<<

Exploring Student Thinking About Addition Reactions Solaire A. Finkenstaedt-Quinn*, Field M. Watts, Michael N. Petterson, Sabrina R. Archer, Emma P. Snyder-White, and Ginger V. Shultz

Department of Chemistry, University of Michigan, Ann Arbor, Michigan 48109, United States

5 ABSTRACT is a required course sequence for many STEM students, however research indicates

that organic chemistry reaction mechanisms are especially challenging for students due to a mixture

of underlying conceptual difficulties, the process-oriented thinking inherent to the discipline, and the

representations commonly used to depict mechanisms. While student reasoning about many of the

10 reaction types covered in the organic chemistry curriculum have been studied previously, there is

minimal research focused specifically on how students think about the mechanisms of addition

reactions. This study addresses that gap by probing first semester organic chemistry students’

thinking using think-aloud interviews as they worked through two different addition reactions. To

elicit a range of thinking, students worked through the mechanisms using either paper and pencil or

15 an app that dynamically represents the molecules. Overall, students were able to identify the steps of

the two addition reactions but did not always successfully apply chemical thinking during the

mechanistic steps. Specifically, both groups of students struggled with the concepts related to

carbocation stability, frequently misapplying stabilization via substitution and demonstrating difficulty

in identifying the potential for resonance stabilization. Our results suggest that instructors should

20 emphasize the conceptual grounding directing mechanistic steps, in particular when determining

carbocation stability.

GRAPHICAL ABSTRACT

KEYWORDS 25 Second-year undergraduate, chemical education research, organic chemistry, computer-based learning, problem solving/decision making, addition reactions

INTRODUCTION The organic chemistry sequence is a two-semester prerequisite for numerous fields of study with

the aim of teaching students the fundamentals of organic chemistry reactions, developing students’

30 ability to reason through reaction mechanisms, and strengthening students’ understanding of

structure-property relationships.1 Unlike in general chemistry, where reactions are presented simply

as reactants and products separated by an arrow, organic chemistry emphasizes reaction

mechanisms, which require students to envision electron flow, denoted by the curved-arrow

formalism, from reactant through intermediate molecules to the desired product.2, 3 However, reaction

35 mechanisms have been identified as one of the most difficult concepts covered in organic chemistry,1

as students must engage in multivariate thinking, weighing considerations of chemical reactivity and

spatial constraints while also interpreting numerous representations.4 Thus, it is important to

characterize how students reason through reaction mechanisms in order to better support their

learning.

40 Prior research focused on how students approach and think about organic chemistry reaction

mechanisms indicates that students use mechanistic steps primarily to move from the reactant to the

product.2, 5-8 Additionally, students have a tendency to focus on the explicit features of the molecules

involved in the reaction.7-10 However, research focused specifically on student reasoning about

addition reactions is minimal.5, 9 Addition reactions are often introduced early in the organic chemistry

45 curriculum and require students to begin applying new content knowledge as they weigh multiple

reaction pathways, utilizing reasoning processes that they will engage in throughout the course. As

such, investigating student thinking on this topic has the potential to inform our understanding of

how students may approach more difficult reactions that similarly require consideration of multiple

reaction pathways. This study explores how students reason about addition reactions during think-

50 aloud interviews where students work through reaction mechanisms on paper or through an app.

Mechanistic Reasoning An understanding of reaction mechanisms is a key component of the organic chemistry curriculum

and requires students to integrate several sources of knowledge including knowledge of chemical

reactivity and mechanistic representations.3-4 With its importance and complexity, a number of studies

55 have been conducted that investigate the processes students use when approaching mechanisms.2, 5-15

These studies indicate that students are not successfully applying the conceptual knowledge they have

developed throughout general and organic chemistry and focus on explicit structural features,

neglecting implicit, electronic features when making mechanistic decisions.2, 7-10, 15 For example, in two

studies using card sort tasks, students tended to sort organic reactions based on two main criteria:

60 similarities in the type of reaction, without any discussion of the underlying mechanistic patterns, and

by surface features, such as the presence of similar functional groups.9, 15 Additionally, students have

been found to rely on memorized mechanisms that lead them to the desired products.2, 5-8, 14-15

These difficulties may be due to underdeveloped knowledge or a poor ability to apply the knowledge

of electronics and electron flow during organic chemistry reactions,2, 6, 10 which is indispensable when

65 solving unfamiliar mechanisms. To facilitate thinking about the electronic properties and relationships

between atoms and molecules, the ability to draw and manipulate Lewis structures is considered key,

yet students are known to have difficulties with this skill.16 Additionally, students are not adept in

their use of the curved-arrow formalism,2, 5-6 where they do not always use the formalism when it

would be useful to predict reaction products,5 and may follow known reaction steps rather than

70 thinking chemically about the process.6 Additionally, mechanistic arrows have little meaning for

students and are not recognized as representing electrons.2, 6 This may be indicative of a profound

learning gap—that students do not fully understand the usefulness of mechanistic arrows or what

these arrows depict.2, 5-6 Lacking an appreciation for curved arrows may promote reliance on

memorization and is especially detrimental as students who use the curved-arrow formalism have

75 been shown to perform better on more difficult problems than those who do not.17

Addition Reactions Pi bond addition reactions are introduced early in the organic chemistry curriculum and start to

extend students’ reasoning to incorporate content beyond what they have learned in general

chemistry, e.g. decisions about acidity and basicity, yet there is little research focused on how

80 students approach and reason through this type of reaction specificially.5, 9 Grove et al. provided a

detailed analysis of students’ drawn responses to an acid-catalyzed addition reaction.5 The percentage

of students who provided a mechanism for the reaction, instead of just drawing the product, increased

from 56 to 70% over the course of two semesters in organic chemistry. When students did show the

mechanism, the diversity of mechanistic pathways depicted, both correct and incorrect, increased over

85 time and only a small percent drew a carbocation intermediate.5 Graulich and Bhattacharyya used

variations of a card-sorting task targeting reactions and found that students

focused primarily on surface characteristics.9 When students were directed to consider mechanisms,

they created categories based on characteristics of the reactions—e.g. reaction intermediates or

steps—but did not always make correct assignments. Throughout the tasks, the most commonly

90 identified mechanistic step was carbocation formation.9 Beyond these two studies, additional research

focused on the concepts students use as they reason through addition reactions is merited.

Analyzing student reasoning about addition reactions is especially important as the reaction

mechanisms begin to be more complex than what students may have seen previously. When reasoning

through addition reactions, students must consider multiple reaction pathways—e.g., which carbon of

95 the double bond to protonate—and begin weighing multiple chemical considerations—e.g. substitution

versus resonance stabilization. Students are known to struggle with these,6, 8 yet considering multiple

factors is especially important in more complex reactions, such as substitution and elimination. This

study sought to develop descriptions of students’ thinking about addition reactions by identifying what

chemical features and concepts students focused on as they worked through two addition reactions.

100 Student reasoning was elicited through think-aloud interviews that utilized two different modalities,

conventional paper-pencil and a touch-screen app. It is important to use multiple modalities to probe

how students reason, as there are an increasing number of interfaces with which students can

interact with mechanistic representations. Additionally, previous research indicates that multiple

modalities can elicit a greater range of reasoning from students.18 This research was guided by the

105 following research questions:

1. How do undergraduate students in organic chemistry think about addition reactions when working through mechanisms on paper or using a touch-screen application? 2. When considering addition reaction mechanisms, what components of mental models do students use to guide their reasoning?

110 THEORETICAL FRAMEWORK This study is grounded in the models and modeling framework as described by Briggs for use in

chemistry education.19-20 Models are defined as mental representations of processes or systems that

allow for scientific reasoning. The development of models requires experiences with related phenomena

and models are used to make sense of the structure or behavior of a system. This framework aligns

115 with the understanding of mechanisms in organic chemistry, as students are exposed to mechanisms

throughout the organic curriculum and encouraged to use mechanisms as a predictive or explanatory

tool for reaction outcomes. Students gather experience from mechanisms learned in various situations

in their coursework (e.g., lecture, homework, studying, etc.) and have the opportunity to develop

mental models of mechanisms, which they can then apply to make sense of the behavior of unfamiliar

120 reaction mechanisms using fundamental chemical reasoning.

The models and modeling framework theorizes models to be made up of five components: (1)

referents, (2) relationships, (3) rules/syntax, (4) results, and (5) operations.21 The first four of these are

static ways to make sense of system components that build upon one another. Referents are the

symbols used to represent the components of a system, and they have a certain set of relationships to

125 other referents. Rules, or concepts, are the meaningful ways in which referents relate to one another

and syntax is how the rules may be utilized in a specific context. Results are the changed referents

after operations, which are the actions performed by the referents. Essentially, referents have specified

relationships and interact with one another through operations to produce results, all constrained by

rules and syntax.

130 In the context of addition reactions, referents are the atoms and molecules involved—e.g., a

molecule of hydronium and an . These molecules are related to each other in that hydronium is

an acid and the alkene is nucleophilic. The rules and syntax of acid-base chemistry dictate the first

operation of the reaction between the two molecules: that the acid will protonate the alkene double

bond. However, a decision must be made to determine which carbon of the double bond will be

135 protonated. This decision is guided by additional rules for alkene addition reactions: namely, that

carbocation stability as determined by degree of substitution or, when applicable, resonance, will

influence which carbon is protonated. Depending on this decision, the result will be a molecule of

water and the protonated alkene with the most stabilized of the two possible carbocation

intermediates. This result now has a new set of properties—i.e., the carbocation is positively charged

140 and electrophilic—and a new relationship to the nucleophilic water molecule in the system. The rules

of how nucleophiles and interact can then be used to make sense of the next step, or

operation, of the reaction to produce the next result.

When working through mechanisms, students may use different representations

to model the mechanistic process. Most commonly, this takes the form of drawn Lewis or line-angle

145 structures to represent molecules and curved-arrows to represent the movement of electrons during

operations. Students can extract knowledge about the relationships between or within molecules from

the molecular representations of referents, and can use this knowledge when considering the pertinent

rules and syntax to predict a mechanistic step. Increasingly, there also exist applications which can

also support students working through organic structures and mechanisms.22-24 The app used in this

150 study, “Mechanisms”, may provide students with additional sources of knowledge about the progress

of the reaction. This may take the form of an operation depicted by a task card to students at the

beginning of a puzzle or a hint in the app directing students to consider how the concept of

carbocation stabilization through resonance applies syntactically to inform the operation they perform.

Students’ ability to use and extract information from models during the mechanism may impact their

155 success with mechanistic reasoning.

METHODS Context and Participant Recruitment This study was part of a larger effort that targeted student reasoning about acid-base and addition

reactions.18 Herein we present results pertaining to addition reactions. The study was conducted at a

160 large, research-intensive university in the Midwestern United States. Approval for data collection and

analysis was obtained from the Institutional Review Board (HUM00156602). Students were recruited

across three semesters from the first of a two-class introduction to organic chemistry shortly after

covering addition reactions in the course. The course content covered prior to the interviews included

a review of pertinent general chemistry content, the curved-arrow formalism, resonance, acid-base

165 reaction mechanisms, and strong and weak acid addition. In the first semester of data collection,

purposeful sampling was used to invite students to participate based on their performance on the first

exam.25 In the second and third semester, convenience sampling was used to recruit students via

course announcement to increase participation in the study.25 During recruitment students were told

that following the interview they could ask the interviewer any organic chemistry content questions,

170 and no additional incentives were provided. In total, thirteen students participated across the three

semesters and consented to participate in the study at the beginning of the interviews. Students were

divided into two groups, where six of the students worked through the problems using the

conventional paper-pencil style of problem solving and the other seven solved problems using the

“Mechanisms” app, identified herein as paper-pencil and app students respectively (pseudonyms are

175 given in Table 1).

Table 1. Student participants by think-aloud interview group type Reaction Participants modality groups Paper-pencil Ana, Aurora, Daisy, Francis, Mary, Perdita App Angela, Belle, Flynn, Jasmine, Pepper, Peter, Tiana

Description of app software The “Mechanisms” app contains a wide range of puzzles where students can work through organic

reaction mechanisms on a touch-screen device.23, 26 At the beginning of a puzzle, students are

180 presented with a task card that depicts the initial reactants, intermediate, or a step in the reaction

mechanism. Students can manipulate and move the molecules using the touch-screen, reveal or hide

implicit hydrogens and lone pairs by tapping on atoms, and break and form bonds by moving bonds or

lone pairs from one atom to another. The app prevents students from making incorrect chemical

moves and when students form a bond that then requires breaking a bond the app will only let them

185 perform bond-breaking moves. Additionally, the app provides information about the reaction in the

form of goals students can access that describe the steps needed to complete the puzzle and hints that

appear in certain cases, such as if they have formed a minor product. The goals for the pertinent

reactions are given in Figure S1.

Problem Selection 190 To select problems for the study, the research team chose five reactions from the app reflective of

those presented in first-semester organic chemistry. We then performed cognitive interviews with three

organic chemistry instructors to validate and select two of the reactions. Of the instructors

interviewed, one was the instructor during the first semester of data collection and two had taught the

material previously. They all agreed that the selected problems mirrored the difficulty level expected of

195 students and provided feedback on the presentation of the reactions to the paper-pencil students.

The paper-pencil students were presented with the line-angle representation of the reactant

molecule along with the molecular formula of the product, with reagents depicted above the reaction

arrow. The reactions were presented this way to mirror the information given to the app students at

the beginning of the reaction. The molecular formula of the product was given to provide students with

200 an end-point to parallel the completion goals provided to the app students. The reactions as seen by

both groups of students are presented in Figures 1 and 2.

Think-aloud interviews Interviews were conducted in the think-aloud style where students verbalized their reasoning as

they worked through the reactions.27 Prior to each interview, students were given a practice think-

205 aloud to familiarize them with verbalizing their thinking. For both groups of students, we would ask

probing questions during the interviews to further elicit participants’ reasoning, such as why they

were making certain moves or what they were thinking about between actions. Examples of these

questions are, “Why did you choose to protonate that carbon?” and “What are you thinking about

now?” Both groups of students were provided with the pKa table used in the course. The order of the

210 reactions was varied for each participant.

To match the format of the app, paper-pencil students were asked two general questions at the

completion of each problem: “Are there any resonance structures for your product?” and “Why do you

think you are done?” The think-aloud interviews for the paper-pencil students were performed using a

LivescribeÔ pen and notebook to concurrently capture students’ drawing and verbalizations. The app

215 students went through a modified tutorial in the “Mechanisms” app to minimize any difficulties arising

from working through the reactions in a new modality. The tutorial showed students how to reveal

implicit hydrogens or lone pairs, how to make or break bonds, and how to manipulate the molecules.

Audiovisual recordings were captured for the app-students, so that we could identify the moves they

made as they worked through the reactions in the app as well as their verbalized thought processes.

220 Development and application of coding scheme After the interviews, the coding scheme was developed using open-coding and constant

comparative analysis.28 The research team read the transcripts individually, noting recurring themes.

The team met to discuss findings and began to develop a coding dictionary based on the features

identified by each member of the team. The coding dictionary is given in Table S2. To establish what

225 sections of the transcript should be coded, all transcripts were divided into units of analysis

corresponding to different stages in the mechanism: planning stages where students vocalized ideas

about electron movements and chemical reactivity, and operation stages where the student

manipulated bonds/electrons to advance the mechanism. Two members of the research team met and

agreed on the units before coding. Following this, two researchers independently coded the interviews

230 of both reactions for four participants (30% of the data), met to clarify code definitions and reached

consensus on application of the coding scheme. The same two researchers then coded the remaining

70% (nine participants). The fuzzy kappa for this data was initially 0.78, indicating substantial

agreement between the two coders.29-30 Furthermore, the two researchers discussed discrepancies in

the codes applied to each transcript and reached consensus agreement for the final application of the

235 codes.

RESULTS AND DISCUSSION In this study, students worked through two addition reaction mechanisms using either paper-

pencil or the app interface. As students depicted the mechanisms, they were prompted to describe

what they were thinking and doing in a think-aloud interview. The interviews were qualitatively

240 analyzed through inductive coding and then interpreted through the lens of the models and modeling

framework.19-20 The results are presented below, and common problematic reasoning students

exhibited across the levels of the models and modeling framework is presented in Table 2.

Table 2. Common student difficulties across modalities Problematic student Problem Level(s) in the models and thinking modeling framework Overgeneralization of Resonance-stabilized Rules substitution addition Focusing on resonance Resonance-stabilized Relationship in the reactant rather addition than the potential product Mixture of resonance Resonance-stabilized Relationship structures addition Inverting substitution Resonance-stabilized Syntax stabilization rule addition and acid- catalyzed addition Ignoring octet rule Resonance-stabilized Rules, relationship addition and acid- catalyzed addition Identifying acid- Acid-catalyzed Relationship, rules and syntax catalyst, but not addition knowing how to apply it

Acid-Catalyzed Addition

Figure 1. Reaction schemes for the acid-catalyzed addition reaction as presented during the think-aloud interviews to A) paper-pencil students and B) app students in the task card prior to beginning the reaction. 245

The acid-catalyzed addition reaction was a two-step addition of water to a carbon-carbon double

bond (Figure 1). For this reaction, students needed to recognize that the conjugate acid, hydronium,

was given to them as a pre-formed referent. They additionally needed to identify the most stable site

for carbocation formation by considering the rules and syntax related to carbocation stabilization due

250 to degree of substitution of the two double-bonded carbons. Following these decisions, students could

perform the associated operations to first form a carbocation intermediate and then form the final

hydroxyl product.

The students either explicitly identified the reaction type as acid-catalyzed before proceeding

through the mechanism (four of the app students and five of the paper-pencil students) or at some

255 point while working through the problem (three app students and one paper-pencil student).

Specifically, students recognized the hydronium as a key reactant, or referent, in the reaction. Belle

said:

I know that this reaction probably wants me to use the H3O+ and then with the . I

see a double bond, so it probably wants me to do something with that, and then regenerate H3O+ 260 at the end, probably.

The ways students referred to the hydronium indicate that they used it as a cue for which steps they

needed to perform during the , similar to findings by Bhattacharyya and Bodner

focused on how chemistry graduate students approached reaction mechanisms.6

Some students, Peter, Pepper, Jasmine, and Perdita, did not initially vocalize their recognition of

265 the hydronium referent. Of these students, it was unclear whether Peter and Pepper only vocalized

their recognition later or if they did not recognize the reaction type until they had regenerated the

catalyst. Jasmine, an app student, did not recognize the role of hydronium at first, determining that

the hydronium would be deprotonated based on the task card for the reaction, and then remembering

and justifying the move by identifying hydronium as a strong acid. This exemplifies how students’

270 knowledge can be prompted by engaging with the reaction through the app. Perdita, a paper-pencil

student, only identified the reaction as an acid-catalyzed addition after they had drawn out the full

reaction mechanism whereupon, looking over the operations they had performed, Perdita recognized

the steps as an acid-catalyzed addition and used that knowledge to confirm the mechanism they had

drawn. Perdita said:

275 I'm pretty sure for this problem there should be ... I think the first step is a proton transfer. And then there would be a complexation, and then there would be clean-up. And I guess this is what I'm thinking just to check that this is kind of what was supposed to happen. I don't know, this is what I did, but I think that's what's supposed to happen in this, so now I'm just looking to see if that's what did happen.

280 It is promising that Perdita was able to reason their way through the mechanism without recognizing

the reaction type, only comparing what they had done to known steps afterwards. However, we only

observed the technique of retrospectively confirming the proposed reaction mechanism by aligning it

with known reaction steps once during all of the think-aloud interviews. Instructors could potentially

facilitate such metacognitive strategies to help students review their mechanisms for feasibility.

285 Some students struggled to use hydronium in their depicted mechanisms despite recognizing it as

a referent in the reaction. Four of the students who identified hydronium at the beginning of the

reaction did not use it appropriately. Flynn and Aurora both exhibited difficulties with the rules and

syntax of the hydronium, despite identifying it at the start of the reaction. Flynn focused on the water

molecule and tried to use the delocalized electrons in the double bond to form a bond with the oxygen

290 in the water molecule, which the app did not allow, before realizing the hydronium was available to be

deprotonated. Aurora, a paper-pencil student, determined that water was added to the reactant during

the course of the reaction by comparing reactant and product formulas, analogous to mapping

strategies described in the literature.2, 13, 31-32 This led Aurora to deprotonate the water as opposed to

the hydronium, using the delocalized electrons in the double bond. Aurora then formed a bond

295 between the resultant hydroxide and carbocation. Hence, Aurora did not use the hydronium during

the reaction and depicted an unfavorable acid-base reaction that formed hydroxide and a carbocation.

Angela and Daisy, on the other hand, had difficulties with the relationship of the hydronium to the

other reactant species. Neither student recognized that the hydronium was already present in its

catalyzed form and both started by trying to protonate it with the available water molecule. This

300 indicates that while they both recognized the reaction type, neither student recognized that the

conjugate acid had already been formed. Thus, they either lacked or did not apply the chemical

intuition to identify the electrophilicity of the hydronium. This could also be due to the instructional

context in which our study took place, where instructors often show forming hydronium as the first

step in this type of reaction before protonating the alkene. Hence, these students might have felt the

305 need to show a step before protonating the alkene even though the hydronium had already been

formed. In Angela’s case, the app did not allow the chemically inaccurate move so they instead used

the double bond to deprotonate the hydronium. Daisy, a paper-pencil student, drew the result of this

step as an uncharged H4O molecule and continued to show a poor understanding of chemical

reactivity and Lewis structures as they continued the mechanism. While Daisy correctly deprotonated

310 the H4O ‘catalyst’ with the delocalized electrons in the carbon-carbon double bond, they then used the

lone pairs on the oxygen in the reformed hydronium to form a bond with the carbocation intermediate,

despite the positive charge on the oxygen. Following this operation, Daisy drew the product structure

without any charges and determined they had completed the reaction. Angela and Daisy’s attempts to

protonate the already positively-charged acidic oxygen, in addition to the unfeasible chemical

315 structures that Daisy continued to form and use, exemplifies previously characterized difficulties that

students have drawing and applying Lewis structures to the more complex molecules in an organic

chemistry context.8, 16

The last step of this mechanism required students to reform the acid catalyst in the process of

forming the hydroxyl group on the product. While all but two of the students did reform the catalyst as

320 part of their reaction mechanisms, most did not discuss the underlying chemistry. Belle, Pepper,

Flynn, Ana, and Perdita all talked about knowing that the last step of the reaction was to regenerate

the acid, but did not specify why this was the case. Jasmine and Peter were focused on forming the

product given as the target on the app goal card (Figure S1A) and regenerating the catalyst

seemed to be a byproduct of that step rather than a recognized part of the reaction. Students’ focus on

325 forming the product when drawing mechanisms has been documented frequently in the literature.2, 4, 6-

8 Aurora and Daisy, both paper-pencil students, did not reform the catalyst. For Aurora this may have

been because they did not use the catalyst during the reaction and so it did not need to be

regenerated. Additionally, Aurora’s drawn mechanism did not result in any unstable products. While

Daisy had recognized the catalyst at the start of the reaction and mis-applied a memorized step to

330 form it, they did not consider regenerating the catalyst at the end of the mechanism. For both of these

students, the lack of explicit cues indicating that another mechanistic step was required in the

referents they had drawn may have hindered their chemical thinking about the final step of this

reaction.

Three students verbalized chemical thinking about the last step of their mechanisms. Tiana knew

335 there was a final mechanistic step and remembered that it involved protonation. Tiana first tried

protonating the water they had added to the carbocation and then, when the app did not allow this,

protonated the free water instead. While Tiana initially showed a lack of chemical thinking, they

subsequently used charge to explain why their original move was incorrect. In juxtaposition, Mary and

Francis were the only students who provided chemical reasoning about the last step of the reaction

340 prior to or as they were performing the mechanistic operations. Mary focused on using the free water

to deprotonate and remove the positive charge on the oxygen as the motivation for their last step,

whereas Francis discussed removing the positive charge on the oxygen and regenerating the catalyst.

Overall, it appears that students are able to recognize the reaction as containing an acid catalyst, but

they have some difficulties with the relationships, rules, and syntax affiliated with acid catalysts that

345 led to incorrect operations during their mechanisms. This may be indicative of a surface-level

understanding of the role the catalyst plays in the reaction, as well as the chemistry dictating the role,

as characterized in the literature.33

The other key element of this reaction was for students to identify the correct site for carbocation

formation following protonation at the double bond. While many students (eight out of thirteen)

350 correctly identified that they should use substitution to determine the most stable cation intermediate,

they did not discuss why carbocation stability increases as the number of non-hydrogen substituents

increases. This is exemplified by Belle who states that tertiary carbocations are more stable than

secondary but, following an interviewer question about why this is the case, could not explain the

phenomenon:

355 A tertiary is attached to three carbons and yeah. I don't know. I just know that it is. Despite correctly identifing substitution as the mode of stabilization for the carbocation intermediate,

Aurora and Perdita did not describe substitution correctly. Aurora inverted the syntax associated with

the concept of carbocation stabilization via substitution, which led to their performing an incorrect

operation:

360 ...so then this carbon is attached to three other carbons, and this one is only attached to two. And sp is the most stable, so that means this one [attached to two carbons is] more stable, this one [attached to three carbons is] less stable... These results indicate that students lack an understanding of the chemical principles that relate

degree of substitution to carbocation stability. Students similarly struggled with carbocation stability

365 in the resonance-stabilized addition reaction, as discussed below.

Five students did not initially consider substitution to determine carbocation stability. Jasmine

and Pepper selected the site for protonation based on the app task card (Figure 1B) and Daisy and

Tiana did not verbalize reasoning behind their protonation decision. However, following prompting by

the interviewer, both Tiana and Pepper talked about using substitution to determine carbocation

370 stability. Lastly, Ana used a steric hindrance argument to justify their decision, but focused on the

sterics in the final product rather than during the protonation step. Ana said:

Where would an H fit better? I'm just gonna say the H fits better here because this is a big [five membered ring]. Ana uses similar reasoning in the resonance-stabilized reaction, which is discussed further below.

375 Think-aloud interviews about this reaction indicated that even when students were familiar with

the reaction type and referents, they did not necessarily know how to correctly use the referents or

apply the related steps. While most students recognized the hydronium, they did not always use it

correctly and may have had a rote understanding of the role it played in the reaction. Additionally,

most students correctly used substitution to determine carbocation stability, but did not exhibit an

380 understanding of the underlying chemistry. While there is merit in students’ use of rules, such as

knowing to focusing on substitution when determining the most stable carbocation in this reaction, it

can lead to syntactic difficulties and improper operations when they do not understand the chemical

principles from which they derive.

Resonance-Stabilized Addition

Figure 2. Reaction schemes for the resonance-stabilized addition reaction as presented during the think-aloud interviews to A) paper-pencil students and B) app students in the task card prior to beginning the reaction. 385

The resonance-stabilized addition reaction involved students protonating a carbon-carbon double

bond with a strong acid, HBr, and then forming a bond between the resultant bromine anion and

carbocation to produce the addition product (Figure 2). In this reaction, students needed to assess

carbocation stability by considering the rules and syntax associated with the available resonance

390 stabilization due to the delocalizable electrons on the oxygen adjacent to the double bond. Then they

needed to recognize the electrophilic-nucleophilic relationship between the carbocation intermediate

and bromine anion and perform the associated operation to form the final product.

The first step of this mechanism was determining the most stable carbocation intermediate. Two

students did not consider carbocation stability. Jasmine, an app student, first tried protonating the

395 oxygen. When the app would not allow the move, Jasmine recognized that protonation should occur at

the double bond and chose a carbon at random. Ana, a paper-pencil student, focused on sterics as

they had in the acid-catalyzed addition reaction, applying the rule of steric hindrance to determine

which carbon the bromine would later add to. Ana said:

Oh but that methyl, or that group [the ethyl], that's a big group. So would a Br want to be next to 400 [an ethyl] or an oxygen? They debate the relative merits of having the bromine adjacent to the oxygen versus the ethyl group on

the opposite side of the double bond, in the end deciding that steric hindrance will favor bromine

placement on the carbon adjacent to the oxygen and thus protonating the other double bond carbon.

While this allowed Ana to form the favored product, it indicates limited chemical reasoning as they

405 neglected to consider the stability of the carbocation intermediate.

The remaining students used the rules and syntax of carbocation stabilization to determine the

most stable carbocation, with half considering resonance stabilization and the other half considering

substitution. Four paper-pencil students and one app student correctly identified that resonance

stabilization would dictate where the carbocation would form. Of the four paper-pencil students, Mary

410 incorrectly applied the rules of resonance stabilization, focusing on the potential resonance forms of

the reactant rather than the intermediate when making their decision. This is similar to results in a

parallel study focused on acid-base reaction mechanisms.18 The one app student who focused on

resonance, Peter, accessed the goals directly from the task card prior to starting the puzzle and used

the goals to guide their movements. Peter paraphrased the goal card, saying:

415 So, my goal is to use resonance structures to show how positive charge with the carbocation can be delocalized to form the major alkyl bromide. The goals directed Peter to consider resonance rather than substitution for the rules to determine

carbocation stability, and Peter then correctly applied the syntax affiliated with resonance

stabilization. The remaining students, four app and one paper-pencil, focused on substitution to

420 determine the site of protonation. Daisy, the paper-pencil student, protonated the correct carbon

without stating their reasoning, but upon prompting recognized that both carbons in the double bond

could be protonated and stated that one would be the major product due to substitution of the

carbons. However, Daisy did not reconcile the fact that this principle was not used to determine the

carbon they protonated and proceeded to carry on with the reaction. Tiana and Flynn, both app

425 students, first identified the two double-bonded carbons as secondary and then started thinking about

how the oxygen could stabilize the adjacent carbon through induction. This allowed Flynn to form the

correct carbocation despite lacking some chemical intuition. Tiana’s reasoning, however, focused on

how induction would impact the ultimate placement of the bromine, causing them to incorrectly

identify that the oxygen would donate electron density to the adjacent carbon. With this reasoning,

430 Tiana decided that the carbon further from the oxygen would be more electron deficient and thus

would more readily accept the electrons on the bromine anion during the second step of the

mechanism, which led them to protonate the carbon adjacent to the oxygen. Pepper and Belle

exhibited incorrect knowledge of substitution syntax, as both identified the sp2-hybridized carbon

bonded to the oxygen as primary. Pepper said:

435 So, this is the secondary carbon where this [carbon bonded to an oxygen] is the primary carbon so I would say [the secondary carbon] is the more stable carbocation... Belle hesitated, having noticed resonance on the task card, but continued with their reasoning despite

this. Aurora, a paper-pencil student, demonstrated similar incorrect knowledge in their think-aloud,

despite correctly focusing on resonance. Aurora called the double-bond carbon adjacent to the oxygen

440 sp-hybridized in a misapplication of hybridization rules, discounting the heteroatom. The syntactic

difficulties leading to the mis-assignment of substitution and hybridization may be indicative of poor

understanding of the affiliated concepts.

The app students who did not initially consider resonance to determine carbocation stability did so

following prompting by the app, whereas the two paper-pencil students did not during the post-

445 reaction interview questions. However, Daisy demonstrated in their response to the question about

resonance they did not understand what the term ‘resonance structures’ meant, conflating it with

major and minor products. The five app students who did not consider resonance when initially

determining carbocation stability were prompted to do so by the app in two different ways. As

described above, Pepper and Belle both selected the incorrect double-bond carbon to protonate,

450 forming the minor product. This resulted in a hint from the app that one of the carbons would be

resonance stabilized which led both students to form the resonance structure of the carbocation

intermediate prior to addition of the bromine. Pepper said:

Okay, so, resonance is always more important than based on what's primary and secondary and tertiary. I guess I didn't consider that just because I only saw those double bonds and didn't see 455 these lone pairs but if we think about putting it on this carbon instead then the lone pairs from this oxygen can make a bond between this oxygen and carbon which would stabilize it because then this carbon wouldn't really be a carbocation all the time. While this prompted Pepper to consider resonance-stabilization, it also elucidated a misunderstanding

that resonance forms are present as a mixture of structures in solution, an error also identified in the

460 parallel acid-base study.18 Although Tiana also received the hint after forming the minor product, they

immediately applied the syntax of resonance stabilization to form the major product and did not depict

the resonance contributor until prompted by the goals card in the app (Figure S1B). The two app

students who chose the correct double-bonded carbon to protonate without considering resonance

were prompted by the goals card to form a resonance structure after they had achieved the major

465 product. Following the operation of forming a resonance structure in the app, only one of the two

students, Flynn, verbalized that they should have considered resonance to determine the intermediate

structure.

In the last step of the reaction, students tried adding the bromine anion formed during protonation

of the carbon-carbon double bond to either one of the carbon atoms or the oxygen atom in the reaction

470 intermediate. Three app students and four paper-pencil students formed the major product, while

three app and two paper-pencil students formed the minor product. As mentioned above, the app

prompted students who formed the minor product to consider resonance stabilization to determine the

major product which led those students to eventually form the major product. After forming the

correct carbocation and resonance structure following the hint, Belle then tried to add the bromine

475 anion to the positively charged oxygen in the resonance structure. When the app did not allow this,

Belle tried to reason through what the next step in the mechanism was, focusing on the intermediate

in its resonance contributor form, before turning to the app goals for guidance:

A negatively charged bromide, I have a positively charged oxygen, and it doesn't ... Okay, oxygen is never gonna wanna have four bonds. I can check what the goals are, though. Major alkyl 480 bromide product. Belle was then able to use both their chemical intuition and the scaffolding provided by the app to

correctly form the major product. Jasmine, another app student, also tried to form a bromine-oxygen

bond based on directions from the goal card, exhibiting a poor understanding of Lewis structures and

nomenclature. After trying a few moves in the app, Jasmine undid the carbon-oxygen double bond to

485 form a bond between the bromine and the carbocation. However, they then exhibited poor

understanding of the forces driving major versus minor product and inductive effects by implying that

the bromine would be stabilized by being located closer to the oxygen. Despite making some

chemically inaccurate moves, students generally justified the last step by focusing on forming a bond

to stabilize the charged species.

490 Generally, it appears that students exhibited a series of difficulties with the first step of this

problem. Only four of the thirteen students identified resonance as the source of carbocation

stabilization without additional direction. With the exception of Jasmine and Peter, the other students

reasoned based on substitution or sterics. This could be due to features such as substitution and

sterics being more explicitly represented to students in the referents for the reaction, compared to

495 resonance being a more implicit feature, which aligns with previous literature indicating that students

tend to rely more heavily on explicit chemical features.7-10 As exemplified by Pepper’s statement above,

the hidden oxygen lone pairs in the app may have limited app students’ ability to identify the oxygen's

potential to provide resonance stabilization to the neighboring carbocation, however the app does

require students to consider resonance to complete the puzzle. Additionally, students’ overapplication

500 of substitution to determine reaction progress indicated some difficulties with the syntax of

substitution when carbon was bonded to a heteroatom. Determining degree of substitution when a

carbon is bonded to a heteroatom would be an unfamiliar situation for students as they would not

normally have to consider this situation when resonance should be the leading factor in determining

carbocation stability. We also identified a few students who focused on the placement of the bromine

505 as a driving force for the reaction, which aligns with a product rather than process-oriented approach

to the reaction.7-10

Limitations This study was qualitative in nature with a small pool of participants and as such the results are

descriptive and may not be generalizable. Additionally, the participants were all recruited from the

510 same institution and so we could not characterize potential differences due to curricular differences.

However, our experimental design, recruiting students with a range of ability levels and using two

modalities to probe student thinking, did capture an array of students’ conceptions. Claims about

differences between the two groups, paper-pencil and app students, may also be limited by our sample

size and as there may have been inherent differences in the think-aloud interviews that could not be

515 controlled for which influenced student responses. However, we incorporated several measures in the

interview protocols to mitigate differences, such as aligning the reaction representations and leading

the app students through a tutorial to minimize added cognitive load from using a new interface.

CONCLUSION In this study we elicited student thinking about addition reaction mechanisms using two

520 modalities: paper-pencil and app-based. Qualitative analysis of the resulting think-aloud interview

data provided insight into how students were utilizing the chemical information contained within the

referents of the reaction, as well as what concepts students focused on and how they applied them to

the given reactions. Additionally, comparing the two groups of students elicited a wider range of

student thinking and indicated the potential benefits and drawbacks of each modality.

525 When approaching the acid-catalyzed addition reaction, many students recognized the hydronium

and used it as a cue to inform the initial steps of their mechanisms. However, some students appeared

to have a surface level understanding of the chemistry influencing the role of the acid-catalyst as they

did not know how to use it correctly during the reaction. Additionally, most students performed the

steps of using and reforming the catalyst with a focus on completing potentially memorized reaction

530 steps rather than discussing the chemistry driving the reaction in a specific direction.

For both reactions, students recognized that they needed to identify which of the double-bonded

carbons would form a more stable carbocation to determine the favored reaction pathway, aligning

with previous literature that carbocation formation is the most commonly recognized step in addition

reactions,9 but they had some difficulties identifying and applying the different modes of stabilization

535 between the two reactions. In the acid-catalyzed reaction, most students successfully identified that

the stability of a carbocation would increase as its substitution increased but did not necessarily know

why and exhibited some difficulties with the associated syntax of applying stabilization via

substitution. In juxtaposition, over half of the students failed to identify the potential for resonance

stabilization in the resonance-stabilized addition reaction. Some of those students then struggled to

540 apply substitution or relied on arguments about induction and steric effects to determine the favored

carbocation intermediate. Of the students who correctly identified resonance as the primary source of

carbocation stabilization for this reaction, most did so successfully. Interestingly, while students

appeared to understand the underlying concepts affiliated with resonance better than those associated

with substitution, they appeared to favor substitution when determining carbocation stability. This

545 may be related to students’ tendencies to focus on explicit chemical features compared to implicit

electronics,7-10 as they can readily determine substitution by counting bonds whereas resonance

requires students to visualize or draw resonance contributors. Delocalizable lone electron pairs could

serve as explicit cues for students to consider resonance, and as such it would be important for lone

pairs to be depicted in whatever modality students use when considering a mechanism. However,

550 more research is needed targeting student conceptions of resonance.

The two groups of students were relatively similar in how they approached the mechanisms. Both

paper-pencil and app students considered multiple reaction pathways, unlike in the companion study

where paper-pencil students primarily considered one pathway.18 Students also demonstrated some

reflective thinking during the think-aloud interviews. Specifically, some of the app students provided

555 chemical reasoning after receiving prompting from the app and some paper-pencil students were

reflective upon completing the reactions. While fewer app students identified resonance without

prompting, they were directed to do so by the app in order to complete the puzzle.

Results from our study substantiate previous research indicating that students tend to recognize

explicit, surface features and use those as cues to apply known steps.2, 6 Additionally, our results

560 indicate students have a surface-level understanding of the concepts relevant to addition reactions

that inhibit their ability to apply them in unfamiliar situations. Thus, it is important that, as

instructors, we include discussions of the chemical principles behind the rules presented to students

in class and provide them with opportunities to apply concepts such as carbocation stabilization to

identify gaps in their understanding before moving on to more complex reaction types. Additionally,

565 instructors can include the explicit cues contained in line-angle representations—i.e., lone pairs on

heteroatoms—that experts recognize inherently but may still be difficult for students to identify at the

introductory organic chemistry level. While a few students exhibited some reflection during the course

of the think-aloud interviews, instructors could help students develop these skills by providing

additional scaffolding or modeling reflective behavior.

570 ASSOCIATED CONTENT Supporting Information Coding Scheme and goal cards presented by the app (PDF)

AUTHOR INFORMATION Corresponding Author 575 *E-mail: [email protected]

ACKNOWLEDGMENTS This work has been supported, in part, through a grant to Alchemie from the National Science

Foundation Small Business Innovation Research program, #1659983, and the Michigan Corporate

Relations Network (MCRN), funded by the Michigan Economic Development Corporation (MEDC) and

580 administered by the University of Michigan Business Engagement Center and the U-M Economic

Growth Institute's Small Company Innovation Program (SCIP). This material is based upon work

supported by the National Science Foundation Graduate Research Fellowship Program under Grant

No. DGE 1256260. Additionally, we would like to thank the students who participated in our study

and Alchemie for creating the graphical abstract.

585 REFERENCES 1. Duis, J. M., Organic Chemistry Educators’ Perspectives on Fundamental Concepts and Misconceptions: An Exploratory Study. Journal of Chemical Education 2011, 88 (3), 346-350. 2. Ferguson, R.; Bodner, G. M., Making sense of the arrow-pushing formalism among chemistry majors enrolled in organic chemistry. Chemistry Education Research and Practice 2008, 9 (2), 102-113. 590 3. Bhattacharyya, G., From Source to Sink: Mechanistic Reasoning Using the Electron-Pushing Formalism. Journal of Chemical Education 2013, 90 (10), 1282-1289.

4. Graulich, N., The tip of the iceberg in organic chemistry classes: how do students deal with the invisible? Chemistry Education Research and Practice 2015, 16 (1), 9-21. 5. Grove, N. P.; Cooper, M. M.; Rush, K. M., Decorating with Arrows: Toward the Development of 595 Representational Competence in Organic Chemistry. Journal of Chemical Education 2012, 89 (7), 844- 849. 6. Bhattacharyya, G.; Bodner, G. M., "It Gets Me to the Product": How Students Propose Organic Mechanisms. Journal of Chemical Education 2005, 82 (9), 1402. 7. DeCocq, V.; Bhattacharyya, G., TMI (Too much information)! Effects of given information on 600 organic chemistry students’ approaches to solving mechanism tasks. Chemistry Education Research and Practice 2019, 20 (1), 213-228. 8. Anderson, T. L.; Bodner, G. M., What can we do about ‘Parker’? A case study of a good student who didn't ‘get’ organic chemistry. Chemistry Education Research and Practice 2008, 9 (2), 93-101. 9. Graulich, N.; Bhattacharyya, G., Investigating students' similarity judgments in organic 605 chemistry. Chemistry Education Research and Practice 2017, 18 (4), 774-784. 10. Graulich, N.; Hedtrich, S.; Harzenetter, R., Explicit versus implicit similarity – exploring relational conceptual understanding in organic chemistry. Chemistry Education Research and Practice 2019. 11. Cruz-Ramírez de Arellano, D.; Towns, M. H., Students' understanding of alkyl halide reactions 610 in undergraduate organic chemistry. Chemistry Education Research and Practice 2014, 15 (4), 501- 515. 12. Flynn, A. B., How do students work through organic synthesis learning activities? Chemistry Education Research and Practice 2014, 15 (4), 747-762. 13. Bhattacharyya, G., Trials and tribulations: student approaches and difficulties with proposing 615 mechanisms using the electron-pushing formalism. Chemistry Education Research and Practice 2014, 15 (4), 594-609. 14. Caspari, I.; Weinrich, M. L.; Sevian, H.; Graulich, N., This mechanistic step is “productive”: organic chemistry students' backward-oriented reasoning. Chemistry Education Research and Practice 2018, 19 (1), 42-59. 620 15. Galloway, K. R.; Leung, M. W.; Flynn, A. B., Patterns of reactions: a card sort task to investigate students’ organization of organic chemistry reactions. Chemistry Education Research and Practice 2019, 20 (1), 30-52. 16. Cooper, M. M.; Grove, N.; Underwood, S. M.; Klymkowsky, M. W., Lost in Lewis Structures: An Investigation of Student Difficulties in Developing Representational Competence. Journal of Chemical 625 Education 2010, 87 (8), 869-874. 17. Grove, N. P.; Cooper, M. M.; Cox, E. L., Does Mechanistic Thinking Improve Student Success in Organic Chemistry? Journal of Chemical Education 2012, 89 (7), 850-853. 18. Petterson, M. N.; Watts, F. M.; Snyder-White, E. P.; Archer, S. R.; Shultz, G. V.; Finkenstaedt- Quinn, S. A., under review. 630 19. Briggs, M., Models and Modeling. In Theoretical Frameworks for Research in Chemistry/Science Education, Bodner, G. M.; Orgill, M., Eds. Pearson Prentice Hall: Upper Saddle River, NJ, 2007; pp 69- 82. 20. Briggs, M.; Bodner, G., A Model of Molecular Visualization. In Visualization in Science Education, Gilbert, J. K., Ed. Springer: Netherlands, 2005; Vol. 1, pp 61-72. 635 21. Lesh, R.; Hoover, M.; Hole, B.; Kelly, A.; Post, T., Principles for Developing Thought-Revealing Activities for Students and Teachers. In Handbook of Research Design in Mathematics and Science Education, Routledge: 2000. 22. Cooper, M. M.; Grove, N. P.; Pargas, R.; Bryfczynski, S. P.; Gatlin, T., OrganicPad: an interactive freehand drawing application for drawing Lewis structures and the development of skills in 640 organic chemistry. Chemistry Education Research and Practice 2009, 10 (4), 296-301. 23. Winter, J. E.; Wegwerth, S. E.; DeKorver, B. K.; Morsch, L. A.; DeSutter, D.; Goldman, L. M.; Reutenauer, L. M., The Mechanisms App and Platform. In Active Learning in Organic Chemistry: Implementation and Analysis, American Chemical Society: 2019; Vol. 1336, pp 99-115.

24. McCollum, B. M.; Regier, L.; Leong, J.; Simpson, S.; Sterner, S., The Effects of Using Touch- 645 Screen Devices on Students’ Molecular Visualization and Representational Competence Skills. Journal of Chemical Education 2014, 91 (11), 1810-1817. 25. Cohen, L.; Manion, L.; Morrison, K., Research Methods in Education. Routledge: 2011; pp 559- 573. 26. Mechanisms, Alchemi Solutions, Inc: 2018. 650 27. Herrington, D. G.; Daubenmire, P. L., Using Interviews in CER Projects: Options, Considerations, and Limitations. In Tools of Chemistry Education Research, American Chemical Society: 2014; Vol. 1166, pp 31-59. 28. Corbin, J. M.; Strauss, A., Grounded theory research: Procedures, canons, and evaluative criteria. Qualitative Sociology 1990, 13 (1), 3-21. 655 29. McHugh, M. L., Interrater reliability: the kappa statistic. Biochemia Medica 2012, 22 (3), 276. 30. Kirilenko, A. P.; Stepchenkova, S., Inter-Coder Agreement in One-to-Many Classification: Fuzzy Kappa. PLOS ONE 2016, 11 (3), e0149787. 31. Galloway, K. R.; Stoyanovich, C.; Flynn, A. B., Students’ interpretations of mechanistic language in organic chemistry before learning reactions. Chemistry Education Research and Practice 660 2017, 18 (2), 353-374. 32. Webber, D. M.; Flynn, A. B., How Are Students Solving Familiar and Unfamiliar Organic Chemistry Mechanism Questions in a New Curriculum? Journal of Chemical Education 2018, 95 (9), 1451-1467. 33. Bain, K.; Towns, M. H., A review of research on the teaching and learning of . 665 Chemistry Education Research and Practice 2016, 17 (2), 246-262.