Scholarship of Teaching: Online Courses as a Means of Publishing

Aug 16, 2016 - One of the authors (Hutchinson) created two of his own open online courses offered on Coursera, Chemistry Concept Development and Appli...
0 downloads 10 Views 722KB Size
Chapter 2

Downloaded by CORNELL UNIV on September 27, 2016 | http://pubs.acs.org Publication Date (Web): August 16, 2016 | doi: 10.1021/bk-2016-1217.ch002

Scholarship of Teaching: Online Courses as a Means of Publishing Innovations John S. Hutchinson*,1 and Carrie A. Obenland1,2 1Department

of Chemistry, Rice University, 6100 Main St., MS-60, Houston, Texas 77005, United States 2Rice Office of STEM Engagement, Rice University, 6100 Main St., MS-100, Houston, Texas 77005, United States *E-mail: [email protected].

Though the scholarship of teaching has been increasingly recognized in the past two decades, teaching as a form of scholarship differs from other forms in that it is much more difficult to publish. We publish lesson plans, we publish innovative ideas, and we publish research on teaching. But publishing the teaching itself requires non-traditional means. The rise in availability of online courses creates exciting possibilities for demonstrating how lesson plans, innovative ideas, and creative approaches are actually implemented. At Rice University, we have designed a novel General Chemistry curriculum based on the development of concepts through inductive reasoning. Teaching using the Concept Development Study approach has been published via our Coursera courses, Chemistry Concept Development and Application. In this chapter, we will discuss our motivations, approaches, observations, and results.

© 2016 American Chemical Society Sörensen; Online Course Development and the Effect on the On-Campus Classroom ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by CORNELL UNIV on September 27, 2016 | http://pubs.acs.org Publication Date (Web): August 16, 2016 | doi: 10.1021/bk-2016-1217.ch002

Introduction “Great teaching is a form of synthesis and scholarship (1).” The rapid rise in the past few years of the “massive open online course” (MOOC) has been met with well justified skepticism. Why would a university offer their courses free of charge to anyone in the world without regard to admission or prior qualification? Why would a faculty member put in the effort to design and create a course which ordinarily they would offer only to their own students and only as part of their assigned job responsibilities? How are these activities consistent with the fundamental missions of a university and of the faculty? There are a great many answers to these questions, and from discussions with other faculty who have created and offered these courses, the motivations of individual instructors vary from person to person. One of the authors (Hutchinson) created two of his own open online courses offered on Coursera, Chemistry Concept Development and Application I and Chemistry Concept Development and Application II, based on his course in General Chemistry at Rice University (2). The course has been recast as an “on demand” course, and is now offered as General Chemistry: Concept Development and Application. The details of these courses will be discussed later in this chapter. In increasing order of priorities, the motivations for creating these online courses were and are as follows: 1. 2. 3. 4.

To provide access to a Rice course throughout the world To “back-flip” the General Chemistry course at Rice University To educate Chemistry teachers, particularly high school teachers To publish our teaching approach as a scholarly work

The first of these, to provide global access to course content, is the most commonly cited reason for offering a MOOC rather than a campus based course (3). University level education is expensive, and admissions are highly selective. The fraction of people who can afford such an education and can be selected to receive it compared to those who could benefit from such an education is miniscule. The number of people who have accessed this online course in the three times it has been offered in the past three years is vastly greater than the total number of students who have taken Hutchinson’s course at Rice in the twenty-eight years he has been teaching it. This alone could be sufficient motivation for the effort involved to create the course, but it is not clear that this is within the scope of the mission of the university. The second motivation, to which considerable discussion will be devoted in this chapter, is to “back-flip” the General Chemistry course at Rice. From the name “back-flip,” it is clear that this approach is related to, but not the same as, the use of video lectures as a component of the resources available to students enrolled in a regular course. The third motivation arises from work done at Rice for over a decade to provide professional development (PD) for high school Chemistry teachers. As 8 Sörensen; Online Course Development and the Effect on the On-Campus Classroom ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by CORNELL UNIV on September 27, 2016 | http://pubs.acs.org Publication Date (Web): August 16, 2016 | doi: 10.1021/bk-2016-1217.ch002

is well known, most such teachers are not trained in Chemistry (4), and they are thirsty for both greater content knowledge and innovative content pedagogy. A major challenge for teacher PD is access for the teachers to attend an on campus program. Given their busy schedules, most teachers can only pursue professional development during the summer. We have found that teachers are more likely to implement new information and new lesson plans when they are presented in “just in time” courses (5). Only teachers who live near enough to campus to attend evening courses have been able to benefit from this “just in time” approach. By extending the course offering to the online open format, teachers from anywhere can sharpen their content knowledge and their pedagogical skills. The most significant motivation and the one which is least obvious is the desire of Hutchinson to publish his teaching, in much the same way that Chemists routinely publish their research. By this term, we do not mean publishing articles about our teaching or publishing our lesson plans or laboratory exercises. We mean actually publishing the teaching itself, as will now be discussed.

Publishing as a Professional Activity Publishing scholarly scientific research is the major professional activity of most Chemistry faculty. Here is a possible list of motivations for faculty to publish their research findings: 1.

To share advances in the state of knowledge This is the core motivation. Research is about answering questions or solving problems, both of which provide understanding and progress for society. This is the basis for government, foundations, and corporations to fund professorial research. Though there may be financial, economic, or productivity gains that result from these advances, the advances themselves have inherent value which can only be realized when the information is shared openly and widely.

2.

To hold work up to the critical evaluation of peers Not all advances are equally valid or valuable. The agreed upon means of establishing validity and value is the peer review process. By subjecting research to the critique of our peers, researchers validate and deepen their own understanding. Furthermore, for the non-expert, the difference between authoritative and non-authoritative studies is hard to perceive. In Chemistry as in all science, all ideas are not equal. Only those methods, findings, applications, and conclusions that have been subjected to critical analysis and found to have validity have the inherent value discussed above. Publishing research is the means by which researchers receive this critical analysis.

9 Sörensen; Online Course Development and the Effect on the On-Campus Classroom ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

3.

To provide a foundation for others to build upon

Downloaded by CORNELL UNIV on September 27, 2016 | http://pubs.acs.org Publication Date (Web): August 16, 2016 | doi: 10.1021/bk-2016-1217.ch002

Science progresses as a cumulative effort. Lewis and Randall wrote that “Science has its cathedrals, built by the efforts of a few architects and of many workers” (6). Most research is the endeavor of building onto the scaffolding of the work which precedes it. Very little work appears de novo. Each scientist’s ideas are owed to the ideas of others, which can be found in the published literature. In turn, by publishing research, scientists provide ever higher scaffolding on which others may build, learn ideas, leverage this understanding, refine the ideas with their additional studies, and create new knowledge. 4.

Institutional prestige National and international reputations of universities are substantially based on the reputations of their research programs and their research faculty. Creation and dissemination of new knowledge, creative work, theory and design are core missions of universities, so the productivity and visibility of faculty in this work are key indicators of the success of a university in achieving these missions. Moreover, there are clear external markers of success in these efforts. Research grants, invitations for keynote lectures or seminars, H-factors for publication records, and induction in the professional academies are all accorded to those who are successful in publication of their research. Measuring these is therefore part of determining an institution’s ranking in any of the various systems which purport to compare universities, and in turn, publication of research enhances these measures.

5.

Personal prestige and professional advancement Given the prestige of research to the institution, it follows of course that the institution rewards those faculty who contribute to its prestige. As such, professional advancement within the institution is strongly tied to successful publication of research. Conversely, scholarly accomplishment without peer-reviewed publication carries no weight at all in consideration of advancement. This is in large part because, as stated above, that publication requires submitting ones scholarship to critical evaluation and validation. But it is also because, without publication, scholarship contributes nothing to the prestige of the individual or the institution. Advancement in the academic scientific community is based on publication of excellent research that advances the state of knowledge.

10 Sörensen; Online Course Development and the Effect on the On-Campus Classroom ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by CORNELL UNIV on September 27, 2016 | http://pubs.acs.org Publication Date (Web): August 16, 2016 | doi: 10.1021/bk-2016-1217.ch002

Publishing Teaching as a Professional Activity Consider then these motivations for publication in the context of teaching. On review, every one of them would apply if one were to publish teaching. Indeed, looking back at the quote from the National Science Board that opened this chapter, “great teaching is a form of scholarship” suggests teaching should be published. Why then has scholarship of research held a position of preeminence over the scholarship of teaching? The answer seems obvious and comes in the form of another question: how would we publish teaching? Alternatively, reflecting on the motivations above, how could these motivations be equally achieved in teaching as they are in research? Historically, the answer empirically has been that publishing teaching cannot be done. How many of us have ever observed the teaching of a colleague at a university other than our own? If called upon to do so, how many of us would be able to evaluate the quality of the content and pedagogy of a colleague’s teaching in the same way that we are called upon to evaluate a colleague’s research? Most people would not know where to begin because there would be minimal or no data on which to do such an evaluation. In assessing this claim, recall that we are not discussing publication of research on teaching and learning. While this is an extremely important area of scholarship, research on teaching and learning is not the same activity as teaching. Nor is publication of lesson plans, classroom demonstrations, laboratory activities, or classroom technology publication of teaching. So, how do we publish the teaching itself? In our view, the advent of the massive open online course provides an answer to these questions. Hence, the most significant motivation for offering the General Chemistry course in the MOOC format was to publish the teaching approach developed and implemented at Rice.

Constructivism in Chemistry The teaching approach discussed in this chapter is based on a constructivist learning model. The name of the model derives from the principle that students will best learn new material when they construct that knowledge in their own minds, guided by the instructor. Good and Brophy (7) explained this well: “Learning is a constructive process that involves “seeking after meaning,” so students routinely draw on prior knowledge as they attempt to make sense of what they are learning.” This is crucial in a scientific context, because scientific reasoning is inductive not deductive. Cooperstein and Kocevar-Weidinger (8) drew the tie together between constructivist learning and scientific reasoning: “Constructivist learning is inductive. Constructivist learning dictates that the concept follow the action rather than precede it. The activity leads to the concepts; the concepts do not lead to the activity.” There are two key ideas here. The first is that the development of a scientific model, concept, or theory is an inductive reasoning process, beginning with controlled experimental observations and leading to general principles. If students 11 Sörensen; Online Course Development and the Effect on the On-Campus Classroom ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by CORNELL UNIV on September 27, 2016 | http://pubs.acs.org Publication Date (Web): August 16, 2016 | doi: 10.1021/bk-2016-1217.ch002

are to learn scientific reasoning, they need to experience this inductive learning process. Teaching science as a deductive process of application of previously stated principles misrepresents the science process and scientific reasoning. The second key idea is that it is well established that most of our students are inductive learners, not deductive learners (9), and this is consistent with using teaching approaches which encourage constructivist learning. Teaching science to students via the traditional deductive method of presenting the concepts and testing them in applications is the reverse of the learning styles of our students. This contrast is illustrated in Figure 1, a diagram relating the inductive process to the deductive process in the scientific method. While examining Figure 1, consider how much time is spent in a standard introductory science course on the right “deductive” side of the diagram versus how much time is spent on the left “inductive” side of the diagram. More tellingly, consider a comparison of how much of the tested material comes from the deductive side of the diagram versus the inductive side of the diagram. Traditional science teaching is not always aligned with a more modern understanding of constructivist learning. A number of approaches have been developed that use inquiry to guide students in the construction of knowledge (for example references (11–13)). However, conventional ineffective methods are often deeply entrenched because these are the only methods that most teachers of science have ever observed.

Figure 1. Scientific process via inductive and deductive reasoning. (Reproduced with permission from ref. (10). Copyright 2104 ACS.)

Concept Development Studies in Chemistry How then do we implement a constructivist learning model in an Introductory or General Chemistry class? In such a course, “everything is already known,” so what is there to construct? Our answer is to construct the fundamental concepts of chemistry themselves, beginning with foundational experimental observations. 12 Sörensen; Online Course Development and the Effect on the On-Campus Classroom ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by CORNELL UNIV on September 27, 2016 | http://pubs.acs.org Publication Date (Web): August 16, 2016 | doi: 10.1021/bk-2016-1217.ch002

Each concept is constructed in a “Concept Development Study” organized with a structure illustrated in Figure 2. We begin each development by reviewing what we already know and then asking questions which need to be understood more fully. We then describe experiments and data which present clues to the answers to the questions. The clues are assembled via inductive reasoning into a preliminary model which accounts for the observations. This model presents additional questions that require additional experimental observation, which in turn leads to a refinement of the model. This process occurs iteratively until we have a satisfactory model or theory which accounts for the experimental observations and which thereby answers the questions posed. Note the process illustrated in Figure 2 mimics the process by which a scientific principle, model or theory is actually developed. In each study, we attempt in every instance possible to use the actual experimental data which were used historically in the development of the model or theory. For example, to demonstrate the existence of atoms, we do not use observations of atoms via scanning tunneling microscopy; rather we use the Law of Multiple Proportions ala Dalton.

Figure 2. Structure and development of a Concept Development Study. (Reproduced with permission from ref. (10). Copyright 2014 ACS.) There are twenty five Concept Development Studies in Chemistry, assembled into an online book which is freely available on Rice University’s OpenStax website (14). These span nearly all fundamental concepts typically introduced in a two semester General Chemistry course. In doing so, these studies provide the almost always missing inductive reasoning side of Figure 1. The collection of studies has been a work in progress since 1993. The method has been presented at conferences regularly since 1995 and was first published in the literature in 2000 (2). The entire collection has been available freely online since 2004 (15). The response to these presentations and publications and to the free availability of the materials has been unfailingly positive, most often enthusiastic. However, accompanying these positive responses has been a consistent theme of uncertainty about adoption. In the absence of observing an 13 Sörensen; Online Course Development and the Effect on the On-Campus Classroom ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

implementation of the Concept Development Studies (CDS) approach, instructors have been reluctant to adopt it. How does one actually implement this approach? It is a truism that the majority of teachers teach how they were taught. Having only taught the deductive reasoning side of Figure 1, how does one learn to teach the inductive reasoning side? Our answer is, by observing the approach being taught. Hence, we need to publish, not just information about our teaching, but the teaching itself.

Downloaded by CORNELL UNIV on September 27, 2016 | http://pubs.acs.org Publication Date (Web): August 16, 2016 | doi: 10.1021/bk-2016-1217.ch002

Response of Students to the CDS Approach Throughout the almost three decades of using and expanding the CDS approach in General Chemistry at Rice University, student responses have been consistent. At first students struggle with the challenge of being forced to go through the inductive reasoning necessary to develop chemical concepts from data. However, through practice and guidance, the majority of students see the value in understanding the scientific basis of chemistry rather than simply memorizing the process of “plug and chug” into formulas and quantitative problems. Over the years, students were probed via surveys for their most strongly held opinions on the CDS approach. Students were offered 12 statements of opinion, both positive and negative that show up in course reviews, and were asked to select their most, second most, and third most strongly held opinions. Figure 3 shows the statements that were selected as strongly held opinions by at least 30% of the students surveyed. While 62% of students do express frustration with the challenging concept-based tests, 56% selected they feel like they are understanding chemistry rather than simply memorizing. Over a third of the students, 35%, preferred the CDS portions of the course, so since this survey in 2007, the text has been expanded. And 30% of the students strongly held the opinion of enjoying the discovery aspect of the CDS approach. The statements that were not selected by students as strongly held opinions included “Just tell me what I need to know and I can do it” selected by 16% and “It takes an awfully long time to figure out how the Concept Development Studies approach works” chosen by 13% of students. Students were also probed for their responses to the CDS text, as shown in Figure 4. The overwhelming majority of students, 96%, agreed or strongly agreed that the Concept Development Studies in Chemistry enhanced their understanding of chemical concepts. Students also agreed the text improved retention (88%), interest in studying (64%), and ease in learning (82%) chemistry. These outstanding responses from students at Rice University motivated Hutchinson to expand the course to more students through the MOOC and to use the online course as an avenue to publish teaching via the CDS approach.

14 Sörensen; Online Course Development and the Effect on the On-Campus Classroom ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by CORNELL UNIV on September 27, 2016 | http://pubs.acs.org Publication Date (Web): August 16, 2016 | doi: 10.1021/bk-2016-1217.ch002

Figure 3. Percentage of students selecting statements as “Most/Second most/Third most strongly held opinion from Fall 2007 General Chemistry student survey, N=111.

Figure 4. Student responses to “Concept Development Studies in Chemistry” by Hutchinson enhanced my ____ of chemical concepts, N=111.

“Back-Flipping the Classroom”: Effect of the MOOC on the On-Campus Students The theme of this volume is about online courses in general and about the impact these have on the corresponding on-campus courses in particular. A common theme in this context is the concept of the so-called “flipped classroom.” Our view is that this is an odd term, the word “flipped” implies that the idea of students preparing for a course by reviewing material in advance is new. This is not at all new. Requiring students to watch an online video before coming to 15 Sörensen; Online Course Development and the Effect on the On-Campus Classroom ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by CORNELL UNIV on September 27, 2016 | http://pubs.acs.org Publication Date (Web): August 16, 2016 | doi: 10.1021/bk-2016-1217.ch002

class is not pedagogically different than requiring these same students to read a chapter or a module before coming to class. The latter approach is quite old, and has essentially universally been regarded as an integral part of what we now call active learning, e.g. Socratic dialog or classroom discussion. Students have been required to read Concept Development Studies for General Chemistry at Rice for more than two decades. The class time is then focused on discussion of the reading, not presentation by lecture. The “flipped classroom” is therefore a very old concept that has never required and still does not require video recorded lectures. We have taken a different approach to use of the online video lectures. These can be of immense value in providing summative reviews of the material discussed during an active learning class. The CDS approach facilitates classroom discussion, including a Socratic dialog. The video lectures provide a consolidated review of that discussion. As such, we provide our video prerecorded lectures to our students after the class, not before. Our students then use the videos to complement their notes or to rehash any concepts that were left unclear during the discussion. In addition, our students use the videos to review the material when studying in groups or before exams. In the most recent course survey, 58% of students self-report that they watch the videos regularly and find them useful for these purposes.

Response of Teachers to the CDS Approach Teachers have been introduced to the CDS approach through professional development (PD) courses at Rice University over the past seventeen years. High school chemistry teachers eagerly learn about the CDS approach and constructivism. Initially, practicing teachers do not have a strong understanding of constructivism, with over 60% of teachers having little to no familiarity. However after the PD course, the majority of teachers report knowing much more about constructivism, as shown in Figure 5. The teachers become students as the CDS approach is modeled for them. Many have limited backgrounds in chemistry, with 29% nationally having a degree in chemistry (4) and 30% from the Rice PD program. Many teachers are very excited to be students again and expand their content knowledge along with pedagogical skill. Teachers report via surveys and interviews that they understand the science better themselves and feel more confident in their abilities to teach chemistry. A recent article illustrates exactly how the CDS approach can be implemented within the College Board’s updated Advanced Placement Chemistry Curriculum Framework (16). Participants from the PD programs have provided insight on the implementation of the CDS approach at the high school level. Of 90 teachers surveyed after participation in PD, 95% agreed that the Concept Development Studies in Chemistry textbook (15) deepened their understanding of chemistry. Data showing teachers’ self-reported use of the CDS approach is shown in Figure 6. More than 90% of the teachers had at least attempted to use the CDS approach, with almost 70% of teachers trying it more than once. The main barriers faced 16 Sörensen; Online Course Development and the Effect on the On-Campus Classroom ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Downloaded by CORNELL UNIV on September 27, 2016 | http://pubs.acs.org Publication Date (Web): August 16, 2016 | doi: 10.1021/bk-2016-1217.ch002

by teachers who were unable to implement the CDS approach were impediments from the hierarchy of schools, lack of pedagogical knowledge, and lack of content knowledge, the last of which could be remedied through the MOOC.

Figure 5. Teacher responses to: “Which of the following best describes how familiar you are with Constructivism as a teaching/learning model?”, pre N=84, post N=73.

Figure 6. Response from 90 teachers to “I have used the Concept Development Studies approach in my class.” Teachers in the PD program were regularly requesting the ability to attend the General Chemistry course at Rice University, which is inhibited due to the timing of the class being during the school day. Instead, the MOOC allows teachers full access to the entire two-semester course. The MOOC is a significant resource that allows teachers to enhance their content knowledge or remedy knowledge gaps. 17 Sörensen; Online Course Development and the Effect on the On-Campus Classroom ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

Response of Students to the MOOC The online course proved to be popular, though whether it would qualify as “massive” is a matter of opinion. Table 1 shows data for the number of people who registered for the class and the various levels at which these people participated. If one were to take the end of the second week of classes as a baseline for genuine initial participation and take watching all lectures as a measure of completion, then the retention rate for the course is roughly 40%.

Table 1. Participation of Students in MOOC Downloaded by CORNELL UNIV on September 27, 2016 | http://pubs.acs.org Publication Date (Web): August 16, 2016 | doi: 10.1021/bk-2016-1217.ch002

Participation

Number of Students (rounded) 16,000

Registered Watched at least one lecture

6900

Watched two weeks of lectures

2850

Watched all lectures

1200

Took at least one quiz

3000

Took all quizzes

550

Received a certificate of accomplishment

380

The course was clearly very well received by these students, as 97% of survey respondents agree that they would recommend the class to others, with half of those strongly agreeing. Other items of note about the students in the online course include that the students were not novices in Chemistry. Nearly half of the students had taken at least one year of college level chemistry prior to taking the course, including 13% one year, 11% two years, and 24% more than two years. However, most of the students had not taken Chemistry in quite some time, including 15% who had not taken Chemistry in 6-10 years and 48% who had not taken Chemistry in more than 10 years. Only 17% of the students were high school or college age. By contrast, 42% were age 40 or older. Students were enrolled from every continent except Antarctica. Of course, it is quite different to hear individual stories from these remote students whom we have never met. One student responded to the CDS approach via the online survey: I just finished this course and can’t say enough great things about it. 17 years ago I gave up on Chemistry in college because a TA told me “I just have to know” the material. The CDS method proved that guy wrong. You don’t know how happy it made me feel to suddenly understand concepts that seemed so foreign to me. Thank you for giving me that excitement. 18 Sörensen; Online Course Development and the Effect on the On-Campus Classroom ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

A second student responded: Unlike other courses, this one really helps the student make a deeper understanding of chemistry and the concepts that built the science!

Downloaded by CORNELL UNIV on September 27, 2016 | http://pubs.acs.org Publication Date (Web): August 16, 2016 | doi: 10.1021/bk-2016-1217.ch002

Finally, here is an excerpt from an unsolicited email sent by another student: Your teaching style enabled me for the first time in my life to begin to understand ideas that I never thought I would, and to get the confidence to tackle a whole range of subjects in the sciences. Not only did it open up my intellectual world by giving me a far deeper understanding of our physical world, it was also the trigger that gave me the confidence to apply for medical school - no mean feat for an Arts graduate who turned pale at the very thought of doing chemistry!

Conclusions What conclusions can be drawn from the approaches and experiences described here? We have argued that a significant, if not the most significant, impact of creating an online course is the publication of teaching. Such publication reaches an audience of both student learners and fellow teachers, each of whom have different things to gain from observing and participating in the online course. The teachers observing the course can adopt or adapt the approaches, can criticize or comment on the approaches with useful feedback, or can build on and improve the approaches with their own innovations. Online courses as a form of publishing teaching therefore open a new capacity to advance innovation in the classroom and beyond. Hutchinson’s current course, General Chemistry Concept Development and Application, serves as one such example. Via this course, we have disseminated the approach developed at Rice, the Concept Development Study approach, based on constructivism. The take-away messages are that, empirically, students learn more effectively via constructivism and that, practically, a primary learning objective of a General Chemistry course should be to teach students how to reason scientifically. This ability empowers students to examine scientific data objectively and to critically analyze conclusions drawn from that data, or even to gather their own observations and data from which to develop new scientific models, concepts, and theories. We close by quoting Sir Harold Kroto, Nobel Laureate in Chemistry, from a recent lecture (17): “I think the most important thing that young people should be taught at school is how they can decide what they’re being told is true.”

19 Sörensen; Online Course Development and the Effect on the On-Campus Classroom ACS Symposium Series; American Chemical Society: Washington, DC, 2016.

References 1. 2.

3.

Downloaded by CORNELL UNIV on September 27, 2016 | http://pubs.acs.org Publication Date (Web): August 16, 2016 | doi: 10.1021/bk-2016-1217.ch002

4.

5. 6. 7. 8.

9. 10.

11. 12. 13.

14.

15.

16. 17.

U.S. Science and Engineering in a Changing World; National Science Board, National Science Foundation: Arlington, VA, 1996. Hutchinson, J. S. Teaching Introductory Chemistry using Concept Development Case Studies: Interactive and Inductive Learning. U. Chem. Educ. 2000, 4, 3–8. Daniel, J. Making Sense of MOOCs: Musings of Myth, Paradox and Possibility. J. Interactive Media Educ. 2012, 3, Art. 18. Smith, P. S. 2012 National Survey of Science and Mathematics Education: Status of High School Chemistry; Horizon Research, Inc.: Chapel Hill, NC, 2013. Feist, L. Removing Barriers to Professional Development. Tech. Horizons Educ. J. 2003, 30, 30–34. Lewis, G. N.; Randall, M. Thermodynamics and the Free Energy of Chemical Substances; McGraw-Hill: New York, 1923. Good, T. L.; Brophy, J. E. Looking in Classrooms, 6th ed.; HarperCollins College: New York, 1994; p 417. Cooperstein, S. E.; Kocevar-Weidinger, E. Beyond Active Learning: A Constructivist Approach to Learning. Reference Services Rev. 2004, 32, 141–148. Felder, R. M.; Silverman, L. K. Learning and Teaching Styles in Engineering Education. Eng. Educ. 1988, 78, 674–681. Nichol, C. A.; Szymczyk, A. J.; Hutchinson, J. S. Data First: Building Scientific Reasoning in AP Chemistry via the Concept Development Study Approach. J. Chem. Educ. 2014, 91, 1318–1325. Farrell, J. J.; Moog, R. S.; Spencer, J. N. A Guided Inquiry General Chemistry Course. J. Chem. Educ. 1999, 76, 570–574. Lewis, S. E.; Lewis, J. E. Departing from Lectures: An Evaluation of a PeerLed Guided Inquiry Alternative. J. Chem. Educ. 2005, 82, 135–139. Sampson, V.; Walker, J. P. Argument-Driven Inquiry as a Way To Help Undergraduate Students Write to Learn by Learning to Write in Chemistry. Int. J. Sci. Educ. 2012, 34, 1443–1485. Hutchinson, J. S. Concept Development Studies in Chemistry 2012. OpenStax CNX. http://cnx.org/contents/[email protected] (accessed July 15, 2015). Hutchinson, J. S. Concept Development Studies in Chemistry. OpenStax CNX. http://cnx.org/contents/[email protected] (accessed October 7, 2013). AP Chemistry: Curriculum Framework 2013−2014; The College Board: New York, 2011. Palca, J. A Discoverer of the Buckyball Offers Tips on Winning a Nobel Prize, 2015. NPR. http://www.npr.org/2015/10/08/445339243/a-discovererof-the-buckyball-offers-tips-on-winning-a-nobel-prize (accessed October 22, 2015).

20 Sörensen; Online Course Development and the Effect on the On-Campus Classroom ACS Symposium Series; American Chemical Society: Washington, DC, 2016.