Student-Guided Three-Dimensional Printing Activity in Large Lecture

Dec 4, 2018 - A summary of this feedback and constructive remarks on the best practices ..... We emphasize this proof-of-concept study was not designe...
0 downloads 0 Views 2MB Size
Activity pubs.acs.org/jchemeduc

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

Student-Guided Three-Dimensional Printing Activity in Large Lecture Courses: A Practical Guideline Denis Fourches* and Jeremiah Feducia Department of Chemistry, North Carolina State University, Raleigh, North Carolina 27695, United States

Downloaded via UNIV OF SASKATCHEWAN on January 1, 2019 at 09:46:47 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

S Supporting Information *

ABSTRACT: Modern technology stimulates the development of innovative classroom activities. We designed a 3D printing activity in two separate Organic Chemistry lectures of at least 200 students each. This assignment required students to 3D print a molecule of their choice, relying on services made available through the university libraries. Data obtained through a survey at the end of the semester provided key information on the students’ experiences with printing 3D models for the first time. A summary of this feedback and constructive remarks on the best practices regarding 3D printing assignments in large lecture courses are presented.

KEYWORDS: Chemoinformatics, Organic Chemistry, Hands-On Learning/Manipulatives, First-Year Undergraduate/General, Second-Year Undergraduate



INTRODUCTION Over the past few years, the increased access to 3D printers has made it easier for instructors across disciplines to create new hands-on activities otherwise unimaginable. This is particularly true for chemistry teachers interested in engaging their students with innovative techniques to illustrate the structural characteristics and properties of molecules. The chemical literature dedicated to the topic of “3D printing” has primarily focused on the creation of visualization aids for instruction,1−7 the printing of experimental tools for laboratory experiments,8−14 or the conception of diverse protocols for converting electronic files to 3D printing-compatible objects.15,16 Reports on student-generated models are related to very small class sizes and activities which inherently limit the student’s choice in what and how to 3D print.17,18 More recent efforts showcase the printing of highly complex molecular models.19−23 This report describes a proof-of-concept 3D printing activity performed in two separate organic chemistry courses.



Herein, an independent 3D printing activity was assigned in both a first- and a second-semester organic chemistry course as a means to enhance student engagement. The general premise was as follows: students were given the opportunity to choose a molecule of interest to them (e.g., drug, metabolite, pesticide, or dye), convert its chemical structure into a three-dimensional object, and print it. In the first-semester course, 208 students were given the option of either 3D printing a molecule, or creating a YouTube video describing that molecule. In the second-semester course, 214 students were asked to create a model of their molecule using a method of their choice, of which 3D printing was an option, and generate information on the molecule to share with the general public. Students in both classes were given the semester to complete the project with several checkpoints to ensure gradual student progress. Since STL files are required as inputs for 3D printers, students were directed to get their STL files by one of the following options: (1) creating the 3D model from scratch, (2) building the STL file using an automated tool such as our RealityConvert Web server,24 or (3) downloading the STL file from the NIH 3D print exchange Web server,25 which provides expert-verified and chemically valid 3D models. Meanwhile, the de novo creation of an STL file is relatively easy for students with minimum computer skills using software programs available to them as freeware or through university libraries (see the Supporting Information). For students wanting to

3D PRINTING: A STUDENT-GUIDED PROJECT

Due to the typical size of an organic chemistry course at a large university, individual projects are obviously not the norm when it comes to extra assignments. This is due to the considerable amount of time required to grade those individual assignments and the high potential for redundancy in project topics. Aside from additional reading and electronic homework, we posit that other types of individual assignments are feasible even in a 200+ student classroom, 3D printing being one of them. © XXXX American Chemical Society and Division of Chemical Education, Inc.

Received: May 10, 2018 Revised: December 4, 2018

A

DOI: 10.1021/acs.jchemed.8b00346 J. Chem. Educ. XXXX, XXX, XXX−XXX

Journal of Chemical Education

Activity

Figure 1. Typical protocol for 3D printing molecules.

Table 1. Distribution of Student Responses to Common Survey Questions Survey Administration Pre-activity Post-activity

First-Semester Course N = 208

Survey Question Characterizations Number of participating students Students who ever 3D printed an object, % Self-declared first time users of 3D printers, % Students who declared activity is good for 3D visualization of chemicals, % Students who declared learning from this project, % Students who declared enjoying working on this project, % Students who declared increased engagement from this project, %

create their own STL file, we suggested the generation of a MOL file using PerkinElmer’s ChemDraw26 or freely accessible Biovia’s Draw27 and then creation of the STL file using freely accessible Blender28 and Chimera.29 Students were encouraged to submit images of their molecules (indicating stereochemistry with dashes and wedges) prior to creating their STL files to ensure that stereochemistry was not lost in the building process. The corresponding flowchart is provided in Figure 1. Perhaps surprising was the minimal amount of class time required to implement this activity. All technical procedures and forum were provided via the electronic platform of the course (in this case, Moodle). In the first-semester course, one class meeting was used to provide students an opportunity to meet in groups to discuss project progress, share troubleshooting tips, and go do printing projects together at the library. In the second-semester course, although no class periods were used for the project, any clarifications to the project prompt or troubleshooting suggestions were communicated to the class by email.



Second-Semester Course N = 214

99 15 85 95

39 13 87 60

99 97 92

90 N/A 49

In both courses, at least 85% of students who 3D printed were first-time users. That could explain why only 48% of enrolled students in the first-semester chose to do this handson activity. Despite the increasing accessibility of 3D printers, a vast majority of students were completely new to this technology and perhaps were apprehensive about it. This activity provided students with an opportunity to gain technical skills outside of their general curriculum. Moreover, the difference between the two classes with respect to how students perceived the activity in relation to their visualization skills could also seem surprising. The much lower percentage in the second-semester course may at first appear counterintuitive to the project goal until one considers the course content, which primarily focuses on planar functional groups. The first-semester course on the other hand requires students to translate two-dimensional representations on paper to a three-dimensional molecular object. This skill is developed extensively in most first-semester courses when students encounter topics such as stereochemistry and substitution reactions (SN1/SN2). Although second-semester students did not overwhelmingly agree that the activity increased their engagement in the course (19 out of 39), first-semester students did (91 out of 99). This result may be important when selecting which course to implement such an activity if increased engagement and enthusiasm is the primary goal. In both classes, most students indicated that they felt that the project helped to increase the connection between the course content and real-world applications. The survey also included a free response section for students to comment on their overall experience and offer recommendations. This compiled data is integrated in the word cloud generated by responses to the first-semester course

RESULTS

Student Survey

At the end of the semester, students were asked to fill out an anonymous survey gauging what impact this project had on their overall experience with the organic chemistry class. The surveys given to both sections contained identical questions asking students to rate the effectiveness of this activity on their overall engagement in the course. A summary of the results of the common questions is shown in Table 1. B

DOI: 10.1021/acs.jchemed.8b00346 J. Chem. Educ. XXXX, XXX, XXX−XXX

Journal of Chemical Education

Activity

CDX, or SDF files. This procedure significantly reduces the risk of molecular models with incorrect geometry. Students should be mindful of their molecule with respect to the complexity and size of the printed product to avoid models being too fragile, due to thinly printed bonds, or too large, requiring printing several parts and gluing them together. For example, a student who chose bilobalide (C15H18O8) was able to generate a high-quality 6 in. × 6 in. × 6 in. model, whereas another student who chose aescin (C55H86O24) ended up with an incredibly fragile model. Printing bed sizes are one of the technical limiting factors for large models. Therefore, we recommend that the instructors warn the students regarding the choice of their structures, especially when it comes to the overall number of atoms, structural complexity, and model size. As most of our students utilized the Lulzbot Mini with a printing bed of 6 in. × 6 in., they had most success with molecules in which their ball and stick model contained fewer than 60 atoms. As shown in Supporting Information Figure 2, it is technically feasible to 3D print a molecule of DNA or a spherical molecule of fullerene (C60). But students need to be aware they will likely face technical difficulties that could dramatically slow down or stop their project. Importantly, the size of the model is directly related to its overall printing cost. Most models printed by our students were less than $10 each, which is rather small compared to other required class resources (textbooks, online homework, and so on). Departments concerned about the cost to their students could consider purchasing filament rolls. Although grading was time-consuming, this responsibility was manageable over a relatively short period of time. For this activity, a student’s grade was based on the content submitted at checkpoints and the quality of the printed model. Several metrics for evaluating the quality of the model (see the Supporting Information) were considered: overall printing quality, correct stereochemistry, molecular complexity, and finishing (e.g., sanding and painting). With a clear grading rubric, faculty efforts can be minimized with the help of teaching assistants. Grading for future semesters will include the originality of the molecule, the protocol to generate the STL file, and a better estimate of the individual role of each student within a group. Finally, one could argue only high-achieving students are likely to choose such activity. Based on our results, even though the average score on the final comprehensive exam of students who printed a molecule was slightly higher (111.5 ± 23.7 points out of 150) compared to the average score for the whole class (106.2 ± 27.4 points), the overall distribution of the final grades was fairly similar for both groups of students. The minimum score received by a student for the 3D printing project was 4 points (out of 10 possible points), whereas the average grade was 8.5 bonus points. Lastly, we would not recommend to start this activity in a large class for the first time and make it mandatory without having tested whether the university infrastructure, printers, staff, and the grading system are appropriate.

survey shown in Supporting Information Figure 1. The words “creativity”, “learned”, and “enjoyed” underline how positively the students characterized and envisioned this project once completed. Examples of full comments are given in the Supporting Information (pp S10 and S11). Some comments referred to the time commitment (5−30 h) necessary to conduct this activity. Obviously, it significantly depends on whether the project was done by one individual or a group of students, whether the STL file was created or downloaded, and whether the printed model was sanded and painted. Proposed Best Practices

In order to effectively increase engagement and help connect organic chemistry to real-world applications, students must have the opportunity to select their own molecules. This freedom of choice is possible as long as a detailed protocol for generating input files for 3D printers is provided. Students commonly chose chemicals for very personal reasons; e.g., a student printed a drug his father discovered, while another student whose grandmother has Alzheimer’s disease printed a molecule currently in clinical trials to treat that disease. Still the most popular molecules were those that students encounter every day: caffeine, aspirin, and ibuprofen. Outlining a clear protocol (Figure 1) allowed students to take advantage of freely accessible software available on university computers. Moreover, working closely with the 3D printer facilities is critical to success for large classes. Students participating in this activity had access to two Makerspace facilities located in two separate libraries, totaling 15 printers. Most students printed models using a Lulzbot Mini 3D printer with PLA filament. The close interaction between faculty and staff was crucial because two projects running simultaneously in large classes resulted in an increased work load for the printers. To streamline access, Makerspace staff offered additional workshops to provide safety training to students, enabling them to independently use the printers. We had no report of hazards or injuries during the printing. Checkpoints throughout the semester aim to (i) reiterate and clarify the expectations of the activity, (ii) provide technical tips and suggestions to the class as a whole, and (iii) ensure that a backlog in the printing facilities was not caused by this activity. A backlog of models to be printed can happen, especially when students wait until the last moment to print their molecules. For example, although checkpoints were clearly established, several students waited too long to schedule a printing time and therefore missed the deadline. Thus, having key checkpoints along the semester helps to ensure the printing process is perfectly scheduled. Additional considerations should be made when considering a project of this scale. First, this 3D printing activity could be more engaging if done earlier in the first semester and if more heavily loaded with instructor-driven tutorials, guidance, and thought-provoking discussions to further enhance the pedagogy content, especially when it comes to molecular geometry, structural constraints, or structure−reactivity relationships. Second is the full accessibility to 3D printers and the appropriate software. Although there are hundreds of STL files available online, some students utilized files in which the 3D coordinates of individual atoms were grossly incorrect. Third, students are encouraged to follow our guidelines encompassing the use of a molecular sketcher (with 2D/3D structural cleaning) and user-friendly tools (such as RealityConvert24) to create the STL file from the standardized MOL,



CONCLUSION The increasing accessibility of 3D printers could have a key role in extracurricular activities. Students overwhelmingly enjoyed working on this 3D printing activity and felt more engaged with such a hands-on, technology-based activity within the organic chemistry class. With a close working relationship between faculty and Makerspace facilities, this C

DOI: 10.1021/acs.jchemed.8b00346 J. Chem. Educ. XXXX, XXX, XXX−XXX

Journal of Chemical Education

Activity

(8) Stewart, C.; Giannini, J. Inexpensive, Open Source Epifluorescence Microscopes. J. Chem. Educ. 2016, 93 (7), 1310−1315. (9) Lu, Y.; Santino, L. M.; Acharya, S.; Anandarajah, H.; D’Arcy, J. M. Studying Electrical Conductivity Using a 3D Printed Four-Point Probe Station. J. Chem. Educ. 2017, 94 (7), 950−955. (10) Porter, L. A., Jr.; Washer, B. M.; Hakim, M. H.; Dallinger, R. F. User-Friendly 3D Printed Colorimeter Models for Student Exploration of Instrument Design and Performance. J. Chem. Educ. 2016, 93 (7), 1305−1309. (11) Porter, L. A., Jr.; Chapman, C. A.; Alaniz, J. A. Simple and Inexpensive 3D Printed Filter Fluorometer Designs: User-Friendly Instrument Models for Laboratory Learning and Outreach Activities. J. Chem. Educ. 2017, 94 (1), 105−111. (12) Grasse, E. K.; Torcasio, M. H.; Smith, A. W. Teaching UV-Vis Spectroscopy with a 3D-Printable Smartphone Spectrophotometer. J. Chem. Educ. 2016, 93 (1), 146−151. (13) Hornung, C. H.; Nguyen, X.; Carafa, A.; Gardiner, J.; Urban, A.; Fraser, D.; Horne, M. D.; Gunasegaram, D. R.; Tsanaktsidis, J. Use of Catalytic Static Mixers for Continuous Flow Gas-Liquid and Transfer Hydrogenations in Organic Synthesis. Org. Process Res. Dev. 2017, 21 (9), 1311−1319. (14) Pinger, C. W.; Heller, A. A.; Spence, D. M. A Printed Equilibrium Dialysis Device with Integrated Membranes for Improved Binding Affinity Measurements. Anal. Chem. 2017, 89 (14), 7302− 7306. (15) Rossi, S.; Benaglia, M.; Brenna, D.; Porta, R.; Orlandi, M. Three Dimensional (3D) Printing: A Straightforward, User-Friendly Protocol to Convert Virtual Chemical Models to Real-Life Objects. J. Chem. Educ. 2015, 92 (8), 1398−1401. (16) Van Wieren, K.; Tailor, H. N.; Scalfani, V. F.; Merbouh, N. Rapid Access to Multicolor Three-Dimensional Printed Chemistry and Biochemistry Models Using Visualization and Three-Dimensional Printing Software Programs. J. Chem. Educ. 2017, 94 (7), 964−969. (17) Meyer, S. C. 3D Printing of Protein Models in an Undergraduate Laboratory: Leucine Zippers. J. Chem. Educ. 2015, 92 (12), 2120−2125. (18) Dean, N. L.; Ewan, C.; McIndoe, J. S. Applying Hand-Held 3D Printing Technology to the Teaching of VSEPR Theory. J. Chem. Educ. 2016, 93 (9), 1660−1662. (19) Paukstelis, P. J. MolPrint3D: Enhanced 3D Printing of Balland-Stick Molecular Models. J. Chem. Educ. 2018, 95 (1), 169−172. (20) Smiar, K.; Mendez, J. D. Creating and Using Interactive, 3DPrinted Models to Improve Student Comprehension of the Bohr Model of the Atom, Bond Polarity, and Hybridization. J. Chem. Educ. 2016, 93 (9), 1591−1594. (21) Jones, O. A. H.; Spencer, M. J. S. A Simplified Method for the 3D Printing of Molecular Models for Chemical Education. J. Chem. Educ. 2018, 95 (1), 88−96. (22) Carroll, F. A.; Blauch, D. N. Using the Force: ThreeDimensional Printing a π-Bonding Model with Embedded Magnets. J. Chem. Educ. 2018, 95 (9), 1607−1611. (23) Penny, M. R.; Cao, Z. J.; Patel, B.; Sil dos Santos, B.; Asquith, C. R. M.; Szulc, B. R.; Rao, Z. X.; Muwaffak, Z.; Malkinson, J. P.; Hilton, S. T. Three-Dimensional Printing of a Scalable Molecular Model and Orbital Kit for Organic Chemistry Teaching and Learning. J. Chem. Educ. 2017, 94 (9), 1265−1271. (24) Borrel, A.; Fourches, D. RealityConvert: A Tool for Preparing 3D Models of Biochemical Structures for Augmented and Virtual Reality. Bioinformatics 2017, 33 (23), 3816−3818. (25) NIH 3D-print exchange webserver; National Institutes of Health, https://3dprint.nih.gov (accessed December 2018). (26) ChemDraw; PerkinElmer, http://www.cambridgesoft.com/ Ensemble_for_Chemistry/ChemDraw/ChemDrawProfessional/ Default.aspx (accessed December 2018). (27) Biovia Draw; BIOVIA, http://accelrys.com/products/ collaborative-science/biovia-draw/draw-no-fee.php (accessed December 2018). (28) Blender; Blender Foundation, https://www.blender.org/ (accessed December 2018).

activity stands as proof that class size is no longer a limitation for innovative undergraduate activities involving 3D printing. We emphasize this proof-of-concept study was not designed to adequately and precisely evaluate the potential impact of such an activity on visualization skills. Because multiple years of survey data are necessary to fully and critically evaluate whether such an activity provides students with long-lasting benefits related to chemical visualization and learning skills, we are currently working toward implementing these activities within our Chemistry majors courses.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available on the ACS Publications website at DOI: 10.1021/acs.jchemed.8b00346.



Word cloud, project prompts, student surveys, 3D printing software and technology, grading metric, and examples of comments from students regarding the activity (PDF, DOCX)

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Denis Fourches: 0000-0001-5642-8303 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We are grateful for the cooperation and feedback provided by the students who participated in this project. We also thank Dan Hawkins and other staff members at the Makerspace facilities at North Carolina State University for assisting students and providing additional workshops to train students using the 3D printers for the first time. D.F. gratefully thanks high school interns Sravya Kuchibhotla and Nyree Baldwin for their help with survey encoding.



REFERENCES

(1) Cooper, A. K.; Oliver-Hoyo, M. T. Creating 3D Physical Models to Probe Student Understanding of Macromolecular Structure. Biochem. Mol. Biol. Educ. 2017, 45 (6), 491−500. (2) Carroll, F. A.; Blauch, D. N. 3D Printing of Molecular Models with Calculated Geometries and p Orbital Isosurfaces. J. Chem. Educ. 2017, 94 (7), 886−891. (3) Smiar, K.; Mendez, J. D. Creating and Using Interactive, 3DPrinted Models to Improve Student Comprehension of the Bohr Model of the Atom, Bond Polarity, and Hybridization. J. Chem. Educ. 2016, 93 (9), 1591−1594. (4) Robertson, M. J.; Jorgensen, W. L. Illustrating Concepts in Physical Organic Chemistry with 3D Printed Orbitals. J. Chem. Educ. 2015, 92 (12), 2113−2116. (5) Rodenbough, P. P.; Vanti, W. B.; Chan, S. W. 3D-Printing Crystallographic Unit Cells for Learning Materials Science and Engineering. J. Chem. Educ. 2015, 92 (11), 1960−1962. (6) Casas, L.; Estop, E. Virtual and Printed 3D Models for Teaching Crystal Symmetry and Point Groups. J. Chem. Educ. 2015, 92 (8), 1338−1343. (7) Griffith, K. M.; de Cataldo, R.; Fogarty, K. H. Do-It-Yourself: 3D Models of Hydrogenic Orbitals through 3D Printing. J. Chem. Educ. 2016, 93 (9), 1586−1590. D

DOI: 10.1021/acs.jchemed.8b00346 J. Chem. Educ. XXXX, XXX, XXX−XXX

Journal of Chemical Education

Activity

(29) UCSF Chimera; University of California, San Francisco (UCSF) Resource for Biocomputing, Visualization, and Informatics, https:// www.cgl.ucsf.edu/chimera/ (accessed December 2018).

E

DOI: 10.1021/acs.jchemed.8b00346 J. Chem. Educ. XXXX, XXX, XXX−XXX