Research Council on forensic science (1). Our students are trained in the theory and practice of the physical, chemical, and biological sciences as well as a wide spectrum of criminalistics topics. During their studies, they develop problem-solving and critical-thinking techniques needed to function in contemporary society. At the same time, they become familiar with professional issues associated with the forensic practice such as evidence handling, chain of custody, expert witness testimony, crime laboratory operations, quality control and assurance, and professional and organizational ethics. Upon completion of their studies, students in our program should demonstrate competence to function as scientists in forensic science–related fields and be prepared to pursue their studies at either graduate schools or health professional schools. They should understand the fundamental physical, chemical, biological, and forensic concepts and be able to engage in scientific research and communications. Our curriculum is designed to achieve such outcomes, in line with recommendations from the National Research Council regarding strengthening forensic science in the United States (1). The versatility of their scientific training and ability to adapt to the demands of a constantly changing market are enshrined.

204 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Table 1. Suggested Course Sequence for the Bachelor’s Program in Forensic Science at Pace University FIRST YEAR First semester

Second semester

ENG 110: Compositiona

ENG 120: Critical Writing

BIO 101: General Biology I

BIO 102: General Biology II

CHE 111: General Chemistry I

CHE 112: General Chemistry II

MAT 131: Calculus Ib

MAT 132: Calculus IIb

UNV 101: Freshman Seminar

CIS 101: Introduction to Computing SECOND YEAR

First semester

Second semester

CHE 223: Organic Chemistry I

CHE 224: Organic Chemistry II

PHY 111: General Physics I

PHY 112: General Physics II

Second language

Second language

BIO 231: Genetics

CHE 200: Math for Physical Chemistry THIRD YEAR

First semester (17 credits)

Second semester (18 credits)

CHE 221: Analytical Chemistry

CHE 331: Instrumental Analysis

CHE 301: Physical Chemistry I

CHE 302: Physical Chemistry II

FOR 252: Crime Scene Processing

FOR 251: Criminalistics

AOK II to IV

MAT 141: Statistics for Life Science

COM 200: Public Speaking

BIO 345: Toxicology FOURTH YEAR

First semester (1516 credits)

Second semester (1115 credits)

CHE 326: Biochemistry

FOR 505: Molecular Biology

ENG 201: Writing in the Discipline

FOR 537: Forensic Biology

FOR 531: Forensic Microscopy

FOR 492: Forensic Seminar

AOK II to IV

CRJ 261: Criminal Investigation

AOK II to IV

Open electivec

a Appropriate English classes determined by placement exam.

b Math classes determined by placement exam.

Students placed in MAT130 (pre-calculus) have to complete the calculus sequence (by taking summer classes) prior to the beginning of their second academic year. c CHE 330 Advanced Inorganic Chemistry is suggested as an open elective for ACS certification.

Our program offers a variety of core scientific subjects, which are conditio sine qua non in the field; in addition, we offer courses that prepare our students to work in the field competently from day one and be able to cross paths with similar disciplines. 205 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

The traditional science components are taught by the faculty members of the chemistry and biology departments. The criminalistics components are taught by adjunct and clinical professors who worked at major forensic laboratories and bring real world experience, current best practice of the field, and their own expertise and professional network to our program. Moreover, forensic science is a rapidly developing field where regulations and standards change often. Our program continuously adjusts the content of the courses offered and occasionally undergoes more structural adjustments. Forensic science undergraduate students receive a liberal arts education designed to provide them with the knowledge and skills for forensic science professionals. This course of study also prepares students for graduate school entrance in forensic science, chemistry, biochemistry, biology, health professions, and other related areas. Since our forensic science program is hosted in the Department of Chemistry and Physical Sciences where our majors are eligible for approval by the American Chemical Society (ACS), our students can get an ACS certified bachelor of science degree in forensic science if they follow our curriculum. Our current undergraduate program curriculum is shown in Table 1. The graduate program prepares students for immediate practice in the forensic field and related fields. Our aim is to maintain a balance between the advanced and specialized topics and the wide scope of the field. The internship requirement extends the classroom experience to the actual applications. Finally, the thesis elective provides research experience in a specific area of forensics, so students can build expertise in the area. Table 2. Suggested Course Sequence for the Master’s Program in Forensic Science at Pace University FIRST YEAR Fall semester

Spring semester

FOR 531: Forensic Microscopy

FOR 537: Forensic Biology

FOR 610: Professional Issues

FOR 620: Analytical Spectroscopy

FOR 635: Forensic Pharmacology

FOR 625: Crime Scene Investigation SECOND YEAR

Fall semester

Spring semester

FOR 615: Separations Chemistry

FOR 621: Forensic Internship FOR 699: Forensic Seminar

FOR 705: Forensic Anthropologyaa

FOR 736 Advanced Topics Criminalisticsaa

FOR 707: Advanced Topics DNAaa

FOR 706: Forensic Toxicologyaa

FOR 798: Research Methodsaa

FOR 799: Thesisaa

aa Electives:

Students need to take two electives. Students who have taken FOR 531 or FOR 537 as undergraduates at Pace need to take two additional electives. Other electives can be offered according to demand.

In both our undergraduate (Table 1) and graduate (Table 2) curricula, there are quite a number of chemistry classes, although not just because the program is hosted in the chemistry department. Drug analysis, together with toxicology, is a big part of forensic science. DNA typing is another important part. Both require a solid understanding of general, organic, physical, and analytical 206 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

chemistry as well as biochemistry. Many chemistry and biochemistry students go on to study forensic science in graduate school. Some forensic science students pursue advanced degrees and careers in chemistry. Both graduate and undergraduate students have the opportunity to learn about the most recent developments by attending seminars delivered by guest speakers and lecturers from different areas of the forensic field. Many of our speakers are forensic chemists, therefore forensic themes are often incorporated in many chemistry classes, especially in Instrumental Analysis, at Pace.

Instrumental Analysis in Forensic Science Methamphetamine is a DEA Schedule II controlled substance. However, it is chiral and its two enantiomers are treated differently in the guidelines published by the U.S. Sentencing Commission: 1 g of d-methamphetamine is equivalent to 10 kg of marijuana, which is 250 times the 40 g of marijuana to which 1 g of l-methamphetamine is equivalent. Pharmaceuticals containing small amounts of l-methamphetamine, such as Vicks Vapor Inhaler, are not controlled, while there are severe consequences for offenses connected to “ice,” which contains at least 80% dmethamphetamine hydrochloride salt by weight. In a hypothetical scenario, a suspected clandestine "meth" lab was raided by a law enforcement unit and a large amount of a white powdery substance was found. However, the defense lawyer claimed that the white powder (the “exhibit”) was just l-methamphetamine, not “ice,” mixed with a lot of caffeine, in order to get lighter sentences for the suspects. A forensic scientist is asked to determine (1) whether the exhibit contains d- or lmethamphetamine; (2) if it contains the d- isomer, whether it is the free base form or the hydrochloride salt form; (3) if d-methamphetamine hydrochloride is present in the exhibit, how much of it is in the exhibit; and (4) whether the weight percentage of d-methamphetamine hydrochloride is high enough that the exhibit can be classified as "ice.” Several tools are needed in order to answer these questions, ideally following a typical forensic approach: narrowing down possibilities until arriving at a conclusion. Initially, observations are done to describe the contents upon accepting the physical evidence, which is followed by a color/presumptive test performed on a small portion of the evidence to determine the category of the substance. Finally, the evidence is subject to instrumental analysis for a definitive identification and quantitation after sample preparation. When the evidence is not pure, gas chromatography (GC), high-performance liquid chromatography (HPLC), or capillary electrophoresis (CE), often coupled with mass spectrometry (MS), are often used to separate the components in the mixture to give insight into the identity and concentration of each component. Identification is generally achieved with the help of spectrometers such as infrared (IR), Raman, nuclear magnetic resonance (NMR), and mass spectrometers. All the authors have been engaged in revising the chemistry piece of the forensic science curricula, as well as our chemistry and biochemistry curricula. We have obtained and maintained U.S. DEA and New York State Department of Health Controlled Substance Licenses for relevant activities. We have been integrating research and laboratory teaching of the undergraduate-level Instrumental Analysis, and the master’s level Forensic Separations Chemistry and Analytical Spectroscopy. Instead of following traditional lab textbooks, the students are asked to consult primary research literature and procedures, then replicate and modify those experiments. Chiral Separations One example is the chiral derivatization using (S)-(−)-N-(trifluoroacetyl)pyrrolidine-2carbonyl chloride (L-TPC) in the identification and quantitation of the enantiomers of ephedrine 207 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

and pseudoephedrine using GC–MS with a nonchiral column (HP-5 ms) (Figure 1), which follows an article in Chirality on the chiral derivatization of amphetamine and methamphetamine (2). Students are also required to run underivatized samples using the same column. In this experiment, students learn firsthand that although enantiomers have the opposite optical rotation, circular dichroism, they cannot be differentiated by nonchiral chromatography. However, they can be separated on a nonchiral column after derivatization by a chiral reagent. Enantio-impurities can be detected after chiral derivatization, as represented by the smaller peaks in Figures 1(a) and 1(b), respectively. Such enantio-impurities do not show up in the GC chromatograms without chiral derivatization or without using a chiral column. At the same time, diastereomers (ephedrine and pseudoephedrine) can be separated by an achiral column without derivatization.

Figure 1. GC chromatograms of ephedrines and pseudoephedrine after derivatization by L-TPC using an achiral HP-5 mc column: (a) (1S, 2R)-ephedrine; (b) (1R, 2S)-ephedrine; (c) a mixture of (1R, 2S)ephedrine, (1S, 2R)-ephedrine, and (1S, 2S)-pseudoephedrine. Chiral analysis using CE is also deployed in the teaching lab, adapting a procedure developed by scientists at the DEA and published in the DEA’s Microgram Journal (3). Dynamically coated capillaries give faster and more robust electroosmotic flow compared to uncoated ones, especially at lower pH, which is required for the chiral CE analysis of basic drugs (4). In the teaching lab, CE analysis using an improved dynamic coating procedure is employed to achieve the chiral differentiation and determination of d- and l-ephedrines, pseudoephedrines, amphetamines, and methamphetamines (Figure 2). A fused-silica capillary kept at 15 °C is treated with rapid sequential flushes of different compositions and functions. It is explained to the students that the capillary is flushed with sodium hydroxide first to remove impurities and coating materials from the previous run. Then water is used to flush out sodium hydroxide. Later, CElixir Reagent A (a buffer containing a polycationic coating agent) is used to coat the originally negatively charged bare fused silica capillary inner wall with a layer of cationic material. Finally, CElixir Reagent B (a buffer containing a polyanionic coating agent, pH = 2.5) is used to flush the capillary to coat an anionic layer containing the chiral selector hydroxypropyl-β-cyclodextrin on top of the initial cationic coating before the sample is injected and analyzed. This process is repeated before the next run, which explains why it is called dynamic coating. Students learn that with a chiral additive serving as a pseudo-stationary phase, enantiomers can be separated by CE. Free zone capillary electrophoresis cannot do that without a chiral selector.

208 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Figure 2. CE electropherograms of ephedrines and pseudoephedrine using a dynamically coated fused silica capillary eventually coated with a modified β-cyclodextrin: (a) (1R, 2S)-ephedrine; (b) (1S, 2R)ephedrine; (c) (1S, 2S)-pseudoephedrine; (d) a mixture of (1S, 2S)-pseudoephedrine, and (1R, 2S)- and (1S, 2R)-ephedrines, (1S, 2S)-pseudoephedrine, d-amphetamine, and d- and l-methamphetamines.

Chiral Analysis Using Spectroscopy Our students also analyze controlled substances using analytical spectroscopy. To better manage such teaching activities, we use only listed compounds (not scheduled ones), such as ephedrine and pseudoephedrine in IR or NMR studies with or without using a chiral shifting reagent (R-mandelic acid) and chiral spectrometry such as polarimetry since larger amounts of substances, sometimes in the form of powder, are generally used in such lab exercises. In polarimetry, samples with known concentration containing different ratios of d-ephedrine and l-ephedrine are used and students are required to calculate the enantiomeric excess from their optical rotation values. After such experiments, students understand why d- or l-ephedrine is so-named. They also now understand that enantiomers produce exactly the same signal when they interact with nonpolarized electromagnetic radiation (EMR), while producing mirror image responses when probed by polarized EMR. When enantiomers interact with a chiral shifting reagent, they form a pair of diastereomeric salts, producing different signals when probed by nonpolarized EMR. An optical rotation experiment is also included in our analytical spectroscopy lab, which is directly related to its corresponding lecture. After the lab, students understand why (1S, 2S)-pseudoephedrine is called (1S, 2S)-(+)-pseudoephedrine, (+)-pseudoephedrine, or d-pseudoephedrine. Since UV-visible spectroscopy can be used to determine the total amount of ephedrine and optical rotation can be used to determine the enantiomeric excess, students are asked to figure out the concentrations of the individual components in a mixture of d-ephedrine and l-ephedrine using both a UV-visible spectroscopy spectrometer and a polarimeter (5), which was acquired using a Verizon Thinkfinity grant. These concepts are also included in teaching organic chemistry, especially in the stereochemistry chapter.

209 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

GC BTEX, Benzene/Cyclohexane and Arson Accelerant Analysis Benzene, toluene, ethyl benzene, and xylenes (m‑, o‑, and p-xylenes), which are collectively known as BTEX compounds, and cyclohexane are important components in many liquid fuels, especially gasoline. Some of these compounds can potentially cause a lot of health problems and some of them are even carcinogenic. Because of incomplete combustion, leaking, vaporization, and illegal dumping, they constitute major pollutants in air, water, and soil. Analysis of these compounds in various samples can be very important to environmental forensics and arson investigation. As a matter of fact, we include such analysis in our undergraduate instrumental analysis for chemistry, biochemistry, and forensic science majors, and in Separations and Analytical Spectroscopy for forensic science and environmental science graduate students. In our lab exercises, BTEX, cyclohexane/benzene mixtures, and arson accelerants are analyzed by GC or GC–MS, while xylenes are also analyzed by Fourier-transform infrared spectroscopy. According to the so-called "like-attract-like" or "like-dissolve-like" principles, polar GC stationary phases will retain polar species more strongly and nonpolar species less strongly, and vice versa, if other factors such as boiling point differences between the analytes are negligible. Figure 3 presents an example from our teaching lab. Benzene (C6H6) and cyclohexane (C6H12) are both cyclic nonpolar hydrocarbons with very similar boiling points (80.1 °C and 81.4 °C, respectively). However, benzene is more polarizable and it has slightly higher polarity. If the GC stationary phase is polar, such as Carbowax, it has a higher affinity to benzene than to cyclohexane. Therefore, benzene is eluted later than cyclohexane as shown on the left in Figure 3. When the stationary phase is very nonpolar, such as HP-1 (polydimethylsiloxane), the elution order is reversed as shown on the right in Figure 3.

Figure 3. Cyclohexane and benzene’s elution orders reversed on polar (Carbowax, polyethylene glycol, left) and nonpolar (HP-1, polydimethylsiloxane, right) GC stationary phases, respectively. The identities of each component can be established directly when a mass spectrometer is used as the detector of GC, while retention times of the components have to be compared with those of the standards when a flame ionization detector is used.

210 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

However, the elution order of the same homologous series of compounds, whose immediately neighboring members differ by a methylene group (CH2) only and have the same functional group and similar chemical properties, generally does not change with the change in GC stationary phases. Parameters that determine the order or time of retention are boiling points (vapor pressure) of the analytes, and the intermolecular forces between the analytes and the stationary phase. Polarity and boiling points of the analytes are considered in concert with the polarity of the stationary phase(s) to determine the elution order. If other molecular characteristics, such as polarity, are similar, more volatile analytes (low boiling points, high vapor pressure), which generally correspond to smaller molecular weights, get into the gas mobile phase more easily and elute earlier. The elution order of homologs generally follows such a boiling point simplification regardless of whether the stationary phase is polar or nonpolar. This is illustrated in our GC BTEX analysis lab since BTEX compounds are homologs. In spectroscopy, xylenes are analyzed by IR to demonstrate the power of IR to “fingerprint” compounds even if they have very similar structures (in this case, isomers). Students find out through their own measurements that the three isomeric xylenes show characteristically strong, yet very different, peaks in the 600–900 cm–1 region: 691 and 769 cm–1 for the meta isomer, 742 cm–1 for the ortho isomer, and 795 cm–1 for the para isomer. For p-xylene, there is another strong peak at 484 cm–1. MS is used as the detector in the GC analysis of BTEX and benzene/cyclohexane. Because of the specificity (identification power) of MS, the identity of each component in the chromatograms can be established without running any standards. When a flame ionization detector is used in the GC analysis of cyclohexane/benzene, pure cyclohexane and benzene have to run through the GC using the same method as standards so that their retention times can be used to identify the cyclohexane and benzene peaks in the mixture’s chromatogram. That the properties of the compounds are determined by their structures reinforces the concepts covered in organic chemistry. The concept and properties of homologs are also covered in organic chemistry. Extraction and Chromatographic Analysis of Illicit Drugs The use of commercially available molecularly imprinted polymers (MIPs) as solid phase extraction materials of illicit drugs such as amphetamine and methamphetamine was performed by a group of students who worked successively. MS and HPLC were utilized to analyze the performance of two commercially available MIPs on water and urine samples that were both spiked with amphetamine. Samples (spiked samples and those that used MIPs) were analyzed using an Applied Biosystems MDS SCIEX API 2000 instrument (AB SCIEX, Framingham, MA). Results show that the MIP increases the percent recovery of the spiked amphetamine, and the effect is observed obviously in synthetic urine where the signal of m/z = 91 (that pertains to amphetamine) is minimal in samples not passed through MIPs. This low signal observed in samples without pretreatment is due to a matrix effect that, in the presence of other chemicals, suppresses the signal of the target analyte. The percent recovery for both commercially available MIPs (SupelMIP from Supelco, Bellefonte, PA, USA and AFFINIMIP from Polyintell, Paris, France) are close to one another. An HPLC (Agilent 1100) isocratic system of solvent (90:10 aqueous phosphoric acid: acetonitrile, pH 2) was also used to separate amphetamine and methamphetamine using a BetaBasic C18 column (2.1 100mm, 3mm particles; ThermoFisher Scientific). Results show the separation of both illicit drugs, with amphetamine eluting before methamphetamine (Figure 4). However, the binding capacity of the MIP posed a problem on its feasibility as an extracting material. Based on

211 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

these results MIPs can be applicable only on instruments with lower detection signal (less than ppm level).

Figure 4. Chromatographic analysis of illicit drugs (amphetamine and methamphetamine). Mass spectra of synthetic urine spiked with amphetamine (first used just spiked sample and the last two are samples that used MIP). HPLC chromatogram of samples containing amphetamine and methamphetamine. Raman Spectroscopy The purchase of a Raman spectrometer in 2012 allows the department to broaden their research field and integrate Raman spectroscopy in teaching (6). This technique together with IR spectroscopy comprises vibrational spectroscopy, used to assess the molecular vibrational motion and to provide a “fingerprint” of a molecule being observed (7). The two techniques complement each another as IR depends on change of the dipole moment while Raman relies on change of polarizability. Being a nondestructive method, Raman has a wide range of applications in the pharmaceutical industry, geology, and life sciences to name a few since it can analyze all types of samples (solid, liquid, or gas). The main application of Raman in forensic science is in explosives and illicit drugs. The availability of a portable instrument make it useful to analyze everyday samples, and for forensics it was used to analyze mock explosives and painkillers. When mentioned during class that this is the same instrument that is being used for on-site analysis, the excitement level of the students increases, knowing that they have the chance to gain firsthand knowledge in using it. During laboratory class, students are given the opportunity to use the instrument so they become familiar with its operation. Explosives usually contain nitro compounds and for the mock explosive experiments, students are given samples containing nitro to be able to observe the peak that pertains to the nitro symmetrical stretching vibration mode (~1310–1397 cm–1). Four nitro samples in liquid and solid forms are used in class: nitrobenzene, nitrotoluene, nitrobenzaldehyde, and 4-nitrobenzyl alcohol. Typical Raman spectra for these samples are seen in Figure 5.

212 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Figure 5. Raman spectra of nitro-containing compounds used as mock explosives. At present, one of society's forensic problems is the proliferation of painkillers, particularly opioids. However, since these materials can be purchased only with prescriptions, students were given over-the-counter drugs instead to simulate the analysis of pain killers. In addition to analyzing the tablets, students are also performing theoretical calculations using Gaussian to obtain the calculated Raman spectra of the active ingredients; for instance Tylenol has acetaminophen and Advil has ibuprofen as active ingredients. This will be then used to compare with experimental results. A book chapter about this activity was published last year (8). The Raman spectrometer was also used to analyze xylene isomers (m-xylene, p-xylene, and oxylene). Different mixtures of xylene isomers were prepared and a method using the portable Raman spectrometer was developed side by side with a confocal Raman spectrometer. Although the method is limited to xylene mixtures, the intention in the near future is to use it to determine not only xylenes but the BTEX compounds in real samples (9). Computational Chemistry in Our Curriculum The experimental IR and Raman spectra often are compared with spectra computationally derived. This approach has been proven extremely helpful for our students in associating the observed frequencies with their corresponding vibrational modes. In addition, they provide a better understanding at the molecular level of the spectroscopic results and occasionally the effect of solvents and concentrations. For the derivation of the computational vibrational spectra we use the Gaussian 09 package (10), which includes Hartree-Fock and discrete Fourier transform methods and a variety of basis-functions.

213 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Interactive Online eTextbook with Forensic Themes and Embedded Questions One of the major challenges in teaching instrumental analysis with a forensic theme emphasis was the lack of excellent textbooks. Many widely used instrumental analysis books mainly use chemistry and biochemistry examples, while many forensic chemistry books do not look at the indepth mechanisms or structure–activity relationships behind forensic analysis examples. Starting in 2017, interactive online eTexbooks have been written and employed in teaching Instrumental Analysis (11), Forensic Separations Chemistry (12), and Analytical Spectroscopy (13) through Top Hat. In these textbooks, real forensic examples taken or adapted with permission from a variety of sources including the DEA’s Microgram Journal are integrated in the main text, interactive questions, or assignments. At the same time, relevant chemistry concepts, reasoning, and structure–property (structure–activity) relationships are elaborated in detail to help students really learn and use what they learn to solve problems not covered in the textbooks. Clicker-like questions are deployed to gauge students’ learning in class, in real time. Embedded questions and assignments help students access their own learning while they read the text. Quizzes and parts of the middle and final exams are administered through the Respondus LockDown Browser to make flexible but rigorous assessments multiple times during each semester so that teaching and learning strategies can be adjusted accordingly. One real-world example, adapted from an article in the Microgram Journal (14), is described in the following pages. An exhibit in the form of a white powder was found by a DEA laboratory to contain cocaine hydrochloride, salicylic acid, and traces of acetylsalicylic acid (aspirin) and benzocaine (Figure 6).

Figure 6. Structures of some components in a seized cocaine exhibit. It was also found to contain a significant amount of another compound, which was at the time not commercially available. However, it has been known that acetylsalicylic acid (aspirin) can participate in some transacylation reactions as shown in many earlier studies. This fact and the general scheme of such a reaction, as shown in the top part of Scheme 1, is given to students. All the components in the exhibit were isolated, purified, and fully characterized. Relevant spectra (mass spectrum, IR, and 1H NMR) of one of the purified components in this sample are shown in the Analytical Spectroscopy eTextbook. Students are asked to figure out the identity (i.e., structure) of this component. Then students are guided to arrive at their solution to the question. The given electron impact ionization (EI) mass spectrum (shown in the eTextbook, but not here) shows a labeled molecular ion peak (m/z) at 207 amu, which is an odd number. Therefore, the component should contain an odd number of nitrogen atoms in its structure according to the socalled nitrogen rule, which is explained earlier in the eTextbook. As a result, it cannot be salicylic acid or aspirin since they do not contain nitrogen atoms. Although benzocaine and cocaine do contain 1

214 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

(an odd number) nitrogen atom each, the exact masses of their molecular ions (M+) should be 165 and 303, respectively. Therefore, this component must be something else. The IR spectrum is shown in the eTextbook, although not shown here. It shows a peak at 1679.77 cm−1, which can be assigned to an amide carbonyl group (–NH–C=O). From the IR spectrum given in the eTextbook we can tell that there is another carbonyl group at ~1700 cm−1 (wavenumber value not printed in the spectrum). Its 1H NMR spectrum, shown in Figure 7, further helps elucidate its structure. Each of the two doublets in the aromatic region (at 8.01 ppm and 7.60 ppm, respectively) represents two H's, which can be obtained by the ratio of peak areas, pointing to a para-substitution pattern on a phenyl/ benzene ring. The relatively "flat" broad hump at 7.53 ppm represents an amide NH that is connected to both a carbonyl and a phenyl group since its chemical shift is quite large. The quartet at 4.35 ppm is CH2 next to an oxygen atom and coupled with CH3, which itself is represented by a triplet at 1.38 ppm. The singlet at 2.20 ppm is CH3 next to a carbonyl. Of course, in the eTextbook, the NMR spectrum is shown initially without any of the annotations shown in Figure 7 when it is presented as part of the question. The annotated version, similar to Figure 7, is shown later in the explanation of the correct and incorrect answers. Therefore, the structure of this component is N-acetylbenzocaine, whose structure is shown in the preceding annotated NMR spectrum. Its formula is C11H13NO3. It has an exact mass of 207, which matches that of the molecular ions in the EI mass spectrum. The transacylation reaction between benzocaine and aspirin indeed produces N-acetylbenzocaine as shown in Scheme 1 (top: the general transacylation reaction; bottom: the specific transacylation that produced the said component).

Figure 7. Analysis of a given 1H NMR spectrum to figure out the structure of a component in a seized cocaine exhibit. The component is found to be N-acetylbenzocaine. Adapted with permission from reference (14). Copyright 2010 United States Drug Enforcement Administration.

215 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Scheme 1. Transacylation reaction. Top: the general transacylation reaction between an amine and an ester; bottom: the specific transacylation between benzocaine and salicylic acid that produced N-acetylbenzocaine. In GC chromatography analysis using a nonpolar stationary phase, this component eluted later than both benzocaine and salicylic acid, which is consistent with the fact that as an amide, it is more nonpolar than both benzocaine, which bears a polar primary amino group (–NH2), and salicylic acid, which bears a polar carboxylic acid group (–COOH). This part, together with the real chromatogram, is integrated into one of the embedded questions in the eTextbook, Forensic Separations Chemistry. In such interactive questions, especially multiple choice questions, even if a student gives the correct answer, an explanation of why it is correct is given. If a student chooses an incorrect answer, hints or an explanation why it is incorrect is given so that the student is directed to the correct answer. Potential pitfalls are pointed out. For example, although cocaine might be able to transacylate benzocaine to give the product shown in Figure 8, whose structure contains two nitrogen atoms, its molecular ion peak should give an even number exact mass. The given mass spectrum and NMR spectrum show that this possible product is not the isolated and purified component.

Figure 8. A potential product of the transacylation reaction between cocaine and benzocaine. This example shows that to be competent forensic scientists, students need to master organic chemistry (the transacylation reaction, the organic structures, structure–polarity relationship) and instrumental analysis, together with physical chemistry (IR wavenumbers, electronegativity, and chemical shift and the splitting pattern in NMR). Although we have been using the online interactive eTextbooks for only one year, limited data shows that this approach produces much better student-learning results. Students who flunk such classes previously got A-range grades after they used the online interactive eTextbooks and completed all the interactive assignments when they retook such classes.

216 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Integration of Forensic Chemistry Research and Teaching To train high-caliber forensic chemists following established scientific criteria, we involve students in our research in a predominantly undergraduate institution, getting them familiarized with state-of-the-art instrumentation and scientific methodologies. Our research has four major educational and workforce development goals: (1) stimulating students’ interest in chemistry through research projects in contemporary issues; (2) providing students with the opportunity to apply modern instrumentation in experimental investigations; (3) fostering collaborative interaction among students; and (4) encouraging students to pursue careers in the discipline of chemistry, especially forensic chemistry. Faculty researchers have a unique opportunity to teach undergraduates and master’s level graduate students about the discipline of chemistry’s rich potential as both an intellectual endeavor and a tool for modern technological breakthrough. Our research covers synthetic chemistry, analytical chemistry, and computational chemistry. Research topics include the search for cures for some neglected diseases, the development of sensors for biologically and environmentally important metal ions (15, 16), the advancement of asymmetric catalysis (15), the search for beneficial compounds from natural products, and the development of better analytical tools for controlled substances and pollutants. Such projects are well suited for students pursuing a liberal arts education because they employ science in the service of criminal justice, environmental justice, and equality in medical care. Our students get exciting results and are coauthors of papers. They also help us modify our research projects and turn them into inquiry-based lab exercises. We use our own research results in lecture so that the most recent examples of some principles and theories can be brought into the classroom. In this way, the teaching of these theories and principles can be more relevant and convincing. For example, we use new nitrogen-containing ligands and the bromine-containing intermediate products we made and their actual EI, electrospray ionization, and matrix-assisted laser desorption ionization mass spectra, to illustrate the nitrogen rule and halogen isotope pattern in teaching MS. Then we apply the nitrogen rule to the mass spectroscopic analysis of many nitrogen-containing controlled substances. We use the fluorescent sensors for the Zn and Hg ions we made to explain chelation-enhanced fluorescence and the theory of photo-induced electron transfer. This is the same mechanism behind sensitive latent fingerprint detection using fluorescence by Zn ions or other metal ions to ninhydrin-based reagents or some other nonfluorescent reagents. The relationship between the structure and properties of fluorescent dyes are explained and their application in DNA typing are introduced.

Conclusion Pace University’s renowned forensic science program is hosted by the Department of Chemistry and Physical Sciences. Our undergraduate and graduate curricula are designed to give our students a sound education and to prepare future scientists who promote the betterment of the administration of justice. In the department, we integrate forensic themes and examples adapted from leading forensic science journals in teaching organic, analytical, physical and biological chemistry, especially instrumental analysis, in both lecture and lab. We have also developed interactive eTextbooks that incorporate forensic-themed examples and assignments for both chemistry and forenisc science students. Our multidisciplinary research, which is integrated with teaching, fuses chemistry and forensic science well. Our effort is well in line with the findings and recommendations of the National Research Council on forensic science. As a result, our students really learn and achieve a lot during their study at Pace and beyond. Many of our alumni and some current students are employed in 217 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

leading forensic laboratories, making great scientific contributions to the sound administration of justice.

Acknowledgments We would like to thank the Verizon Foundation for the purchase of a portable Raman spectrometer and a polarimeter (Thinkfinity grants through Pace University) and the National Science Foundation for the acquisition of an accelerated solvent extraction system. We thank the Northeast Laboratory of the DEA, the NYC Office of the Chief Medical Examiner, the NYC Police Laboratory, and other forensic laboratories and institutions for their support to our forensic science program. We are grateful that Pace University Scholarly Research Fund continuously supports our research and our students, especially Jonathan Oswald, who helped us turn some of the research projects into inquiry-based lab exercises. Z. D. would like to thank the Research Corporation for Science Advancement and the Petroleum Research Foundation for supporting his research involving undergraduate students. D. A. would like thank the National Institute of Justice for supporting his research.

References 1. 2.

3.

4.

5.

6.

7. 8.

9.

National Research Council. Strengthening Forensic Science in the United States: A Path Forward; The National Academies Press: Washington, DC, 2009; p 348. Sievert, H.-J. P. Determination of Amphetamine and Methamphetamine Enantiomers by Chiral Derivatization and Gas Chromatography–Mass Spectrometry as a Test Case for an Automated Sample Preparation System. Chirality 1994, 6, 295–301. Lurie, I. S.; Bozenko, J. S., Jr.; Li, L.; Miller, E. E.; Greenfield, S. J. Chiral Separation of Methamphetamine and Related Compounds Using Capillary Electrophoresis with Dynamically Coated Capillaries. Microgram J. 2011, 8, 24–28. Lurie, I. S.; Hays, P. A.; Parker, K. Capillary Electrophoresis Analysis of a Wide Variety of Seized Drugs Using the Same Capillary with Dynamic Coatings. Electrophoresis 2004, 25, 1580–1591. Dai, Z.; Valdez, L.; Dumit, M. Chiral Analysis of Methamphetamine and Related Controlled Substances in Forensic Science. In Recent Advances in Analytical Techniques; Atta-ur-Rahman, Ed.; Bentham Science Publishers: Sharjah, United Arab Emirates, 2017; Vol. 3, pp 1–39. Dai, Z.; Javornik, A.; Sobolewski, C.; Batte, T.; Viola, J.; Rizzo, J.; Athanasopolous, D.; Mojica, E.-R. E. Integration of Raman Spectroscopy in Undergraduate Instruction and Research at Pace University. In Raman Spectroscopy in the Undergraduate Curriculum; Sonntag, M., Ed.; American Chemical Society: Washingon, DC, 2018; Vol. 1305, pp 199–219. Schrader, B. Infrared and Raman Spectroscopy; VCH Publishers: New York, NY, 1995. Mojica, E. R. E.; Zapata, J.; Vedad, J.; Desamero, R. Z. B.; Dai, Z. Analysis of Over-theCounter Drugs Using Raman Spectroscopy. In Raman Spectroscopy in the Undergraduate Curriculum; Sonntag, M. , Ed.; American Chemical Society: Washington, DC, 2018; Vol. 1305, pp 69–91. Vedad, J.; Reilly, L.; Desamero, R. Z. B.; Mojica, E. R. E. Quantitative Analysis of Xylene Mixtures Using Handheld Raman Spectrometer. In Raman Spectroscopy in the Undergraduate

218 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

10.

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

Curriculum; Sonntag, M. , Ed.; American Chemical Society: Washington, DC, 2018; Vol. 1305, pp 129–151. Frisch, M. G.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Petersson, G. A.; Nakatsuji, H.; Li, X.; Caricato, M.; Marenich, A.; Bloino, J.; Janesko, B. G.; Gomperts, R.; Mennucci, B.; Hratchian, H. P.; Ortiz, J. V.; Izmaylov, A. F.; Sonnenberg, J. L.; Williams-Young, D.; Ding, F.; Lipparini, F.; Egidi, F.; Goings, J.; Peng, B.; Petrone, A.; Henderson, T.; Ranasinghe, D.; Zakrzewski, V. G.; Gao, J.; Rega, N.; Zheng, G.; Liang, W.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Throssell, K.; Montgomery Jr., J. A.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Keith, T.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Millam, J. M.; Klene, M.; Adamo, C.; Cammi, R.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Farkas, O.; Foresman, J. B.; Fox, D. J. Gaussian 09, Revision A.02; Gaussian, Inc.: Wallingford, CT, 2016. Dai, Z. Instrumental Analysis; Top Hat: Toronto, ON, 2018. Dai, Z. Forensic Separations Chemistry; Top Hat: Toronto, ON, 2017. Dai, Z. Analytical Spectroscopy; Top Hat: Toronto, ON, 2018. Casale, J. F.; Nguyen, M. C. N-Acetylbenzocaine: Formation via Transacetylation of Benzocaine and Acetylsalicylic Acid in a Cocaine Exhibit. Microgram J. 2010, 7, 7–11. Dai, Z.; Lee, J.; Zhang, W. Chiroptical Switches: Applications in Sensing and Catalysis. Molecules 2012, 17, 1247–1277. Carney, P.; Lopez, S.; Mickley, A.; Grinberg, K.; Zhang, W.; Dai, Z. Multi-mode Selective Detection of Mercury by Chiroptical Fluorescent Sensors Based on Methionine/Cysteine. Chirality 2011, 23, 916–920.

219 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Chapter 12

Downloaded via MIAMI UNIV on November 5, 2019 at 01:55:17 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Using Expert Witness Testimony with an Illicit Substance Analysis To Increase Student Engagement in Learning the GC/MS Technique Albert D. Dukes III,* Jeffrey M. Hollifield, and David E. Gardner Department of Physical Sciences, Lander University, 320 Stanley Avenue, Greenwood, South Carolina 29649, United States *E-mail: [email protected].

Authenticity in a forensics lab activity is an important tool in combating the CSI effect that has created unrealistic expectations for forensic science. In this chapter, we provide a description of a forensics lab activity that analyzes a simulated illicit compound using gas chromatography-mass spectrometry (GC/MS). In addition to teaching students how to perform GC/MS analysis, we expand this activity by having students testify about the results of their analysis in a simulated courtroom environment. By combining the teaching of standard analytical technique with the experience of testifying about the results, students are able to have a more authentic experience of being a forensic analyst, both inside and outside the laboratory setting.

Forensics as a Medium To Teach Chemistry While television shows focusing on the criminal justice system have been popular for many years, CBS launched a new kind of crime program, CSI: Crime Scene Investigation, in 2000. What set this show apart from many of its predecessors was its focus on gathering evidence from a crime scene and scientifically testing that evidence, rather than the more traditional detective work of interviewing witnesses. The success of the show, and its subsequent spinoffs, have helped bring forensics to the forefront of the criminal justice system in the United States (1). Due to the popularity of CSI and other similar television shows, forensics has become a popular tool for engaging students in chemistry (2–11). Forensics is useful for teaching chemistry because of the chemical principles utilized in performing many forensic tests (12). A well-designed forensics-based activity can be very engaging and effective in teaching students fundamental chemical concepts (3, 10–14). An unfortunate side effect of using forensics to teach chemistry is that the activity can become too focused on having the student figure out who committed the crime; thus, the intended chemistry lesson becomes lost in the background of solving the case (5). © 2019 American Chemical Society Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Our students’ expectations of forensic analysis are often distorted by the CSI effect. The CSI effect is the term given to the perception among law enforcement, court officials, and the general public that forensic evidence is necessary in order to obtain a conviction at trial (1, 15, 16). An additional aspect of the CSI effect is the belief that forensic evidence, when presented at trial, is infallible (15, 16). One method for alleviating this problem is to design forensics laboratory activities for students that are as authentic as possible. Newmann and Wehlage define authentic educational experiences as those that meet the following three criteria: (1) Students construct meaning and produce knowledge; (2) The knowledge they construct is the result of disciplined inquiry; and (3) The work that they produce has value and meaning beyond their success in an individual course or program of study (17). By limiting the scope of the laboratory activities to teaching a single technique, and allowing for analysis and interpretation of the results of measurements, students will be able to better understand how and why the chosen technique is used to analyze the given evidence. Providing students with an authentic laboratory experience allows them to develop their professional identity and makes them less apprehensive about the tasks they are asked to accomplish (18, 19). In order to provide our students with an authentic forensics experience, and to teach them how to properly use gas chromatography-mass spectrometry (GC/MS), we have developed a laboratory activity that simulates the analysis of a suspected illicit drug. This laboratory activity is intended for upper-level students in an Instrumental Analysis course, a course typically required of forensic scientists who work in drug analysis (20). Drug analysis is a routine forensics test with a solid chemical foundation. In South Carolina, where our institution is located, the number of drug arrests has remained relatively unchanged from 2008–2012, the most recent five-year period for which data are available. There were approximately 35,000 drug arrests in 2012 (21). This high number of arrests highlights the need to correctly perform a large number of forensic drug analyses in a given year, which makes a laboratory activity based on forensic drug analysis an ideal candidate for using forensics to teach chemistry. Analysis of a suspected illicit compound provides our students with an authentic experience in learning how to perform GC/MS analysis to identify an unknown compound. A classic laboratory activity for teaching this type of instrumental analysis is to analyze for the aromatic compounds in gasoline (22). While this laboratory activity provides instructors with an easily reproducible activity for teaching GC/MS analysis, its ability to engage students is limited. Students do not develop the separation method that is utilized in separating the aromatic compounds, nor are they asked to speculate on the purpose of including these various aromatic compounds identified in the gasoline. While a traditional GC/MS laboratory activity has its limitations, many instructors have used a variation of the gasoline analysis laboratory activity to teach students how to use GC/MS. It has also been our observation that students who have completed the Instrumental Analysis course where they were taught GC/MS analysis and who subsequently conduct undergraduate research projects involving GC/MS analysis often retain very little of what we have taught them. This should not come as a surprise since the gasoline analysis laboratory activity does not fit the criteria for authentic experiences as defined by Newmann and Wehlage. In the traditional gasoline analysis laboratory activity, students do not construct meaning nor does the work they produce have any meaning beyond completing the assigned activity. This lack of authenticity likely hinders student understanding of the principles that the gasoline analysis laboratory activity is intended to teach. While there are certainly times where traditional instructional methods are preferred by instructors, we contend that when students are in a laboratory setting, the activities that we ask them to complete should be as authentic as possible so that students understand the value of what we are teaching them. 222 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

To that end, we have developed a laboratory activity in which students utilize GC/MS in order to identify a simulated illicit compound. The students were presented with an evidence bag containing a suspected illicit compound. While it is possible to apply for a permit from the Drug Enforcement Administration to use actual illicit compounds for academic purposes (2), the burden of obtaining such a permit is sufficiently large that for many smaller institutions, it is not worth the time and expense. As such, we have utilized common analgesic compounds to simulate illicit compounds. Students are provided the results of an initial colorimetric screening test that was performed in the field when the substance was confiscated. They must use the GC/MS to analyze the evidence and determine if the substance in question is one of the simulated illicit compounds. Upon completing their analysis, the students must complete a drug analysis worksheet similar to the ones employed by forensics labs to document their results. Additionally, the students will take the witness stand in a mock testimony activity where they will be examined and cross-examined by practicing attorneys. By adding the mock testimony to the laboratory activity, students have the added realism of defending their conclusions in a trial setting, just as practicing forensic scientists would.

Experimental Procedure for Simulated Drug Analysis Materials Required Tablets of an analgesic compound (ibuprofen) serving as the simulated illicit compounds were purchased from a local pharmacy and ground into a powder. The powder was wrapped in plastic and then placed in a sealed evidence bag before being provided to the students. The standard for the simulated illicit compound was ibuprofen (Acros Organics, 99%). Ammonium hydroxide (Fisher Scientific, 28–30%), and chloroform (Acros Organics, >99.8%) were used as the solvents for the extraction. Processing of Sample The simulated illicit compound was extracted and separated with a mixture of 25.0 mL of chloroform and 25.0 mL of 1M aqueous ammonium hydroxide solution. Students added a measured mass (~0.1g) of the suspected illicit compound to a beaker containing the chloroform and ammonium hydroxide mixture stirring the solution well. The mixture was allowed to stand for five minutes and then separated into two layers. The students then removed an aliquot from the chloroform layer for GC/MS analysis. The sample in the chloroform layer should be the free base form of the analyte; any unreacted acid-salt or other impurities should remain in the aqueous layer. Both the simulated illicit compound and the standards were prepared in the same manner. Parameters for GC/MS Analysis • Injection port temperature: 250 °C • Detector temperature: 280 °C • Column length: 30.0 m; HP-5MS: 5% phenyl methyl siloxane; inner diameter: 250 μm; film thickness: 0.25μm (Agilent model no. 19091S-433E) • Hold at 60 °C for 2 min, ramp at 10 °C/min until 200 °C, and hold for 4 min; scan mode and solvent delay for 2 min • Mobile phase: He (ultra-high purity, Praxair) with a flow rate of 1.0 μL/min • Manual injection (sample volume: 5.0 mL) • Electron impact ionization 223 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Procedure for Mock Trial As with all other laboratory activities in the course, students submitted a formal lab report so that the instructor can assess how well the students had learned the material presented in the laboratory experiment. However, for this activity students also submitted a forensic worksheet similar to the one used by the State Law Enforcement Division crime lab in South Carolina (Figure 1). The drug analysis worksheet requires students to record routine information such as the mass of the sample, the balance used to determine the mass, and the type of instrument used for the analysis; additionally, there are options for the results of various color tests. These color tests provide the instructor with opportunities to assign identities to the simulated illicit compound. For example, if the instructor wished for the simulated illicit compound to be a substitute for cocaine, then the students would be provided with the information that the COBTHIO SnCl2 test (Figure 1) gave a blue result. The results of several common color tests are provided in Table 1 (23). However, if the instructors want the students perform the colorimetric tests as part of the laboratory activity, then it would be possible to substitute lidocaine as the simulated illicit compound instead of ibuprofen. By using lidocaine, the students would observe the blue color in the COBTHIO SnCl2 test that is characteristic of cocaine. During the pre-lab briefing, the students were reminded of the importance of recording the mass of the sample on the worksheet prior to performing the analysis as this mass can be crucial for determining what charges a defendant would face (i.e. simple possession, intent to distribute, or trafficking). In this lab, the instrument used for the analysis was GC/MS; if a GC/MS is not available, FTIR analysis could also be used to identify the simulated illicit compound (13, 23).

Figure 1. This is an example of the blank drug analysis worksheet that students completed as part of the laboratory activity. The worksheet has blanks for basic information about the sample, the results of screening color tests, and the type of instrument used for analysis. For the mock testimony portion of the lab activity, we secured the services of two local criminal defense attorneys. Our attorneys alternated in assuming the roles of the prosecution and defense. Prior to examining each student, both attorneys were provided with the drug analysis worksheet (Figure 1) and the plots of the GC/MS data that were generated by the student. Students who had yet to give their mock testimony remained outside of the mock courtroom to ensure that their 224 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

answers were not influenced by the answers given by their peers. When students had completed their testimony, they were allowed to remain in the courtroom and observe the testimony of their classmates. Table 1. Results of Common Colorimetric Tests for Illicit Compounds Colorimetric Test

Illicit Compound

Color Result

Cobalt Thiocyanate/Stannous Chloride

Cocaine and Cocaine base

Blue

Dille-Koppanyi

Barbiturates

Blue/Purple

Erlich

LSD and Psilocin

Purple

Froehde

Heroin

Purple to Green

Froehde

Codeine

Green to Red/Brown

Froehde

Morphine

Purple to Grey

Froehde

Hydrocodone

Yellow

Froehde

Oxycodone

Yellow to Blue

Marquis

Opiates

Purple

Marquis

Amphetamine/Methamphetamine

Orange

Marquis

MDMA/MDA

Black

Marquis

Fentanyl

Deep Orange

Results and Discussion of Student Analysis of Simulated Illicit Compound

Figure 2. The gas chromatogram generated by the analysis is shown above. The peak in the chromatogram that corresponds to a suspected illicit compound is identified with a star. Prior to beginning the analysis, students were provided with the results of a colorimetric screening test that would normally be performed in the field by the arresting officer. While the colorimetric tests in Table 1 serve as useful screening methods, the results of these tests alone are not sufficient for positively identifying an illicit compound (23). Because of this, students must perform 225 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

a more sophisticated analysis (in our case, GC/MS) in order to definitively determine if the evidence contains an illicit compound. A representative chromatogram generated by the students that performed the analysis of the simulated illicit compound is shown in Figure 2. As indicated by the chromatogram, several compounds other than the simulated illicit compound were observed in the analyzed sample. Students relied on the automated library search to serve as an initial screening of the compound’s identity. If the mass spectrum of any of the separated peaks matched that of the simulated illicit compound, the students would test a known standard of that compound to ensure that the library match had correctly identified the compound. In this example, the peak marked by a star in Figure 2 was identified as ibuprofen by the software’s library. The mass spectrum for the simulated illicit compound (the peak marked by a star in Figure 2) is shown in Figure 3. A computer library search suggested ibuprofen as the identity of the simulated illicit compound. There are two generally accepted methods for positively establishing the identity of a compound based on the mass spectrometry data. One way is to compare the experimentally determined mass spectrum to the mass spectrum published in the peer-reviewed literature. One of the options available for comparing the measured spectra to peer-reviewed data is to compare the measured mass spectrum of a suspected compound to the one published in the peer-reviewed literature (such as the NIST Chemistry WebBook) (24). Students can then make a visual comparison between the mass spectrum that they have measured and the reference spectrum available from NIST to determine if the computer has correctly matched the experimental spectrum with the spectrum in the computer’s library.

Figure 3. The mass spectrum (electron impact ionization) of the simulated illicit compound that was identified as ibuprofen. The other method available for positively establishing the identity of the compound in question is to compare the experimentally determined mass spectrum to the measured mass spectrum of a known reference standard. In order to confirm that the compound identified in Figure 3 is ibuprofen, the students performed an analysis of a known standard of ibuprofen. If the computer has correctly identified the simulated illicit compound, then both the retention times and measured mass spectra of both the standard and the simulated illicit compounds should match. A simple method for determining if the mass spectra of the reference and sample compounds match is to plot both spectra on the same graph (in different colors) and examine the location of the peaks. The absolute intensity of the peaks in the two samples may differ due to differences in purity between the simulated illicit sample and the reference standard or due to differences in the volumes of samples injected into the 226 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

instrument for analysis. However, the m/z value for the peaks are unaffected by the sample purity or injection volume. If there were missing or extra peaks in the mass spectrum, this would raise serious concerns as to whether the computer software had correctly identified the simulated illicit compound.

Discussion of Students’ Mock Testimony Prior to entering the lab to begin the activity, the students were briefed on the typical process for being declared an expert witness. In South Carolina, the standard for being declared an expert witness is established by Rule 702 of the South Carolina Rules of Evidence (25). To be classified as an expert witness in South Carolina, the witness must have the knowledge to assist the jury in understanding the evidence and determining the facts at issue. This can be established either through education or experience in the field, and the level of knowledge must go beyond what would be considered common knowledge of the jury. In South Carolina, a bachelor’s degree in chemistry or a related field would typically be sufficient for someone to be declared an expert witness on drug analysis. This standard may differ in other states. At the federal level, the standards for expert witnesses are described by the Daubert standard, which was established by a Supreme Court decision (26). In addition to explaining the certification process for becoming an expert witness, a faculty member who previously testified as an expert witness walked the students through some typical responses to common questions posed to witnesses. For the purposes of the mock testimony, we instructed the students and the attorneys to assume that the students would be qualified as expert witnesses. While the sample size of students is small (11 total students enrolled and completed the lab activity), our students experienced a wide range of reactions to being questioned by the attorneys. Many of our students were surprised by the limited questioning that they received from the prosecuting attorney. The questions asked during direct examination by the prosecution were typically limited only to establishing that the evidence had been correctly received (ensuring that chain of custody was established), determining that a chemical analysis had been performed, and confirming the identity of the compound in question. In general, students had the expectation that they would be asked to explain their work to the prosecution in much greater detail than was asked. In fact, when they were asked how the evidence was tested, several students attempted to recall the step-by-step process (with varying degrees of success) of preparing the sample, as well as the details of operating the software on the GC/MS instrument. Under cross-examination, the students realized the hazards of giving an overly detailed answer to the question of how the analysis was performed. The defense attorney took the opportunity to question the students as to why they had performed each step of the process in the way that they described. For example, one popular question was why was the sample extracted into the basic chloroform solution prior to the GC/MS analysis. Our students displayed varying degrees of success in attempting to explain that the extraction was performed in basic chloroform in order to convert any acid salts into neutral molecules so that they could be analyzed by GC/MS. The degree to which the students could correctly answer these questions varied from student to student. After the students had all completed their testimony, the attorneys discussed with the students that the manner with which they delivered their testimony would impact the jury as much as their answers being factually correct. Students who were hesitant in offering explanations as to why certain steps were performed were told that they would likely be perceived as not believable by a jury, even if they had ultimately provided correct answers. 227 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

In addition to providing overly detailed answers, several students were nervous when giving their answers both under direct examination and cross-examination. This manifested itself most often as speaking softly, mumbling, including the word “um” often in answers, and occasionally laughing nervously. These are not unexpected reactions to being placed in an unfamiliar and uncomfortable situation; however, not all students reacted in such a manner. Several students gave confident, correct, and appropriately detailed answers to the questions that they were asked during their mock testimony. Prior to the mock testimony, we would have assumed that the students most likely to perform well during their testimony were the students with higher grades and more outgoing personalities. However, this was not always the case. Students who had earned high grades on written work throughout the semester perhaps simply incorrectly assumed that their good performance on written assignments would translate to a good performance during oral testimony, or is it possible that the students in this particular year’s class had little experience with public speaking and were, as a result, nervous? The debriefing discussion did not focus on how the students prepared for their testimony; in the future, this should be included as part of the laboratory activity, as such a discussion may help students identify successful preparation techniques. Prior to beginning the laboratory activity, students were apprehensive about testifying as they were not certain of what to expect. Most of the expectations were based on what they had seen in crime dramas on television or on arbitration-based reality court shows. One of the biggest lessons that the students took away from this experience was what the process of testifying in court was actually like. Much of the students’ apprehension stemmed from thinking that the aggressive questioning seen in television shows was a reasonable approximation of what they would see from practicing attorneys. While the defense attorney asked questions that were intended to raise doubt about the conclusions of the student’s analysis, the manner in which they asked questions was substantially different from what our students had expected based on what they had seen on television. At the end of the experience, students felt that they had a better understanding of the process for testifying as an expert witness, as well as how forensic scientists present the results of their analysis as evidence at trial. For those interested in replicating this experience for their students, it was our experience that the legal community is willing to volunteer their time to assist in preparing students to be forensic scientists. We recommend reaching out to your state, county, or city bar association to find attorneys who specialize in criminal defense in your area. For the mock testimony, we asked the attorneys to refrain from any questioning that would lead to an objection by the opposing council as we were unable to secure a judge to preside over the mock trial (this was due to a scheduling conflict with a local judge who was initially willing to assist with the mock trial). If your university has a law school, it may also be possible to secure assistance from law school faculty in conducting the mock testimony. Law students may also benefit from assisting the volunteer attorneys in preparing for the questioning. Based on our conversations with the attorneys who volunteered to help with the mock testimony, it is clear that many attorneys do not understand how GC/MS analysis is performed or how to interpret the data that they receive when they subpoena these test results. By engaging with attorneys, and possibly law students, in analyzing the results presented in the mock testimony, we are engaging them in understanding chemistry in a context that is relevant to them. Engaging in discussions with nonscientists about scientific concepts that they find relevant has been identified as one of the most useful methods of increasing scientific literacy that scientists can utilize (27). While it is not reasonable to expect attorneys to be experts on conducting drug analysis, gaining a better understanding of how the analysis is conducted and what the analysis is capable and not capable of telling them will enable them to better represent their clients. 228 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Creating an Authentic Experience for Students The mock testimony that students must give as part of this laboratory activity helps add to the authenticity of the experience for the students. In preparing their mock testimony, students must take the information that is generated by the GC/MS analysis and determine what this data tells them about the sample. In doing this, students have produced knowledge and satisfied the first criteria of an authentic experience as defined by Newmann and Wehlage (17). The knowledge that the students generated was the result of the GC/MS analysis thus satisfying the criteria that the knowledge be the result of disciplined inquiry. It should be noted that simply performing the GC/MS analysis alone does not necessarily result in a detailed inquiry; the detailed inquiry arises from the need to pair the GC/MS experiment with the mock testimony that follows. The mock testimony portion of the activity satisfies the third criteria for an authentic experience (that the activity have value beyond the course) (17). In the traditional curriculum, communication in the form of written lab reports is most often emphasized. However, traditional lab reports are of little value in a forensic science laboratory as most of the data is reported on worksheets and in oral testimony. Traditional curriculum may emphasize oral communication in the form of a scientific talk, but this is still a very different form of oral communication than testimony in a courtroom. One of the most obvious differences between traditional scientific communication and testimony in a courtroom is the intended audience. Often the audience (i.e. the jury) of communication in a courtroom lacks an in-depth understanding of the subject. This type of communication is further complicated by the fact that the questions posed by the attorneys will limit what the forensic scientist can tell the jury. As a result, forensic scientists face the unique challenge of trying to communicate sometimes complex information in their testimony when they do not have control of how they share information. While forensics represents a relatively small fraction of the career opportunities available to chemists, it is not unreasonable to think that a chemist could be called upon to serve as an expert witness if they were to pursue a more traditional career in industry or academia. Thus, it is beneficial for students intending to pursue a career in forensics, as well as those intending on a more traditional chemistry career, to have the opportunity to learn how to communicate effectively in a courtroom environment. The laboratory activity that we have described here provides students with an authentic educational experience wherein they learn how to operate a GC/MS instrument and analyze the data generated. Additionally, students use their collected data to prepare their testimony in a mocktrial setting where they are questioned by practicing attorneys. After the mock testimony, students reported a better understanding of the forensic scientist’s role in the criminal justice system as well as a better understanding of what to expect should they be called to testify in court as an expert witness in their future career. By performing the analysis and preparing to testify, students gain a more realistic understanding of what forensic science is and, more importantly, what it is not capable of determining in a typical drug case.

Acknowledgments We would like to thank our local attorneys, C. Rauch Wise and Joshua S. Nasrollahi, for volunteering their time to serve as the attorneys for the mock testimony. We would also like to thank the 2018 CHEM 331 students at Lander University for their feedback in developing this laboratory activity.

229 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

References 1. 2. 3.

4.

5. 6.

7. 8. 9. 10. 11. 12.

13. 14. 15. 16. 17. 18.

McKay, J. CSI Effect. Gov. Technol. 2008, 21, 18–23. Heimbuck, C. A.; Bower, N. W. Teaching Experimental Design Using a GC-MS Analysis of Cocaine on Money: A Cross-Disciplinary Laboratory. J. Chem. Educ. 2002, 79, 1254. Hasan, S.; Bromfield-Lee, D.; Oliver-Hoyo, M. T.; Cintron-Maldonado, J. A. Using Laboratory Chemicals To Imitate Illicit Drugs in a Forensic Chemistry Activity. J. Chem. Educ. 2008, 85, 813. Ravgiala, R. R.; Weisburd, S.; Sleeper, R.; Martinez, A.; Rozkiewicz, D.; Whitesides, G. M.; Hollar, K. A. Using Paper-Based Diagnostics with High School Students to Model Forensic Investigation and Colorimetric Analysis. J. Chem. Educ. 2014, 91, 107–111. Ahrenkiel, L.; Worm-Leonhard, M. Offering a Forensic Science Camp to Introduce and Engage High School Students in Interdisciplinary Science Topics. J. Chem. Educ. 2014, 91, 340–344. Coticone, S. R.; Van Houten, L. B. DNA, Drugs, and Detectives: An Interdisciplinary Special Topics Course for Undergraduate Students in Forensic Science. J. Coll. Sci. Teach. 2015, 45, 24–30. Menshew, D. Using Biotechnology, CSI, and Zombies to Promote Science Education in One of America’s Most Challenging Regions. J. Commer. Biotechnol. 2015, 21, 69–77. Cresswell, S. L.; Loughlin, W. A. An Interdisciplinary Guided Inquiry Laboratory for First Year Undergraduate Forensic Science Students. J. Chem. Educ. 2015, 92, 1730–1735. Friesen, J. B. Forensic Chemistry: The Revelation of Latent Fingerprints. J. Chem. Educ. 2015, 92, 497–504. Zuidema, D. R.; Herndon, L. B. Using the Poisoner’s Handbook in Conjunction with Teaching a First-Term General/Organic/Biochemistry Course. J. Chem. Educ. 2016, 93, 98–102. Hamnett, H. J.; Korb, A. S. The Coffee Project Revisited: Teaching Research Skills to Forensic Chemists. J. Chem. Educ. 2017, 94, 445–450. Charkoudian, L. K.; Heymann, J. J.; Adler, M. J.; Haas, K. L.; Mies, K. A.; Bonk, J. F. Forensics as a Gateway: Promoting Undergraduate Interest in Science and Graduate Student Professional Development Through a First-Year Seminar Course. J. Chem. Educ. 2008, 85, 807–812. Schurter, E. J.; Zook-Gerdau, L. A.; Szalay, P. Analysis of a Suspected Drug Sample. J. Chem. Educ. 2011, 88, 1416–1418. Morra, B. The Chemistry Connections Challenge: Encouraging Students to Connect Course Concepts with Real-World Applications. J. Chem. Educ. 2018, 8–11. Maeder, E. M.; Corbett, R. Beyond Frequency: Perceived Realism and the CSI Effect. Can. J. Criminol. Crim. Justice 2015, 57, 83–114. Podlas, K. CSI Effect: Exposing the Media Myth. Fordham Intell. Prop. Media Ent. LJ 2005, 16, 429–465. Newmann, F. M.; Wehlage, G. G. Five Standards of Authentic Instruction. Education 1994, 21, 206–210. Sutherland, L.; Markauskaite, L. Examining the Role of Authenticity in Supporting the Development of Professional Identity : An Example from Teacher Education. High. Educ. 2012, 64, 747–766.

230 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

19. Dukes, A. D., III; Gardner, D. E. An Unconventional Approach to Procedural Development in Analytical Chemistry using Food Coloring and Absorption Spectroscopy. Chem. Educ. 2017, 22, 208–211. 20. Almirall, J. R.; Furton, K. G. Trends in Forensic Science Education : Expansion and Increased Accountability. Anal. Bioanal. Chem. 2003, 376, 1156–1159. 21. Office of Highway Safety and Justice Programs. High Crimes and Misdemeanors: A Five Year Overview of Indicators of Illegal Drug Activity in South Carolina, 2014 Edition; South Carolina Department of Public Safety: Columbia, SC, 2014. 22. Kostecka, K. S.; Rabah, A.; Palmer, C. F., Jr. GC/MS Analysis of the Aromatic Composition of Gasoline. J. Chem. Educ. 1995, 72, 853–854. 23. Hilton, R. Charleston Police Department Forensic Services Division Controlled Substances Procedure Manual, 3rd Ed.; Charleston Police Department: Charleston, SC, 2016. 24. Linstrom, P. J.; Mallard, W. G. NIST Chemistry WebBook, NIST Standard Reference Database Number 69; National Institute of Standards and Technology: Gaithersburg, MD, 2018. 25. South Carolina Judicial Branch Rule 702, Testimony by Experts. https://www.sccourts.org/ courtReg/displayRule.cfm?ruleID=702.0&subRuleID=&ruleType=EVD (accessed Dec 14, 2018). 26. Blackmun, H. Daubert et. al. v. Merrell Dow Pharmaceuticals, Inc.; 1993; Vol. 509, pp 579–601. 27. Smith, D. K. From Crazy Chemists to Engaged Learners Through Education. Nat. Chem. 2011, 3, 681–684.

231 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Chapter 13

Generative Learning Strategies and Prelecture Assignments in a Flipped Forensic Chemistry Classroom Downloaded via CHALMERS UNIV OF TECHNOLOGY on November 17, 2019 at 00:05:24 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Tamra Legron-Rodriguez* Department of Chemistry, University of Central Florida, Orlando, Florida 32816, United States *E-mail: [email protected].

The purpose of this chapter is to show how several instructional strategies were incorporated into a senior-level forensic chemistry course, and how the strategies work together in a complementary fashion. Implementing a flipped classroom model allowed for class time to be used for team-based learning such as projects, activities, and quizzes. Class time was also used for instruction on and the exploration of generative learning strategies such as concept maps, drawing activities, practice testing, and summarizing activities. Generative learning strategies have been shown to promote the generalization and transfer of knowledge. The generative learning activities were designed in a way to help students understand the importance of constructing and interpreting information outside of class time as they study. These generative strategies were then used by students to complete weekly prelecture assignments. These instructional strategies were chosen in an attempt to maximize student engagement in the process of learning.

Introduction The forensic science program at the University of Central Florida organizes the upper-division courses by subject specialization. For example, students complete courses in microscopy, forensic biochemistry, trace evidence, and courtroom testimony. This chapter is in reference to a forensic chemistry course specialized in the analysis of controlled substances and seized drugs. The Forensic Analysis of Controlled Substances is an upper-division, required course for all forensic science bachelor of science students. Its prerequisites include General Chemistry, Organic Chemistry, Analytical Chemistry, Introduction to Forensic Science, and Forensic Microscopy. The course meets for three hours of lecture per week and has a required laboratory component for which students are coenrolled. A large research base has demonstrated the benefits of flipped classrooms and active learning strategies for academic success (1–3). A flipped class allows instructors to move the delivery of © 2019 American Chemical Society Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

content (lecturing) outside the classroom, creating time during class meetings to engage students in active learning. One model for active learning is the team-based approach offered by Michaelsen, Bauman Knight, and Fink whereby instructional strategies are chosen to “support the development of high performance learning teams” and “provide opportunities for these teams to engage in significant learning tasks (4).” The use of collaborative learning (such as team-based learning) allows students to learn from one another and clarify their own thinking through the argumentation process (5). Therefore, classroom environments where students learn from one another can have significant benefits for learning. While instructional strategies are important in creating opportunities for collaboration within the classroom, helping students develop effective learning strategies outside the classroom should also be considered. In an effort to engage students in evidence-based strategies outside the classroom, several generative learning activities were incorporated as part of prelecture activities throughout the semester. Generative learning activities that prompt students to build relationships and construct meaning between concepts have been shown to promote the generalization and transfer of knowledge (6, 7). These generative learning strategies were assigned as part of weekly prelecture activities. Prelecture activities have been shown to have many benefits to student learning by giving students an initial exposure to course content before class time (8–10). The most current offering of the Forensic Analysis of Controlled Substances course utilized three research-based practices. Course content was delivered in a flipped format, whereby lectures were delivered outside of class using instructional videos and class time was used for active learning. Active learning in the classroom followed the model of team-based learning (4). Students were assigned prelecture activities that included generative learning strategies. Though content delivery, use of class time, and unit tests changed significantly, the final exam has remained relatively unchanged, with slight modifications to Year 2 and Year 3 to reflect changes in content coverage. Comparing the final exam scores of three iterations of the course, called Year 1, Year 2, and Year 3, will serve as evidence for future changes in instructional strategies.

Course Format and Instructional Strategies Over the past several years, the instructional strategies used in the Forensic Analysis of Controlled Substances course have changed significantly. Year 1 represents the first offering of the course, which included a traditional approach of lecturing during class time. Year 2 and Year 3 involved significant changes to the instructional strategies, which are described below. The changes in instructional strategies are outlined in Table 1. Modifications to the course format have been guided by the implementation of a flipped classroom approach, team-based learning, and the incorporation of generative learning strategies outside of class time. In Year 1, the majority of class time was used for instructor lecturing with a small amount of time each class designated to students solving problem set worksheets. Students were encouraged by the instructor to work together to complete the problem sets, though group work was not formalized or required. As such, Year 1 serves a baseline with no alternative instructional strategies implemented. The most significant instructional change occurred between Year 1 and Year 2. In Year 2, content delivery changed to the flipped classroom format, whereby content was delivered through instructional videos accessed through the online learning management system. These instructional videos were recorded by the instructor as voiced screen casts of PowerPoint slides, lecture notes, and problem solving. The content of the videos was comparable to the coverage of topics presented during lecture time in Year 1. The flipped format was chosen for the Forensic Analysis of Controlled 234 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Substances course to allow for class time to be used to engage students in team-based learning. Class time was used for activities such as team quizzes, team problem sets, student presentations, collaboration on research projects, and peer review. Also, in Year 2, six unit tests were given, prelecture activities were assigned, and a team-based approach was implemented. The team-based learning approach was modeled according to the text Team-Based Learning: A Transformative Use of Small Groups in College Teaching by Michaelsen, Bauman Knight, and Fink (4). Table 1. Instructional Strategies Used in Course Instructional Strategy

Year 1 (N = 22)

Year 2 (N = 12)

Year 3 (N = 22)

Content delivery and use In-class lecturing, Flipped classroom with unstructured problem sets team-based learning of class time

Flipped classroom with team-based learning

Prelecture assignment

None

Copies of handwritten notes

Copies of handwritten notes including generative learning strategies

Number of unit tests

Three unit tests

Six unit tests

Six unit tests, with post-test debriefing

Practice tests provided

Yes

Yes

Yes, with a help guide

Final exam content

Same

Same

Same

The instructional strategies in Year 3 included three changes based on the Year 2 format. These changes are italicized in Table 1. The prelecture assignments included generative learning strategies (which are described below), a help guide was given for each practice test, and, following the return of each unit test, class time was allotted for a debriefing session that prompted students to compare answers within their teams and ask the instructor questions about grading comments. Final Exam Scores As stated previously, the content delivery, use of class time, implementation of prelecture assignments, and unit tests changed significantly between Year 1, Year 2, and Year 3. However, the final exam content only changed slightly between Year 1 and Year 2 and was identical in Year 2 and Year 3. The changes to the final exam between Year 1 and Year 2 were a result of changes to course learning objectives and content covered in the course. Two questions from the Year 1 final exam were removed related to the theory of ultraviolet–visible spectroscopy as this content is now covered in a prerequisite course. One question, which asked students to draw chemical structures of drugs, was removed because it was covered in unit tests for Year 2 and Year 3, and one question, which asked students to define pharmacology terms, was removed. The same instructor graded the final exams for all three years. Table 2 shows a comparison of final exam scores for the three iterations of the course. In Year 1, instructional strategies were traditional and included lecturing during class time, three unit tests, and no prelecture assignments. As expected, the average score on the final exam was the lowest at 65.9 ± 14.7%. In Year 2, significant modifications to the instructional design were implemented, including a flipped format with team-based learning, six unit tests, and prelecture assignments. As expected, the average student score on the final exam increased for Year 2 (74.2 ± 12.2%). The average student score on the final exam for Year 2 is within one standard deviation of the average score for the final 235 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

exam in Year 1, so it is not likely that the increase in scores were statistically significant. Year 3 had only minor changes to the instructional strategies compared to Year 2. These included the addition of generative learning strategies as part of the prelecture assignments as well as an in-class, post-test debriefing. A help guide was also provided for each of the practice tests in Year 3. The average score on the final exam for Year 3 was 88.7 ± 9.2%; this average is greater than one standard deviation above the average final exam scores for Year 1 and for Year 2, indicating that the increase in the final exam score for Year 3 may be statistically significant. However, given the small sample sizes, it is difficult to make conclusions about the statistical significance of the final exam scores. Table 2. Average Scores on Final Exam Year 1 (N = 22) 65.9 ± 14.7%

Average Score

Year 2 (N = 12) 74.2 ± 12.2%

Year 3 (N = 22) 88.7 ± 9.2%

Team-Based Learning Team-based learning was chosen as an instructional strategy as it offers two advantages: it has been shown to improve student learning, and it gives students opportunities to develop the soft skills desired by employers. Team-based learning was modeled according to the text Team-Based Learning: A Transformative Use of Small Groups in College Teaching by Michaelsen, Bauman Knight, and Fink (4). This text was chosen as a model for instructional change as it offers instructors practical guidelines for implementing team-based learning and contains examples for implementation, resources for discussing the team-based approach with students, and a plethora of references. Buy-In Day The instructional strategies for Year 2 and Year 3 differ significantly from a traditional lecturebased format. To prepare students for the flipped classroom and team-based format, the first day of the semester is reserved for a “buy-in day.” The instructor explains the choice of instructional strategies and offers research-based evidence for improved student performance in active learning classrooms where lecturing is minimal or omitted. The team-based format is modeled through an icebreaker activity and a team syllabus quiz, and by allowing the teams to work together to set the grade weights for the semester. The activity of setting grade weights is explained in detail in the text by Michaelsen, Bauman Knight, and Fink (4). This buy-in day helps put students at ease about the choice of instructional strategies and how they will be assessed in the course.

Prelecture Assignments The flipped course format with team-based learning implemented in the Forensic Analysis of Controlled Substances course required that students be responsible for their initial exposure to the course content. Instructional materials were composed of various items such as online videos and assigned reading. Prelecture assignments were utilized to motivate students to engage with the course content prior to attending class. To provide an incentive for students to complete the assigned reading and instructional videos, a small amount of the course grade (5%) was allotted for prelecture assignments. These prelecture assignments were due before the first class meeting of the week, and class activities were designed based on the assumption that prelecture assignments were completed. For Year 2, the prelecture assignments included an upload of students’ handwritten notes on assigned reading and videos to the online learning management system. Students submitted these as 236 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

scanned copies or digital photos. Handwritten notes were chosen to minimize “copy and paste” and plagiarism. Additionally, typed notes are not conducive to drawing instrumental diagrams, chemical structures, flow charts, etc. It was observed that most students simply copied the information directly from the instructional videos and reading with little effort to engage in the cognitive process. Therefore, in Year 3, the prelecture assignments included an upload of students’ handwritten notes as well as generative learning strategies. These generative learning strategies were implemented with the intent of getting students to engage with the course content in a meaningful way, as opposed to rote copying. The strategies are discussed below.

Prelecture Assignments with Generative Learning Strategies Generative learning activities that prompt students to build relationships and construct meaning between concepts promote the generalization and transfer of knowledge (6, 7). Engle found that the degree of transfer was directly related to the time students spend constructing meaning from information (11). Therefore, the more experience students have at constructing meaning from information, the greater the transfer. One of the ways students were prompted to construct meaning in the course was through the use of generative learning activities. Fiorella and Mayer offer two invaluable resources for instructors desiring to incorporate generative learning into the classroom. Their text, Learning as a Generative Activity: Eight Learning Strategies that Promote Understanding, and article, Eight Ways to Promote Generative Learning, offer background information about the cognitive processes of generative learning as well as considerations for implementing each of the strategies (6, 12). Of the eight strategies outlined, four were chosen to be included as part of the prelecture assignments: concept maps, drawing, practice testing, and summarizing. Each week, one of the aforementioned generative learning strategies was assigned as part of the prelecture assignment. Students uploaded the completed generative learning strategies as part of the prelecture assignment to the online learning management system. Concept Maps Concept maps were the most commonly used generative learning strategies in the course. Scaffolded support and training were offered to students on how to construct a concept map prior to the activity being assigned as part of the prelecture activity. For example, before a concept map is assigned as part of the prelecture assignment, a priming activity was completed during class time. The goal of the priming activity is to teach students the skills needed to generate a concept map. The priming activity consists of a short passage about intermolecular forces and instructions on how to create a map. Instructions for the concept map are: What is a Concept Map? A concept map is a spatial map containing key terms and the relationships among them. The key terms are written inside boxes. The lines connecting the boxes have descriptions that explain the relationship between the terms. To create a meaningful concept map (and one that helps you learn effectively), follow these steps: select, organize, and integrate. Select the important information (what are the key concepts from the reading?). Then organize the information by placing it spatially (how will the key concept boxes be arranged on the page; how will the lines/connections be drawn?). Finally, integrate the new concepts with prior knowledge (how do these concepts relate to what I already know?). 237 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Practice—concept map instructions: The main topic for the concept map is “Intermolecular Forces.” You need at least 10 key terms (words inside boxes). Remember to show how the concepts are related by labeling all of the lines between the boxes and integrate with prior knowledge. The topic of intermolecular forces was chosen as it is a familiar topic, and it is assumed that students can easily make connections to how these terms are related. Students were given approximately 10 minutes to read the passage individually and create a map. Then students worked in teams to generate a team map. Half of the teams presented and explained their intermolecular forces maps to the class. Then students were tasked with creating a concept map on a topic of their choice, and the remaining teams presented and explained their maps to the class. The instructor offered support and guidance during the generation of the maps as well as during the presentations. The supports were modified and decreased as the concept map activity was repeated throughout the semester. Concept map #1 instructions: The concept map needs at least 15 key terms (6 of the 15 must be: designer drugs, analogs, precursors, cutting agents, diluent, adulterant). The remainder of the key terms are your choice. It is recommended that you choose a few of the terms from past course readings/videos or other courses such as general chemistry, organic chemistry, or analytical chemistry. Also remember to label the lines between the boxes (the lines must explain the relationship between the key terms in the boxes). As the semester progresses, less support is offered for the concept map activities: Concept map #2 instructions: The main topic for the concept map is gas chromatography. At least 15 key terms (words inside boxes) are needed; at least 5 of the 15 key terms should relate gas chromatography to thin-layer chromatography and liquid–liquid extraction. Remember to show how the concepts are related by labeling all of the lines between the boxes. Drawing When drawing was chosen as a generative learning activity for the prelecture assignment, scaffolded supports for students were also provided. Students were given partially completed drawings for drawing activity #1 and drawing activity #2 with instructions on what to draw and explain. For drawing activity #3, a partially completed sketch was not provided, however, support was still provided as to the important aspects of the drawing. Example drawing activity #1. Diagram and explain how light interacts with a sample when attenuated total reflectance is used as the sampling technique for IR spectroscopy analysis. Make sure to answer in your own words. Example drawing activity #2. Label the missing terms inside the dashed boxes on the gas chromatography diagram: carrier gas, column, detector, injector port, oven. Explain the function/purpose of each of the missing terms as they pertain to gas chromatography analysis.

238 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Example drawing activity #3. Diagram the components of a mass spectrometer (electron impact ionization, quadrupole) and explain the purpose of each component. In your sketch and explanation, be sure to include the ionization chamber, the mass analyzer, and the detector. Practice Testing When practice testing was included as part of the prelecture assignment, students uploaded a copy of the completed practice test. Practice testing (or self-testing) promotes metacognition by making evident to students what they do and do not know. Metacognitive practices have been shown to increase students’ ability to transfer knowledge to new settings (5) and integrate information by building new connections (12). Practice testing was used several times throughout the semester, particularly when a prelecture assignment was due a short time before a unit test. Similar practice tests were provided for Year 1, Year 2, and Year 3. Detailed answer keys were not provided for practice tests, however, students were encouraged to work the practice tests and ask questions during class time or office hours. In Year 3, a help guide was provided. The help guide contained numerical answers for questions involving calculations and key points to consider for essay questions. For example, one question on a practice test asked students to calculate the partition coefficient for an extraction involving water and toluene. The help guide for this question reads: First find the difference between the total moles and the moles in the organic phase. This gives the moles in the aqueous phase. Then find the partition coefficient. The final answer is 5.14. Reading and resources for this concept can be found on the week 2 module page. Instead of giving a detailed answer key with problems solved step-by-step, the help guide supports students in identifying key information and processes needed to solve the problems. The help guide also supports students by identifying where helpful resources can be found to solve the problem. Students were also given the framing language below to encourage them to complete the practice tests. Research related to practice testing indicates that practice tests are best taken 2–5 days before the actual test and in the same test environment. It is recommended to print out the test and take it without looking at any notes/resources in a quiet place with only the coversheet and a scientific calculator. Give yourself the same amount of time that you have on the actual test. Summarizing The fourth and final generative learning strategy that was included as part of the prelecture assignments was summarizing activities. Similar to the concept map and drawing activities, students were given support to complete the summarizing activities. A template document was provided for students to complete, which was then uploaded as part of the prelecture assignment. Students were assigned a passage from a text and given the following framing language:

239 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

After reading the assigned passage, make a list of at least 10 main idea words/key terms. Then write a five-sentence summary about your main idea words/key terms. Your summary should capture the main ideas and the crucial details necessary for supporting them. Think of your audience as a fellow student that did not read the passage but needs to know the most important information. Concisely explain the main ideas in your own words.

Conclusion While final exam scores were offered as a comparison of student performance with various instructional strategies, the intent of this chapter was to provide an example of how the instructional strategies can be used together to facilitate student learning. Since several instructional changes were implemented between Year 1 and Year 2 as well and between Year 2 and Year 3, it is difficult to pinpoint which of these strategies had the greatest effect on student performance. Using the flipped approach enabled class time to be allocated for active learning within the framework of teambased learning. Prelecture assignments provided incentive for students to complete the reading and online lecture videos before class. Furthermore, the generative learning strategies prompted students to build relationships and construct meaning between concepts, and the strategies promoted metacognition through practice testing. The largest benefit of generative learning is that it promotes the generalization and transfer. Evidence of this may be found in the increased final exam scores for Year 3. However, due to the small sample size, more data should be collected in future iterations and a mixed-method approach used.

References 1.

Seery, M. K. Flipped Learning in Higher Education Chemistry: Emerging Trends and Potential Directions. Chem. Educ. Res. Pract. 2015, 16, 758–768. 2. Gross, D.; Pietry, E. S.; Anderson, G.; Moyano-Camihort, K.; Graham, M. J. Increased Preclass Preparation Underlies Student Outcome Improvement in the Flipped Classroom. Life Sci. Educ. 2015, 14, 1–8. 3. Abeysekera, L.; Dawson, P. Motivation and Cognitive Load in the Flipped Classroom: Definition, Rationale, and a Call for Research. High. Educ. Res. Dev. 2015, 34, 1–14. 4. Michaelsen, L. K.; Bauman Knight, A.; Fink, L. D. Team-Based Learning: A Transformative Use of Small Groups in College Teaching; Stylus Publishing: Sterling, VA, 2002. 5. How People Learn: Brain, Mind, Experience, and School; Bransford, J., Ed.; National Research Council, National Academies Press: Washington, DC, 2000. 6. Fiorella, L.; Mayer, R. E. Learning as a Generative Activity: Eight Learning Strategies that Promote Understanding; Cambridge University Press: New York, 2015. 7. Wittrock, M. Generative Learning Process of the Brain. Educ. Psychol. 1992, 27, 531–541. 8. Moravec, M.; Williams, A.; Aguilar-Roca, N.; O’Dowd, D. K. Learn Before Lecture: A Strategy that Improves Learning Outcomes in a Large Introductory Biology Class. Life Sci. Educ. 2010, 9, 473–481. 9. Kinsella, G. K.; Mahon, C.; Lillis, S. Using Prelecture Activities to Enhance Learner Engagement in a Large Group Setting. Active Learn. High. Educ. 2017, 18, 231–242. 10. Seery, M. K.; Donnelly, R. The Implementation of Prelecture Resources to Reduce In-Class Cognitive Load: A Case Study for Higher Education in Chemistry. British J. Educ. Technol. 2011, 43, 667–677. 240 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

11. Fiorella, L.; Mayer, R. E. Eight Ways to Promote Generative Learning. Educ. Psychol. Rev. 2016, 28, 717–741. 12. Engle, R. A. Framing Interactions to Foster Generative Learning: A Situative Explanation of Transfer in a Community of Learners Classroom. J. Learn. Sci. 2006, 15, 451–498.

241 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Editors’ Biographies

Downloaded via CHALMERS UNIV OF TECHNOLOGY on November 17, 2019 at 00:09:57 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Amanda S. Harper-Leatherman Amanda S. Harper-Leatherman received her B.S. in Chemistry from St. Olaf College in Northfield, MN and her Ph.D. in Analytical Chemistry from the University of North Carolina at Chapel Hill. After a position as a National Research Council Research Associate at the Naval Research Laboratory in Washington, DC, she joined the chemistry faculty of Fairfield University in 2006 and is currently an associate professor. Her research interests include nanomaterials, bioanalytical chemistry and chemical education. She has authored or co-authored 18 peer-reviewed articles or chapters and was co-editor of the ACS Symposium Series Volume, The Science and Function of Nanomaterials: From Synthesis to Application. Dr. Harper-Leatherman has regularly taught an Introduction to Forensic Science course for non-science majors for the past 10 years.

Ling Huang Ling Huang is an Associate Professor of Chemistry at Hofstra University. He received his Ph.D. in Chemistry from the University of Virginia, M.S. in Analytical Chemistry from the University of Oklahoma, and B.S. in Chemistry from Fudan University in Shanghai, China. His current research at Hofstra includes rapid identification and quantification of designer drugs, e-cigarette fluid analysis, ink analysis, and forensic elemental analysis of gunshot residues. In 2013 he also published a review on chemistry smartphone apps in the Journal of Chemical Education.

© 2019 American Chemical Society Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

Downloaded via 37.9.47.157 on November 5, 2019 at 01:56:13 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Subject Index C

E

Chemistry and crime, investigating chemistry analytical chemistry, 19 biochemistry, 23 DNA profiling, 25 conclusion, 27 forensic science in the arts and in outreach programs, role, 17 forensic sciences, importance of diversity and inclusion, 26 forensic science teaching, brief history, 15 general/introductory chemistry, 17 inorganic chemistry, 21 deprotonated gallate ligands, 22 inkless fingerprinting system, structure of complex formed, 21f reactions, steps originally proposed, 23f introduction, 13 Locard Exchange Principle, 15 non-science major or the chemistry major, forensic science, 16 organic chemistry, 18

Engage nontraditional learners, using forensic science case studies, 64 conclusions, 76 course assessment, success, and challenges, 75 course overview and content, 62 forensic science lab, example weekly schedule, 63t engaging interest with forensic science, 60 engaging thinking with the scientific method, 61 scientific method steps, example of the list, 62f introduction, 59 lab experiments, 65 crime scene narrative, example, 69t ethanol, example infrared spectrum, 68f example lab notebook page, 70f final crime scenes, example set, 69f lab activities and their safety hazards, 67t mini lectures and video clips, 63 student learning, assessment, 71 grades and SLOs, breakdown, 71t lab notebook rubric, 74t partial rubric, 72 postlab questions, grading rubric, 73t scientific method assignment rubric, 73t week 5 crime scene rubric, 75t

D Drugs and DNA, analytical chemistry curriculum forensic DNA typing, 162 abbreviations, 164t DNA quantification, 163 forensic DNA typing and associated analytical techniques, 162f illicit drugs, analysis, 156 amphetamines, sample toxicological LCMS TIC and EICs, 159f controlled substances, structural examples, 160f OTC drugs, structures, 158f synthetic cannabinoid AM-2201, ATRFTIR and Raman spectra, 161f introduction, 155

F First-semester general chemistry laboratory, on utilizing forensic science to motivate students, 93 introduction, 94 murder mystery redox experiment, 94 experimental methods and results, 95 summary, 100 teaching chemistry with forensic science, 96 analytical methods and instrumentation, 99 forensic science materials, 98 249

Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

forensic science scenarios, 96 Flipped forensic chemistry classroom, generative learning strategies and prelecture assignments conclusion, 240 course format and instructional strategies, 234 course, instructional strategies used, 235t final exam, average scores, 236t generative learning strategies, prelecture assignments, 237 introduction, 233 prelecture assignments, 236 Forensic science, introduction to teaching chemistry background, 1 chapters related to using forensic science to teach lower level chemistry, overview, 5 chapters related to using forensic science to teach upper level chemistry, overview, 5 conclusion, 6 teach chemistry, using forensic science, 2 Forensic science and forensic nursing, interdisciplinary learning activity, 137 alternative forensic laboratory analyses, 151 conclusion, 152 fiber evidence, forensic laboratory analysis, 145 burn fiber analysis experiment, example data, 146t control group of students, pre-survey and post-survey results, 149f forensic evidence, 150 nylon and polyester fibers, example fiber ATR-FTIR spectra, 148f standard multifiber fabric strips, examples of the colors formed, 147f students participating, pre-survey and postsurvey results, 149f interdisciplinary learning activity description, 139 evidence collection, hospital/clinical simulation, 143 Fairfield University simulated forensic evidence collection kit, document, 144f mock crime scene scenario, 142 student learning assessment, select preevent and post-event survey questions used, 140t

three-hour interdisciplinary learning activity event, schedule of activities, 141t introduction and background, 138 Forensic themes, integration, 203 conclusion, 217 forensic science, instrumental analysis, 207 cyclohexane and benzene's elution orders, 210f ephedrines and pseudoephedrine, CE electropherograms, 209f ephedrines and pseudoephedrine, GC chromatograms, 208f illicit drugs, chromatographic analysis, 212f illicit drugs, extraction and chromatographic analysis, 211 nitro-containing compounds used as mock explosives, Raman spectra, 213f seized cocaine exhibit, analysis of a given 1H NMR spectrum, 215f seized cocaine exhibit, structures of some components, 214f transacylation reaction, 216 transacylation reaction between cocaine and benzocaine, potential product, 216f introduction and forensic science curricula, 204 Bachelor's program in forensic science, suggested course sequence, 205t Pace University, suggested course sequence, 206t From DUIs to stolen treasure, 169 conclusion, 199 forensics upper-division elective course, 185 caliber and land and groove marks, bullet samples for students, 197f case background, 188 chemicals and supplies used, manufacturer sources, 192t crime occurred and analyses, evidence collected immediately, 189t crime scene scenario, 187 crime scene scenario, identified and tagged pieces of evidence, 194f drug analysis, 195 evidence analysis related to student D and student E, results, 191t forensic serology, 196

250 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

lecture and laboratory portions of the course, forensic topics covered, 186t Mikrosil putty, tool mark impressions, 197f secondary sites and analyses, evidence collected, 190t shoeprint impression left, 198f instrumental analysis, 170 acetaminophen, sample student calibration curve data, 184f 1:20 dilution, sample student chromatogram, 183f ethanol, headspace GC-FID analysis, 177 fluorescence measurements, sample student external calibration curve, 174f forensic rotation, summary, 172t forensic science laboratory manual, 179 fraction C, sample student total ion chromatogram, 180f fraction E, sample student total ion chromatogram, 181f GC-MS, 178 gentle heating, M HCl, 175 laboratory procedure, 173 noninstrumental approach, 182 sample student standard addition calibration curve, 176f students, evidence, 172f introduction, 170

G GC/MS technique, using expert witness testimony forensics as a medium to teach chemistry, 221 mock trial, procedure, 224 analysis, gas chromatogram generated, 225f blank drug analysis worksheet, example, 224f common colorimetric tests for illicit compounds, results, 225t simulated illicit compound, mass spectrum, 226f simulated drug analysis, experimental procedure, 223 students, authentic experience, 229 students' mock testimony, discussion, 227

I Interdisciplinary learning communities conclusions, 132 CEM 1620 and ENL 2015, course-specific goal, 133f chemistry and laboratory techniques, foundational knowledge, 133 future assessment, 134 culminating final project, 125 Belmont University interdisciplinary learning community assessment rubric, 132t bloodstain pattern analysis, sample of student, 127f crime story rubric, 129t presentation rubric, 131t scientific evidence, sample of student, 126f scientific evidence rubric, appendix, 130t strychnine poisoning, scientific evidence, 128f curriculum, introducing nontraditional chemistry texts, 119 authenticating paintings, analysis techniques, 120t timeline overview, 119t interdisciplinary experiences, 123 CEM 1620-ENL 2015 class, 123 interdisciplinary learning communities, 110 BELL core program requirements, 111t CEM 1620, comparative laboratory schedule, 117t fall 2018 semester, traditional cem 1620 lecture schedule, 114t General Chemistry II, assessment items, 113t General Chemistry II, current course content, 118t recent interdisciplinary learning community course offerings, examples, 112t spring 2018 semester, traditional cem 1620 lecture schedule, 115t introduction, 109 scientific analysis techniques, engaging students in an exploration, 121 bloodstain patterns, samples, 123f forensic-themed lab activities, 122

251 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.

mysterious death dilemma, 49 Chemist's challenge, 51 four drugs being studied, HPLC working ranges, 49t Hardware's challenge, 50 nonmajors or introductory chemistry, 37 alcohol lab, 38 ATR-FTIR instrument, cloth sample being analyzed, 40f blood lab, 40 crime scene processing, 37 DNA fingerprinting lab, 38 drug lab, 39 fiber lab, 39 fingerprinting lab, 39 flame AAS used, 41f ink lab, 38 narcotics identification kit pouch used, 39f poison lab, 41 solid-state breath alcohol detector, 38f overdose lab, 43 absorbance data, UV/VIS spectrometer for collection, 45f analytical chemistry, forensics, 47 diphenhydramine HCl, chemical structure, 44f nonmajors course, summary table for lab experiments, 46t procainamide, lidocaine, and disopyramide used, separation, 48f paint analysis, 41 CIELAB color space, 42 CIELAB color space, American Chemical Society, 43f senior-level analytical chemistry, 52

O Open educational and digital resources, teaching introductory forensic chemistry case studies and research and development reports, 87 conclusion, 88 online and open educational resources, 89 course development, 81 course learning outcomes and assessment methods, 82t introductory forensic chemistry, weekly schedule, 82t digital and open educational resources, 79 Arson evidence collection open-source quiz, 81t article review assignment, 80t future work and next steps, 88 general course themes, 83 introduction, 79 introductory forensic chemistry labs, 84 forensic chemistry labs, 84t specific course details, 85

U Undergraduate analytical curriculum, incorporating forensic science, 35 analysis of matches, ICP-OES, 53 each element in match heads, rough concentrations, 54t matchhead sample preparation, items needed, 53f conclusions, 55 introduction, 36

252 Harper-Leatherman and Huang; Teaching Chemistry with Forensic Science ACS Symposium Series; American Chemical Society: Washington, DC, 2019.