E) Structuring your CURE

Designing a CURE is a complex endeavor that requires coordination with staff members, administration, and other instructors. Fortunately, there is now an extensive literature on CUREs that enables the identification of best practices and recommendations for an effective experience. The elements identified in Table 11 should be revisited throughout the process of designing a CURE.

Critical elements of a successful CURE distributed across categories of topic and focus, research design, pedagogical design, communication, and reflection.
Table 11. Elements of a successful CURE (modified from Dolan, 2016; Hanauer et al., 2022; Kortz & van der Hoeven Kraft, 2016; Wiley & Stover, 2014; Turner et al., 2021; Kloser et al., 2011; Hatfull et al. 2006)

The goal of this section is to use the framing questions presented above (p. 33-34) to guide prospective CURE instructors through the development of the research experience. These questions are not just prompts for reflection; together, they also form a worksheet for the design process.

  • For each question, the “Overview” provides an overview of the critical design elements that are identified and developed in this section of the worksheet.
  • The “starts here” units guide you through these elements of design. Although the pedagogy-centered issues are presented before the research-centered ones, they are better thought of as parallel tracks.
  • The “research and education together” section combines the elements of design from both aspects of the CURE. This is an opportunity to identify and resolve conflicts.
  • The “boxes” help remind instructors of the important elements of the question and critical points of contact across (and beyond) campus.

1. How will the CURE be integrated into the curriculum?

Overview: Despite the desire to design the CURE solely around learning goals and research questions, reality requires first the consideration of the place of the CURE in the students’ slate of courses. This section enables you to determine:

  1. The audience of your course
  2. The level of preparation and prior knowledge you can expect from students
  3. The duration of the experience
  4. The integration of program requirements into the course
  5. The scope and intent of the research

Education starts here: Many of the questions below are best answered in communication with the chair of the undergraduate curriculum committee of your department or unit.

  • Will the CURE be a new course or integrated into an existing course?
  • If developing a new course:
    • Will this course be aimed at majors or non-majors?
    • Will this course be an upper-level class with prerequisites or open to freshmen and sophomores?
    • Will this course become a requirement for any minor or certificate?
    • How many course credits will the CURE represent? Are there constraints on this number? How many hours of contact hours does this correspond to? How will they be divided between lectures, recitations, and labs? What is the total time commitment you can expect from students including homework?
    • Will this course be taught as a summer course? A half-semester course? Through two courses over an entire academic year?
    • What is the expected enrollment for the course?
    • Will this course become a requirement for other courses or be an elective?
      • If becoming a requirement:
        • What is the knowledge base and skillset that students need to master to continue with the next course?
      • If an elective:
        • What are the goals and expected learning outcomes of electives in your department, college, or unit? More broadly, what are expectations in your field and related professional avenues?
  • If revising an existing course:
    • What is the place of this course in the curriculum? The major? Minors or certificates?
    • Will the CURE be replacing traditional labs or are you also integrating the lectures and/or recitations into the experience?
    • What are the expected learning outcomes of the current version of the course? Are they being revised?
    • What is the current enrollment in the course?
    • How does the course fit into the curriculum and how does it prepare students for the next steps in their education and career?

 Scholarship starts here: When developing a researcher-driven CURE, this element of design is led by the researcher and his team. When developing a researcher-independent CURE, this element of design is driven by the CURE program goals that you are joining.

  • What discipline/field of research does the research topic fall under?
  • Is the research project novel? What is already known about the possible outcome(s) of the research?
  • Are you planning on focusing on a single question/goal or will the class tackle several research questions?
  • Is this research hypothesis-driven or exploratory?
  • What is the current stage of the research?
  • Who are the stakeholders of the research that should be involved? Any collaborators?
  • How does the research align with the topic, scope, and knowledge goals of the class (particularly if revising a pre-existing course)?

Research and Education together: Any incompatibility between research and education goals at this point of the development process, including incompatibilities between research goals and class enrollment or number of contact hours, may be fatal. Although revisions to the focus of the research or learning goals of the course are sometimes possible without compromising the student experience and learning, the nature of the course and the scope of the research must come together for the CURE to be possible.

 

Box 1: Important points

  • Contact the chair of the undergraduate curriculum committee of your department or unit.
  • Verify the guidelines and requirements of the course approval process when proposing new courses.
  • Engage colleagues (staff or faculty) involved in the implementation of existing courses if you are revising them.
  • Identify the relevant expected learning outcomes of the department, program, major, minor, or certificate.
  • Do not forget the research stakeholders. How can postdocs, graduate students, and undergraduate students in your lab contribute to and benefit from the CURE?

 

2. To what extent will students have intellectual responsibility and ownership of the research?

Overview: Research shows that student ownership of the research project undertaken as part of the CURE positively impacts their experience of the course and their learning gains (Hanauer et al., 2012; Harrison et al., 2011; Hatfull et al., 2006). Project ownership is a central component of the role of CURE instructors (Hanauer et al., 2022). Different degrees of project ownership may be possible depending on the scope, depth, and conditions of the CURE. They are intimately associated with the level of inquiry that students engage in.

Several levels of inquiry have been defined in the context of laboratory courses (Buck et al., 2008) and revised to fit the framework of CUREs (Brownell & Kloser, 2015). Because CUREs require the investigation of novel questions with no-known outcomes and the communication of the research results, they may only fall under some of these categories (Table 12). The degree of responsibility of the students in the design of the research project is associated with the degree of inquiry desired by the instructor and should be guided by the expected learning outcomes of the course. Together, they will open a range of opportunities for the student’s ownership of their research.

 

There are four modes of inquiry: structured inquiry, guided inquiry, open inquiry, and authentic inquiry. CUREs are most appropriate in inquiry modes where the answer is not known.
Table 12. Levels of inquiry in undergraduate courses (modified from Buck et al., 2008 and Brownell & Kloser, 2015). It is entirely possible to imagine hybrid situations in which for example only parts of the research protocol are provided to students.

 

This section enables you to determine:

  1. The level of inquiry you are aiming for in the CURE.
  2. The choice between students exploring researcher-chosen questions/hypotheses and student-driven questions/hypotheses
  3. The amount of input students will have on the design of the research project
  4. The ability of students to communicate the research upon its completion

Education starts here: The ELOs for the course are critical in determining everything from the nature of the research activities included in the experience to the scope of the research questions. They should be referenced throughout the course design process.

  • What level of inquiry is dictated by the ELOs?
    • Are students expected/able to develop original questions directly from observations? Will students be choosing from a range of predetermined questions and topics?
    • How will you foster peer discussions and class-wide activities if several questions/hypotheses will be studied by different groups or individuals?
    • Will you be giving students an introduction to the existing knowledge on their research topic or are they expected to collect this information directly from the published literature? If the latter, how will you guide this work and/or assess it?
    • Are students allowed/capable of collecting their own data based on permitting requirements (e.g., IRB) and ethical best practices?
    • Will appropriate analytical procedures be suggested to students or should they propose particular approaches based on their reading of the literature?
  • Is the level of inquiry selected compatible with the amount of time given to students?
  • Is the level of inquiry selected compatible with the expertise of the students?
    • What is the efficacy of students with reading the primary literature? Will you be providing training on this topic?
    • What is the familiarity of students with the field of research chosen for the CURE? How much background information is necessary to be able to comfortably initiate the research process?
    • Can you expect the students to understand and implement the analytical procedures likely to be involved in the research? Will you be providing the necessary knowledge as part of the class?
    • What are the computational skills of the students? Does the analytical protocol require coding skills or can it be implemented in a software program with a GUI?
    • In what prior classes or context would the students have acquired the required expertise/competence for the CURE?
  • If students will explore their own questions:
    • Will you be placing constraints of the breadth and depth of the questions chosen by students?
      • What is too narrow a question?
      • What is too broad a question?
      • Can you mentor the diversity of questions and datasets entailed by this structure?
      • Are there topics that could be emotionally difficult for some students? Will these students have the opportunity to work on a different projects or will these projects simply not be validated by the instructor?
    • How will students select their hypothesis/question from all possible options?
      • What is the process by which a student question/hypothesis is approved for the CURE?
      • Can a mock panel (for example composed of the instructor, graduate student researchers, and colleagues) review student proposals?
    • If students will be working in groups, how will you balance the group structure with individual questions?
  • If students will be exploring researcher-chosen questions:
    • Will the questions be prescriptive, or will there be an opportunity for students to refine their question, narrow it, or define a question from a broad topic?
    • Is the question chosen compatible with the time frame of the CURE and the skill level of the students?

Scholarship starts here: It is important to recognize that there are constraints on scholarship stemming from regulatory and ethical requirements. Additionally, and no less important to the success of the CURE, its place within the broader research project it is part of, may drive the nature of the deliverable, the breadth and depth of the work, and the level of prescription imposed by the instructor(s).

  • What are the existing constraints on the project from other stakeholders (e.g., other lab members, external collaborators, and national programs or networks)?
  • What are the existing constraints on the project from existing data, commitments to funding agencies, museum collections, permitting agencies, university rules (including EHS and IRB), formal and informal agreements, and ethical concerns?
  • How do these different types of constraints narrow the nature, breadth, and depth, of the questions that can be investigated by the students? How about the analytical protocols?
  • If the CURE is researcher-driven:
    • What question are you interested in exploring?
    • What question/topics are you comfortable/uncomfortable mentoring?
    • How much of the research process are you comfortable leaving up to the students?
    • What is the deliverable you are hoping for upon completion of the CURE?
    • How critical is the success of the research (i.e. the obtention of the deliverable defined above) to you?
    • How much time will you be dedicating to the CURE outside of contact hours?
    • Is the research goal of the CURE compatible with the time and means given to students?
  • Does the CURE represent a standalone project publishable upon its completion or a portion of a bigger project? How does this impact the format of research communication upon completion of the CURE (i.e. does the end product represent a manuscript? A poster presentation? An element of a grant proposal?)

Research and Education together: Legal, ethical, and professional requirements dictated by the field of research, necessary resources for the project, and the research program context of the CURE are the starting point for the level of inquiry that is possible for the CURE. However, there are several existing strategies to increase student engagement and ownership of the project that should be considered:

  • Hanauer et al. (2022) presented a model of CURE mentoring that includes a project ownership strategy centered around fostering personal responsibility starting with teaching a scientific protocol and includes promoting research ethics, facilitating peer collaboration, encouraging independence, encouraging engagement and enthusiasm, creating opportunities for presentations, and fostering future educational and career opportunities.
  • Students who collect their own data are more invested in the research project and display a greater research identity than students working with pre-existing datasets they did not collect (Cooper et al., 2020a).
  • Undergraduate students can work with graduate students or undergraduate researchers who have benefited from mentored research experience or previous iterations of the CURE to contribute to the research questions and methods within a framework (Hanauer et al., 2012).
  • Emphasizing to students the significance of the research to the broader community of scholars in the field is important (Hanauer et al., 2012). Research shows that broadly relevant novel research leads to higher ownership of the project by students (Cooper et al., 2019).
  • It is sometimes possible to pursue research questions relevant to the community or the students themselves (e.g., Malotky et al., 2020; Silvestri, 2018; Valliere 2022a). Students enjoy the opportunity to choose their own research topic (Amir et al., 2022) and those who are given opportunities to investigate questions relevant to themselves or their community show increased ownership of the research (Hanauer et al., 2012).
  • Incorporating meetings that mimic research group meetings in the discipline and poster presentations can promote project ownership (Satusky et al., 2022).
  • A CURE design that creates intellectual challenge and encourages problem solving deepens the engagement of students and the significance of the research experience (Hanauer et al., 2012). Challenges and iterations are critical to increasing the ability of students to navigate research obstacles (Gin et al., 2018; Light et al., 2020). In fact, students report valuing the ability to learn from their mistakes (Harrison et al., 2011) and view challenges and iterations as more representative of a real research experience (Goodwin et al., 2021). CUREs should deliberately incorporate iteration and discussions about the importance of iteration in research as part of student development (Light et al., 2020). Iteration enables the development of adaptive strategies by students that benefit them beyond the course (Cooper et al., 2022). A review of the best practices to engage students in problem solving is provided by Frey et al. (2022) and at https://lse.ascb.org/evidence-based-teaching-guides/problem-solving/.
  • The CURE should be challenging without being overwhelming; this level of difficulty, paired with instructor support, has previously be showed to foster motivation (Dolan, 2016).
  • If students will be developing their own questions, they should be guided through the process of identifying their own research question including the following critical issues:
    • What questions have already been asked?
    • What is the existing knowledge on the topic?
    • What questions have not yet been answered or even asked?
    • What are the gaps in the conversation?
    • What questions are relevant to the field of research?
  • An activity leading students through the process of identifying scholarly significant questions can be followed up with an activity helping them narrow big research questions into manageable CURE projects (see [Activity 2]).

 

Box 2: Important points

  • Students should tackle questions and test hypotheses relevant to the broader (research) community.
  • Students should be allowed to make decisions throughout the design and implementation of the research protocols.
  • Instructors should rethink their place in the classroom to that of mentors and not merely supervisors.
  • There should be multiple opportunities for students to develop their own hypotheses and defend them with evidence from the literature and their own work.

 

3. Which components of the research process will be integrated into the cure?

Overview: The traditional view of the scientific method involves a series of steps starting with observation and the formulation of a hypothesis followed by the test of this hypothesis and the dissemination of the findings (Voit, 2019). In the social sciences, qualitative, quantitative, and mixed approaches may be adopted to probe questions pertaining to human beings and their societies, but the overall process of inquiry remains the same (Creswell, 2014). The nature of the data collection process and data analyses differs in the humanities, but the scholarly endeavor is still the “evidence-based exploration of a question or hypothesis that is important to those in the discipline in which the work is being done” (University of Washington English Department). Thus, across disciplines, the research process requires the identification of an appropriate research question, whether it be from direct observation or a review of the primary literature, the collection of data, their analysis, the interpretation of outputs, and the communication of the knowledge gained from the research. It is critical to identify which of these elements of the research process will be included in the CURE while considering the constraints of the undergraduate classroom. This section enables you to determine:

  1. Your approach in defining research to students and framing their experience as scholars
  2. The background information that needs to be provided to the students
  3. The role of students in the data collection process
  4. The engagement of students with the primary literature
  5. The role of students in the data analysis process
  6. The role of iteration and impact of failure on the CURE

Education starts here:

  • How would your students define “research”?
    • What does the research process look like to them?
    • What is their understanding of the concept of “Research as Inquiry?”
    • What is your understanding of the students’ view of research?
  • What information is necessary to understand the basis of the question(s) the students will investigate?
    • Are there recently published reviews of the field/topics that students will be researching? Are there other sources of information that can help draw students into the literature (including secondary and tertiary literature, videos, and journalistic writings)?
    • Are there important conflicts or controversies in the field that students should know about? What is the current paradigm and are we on the edge of a new one?
    • Will you provide a brief introduction to the study system in class drawing upon your knowledge, past iterations of the CURE, and ongoing research on campus (and beyond)?
  • Will students be collecting the entire dataset they will analyze? Alternatively, will they be working from pre-existing datasets in whole or in part?
    • Are there pilot or example datasets that enable students to visualize their objective while collecting most of the data they will analyze?
    • Will specific data collection protocols be enforced by permitting requirements (e.g., IRB) or ethical best practices?
    • Are there professional, legal, or ethical training requirements for students to collect data? Consult and analyze data?
      • How can you incorporate responsible and ethical conduct of research training into the course to satisfy needs and train reflective and responsible professionals?
  • Will students be required to read certain publications or a minimum number of freely chosen publications? At what stages of the research process will engagement with the literature be suggested/enforced?
    • How will you guide students towards relevant publications and/or vet their choice of readings?
    • How will you make sure that students engage with a variety of sources and read the work of multiple authors representing different points of view?
    • Students should be prompted to reflect upon questions like: (1) How does the source contribute to the scholarly conversation? (2) How do others in the field perceive the value of the source? (3) How does the source guide or support my work?
  • Will specific analyses be required of students in the form of a detailed protocol or as a list of milestones? If not, how will students choose the analyses they will use? How will they be guided through this process?
    • What is the role of published studies as models for the students’ work?
  • Will students identify analyses that will be run by scholars with greater computational skills or are they able to run some or all of these analyses themselves?
  • How would experimental failure or faulty data collection impact the learning process?
    • Are there backup datasets?
    • Are there alternative protocols and analyses?
    • What is the relevance/significance of null results or a failed protocol to the scholarly community?
  • Is there enough time in the CURE to implement alternative protocols? Revise the data collection? Add to the dataset? Repeat an analysis?

Scholarship starts here:

  • Is there a need for a formal review of the field of study the students will engage with?
    • Can the CURE be the impetus to publish a peer-reviewed review of a topic?
    • Would a more accessible summary benefit audiences beyond the CURE including undergraduates starting mentored research experiences in the lab or the general public?
  • Is the CURE an opportunity for the instructor/researcher to engage or reengage with specific aspects of the primary literature?
  • What is an appropriate sample (including quantity and quality parameters) to address the question of the CURE?
  • Will student-collected data represent the entire set of relevant data for this project or are there other elements of the research project they will not engage in?
  • Does the use of preexisting data dictate a specific data collection protocol to enable comparisons/integrations?
  • What work is necessary ahead of the CURE to make data collection by students possible?
  • What checks and balances will be implemented to validate the quality of the student-collected data?
    • Do those checks and balances involve peer-review? Instructor review? Outside researcher review?
  • Are there analyses that are standard/expected in the field?
    • Are some of these analyses computationally too demanding for the CURE?
    • Do some of these analyses require coding or statistical training beyond what can reasonably be expected of students?
  • Are there critical failure points in the research protocol? Are there data or analyses whose failure would hobble or interrupt research progress?
  • What would be the consequence of failure (of a specific analysis, experiment, or data collection) on proposal submissions, publication completions, and professional advancements (including promotion and tenure)?
  • What is the role of the CURE in the development of the research program/project(s) of the researcher/instructors/collaborators/graduate students/mentored undergraduate researchers/ etc.?

Research and Education together:

  • Do not underestimate the importance of framing the entire CURE experience with a conversation about the nature of research. Students may not always recognize research as information creation and inquiry. “Research as Inquiry” is a core concept underlying the practice of research. Defining research as inquiry means explaining to students that research is an open-ended and messy exploration process focused on information gaps, unanswered questions or problems, involving multiple sources and the interpretation of information, often with no clear right answer, leading to new questions, generating ambiguity, and requiring an open mind, persistence, and flexibility (e.g., [Activity 3]). Research shows that engaging in this type of conversation about the nature of research can lead to higher student outcomes (Lopatto et al., 2022).
  • CUREs are designed as authentic engagements in research and as such should incorporate responsible conduct of research (RCR) training to form conscientious professionals (Diaz-Martinez et al., 2021). Many institutions and organizations require ethical conduct of research training. At Ohio State, RCR training is provided through the Collaborative Institutional Training Institute (CITI) course, which can be assigned as homework.
  • Discipline-specific best practices in ethical conduct of research should also be implemented as part of the CURE (e.g., Hills and Light 2022). They can be integrated in the grading scheme for the course.
  • Consider using mind-mapping exercises to help students identify the existing knowledge of the topic of the CURE and the current debates and controversies in the field of study [Activity 4].
  • Established databases provide sources of data that can be used in CUREs (e.g., Gastreich, 2020). Those include museum databases, citizen science databases, and professional databases (Table 13).

Examples of databases that can be leveraged for research projects in CUREs

Research databases across disciplines including Art History, Biology, Earth Sciences, Machine Learning, Physics and Astronomy, as well as Political and Social Sciences
Name URL
Art History
Index of Medieval Art https://theindex.princeton.edu/
Wax collection of Islamic art https://minicomp.github.io/wax/collection/
Biology
AntMaps https://antmaps.org/
AntWeb https://www.antweb.org/
Arctos https://arctos.database.museum/
CalFlora https://www.calflora.org
CalPhotos https://calphotos.berkeley.edu/
COSMIC https://cancer.sanger.ac.uk/cosmic
Diatoms of North America https://diatoms.org/
eBird https://ebird.org/home
featherbase https://www.featherbase.info/en/home
iNaturalist https://www.inaturalist.org/
Madagascar Terrestrial Camera Survey doi/10.1002/ecy.3687
Movebank https://www.movebank.org/cms/movebank-main
NEON https://www.neonscience.org/
NEOTOMA https://www.neotomadb.org/
NOW https://nowdatabase.org/
Paleobiology Database https://paleobiodb.org/
Protein Data Bank https://www.rcsb.org/
VertNet http://vertnet.org/
Earth Sciences
Macrostrat https://macrostrat.org/
Soil grids https://soilgrids.org/
WorldClim https://www.worldclim.org/
USGS Current Water Data https://waterdata.usgs.gov/nwis/rt
LAGOS https://lagoslakes.org
EarthData https://search.earthdata.nasa.gov/search
Stone Lab Algal and Water Quality Laboratory https://ohioseagrant.osu.edu/research/live/water
Mauna Loa Trends in Atmospheric CO2 https://gml.noaa.gov/ccgg/trends/data.html
NASA GISS Surface Temperature Analysis https://data.giss.nasa.gov/gistemp/station_data_v4_globe/
North Temperate Lakes LTER https://lter.limnology.wisc.edu/data
Machine Learning
Machine Learning Repository https://archive.ics.uci.edu/ml/index.php
Physics and Astronomy
SIMBAD http://simbad.u-strasbg.fr/simbad/
Astrophysics Data System https://ui.adsabs.harvard.edu/
Political and Social Sciences
ICPSR https://www.icpsr.umich.edu/web/pages/
Various
Smithsonian Open Access https://www.si.edu/openaccess
Hathi Trust Digital Library https://www.hathitrust.org/datasets
GeoPlatform https://www.geoplatform.gov/
Omeka Project Directory https://omeka.org/classic/directory/

Table 13. Examples of databases that can be leveraged for research projects in CUREs.

  • Iterative CUREs enable students to access data from previous years and other lab groups and as such the opportunity to analyze realistic sample sizes that may be difficult or impossible to gather in the time of a CURE (Kloser et al., 2011; see Satusky et al., 2022 and Sun et al., 2020 for examples). The assembly of large datasets also enables long-term outlooks (e.g., Potter et al., 2009).
  • It is important to define “Scholarships as Conversation” for students. Scholars and researchers engage in ongoing conversations in which new ideas and research findings are continually being discussed (ACRL) (e.g., [Activities 5-6]).
    • Being part of the scholarly conversation is an expectation of academic research and is integral to research and therefore of a CURE.
    • The scholarly conversation takes place in the peer-review literature, including books and journal articles.
    • Scholarly conversations also take place at conferences.
    • The classroom is an environment for scholarly conversations.
  • There is an extensive literature on teaching the primary literature. There also exists multiple activities and exercises that enable instructors to introduce the primary literature to undergraduate students. Examples can be found in [Activity 7], Beck (2019), Hammons (2021), Howard et al. (2021), Chen (2018), Carson & Miller (2013), Hartman et al. (2017), Mitra & Wagner (2021), and Smith (2001) as well as the C.R.E.A.T.E. strategy (https://teachcreate.org/roadmaps/) and the science education resource center at Carleton College (Egger; Mogk). The collaborative reading and annotating of articles from the primary literature by students can also be helpful (Cafferty, 2021).
  • Providing guidelines and help on exploring the primary literature through formal (or informal) bibliography exercises (e.g., [Activity 8]) helps students find additional resources on their own (e.g., [Activity 9]).
  • Encourage students to identify the consensus that have developed, but also competing perspectives and approaches, and how new voices and evidence emerge (e.g., [Activities 9-11]).
  • Do not overlook the importance of explaining to students how to evaluate and select sources as well as how to provide citations. Many students are not familiar with the process of searching for references (video).
  • You should also engage students to reflect upon best practices of literature review, including:
    • Using existing research to guide one’s own work (including on the issue of “tracing the scholarly conversation” [Activity 9]).
    • Considering context when evaluating sources. One may not be able to understand the value of a particular piece of scholarship unless they consider the broader conversation (see [Activity 10]).
    • Providing attributions.
    • Needing to learn the “language” of the conversation before being able to fully participate.
    • Acknowledging that joining in the conversation confers both rights and responsibilities.
    • Recognizing that one is most likely entering into an ongoing conversation and not a finished one
  • Are there freeware programs or programs with user-friendly GUIs that can be used by students to engage in the data analysis process without the need for extensive coding or other computationally difficult tasks (see Acuna et al., 2020 for an example; Zelaya et al., 2022 for an example in a CURE context)? An extensive collection of freeware programs for data visualization and analysis, network analysis, mapping, text analysis, etc. is available at https://guides.osu.edu/DH/digitalhumanities (see also Table 14).

 

Examples of freeware programs and programs with user-friendly GUIs that can help engage students in analyses

Select freeware programs enabling image analyses, statistical analyses, concept mapping, and visualization of data
Program name URL Type of analyses
ImageJ https://imagej.net/software/fiji/ Image analyses
PAST https://www.nhm.uio.no/english/research/infrastructure/past/ Statistical, time-series, and spatial analyses
jamovi https://www.jamovi.org/ Statistical analyses
VUE https://vue.tufts.edu/ Concept mapping
Google Jamboard https://jamboard.google.com/ Concept mapping / Interactive whiteboard
SwissADME http://www.swissadme.ch/ Drug discovery
UCSF Chimera https://www.cgl.ucsf.edu/chimera/ Visualization and analysis of molecular structures

Table 14. Examples of freeware programs and programs with user-friendly GUIs

that can help engage students in analyses.

  • It is important to reflect on the role of the CURE in the development of the research program of the researcher(s). CUREs can help enable the research program of a research group (see figure 2 in Sun et al., 2020), but careful planning and circumscribing of the role(s) of the CURE(s) are critical.

Box 3: Important points

  • CUREs should integrate responsible and ethical conduct of research into the student training.
  • Consider the use of published data, museum and research databases, data collected by collaborators, and data collected by previous iterations of CUREs.
  • Set-up checks and balances for student-collected data.
  • Formally training students in the reading and analysis of the primary literature is critical to their engagement with scholarship.
  • Build room in the course structure for failure and iteration; they are important elements of the learning process.

 

4. How will research progress be balanced with student learning and development?

Overview: The strength of CUREs is their combination of research and teaching in a unique pedagogical experience for both students and researcher-instructors; it is also an important source of challenges. The ideal CURE would involve a steady progress of the research process that is accompanied by student gains and learning. These pedagogical goals can sometimes conflict with the research advancing at a necessary pace. Similarly, the required validation and repetition steps of many research protocols may not be needed to fulfill many learning goals. The structure of a course itself with the associated time-constraints can place limits on repeating experiments, increasing sample size, and even fully exploring a particular dataset or question.

This section enables you to determine:

  1. The pace of the research process
  2. The appropriateness of lessons on replication and statistical power to address the limitations of time in a CURE
  3. The possibility to assign some analyses and research tasks to outside researchers
  4. The role of peer-review as a combination of the scholarship review process and a pedagogical feedback
  5. Inclusivity issues to consider alongside research progress and student development

Education starts here:

  • How will you make explicit to students the pace of the CURE?
    • Will there be class meetings dedicated to specific data collection efforts, analyses, group brainstorming, peer-discussions, instructor conferences, etc.?
    • What activities will be assigned to students as homework?
    • What is the schedule of writing of the different sections of the deliverable?
    • Can you identify weekly goals for the project?
    • When will students be prompted to reflect upon their work? When will they be asked to discuss their work with peers or mentors? When will they be asked to formally review and critique their work or the work of others?
  • What are the elements of the research process that will not be authentically included in the CURE?
    • Can any of those elements be modeled on smaller datasets by the students?
    • Can any of those elements be modelled on a different dataset of the same nature (published or not) as part of a class activity?
    • Can some complex analyses demanding high computational skills be demonstrated by the instructor or guests (e.g., a graduate student working on this type of analysis)?
    • Are there elements of the research that can be outsourced to the instructor or other researchers and merely explained or showed to students?
  • What activities and requirements could represent obstacles to the engagement of all?
    • Are some specific lab tasks, field work, data collection protocols, etc. in conflict with student accommodations?
    • Are there activities that take place outside of the normal class time but need to happen on a specific day or time (e.g., attending to something in the lab, recording field observations at specific events)?
    • Is the literature that students need to consult to complete the work accessible and affordable to them?
    • Are there graphical representations of data or media used as part of the research that are not accessible to all?
    • Are the tools and support necessary for the success of all available on campus, including IT resources?
  • How can you develop a supporting environment in which students explore research, manage their work, and fail safely?

Scholarship starts here:

  • Are there specific deadlines associated with the research project including proposal deadlines, abstract deadlines, theses deadlines, manuscript deadlines, and commitments to collaborators or research students (graduate or undergraduate)?
    • Is the CURE work necessary for another step of the research process outside of the classroom?
    • Is the CURE work integrated with the research of a graduate student or mentored undergraduate student?
  • Does the work that students will undertake in the CURE represent the work necessary to complete the research deliverable chosen? Alternatively, is the CURE research only one component of the deliverable?
  • What analyses, if any, typically reported in supplementary information, are necessary for validation but cannot be fully incorporated in the CURE?
  • Can the research questions and/or tasks be easily divided among groups or students?
  • Can you collect the literature relevant to the project (or a representative subset of it) ahead of the CURE to share with students?
    • Can you share with students during the CURE examples of papers being published that are relevant to the research?
  • What are important research concepts and analyses for the CURE that may require some explanation? Have you completed [Activity 1] yourself?

Research and Education together:

  • The milestones of the CURE should mimic the different elements of the process of scholarly work from inception to presentation/publication. Thus, students should produce deliverables that are equivalent (at least in format) to the deliverables of professional scholars (e.g., Bakshi et al., 2016; Gastreich, 2020; Ramírez-Lugo et al., 2021).
  • Peer-review and instructor feedback are models for the scholarly review process. Consider using the review framework of a prominent journal in your field or alternative rubrics (see [Activities 12A and 12B]) to guide students through this process and make peer-reviews valuable educational experiences.
  • A consistent structure helps instructor and students keep track of the research and learning processes. One such structure is presented in [Appendix 1]. The structure is designed in eight different pedagogical steps that lead to a research deliverable as an integration of the two aspects of the CURE. Other examples of course structures are presented in the published literature (Bakshi et al., 2016; Bell, 2011; Bowling et al., 2016; Carr et al., 2018; Chen, 2018; Colabroy, 2011; DeHaven et al., 2022; Hekmat-Scafe et al., 2017; Mills et al., 2021; Murren et al., 2019; Ramírez-Lugo et al., 2021; Satusky et al., 2022; Sewall et al., 2020; Sweat et al., 2018; Thompson et al., 2016; Waddell et al., 2021; Wilczek et al., 2022; Zelaya et al., 2020).
  • Dedicating class time to tutorials associated with formative assessments and/or discussions enables instructor(s) to verify that students are ready to undertake the homework and activities of the CURE.
  • Dedicating class time to group meetings enables the instructor(s) to check in on all students/groups of students and address concerns and difficulties. Such meetings should be structured and scaffolded to promote constructive conversations. An example of the possible structure for one of these meetings is presented in [Activity 13].
  • Field work can sometimes be implemented through asynchronous self-led field trips to overcome logistical obstacles (Washko, 2021).
  • Lessons and interventions on specific topics, including replication and statistical power, may help students understand issues and concepts in research that cannot be undertaken in the CURE itself and overcome misconceptions (Schwartz et al., 2004).
  • Conversations with staff members in the Office of Academic Enrichment, Disability Services, the University Libraries, and the Teaching and Learning Center (Drake Institute at the Ohio State University) can help instructors find compromise and solutions to the challenges of implementing research in the classroom.
  • The TILT framework (Winkelmes et al., 2016) enables instructors to increase transparency in assignments. Including explicit links between assignments and tasks as part of the purpose of the assignment helps make explicit the progress through the scholarly process. An introduction to the TILT framework is available here.
  • Transparency in the assignments can also be improved by working with students through “understanding assignments” activities. Such activities can be done individually or as a group and enable students to make sure they are meeting the expectations of the instructor without being impeded by jargon. It encourages students to think about what it is they are expected to do (examples of assignments available at UNC Chapel Hill).
  • Universal Design for Learning (UDL) is a framework that helps design teaching products and structures that can be used by all without the need for accommodations. UDL is a proactive approach to adaptation in the classroom and benefits all regardless of their needs for specialized designs. UDL enables instructors to meet the diversity of their classroom. It includes several key components reviewed in Burgstahler, 2013. UDL includes the implementation of different media, additional scaffolds, technology, etc. (Burgstahler, 2009; Griful-Freixenet et al., 2017), but also expands to other issues. Consider how you can equitably support students through flexibility in deadlines, meetings, and schedules. This can include letting students set their own work hours or provide students the opportunity to attend remotely certain meetings. Learn what users need by conducting check-ins for group access needs during lab meetings and surveying users about the overall accessibility of the course experience. Consult with accommodation services staff to determine how you can optimize research and teaching practices and spaces and how to make them more inclusive for individuals with disabilities and specialized needs. Advocate on behalf of students with disabilities by communicating with accommodation services staff and encourage students to reach out to those services to get the support they need and deserve.
  • Mentorship structure can involve peer-researchers and graduate students.
  • Any mentorship structure can benefit from direct feedback from students. Communication is key to the success of the CURE.

Box 4: Important points

  • Mini-workshops and group or class-wide activities on subsets of the dataset analyzed or published data comparable to those studied by the students can help students understand critical concepts that cannot be authentically explored in the CURE because of time.
  • CURE students are integral members of the research team and can benefit from intellectual or data exchanges with other members of the research team, including graduate students, postdocs, and undergraduates in mentored research experiences.
  • Consider presenting to students the trajectory of a research project from inception to publication, including the role of formal and informal peer-reviews of ideas, presentation conferences, and manuscripts.
  • Consider how universal design for learning can be implemented in the CURE to facilitate the engagement of all students in the experience.
  • Implement a mentorship model of the CURE students that fosters a safe environment for self-exploration and self-management (Palmer et al., 2015).

 

5. How will the research learning tasks be structured to foster students’ development as scholars?

Overview: The design of a CURE requires the incorporation of a carefully thought-through scaffold that enables students to engage with an intellectual challenge that is often beyond any they have encountered before. Well-constructed scaffolds are critical to the success of CUREs because they enable students to tackle appropriately challenging tasks (Lopatto et al., 2020) and support students through failures, leading to greater perseverance (Corwin et al., 2022) . Scaffolds are relevant to the design of the entire CURE, which should include a series of activities and assignments that guide students through the entire project (e.g., Bakshi et al., 2016; Delventhal & Steinhauer, 2020; Hills et al., 2020; Makarevitch et al., 2015; Peyton & Skorupa, 2021; Rennhack et al., 2020). They are also relevant to the design of individual assignments because they help guide students through the tasks of research. This section enables you to determine:

  1. The appropriate framework for the CURE assignments
  2. The structure of CURE tasks as individual or group assignments
  3. The activities that require detailed tutorials or mini workshops
  4. The teaching interventions and writing assignments necessary to guide student learning
  5. The role of reflective activities in the development of student deliverables
  6. How the learning activities will foster an inclusive experience

Education starts here:

  • What are the knowledge/skill bottlenecks students will encounter during the CURE? What tasks or expectations might represent obstacles?
  • What does group work contribute to this activity? Which activities require group work?
    • Is the amount of work simply too much for a single person?
    • Is the interaction among students necessary to generate answers?
  • How does students’ ability to work well with others factor into the goals of the CURE?
  • What does an authentic research team structure resemble in the field of research associated with the CURE? How can the CURE mimic this?
  • How will you design groups whose structure promotes inclusivity and constructive intellectual exchanges?
  • Are there activities that require tutorial or workshops to be explained?
    • Can students perform all necessary analyses for the CURE without particular training?
    • Will students need to be introduced to a particular software program to collect or analyze data? To construct figures or tables?
    • Can you feasibly guide individual students/groups of students through the experimental components of the CURE without prior training on a smaller/published/example dataset?
  • How are you introducing the expectations for graded assignments and deliverables to students? How do they know what is expected of them?
  • What are the necessary or helpful intermediate steps to completing all deliverables and graded assignments? What steps do you go through to complete these activities yourself? What about a graduate student? How would you guide a mentored undergraduate researchers outside of the course through this activity? Based on all of this information, ask yourself: “What would a novice do?”
  • How can reflection help students develop better deliverables? What is the role of reflection in the development of student deliverables?
  • Can a think-pair-share structure facilitate progress or foster the development of deliverables?
  • How can you design learning activities that foster an inclusive experience?

Scholarship starts here:

  • What are the different steps of the research process that students will be going through? What numbered elements of Figure 2 are necessary or, instead, need to be edited?
  • What foundational skills and knowledge of the field of research should be incorporated in the CURE? Does that include skills or knowledge that may not in fact be strictly necessary for the specific research question/hypotheses tested in the CURE?
  • Which of the CURE tasks is typically achieved by individuals in a research setting? By groups of people?
  • Are certain elements of reflection or of the scaffolding of the research project typically shared through publication (for example as appendices) in professional deliverables?

Research and Education together:

  • Instructors often do not consciously realize that the process of research relies upon underlying concepts and practices. Researchers familiar with the research process skip over some of these foundations in their daily engagement in scholarship. Instructors can help students become researchers by explaining the critical concepts underlying the practice of research ([Activity 3]).
  • Engaging in research can be daunting for students. Reflection workbooks and journaling can help students overcome the challenge of the research process, reduce their anxiety, increase their appreciation of research, and facilitate exchanges between students and instructors (Apgar, 2022). Research shows that training significantly improves students’ reflection skills (Devi et al., 2017).
  • The entire course would benefit from a scaffold that guides students (and instructors) through the learning process. Sabel (2020) provides an example of a framework to develop the scaffold of a course. Another example is showed by the Decoding the Disciplines framework (https://decodingthedisciplines.org/) introduced here. [Appendix 1] shows yet another example in which students, for each element of the CURE (e.g., searching the literature, data collection, data analysis, writing the material and methods) engage in a series of activities that lead them to produce a research deliverable (Fig. 2).
  • Group meeting agendas can help students prepare the meetings they will have with their teammates during class time. Example of questions and guidelines to prepare various group meetings are provided in [Activity 14].

 

This proposed scaffold guides students through the different steps of research and the composition of a research paper, including framework, bibliography, data collection, data analysis, material and methods, figures and tables, results, discussion, and introduction. Students engage in tutorials, scaffolding activities, group meetings preceded by an agenda activity, the writing of a deliverable, and the peer-review of deliverables in alternance with reflection activities.
Figure 2. Possible CURE scaffold. The colors denote the location of the work, the shape of the boxes whether the work is undertaken individually, as part of the research group, or class wide. Arrows show the flow of work through deliverables for which prompts are provided in [Activities 15-17].
  • There is an extensive literature on the design of groups emphasizing the need to carefully consider the group composition as well as the framing of the group work by the instructor(s). Best practices for group work are summarized in Wilson et al. (2018) and at https://lse.ascb.org/evidence-based-teaching-guides/group-work/
    • Consider carefully the size of the groups (Heller & Hollabaugh, 1992).
    • Consider carefully the gender and ethnic minority status of students in the composition of the groups (Adams et al., 2002; Micari & Drane, 2011).
    • Assigning specific roles to students is helpful to encouraging and structuring the conversation (Heller & Hollabaugh, 1992). Enforcing a rotation between the roles of the students, if possible, can be helpful. There are several strategies for distributing roles presented in the literature (e.g., Olimpo et al., 2016; Winkelmann et al., 2015).
    • Including discussions of the nature of intelligence, academic failure, systemic biases, as well as fostering a growth-mindset in students can greatly help students overcome social-comparison concerns (Micari & Pazos, 2014).
    • Setting group goals can also be helpful (Werth et al., 2022) and strategies for effective group work are important (Washington University in St. Louis). In addition to assigning specific roles, instructors should consider using group contracts [Activity 18], and peer evaluations (see [Activity 19], which have been showed to encourage student participation (Chang & Brickman, 2018), can facilitate the assessment of the contract compliance, and make easier the identification of conflicts and problems leading to a more rapid resolution. Group contracts and peer evaluations can be used as part of the grading scheme to determine individual contributions.
    • The Peer Assessment Factor is a quantitative assessment of the students by themselves and peers that can be used as a formative assessment and a guide for interventions.
    • Examples of other activities, and interventions are provided in the PETS Process Manual.
  • Frameworks for mini-workshops are presented in the primary literature for many tools and protocols. Some are specifically aimed at CUREs. A statistical workshop is presented by Olimpo et al. (2018). Sewall et al. (2020) presents several mini-workshops on computational tools (R and QIIME2). Alternatively, published protocols might be sufficient to guide students through an experiment or research task without the need for a prior tutorial or workshop (Acuna et al., 2020; Buser et al., 2020; Chen, 2018; Craig, 2017; Goeltz & Cuevas, 2021). Several published CUREs include within the supplementary information files useful protocols and tutorials (e.g., Bucklin & Mauger, 2022; Jurgensen et al., 2021; Poole et al., 2022; Roberts et al., 2019; Zelaya et al., 2022).
  • There are numerous published strategies to help scaffold inquiry activities such as authentic research. Several are summarized in a TLRC teaching topic. Additional information on scaffolding is provided in Killpack et al. (2020).
  • Writing-to-Learn (WTL) activities can help students build their understanding of the existing literature on the CURE topic, work through their project, and complete their deliverables (Balgopal et al., 2018; Bangert-Drowns et al., 2004; Fry & Villagomez, 2012). Many examples of WTL activities are available online (see in particular the Center for the Study and Teaching of Writinghttps://cstw.osu.edu/writing-learn-critical-thinking-activities-any-classroom). A selection of scaffolding activities derived from WTL concepts that can be implemented in a CURE are presented here, including an activity guiding students through the process of deciding on their data analysis protocol [Activity 20], an activity helping students summarize the existing literature [Activity 21], two sister activities helping students compare their findings to data from the primary literature [Activities 22-23], and an activity meant to help students prepare the discussion section of their paper [Activity 24].
  • Tutorials for writing can be developed for each section of the deliverable, including the elements of an IMRaD paper (Introduction, Methods, Results, and Discussion). An example of the structure of such tutorials are showed in [Activity 16]. An example activity to develop students’ graphical skills is showed in [Activity 25].
  • Other activities can be designed to scaffold the students’ work including activities encouraging students to predict their results and represent them graphically or activities comparing the introduction and discussion from a single publication to emphasize their roles as bookends of the deliverable.
  • Figure 2 shows a possible framework by which reflection can be incorporated in the scaffolding of the course to enable students to build their metacognition while they engage in scholarship (see Denke et al., 2020). Student reflection can be guided by a set of questions and an information literacy framework. An example of such guide is provided in [Activity 26].
  • Best practices to build student metacognition are reviewed in Stanton et al. (2021) and at https://lse.ascb.org/evidence-based-teaching-guides/student-metacognition/.
  • Think-Pair-Share activities, whereby students first reflect upon their work and then discuss their thoughts with group members before to engage with the rest of the class and the instructor(s), may be helpful in promoting reflection as well as conversations among students; they can also be used as formative assessments by the instructor(s) to improve student performance (Akhtar & Saeed, 2020). Best practices for this active-learning activity have recently been reviewed in Cooper et al. (2021) and Prahl (2017).
  • The TILT framework (Winkelmes et al., 2016) enables instructors to increase transparency in the assignments by providing clear descriptions of the purpose, tasks, and evaluation criteria for each activity the students engage in. Examples of applications of this framework are provided at: https://tilthighered.com/tiltexamplesandresources.
  • There are numerous sets of best practices and recommendations (e.g., Cooper et al., 2020b; Faulkner et al., 2021; Linder et al., 2015; see question 10 on p. 104) that have been developed to promote an inclusive classroom environment (whether in a CURE or not). Miller et al. (2022) presented a framework to incorporate “antiracist, just, equitable, diverse, and inclusive principles” in the design of a CURE. Best practices for inclusive teaching are presented in Dewsbury and Brame (2019) and at https://lse.ascb.org/evidence-based-teaching-guides/inclusive-teaching/.

Box 5: Important points

  • The prior knowledge of the students informs the first rungs of the learning scaffold.
  • The ELOs for the course provide the framework for the endpoint of the scaffold.
  • Group work should be intentionally designed, supervised, and assessed to lead to accountability and personal learning milestones.
  • The scaffold adopted should provide students all required background information, guide them through challenging tasks in an accessible context, provide opportunities for reflection, and enable students to repeat or expand on their work to advance the research process.
  • Writing-to-learn activities can facilitate student learning and development as writers.
  • Information literacy is critical to student success and can be fostered through reflective work as well as activities promoting metacognition.
  • The Transparency framework of TILT enables instructors to articulate the purpose, tasks, and criteria for evaluations of each activity to students, thereby promoting metacognition and reflection.

 

6. How will students communicate the results of their research?

Overview: The communication of research findings is critical to the process of scholarship, including undergraduate-led research (Spronken-Smith et al., 2013). It is also an important element of CUREs. Many traditional outlets of professional scholars are also appropriate for student research derived from CURE. Thus, presenting research findings at conferences has been showed to greatly benefit undergraduate students (Little, 2020). Publishing the findings of a CURE on a database used by researchers has been demonstrated to lead to increased student motivation (Wiley & Stover, 2014). Finally, publications resulting from CUREs have been associated with both personal and professional benefits for students (Turner et al., 2021). In all forms of deliverables, it is important to consider how the contributions of all members of the CURE will be recognized. Research has showed that there exists a gap between what an instructor or professional researcher might consider necessary for authorship versus what students consider necessary (Turner et al., 2021). This section enables you to determine:

  1. The appropriate mode(s) of dissemination of the CURE’s findings
  2. The division of labor among students in preparing the deliverable
  3. Authorship rules

Education starts here:

  • What is an appropriate expectation in terms of deliverable given the time constraints of the CURE?
  • Would a specific deliverable be more motivating for students? Have you asked students?
  • Is one format of deliverable going to benefit students professionally more than another?
  • Which mode of dissemination would be appropriate for a professional scholar at this point in the research? Would that be any different for a student participating in a mentored research experience?
  • Will the deliverable be evaluated for a grade?
  • Will each student be expected to produce a final document (be it a dataset, poster, manuscript, or proposal) or will the deliverable be prepared by a group of students?
    • Will each student prepare one section of the deliverable? Alternatively, will all students be contributing to all sections of the document?
    • Will all students in the group be evaluated as a group with a single grade for the document applying to all students? Alternatively, will the students receive an individualized grade?
    • If students will be individually graded, what criteria will be retained to differentiate between group members? Are those going to impact authorship order?
  • How will the intellectual and practical contributions of the students participating in the CURE be recognized in products of the project?

 Scholarship starts here:

  • What deliverable are you expecting/hoping for at this stage of the research?
    • Are the students completing a project leading to a publication manuscript? Will the students produce a version of the manuscript near-ready for submission or will there be large amounts of work undertaken after the conclusion of the course to bring the manuscript to a submittable form? Who will undertake this work?
    • Are the students helping start a new project, developing preliminary results for a proposal or contribute to a database?
  • Will the findings of the CURE be appropriate for a conference presentation?
  • What are the standards for manuscript authorship in your field? At your institution? In your research group? Alternatively, when are contributors merely acknowledged?
  • How will the contributions of instructors and outside researchers be considered for authorship?

Research and Education together:

  • Different deliverable formats are not mutually exclusive.
  • The format of the CURE deliverable should reflect the authentic mode of delivery of research findings in the field (Kloser et al., 2011).
  • Students enjoy the opportunity to present the results of their research at conferences and/or publish their work (Amir et al. 2022).
  • Participation in research conferences has been showed to benefit students both by helping students develop their communication and research skills, but also by positively impacting their career and engagement in other activities (Little, 2020).
  • The opportunity to publish the findings of the CURE enhances student motivation (Wiley & Stover, 2014).
  • Publications can be powerful deliverable for a CURE because:
    • They can act as scaffold with the different sections of a manuscript corresponding to elements of the research process
    • They represent the scholarly endeavor and can help students gain an appreciation for the process of research.
    • They can help build the students’ identity as researchers through co-authorship of a published document.
    • They positively impact the careers of students (Turner et al., 2021).
    • They can be motivating and thus increase student engagement in the CURE.
    • They provide a tool for accountability among students and with the instructor.
  • Consider how the participation in a national or university-wide CURE program may dictate the deliverable if the CURE is not researcher-driven.
  • Undergraduate research symposia can be great venues for the students to present the results of the CURE. Many colleges and universities as well as some departments at larger institutions organize symposia at least once a year (Presentation Opportunities).
  • There are also numerous national conferences dedicated to student research (e.g., SigmaXi; Council on Undergraduate Research).
  • In addition to discipline-specific publication venues and journals aimed at professional scholars, there are also open-access online undergraduate research journals (Sun et al., 2020) that may be appropriate outlets for the CURE’s findings.
  • The Knowledge Bank of Ohio State (https://library.osu.edu/kb) enables the archiving of intellectual outputs of the university community in an accessible digital format. Other digital repository, some discipline-specific, may also be appropriate (e.g., the Environmental Data Initiative)
  • There are online repositories of research products that are citable and can be appropriate for the results of CUREs (e.g., https://figshare.com/).
  • Digital and artefactual exhibits can also appropriate venues for the presentation of archival and object or specimen-based research (e.g., Donegan et al., 2022; ENG5612 at Ohio State).
  • Multiple online scholarly projects and databases (Table 13) may be appropriate venues for student contributions (e.g., Map of Early Modern London).
  • Establishing clear authorship guidelines is critical to a smooth and rewarding experience:
    • Students should buy into the rules. It is important to explain the rationale for authorship rules to students.
    • Students may also have valuable input on the rules. Consider developing authorship rules specific to the CURE, with the students, that incorporate guidelines and recommendations from relevant institutions as well as best practices in the CURE’s field of research.
  • Many universities and colleges have developed authorship guidelines including The Ohio State University (Authorship Guidelines) and other institutions (e.g., Harvard University, University of Michigan, Yale University).
  • Many professional societies have also developed best practices documents relating to authorship (e.g., American Physical Society, American Psychological Association, British Sociological Association).
  • Some journals and publishers also have rules for authorship (e.g., Nature, British Medical Journal, Taylor & Francis).
  • If other instructors or researchers (e.g., graduate students, peer teaching assistants, research collaborators at other universities) will be involved in the project, they should be part of the conversation and agreement on the authorship rules. Consider the authorship rules you employ in your research group.
  • There are several published sets of guidelines for authorship that can also be used. The CRediT taxonomy (Honoré et al., 2020), which has been adopted by PLOS journals (PLOS Authorship Guidelines), can provide a starting point to develop guidelines specific to your CURE.
  • The copyright services department of the Libraries (https://library.osu.edu/copyright) can help with issues of students’ intellectual property rights.
  • In the case of CURE-specific requirements for authorship, consider whether or not it is appropriate to tie authorship to course requirements and achievements (e.g., fulfilling the contract if contract grading, obtaining a certain minimum grade on the deliverable).
  • Students should be offered the opportunity to approve the final version of the manuscript they co-author, even after completion of the course.

Box 6: Important points

  • Different deliverable formats are not necessarily mutually exclusive.
  • Establishing clear guidelines and requirements for authorship is critical.
  • Communicating the results of the CURE should reflect authentic scholarly communication in the field.
  • Publication manuscripts can serve as scaffold, provide motivation, ensure accountability, and shape student identity.

 

7. How will the progress and experience of students be assessed?

Overview: The assessment of the students’ work in a CURE enables the instructor and the entire class to stay up to date on the research progress, including the findings of the project and the setbacks. Well-designed assessments also enable students to reflect on their learning, help them identify their difficulties, and provides tools for overcoming obstacles. Because CUREs are elements of formal courses, it is also often necessary to evaluate the students’ work qualitatively and quantitatively for the purpose of a grade. This section enables you to determine:

  1. The goals of assessing the CURE work
  2. The mode of grading to adopt in the CURE
  3. The proportion of the evaluation represented by group work and individual work
  4. The proportions of formative assessments and summative assessments
  5. The roles of instructors, peers, and self in evaluation.
  6. The nature of the assignments graded
  7. The need for specific rubrics and grading criteria
  8. The importance of an inclusive mode of grading

Education starts here:

  • Does all evaluation work have to translate to a grade?
    • Can some assignments be evaluated simply for completion/genuine participation?
    • Can some scaffolding assignments be reviewed to provide feedbacks without associating the work with a grade?
  • What are the goals of the grading in the CURE?
    • Will it be used to determine authorship eligibility?
    • How will the grading be aligned with the expected learning outcomes (ELOs) for the CURE?
  • Which assignments will be graded?
    • When you assess student performance, what are you rewarding? Is the grading based on the mastery of the ELOs or on other criteria? Are those criteria made explicit to students?
    • Are students given a chance to revise their work and correct their mistakes?
    • How do you give students practice with and feedback on performance items that you are rewarding on the final assignment?
    • Do you give students the opportunity to reflect on what they are learning or how they are growing?
  • What percentage of the class grade will be represented by the CURE grade?
    • How does this compare to the time investment made by students?
  • Will group assignments be graded or will only individual assignments be graded?
    • When group assignments are graded, do all members of the group get the same grade?
    • Does the mode of grading lead to increased intra-group conflicts?
  • What mode of grading will be adopted in the CURE?
  • What proportion of the final grade for the CURE is represented by formative assessments versus summative assessments?
  • Are summative assessments based on the formative assessments on which the students got feedback?
  • Is there room for self-evaluation in the CURE? What about peer-evaluation beyond peer-reviews of deliverables?
  • Can students grade some of their own assignments with the help of a rubric as a metacognition tool? Can this exercise be combined with instructor feedback and a redo enabling students to gain missed points?
  • How often will students be completing work reviewed by the instructor(s)?
    • How often do students get instructor feedback?
    • How many iterations of a given document, section of deliverable, research product will students be producing?
  • Will students get rubrics for each assignment?
    • How will explicit grading criteria be provided?
    • Is it possible to hand out examples of successful products to students as models and examples of works that do not meet expectations to guide their own writing?
  • What approaches can you adopt to ensure an equitable and inclusive mode of grading? How can assessments be used to reduce the equity gap?

Scholarship starts here:

  • How are scholars evaluated in your field? Can this mode of evaluation be mimicked in the CURE?
  • What determines the quality of a deliverable in your field and what aspects of that document quality do students have control over?
  • What is a reasonable quality standard to expect for the deliverable of choice? Is it realistic to expect the students to produce a document nearly ready for publication submission or presentation?
  • How would you assess the work of a mentored research student engaged in a similar project? What expectation would you have of this student’s deliverable?
  • How will the success of the research impact assessment and grading?

 Research and Education together:

  • The nature of CUREs as authentic research may lead to failures and mistakes; in fact the novel aspect of the CURE will almost certainly lead down unpredictable paths and outcomes. Designing the assessment of the CURE to mitigate the effects of technical problems, negative results, and the steep learning curves of scholarly endeavors is critical to student engagement. The goal of assessments should be to inform students and instructors alike of the progress of students as researchers, not the success of the observations or experiments. The ability to troubleshoot, the resilience of students in the face of failures, the understanding and explanation of errors, and the repetition of tasks should be integrated in the grading scheme. Finding the solution to a problem and understanding why an experiment did not work are progress. A discussion of failure, how to approach failure with students, and the point of view of the instructor in the context of CUREs is provided in Townsend (2022).
  • Backward-design requires the alignment of the assessments with the ELOs for the CURE. Instructors should rigorously verify that all ELOs for the CURE are assessed by the end of the course, but also that assessments do not evaluate expectations that are not made explicit to students. Different assessment modes are suitable for diverse learning goals (Verb wheel).
  • Transparency in assessment is critical to student success and inclusive teaching. The TILT framework (Winkelmes et al., 2016) presented earlier in this document is particularly helpful: https://tilthighered.com/tiltexamplesandresources.
  • Beyond traditional point-based grading, there are other modes of grading that an instructor can consider.
    • Criterion-referenced grading provides students with flexibility in the weights of the various components of the course. It enables students to mitigate test anxiety and emphasize formative assessments; it can also be used by students to lessen the consequences of out-of-school responsibilities on their academic endeavors. In practice, you should determine bounds for the weights of the different categories of course assessments and enable students to develop their own formula for the CURE grade. Every student will be graded according to their unique formula (which can easily be setup in a spreadsheet program).
    • Contract grading (Inoue, 2019) offers the opportunity to combine flexibility for students and for instructors while maintaining rigor in expectations and avoiding conflicts over grades. Contract grading is a format of grading that does not involve points or letter grades, apart from the final course grade. At the start of the semester, students choose/agree to/sign a contract that sets their path for the semester. Each contract lists the work required of the students for a particular outcome (e.g., an A, a B). If the student satisfactorily completes the work associated with their contract, they will earn that grade. Contracts are often associated with a mid-semester conference to check on progress towards the completion of the contract and reevaluate commitments if necessary. This conference is repeated at the end of the semester to assign the students’ grades. The onus of tracking contract compliance may be placed on the students who are responsible for writing self-evaluations ahead of each conference in which they need to address their work, its quality, their engagement with the class, etc. You should respond to these self-assessments during the conference. You can always disagree with the student in their final assessment of their performance. Students can also be asked to keep track of their time using a labor log. An example of contract grading for writing courses is provided by Inoue (2019). Another example been published online by Cathy N. Davidson. Contracts can be more or less complex incorporating aspects of criterion-referenced grading (Hiller and Hietapelto 2001) or including specifications grading elements (Lindemann and Harbke 2011; see Appendix 2 for a model applied to CUREs).
    • Specifications grading can help uphold academic standards and motivate students while reducing grade anxiety and cheating. In specifications grading, the assignments are graded on a pass/fail basis, a check/check minus/check plus/unsatisfactory basis, or an excellent/meets expectations/needs revisions/fragmentary basis. Instructors should provide clear specifications of what constitutes an acceptable work, what qualifies for a check (or earns a check plus), what leads one to meet (or exceed) expectations, etc. Students may also be given the opportunity to revise their work. Assignments may be assessed individually or grouped into modules that are assessed holistically. Modules can be weighed to reflect the complexity of the tasks, their relevance to learning outcomes, and/or their status as formative/summative assessments. Several models of specification grading for different fields have been published (e.g., Blackstone & Oldmixon, 2019; Carlisle, 2020; Elkins, 2016; Howitz et al., 2021).
  • Group reviews can be used to assess the contributions of all members of a group to the outputs of the CURE, including specific documents. A group review template is provided in [Activity 19]; others exist in the literature (e.g., Bucklin & Mauger, 2022; Waddell et al., 2021).
  • Self-evaluation, also called self-assessment or self-grading can be a powerful tool to promote student pacing and success, primarily as part of formative assessment (Andrade, 2019).
  • Many rubrics have been published online (University of Minnesota, Chicago State University) and in the peer-reviewed literature for both writing assignments and CUREs (Bakshi et al., 2016; Kishbaugh et al., 2012; Lee & Le, 2018; Murren et al., 2019; Ramírez-Lugo et al., 2021; Sewall et al., 2020; Waddell et al., 2021).
    • An example of a rubric for discussion boards in provided in Appendix 3.
    • An example rubric for a manuscript-type deliverable is provided as part of the peer-review presented in [Activity 12].
    • Bucklin & Mauger (2022) as well as Merrell et al. (2022) both include rubrics for a poster presentation stemming from CUREs.
    • An example rubric for a grant proposal-type deliverable is provided in Rennhack et al. (2020).
  • When designing a rubric, consider the following best practices (from Bean, 2011):
    • Numbers on the rubric do not add up to 100 and do not represent directly course points.
    • The grade associated with the rubric is always presented as a letter or a check/check plus/check minus, not a number.
    • Students are explained how the rubric is used in grading.
  • A good rubric should include a detailed explanation of the task students are expected to perform, the characteristics of the work that will be evaluated, the levels of mastery that will be considered, and descriptions of the characteristics for each level of expertise. Additional guidelines for the development of rubrics can be found in a number of different publications (Boston University, Brown University, and Arizona State University). You can also consider co-creating the rubric you will use with the students (University of Colorado Boulder).
  • Highly structured courses with frequent low stakes assignments increase student engagement (Cavinato et al., 2021), lead to higher performance (Freeman et al., 2011), and help reduce the equity gap (Eddy & Hogan, 2014; Haak et al., 2011).
  • Consider using concept inventories to test for knowledge acquisition in the field of research (Table 6).
  • (More) authentic formative assessments can be undertaken by reviewing and grading lab notebooks, periodic research updates akin to those a research student would provide in a lab/research group meeting, conference-style abstracts, elevator speeches, or chalk board presentations.

Box 7: Important points

  • Assessments should be aligned with ELOs for the CURE.
  • Rubrics and clear criteria for success help communicate the expected level of mastery to students.
  • Student practice and instructor feedback on formative assessments should be aligned with summative assessments.
  • Contract grading, specifications grading, peer-grading, and self-assessment are valid alternatives to traditional numerical grading by instructor.
  • Frequent low-stakes assignments can help reduce the equity gap.
  • Accountability for personal tasks and peer/self-review facilitate grading of group work.

 

8. How will  research learning tasks change as discoveries are made and initial research questions are answered?

Overview: The process of research is inherently iterative and involves sequential hypothesis testing. As results emerge from observations, experiments, and analyses, new hypotheses are developed and require testing. This is often the case in CUREs. Unlike a scholar’s research program, however, CUREs have strict curricular goals and involve time limits, including a comparatively small number of weekly hours of research engagement and an end to the research process imposed by the end of the instructional period be it a quarter, semester, or other. This section enables you to determine:

  1. Whether to include such sequential aspect of research within a single implementation of a CURE or across repetitions of a CURE.
  2. How to bring about the transition from one hypothesis to another
  3. The need to revise the scaffold and learning tasks of the CURE to fit new hypotheses and research paradigms
  4. The importance of planning the research course of the CURE to offer an original experience to successive cohorts of students

Education starts here:

  • Will students have time to explore only one hypothesis/question, or will they be able to at least partially engage in a second one?
  • For students exploring researcher-chosen questions/hypotheses:
    • Will they be able to suggest their own follow up hypothesis/question?
    • How will you explain to them how the question/hypothesis they are investigating was developed?
  • When revising or repeating a CURE:
    • What elements of the CURE are outdated? Any recommended paper or analysis no longer reflecting the knowledge or best practices in the field?
    • How will you ensure that the upcoming CURE is not merely a variation on a previous iteration, in which all novelty has been lost?
    • What should you change to make sure that the CURE does not become a “cookbook CURE”, a research experience for which the approach and protocols are fully developed and the result guaranteed, leading students to engage in research novel only to them that does not significantly contribute to the field?
    • Is the new iteration of the CURE following-up on questions or hypotheses developed in previous versions of the course?
    • What advances in the field have been made since the last time the CURE was run? Should planned questions/hypotheses be revised as a consequence?
  • Have new best practices in pedagogy and CURE implementation been published since the last iteration of the CURE? Do they mandate revisions to the structure, scaffold, or assessment of the CURE itself?

Scholarship starts here:

  • Does the scope of the CURE represent a publishable manuscript in the field of research of the CURE? Will more than one iteration of the CURE be necessary to answer all necessary questions/test all necessary hypotheses?
  • Are adjacent or related questions/hypotheses being tested by colleagues, collaborators, research students, thus limiting the scope of the work the students can follow-up on?
  • When developing a new researcher-driven CURE:
    • What is the place of this CURE within the research program of the researcher(s) involved?
    • What are the natural follow-up hypotheses or questions already known?

Research and Education together:

  • It is critical to introduce students to the nature of the scholarly endeavor including the fact that research generates more questions than answers and the need for researchers to critically select the questions that they will in fact explore. This discussion can take place as part of the process to select the question investigated in the CURE or not (see [Activity 2]). The roles of funding agencies, other scholars, institutions, and systemic biases in this selection should also be presented. Some aspects of these issues are discussed in https://opentextbc.ca/researchmethods.
  • It is possible to demonstrate to students the nature of research and the succession of investigations that lead scholars to build piece-by-piece the puzzle of a particular topic without requiring students to walk these steps themselves through retrospectives and walkthroughs of the history of the current paradigms and questions they will explore.
  • CUREs can build on each other over time to enable the investigation of particular topic through a series of studies (Satusky et al., 2022; Sun et al., 2020: figure 2). It is important to incorporate this history when presenting the CURE to students.
  • Extensive notetaking during the design and implementation of the CURE enables an instructor to revise the CURE. There are published guides to revising courses and reflecting on pedagogy (see for example McGahan, 2018).
  • Instructors teaching researcher-independent CUREs should consider following the research advances in the field of the CURE as they do their scholarship area to maintain up-to-date knowledge of the field, active areas of research, literature reviews, and methodological developments.

Box 8: Important points

  • Just like a research program, a CURE changes over time following research inquiry, progress, and setbacks.
  • Showing/explaining the development of the CURE overtime to students is integral to showing/explaining the research process.
  • Just like the work of students during the CURE is iterative, so is the work of (re)designing the CURE itself.
  • Beware of creeping away from a CURE towards a “cookbook lab” with successive iterations.

 

9. What are the logistical obstacles and solutions for the different steps of the CURE?

Overview: The implementation of a CURE requires students to access the tools of the research trade. These are discipline-dependent, but may include lab space, consumables, specific technologies or equipment, library and documentary resources, research specimens, and computing facilities. Certain CUREs may also involve field experiences, which introduce their own set of logistical challenges. Planning the needs of the CURE is critical to its success and may influence the nature of the research questions investigated with students. This section enables you to determine:

  1. The resources necessary for data collection
  2. Alternative sources of data
  3. The needs of the data analysis
  4. The potential for crowd-sourcing the CURE’s support
  5. Possible sources of funding to support CUREs

 Education starts here:

  • What is the budget of the course?
  • Is the CURE part of the broader impacts of a grant?
  • Is the CURE part of a creative teaching or curriculum redesign effort that can receive funding or logistical support from the institution, professional societies, or funding agencies?
  • What help can support staff, including lab technicians, research and teaching assistants, lab coordinators, librarians, IT staff, and museum staff, provide with data collection and analysis? What do they need to be able to help?
  • Who can help me think about my CURE development process and the pedagogy of this model of research?

Scholarship starts here:

  • What are the consumable needs of the CURE?
  • Do loans of specimens need to be secured prior to the start of the CURE? What are the restrictions on the handling of research specimens and materials?
  • What equipment, including laboratory equipment and computational resources inclusive of hardware and software programs, is necessary to not only collect and analyze the data, but also prepare deliverables? Is this equipment accessible in teaching or research facilities on campus?
  • What permits, certifications, and approvals are necessary for students and instructor(s) to undertake the research?
  • What existing databases and online datasets of images, observations, measurements, etc. can be leveraged to facilitate the data collection for the CURE?
  • For researcher-driven CUREs:
    • Can the CURE be integrated with other data collection efforts of the research group? What are the benefits for the research students involved in contributing their data to the CURE?
    • What is the role of research collaborators in the project?
  • Are there sources of funding associated with the research ongoing in the laboratory/research group that can legitimately support the research undertaken in the CURE?

Research and Education together:

  • Although the cost of a CURE has been reported to be higher than that of a traditional introductory science laboratory (Rodenbusch et al., 2016; Spell et al., 2014), it is much lower than the cost of mentored research experiences or summer research experiences (Rodenbusch et al. 2016; Smith et al., 2021). Some estimates are around $400 to $500 per student per course and many CUREs are cheaper (Poole et al., 2022) or even approach no financial costs.
  • Online databases (Table 13) and freeware programs (Table 14) enable data collection and analyses at low costs.
  • Collaborators and colleagues may be able to share equipment and supplies at low costs. CUREs can be departmental/institutional resources that spur enrollments and raise the profile of the teaching/research unit. Discuss with stakeholders (e.g., department chair, associate dean, and colleagues) the benefits of supporting the CURE.
  • Many institutions have some equipment and resources (e.g., supercomputer, computer labs, greenhouse space, imaging facility) whose costs and access are mutualized or free for in-house projects.
  • The integration of research and teaching missions of the CURE may enable different sources of funding to support the work (e.g., CUREnet). Those may include grants from professional societies, government agencies, internal competitions at the host institution, etc. (e.g., Council on Undergraduate Research). There are also calls for proposals appropriate for CUREs (NSF Division of Undergraduate Education). Some network CUREs may provide seed funding to implement the course (DeChenne-Peters & Scheuermann, 2022).
  • Trainings, permits, and approvals should be secured ahead of the CURE as much as possible. Student-specific trainings should be integrated in the course. Logistical obstacles of research are important part of the curriculum. Students learn of the realities of the process of research and the important legal and ethical regulations that are associated with it. Additionally, students who navigate through these obstacles may gain important skills in project management.
  • The involvement of research group members in the CURE should be designed to benefit them in the form of co-authorship, mentoring experience, opportunities for career advancement, funding, etc.
  • Teaching assistants, including graduate and undergraduate students, have been showed to be very helpful in supporting undergraduate students enrolled in CUREs (Olson et al., 2022). They should receive necessary training to learn how to teach a CURE (Kern & Olimpo, 2022)
  • There are multiple offices and resources at the Ohio State University to support the development of CUREs that are summarized in Table 15.

 

List of workshops, pedagogical resources, and offices that can assist instructors in the development of a CURE at the Ohio State University.
Table 15. Resources at the Ohio State University for the development of CUREs.

 

Box 9: Important points

  • Obstacles to data collection can be overcome with research databases, citizen science data, museum databases, and online sets of images and observations or measurements.
  • Campus resources including IT, the Libraries system, as well as research lab members and collaborators can help overcome data collection and analysis obstacles.
  • CUREs can be integrated with mentored student research (of graduate or undergraduate students) to facilitate funding, mentoring, data collection, data analysis, and the professional development of research students.

 

10. What are the roles of instructional and support staff?

Overview: Although it may be tempting to think of the instructor (or instructors) as the linchpin of both the research and pedagogical processes of the CURE, the structure of CUREs is inherently student focused. As such, CUREs provide the opportunity for instructors to rethink their identity in the classroom to include a critical mentorship component built around providing a supportive environment in which students are assisted in acquiring knowledge and skills, but also learn the significance of their work (Figure 3).

 

Research mentors should promote project significance and explicit tasks, foster a supportive environment, and enable high student efficacy.
Figure 3. The three components of student attitudes that should be fostered by research mentors (redrawn and modified from Ambrose et al., 2010).

A successful CURE is easier to implement with the support of other members of the research team, educational support staff, or campus community. A researcher-independent CURE may rely on the resources and structure of a national program, but the success of the CURE will also likely involve campus members who can complete and enhance the work of the instructor. This section enables you to determine:

  1. The role(s) of the instructor(s) in mentoring the CURE
  2. The fostering of a supportive, motivating environment
  3. The role of the instructor and other experienced scholars in the research
  4. The expectation and needs of graduate or peer-teaching assistants
  5. The support provided by lab managers, technicians, and other staff members

Education starts here:

  • How can the experience of CURE students be improved through interactions with students who have previously completed the course, mentored undergraduate researchers, graduate students, or postdoctoral researchers?
  • What challenges are students likely to encounter during the CURE?
  • What are the obstacles to the mentor-mentee relationship you foresee for the CURE? Consider expert-novice gap, communication skills, and cultural differences as well as “social-distance” (Shanahan, 2018).
  • What is the ability and willingness of instructors to commit time to mentoring?
  • What is the prior experience of the instructor(s) in mentoring research?
  • What training and support is necessary to foster the success of postdoctoral researchers, graduate student instructors, or peer undergraduate assistants as educators?
  • How do the ELOs of the CURE impact the model of mentoring that instructors should approach?
  • What are the expectations for postdoctoral researchers, graduate student instructors, or peer undergraduate assistants of the instructor-of-record? Of the department/college/graduate school/HR/ etc.?
  • What are the ELOs as well as professional and personal goals of postdoctoral researchers, graduate student instructors, or peer undergraduate assistants?
  • How can the mentoring of postdoctoral researchers, graduate student instructors, or peer undergraduate assistants by experienced researchers and teachers help them reach their goals?
  • What are the roles of the support staff and outside researchers in the CURE? Can these professionals free up instructor time for mentoring by assuming some technical responsibilities?

Scholarship starts here:

  • What is the commitment (time and effort) that the instructor will dedicate to the research, particularly in class, that cannot be directed to mentoring?
  • What is the responsibility of postdoctoral researchers, graduate student instructors, or peer undergraduate assistants in the research tasks of the CURE?
  • What training and support is necessary to foster the success of postdoctoral researchers, graduate student instructors, or peer undergraduate assistants as researchers?
  • What mentorship model(s) do you adopt in mentored student research experiences in your own research group and how can it/they be transposed to the classroo

Research and Education together:

  • CUREs require a wider vision of the role of the instructor than traditional classrooms because of the need for emotional and research support of students (Cooper et al., 2022; Goodwin et al., 2021; Linn et al., 2015; Shortlidge et al., 2016).
  • Different models of the role(s) of instructors in research experiences have been developed, but the consensus is that mentorship covers five critical categories of skills: research, diversity/culture, interpersonal, psychosocial, and sponsorship (Gentile et al., 2017). These five branches of mentorship lead instructors to promote the development of adaptive attitudes in students.
  • Instructors can encourage constructive attitudes through (1) framing and scaffolding (designing transparent tasks associated with an explicit awareness of both the pedagogical and research significance of the work), (2) interventions and pedagogical activities (promoting high student efficacy), and (3) explicit integrations of inclusive teaching practices (creating a supportive learning environment).
  • Just like explicit expectations and relevance of CURE tasks need to be communicated to students, clear expectations and ELOs should be developed for the work of postdoctoral researchers, graduate student instructors, or peer undergraduate assistants who are engaged in the CURE. These goals should incorporate the personal and professional goals of the postdoctoral researchers, graduate student instructors, or peer undergraduate assistants (see Mabrouk, 2003).
  • The TILT framework (see above) can be useful in framing both student activities and postdoctoral researchers, graduate student instructors, or peer undergraduate assistants’ tasks.
  • Instructors should strike a balance between “being overly prescriptive, which inhibits creativity and agency, and being insufficiently supportive, which leads to uncertainty, frustration, and a sense of failure” (Hanauer et al., 2012:384).
  • Many activities and learning exercises are helpful in engaging students in experiences that promote student efficacy, including:
    • Sharing their personal goals for the CURE and expressing their expectations for the CURE.
    • Participating in the social network of their research team through meetings, discussion boards, etc.
    • Explicitly associating elements of the research process with valuable skills and attitudes for their career aspirations.
    • Articulating hypotheses and questions in the context of the CURE’s research goals.
    • Conducting experiments, collecting and organizing data.
    • Analyzing and interpreting data. Evaluating evidence. Critiquing conclusions.
    • Becoming aware of the necessity of experimental failure.
    • Understanding that sometimes discoveries emerge from iterative processes.
    • Considering the quality of evidence and their relevance to the argument.
    • Synthesizing results and drawing conclusions; planning next steps.
    • Reading the primary literature, attending relevant seminars, and discussing their work with others.
    • Presenting progress reports and comparing ideas in group setting.
    • Reflecting on how the process of critique contributes to research progress.
  • Consider guiding the cognitive exploration of students by implementing these recommendations modified from Cooper et al. (2022):
    • Challenging students to check if their hypothesis is explanatory, clearly stated, and distinguishes between multiple ideas.
    • Pushing students to fully explain ideas by asking follow-up questions and asking for clarification.
    • Highlighting potential pitfalls of hypotheses and protocols encouraging student to identify the problem and its solution.
    • Asking students to explicitly articulate links between research goals, hypothesis, and experiment.
    • Encouraging students to consider alternative experimental outcomes or explanations for their predictions or results.
    • Redirecting student ideas that are unproductive by bringing student attention back to their original hypothesis/goal.
    • Reminding students to think about controls they need to consider in their experiments/analyses.
    • Providing assistance with analytical tools or protocols that are merely means to an end and not critical elements of the learning goals, to enable students to focus on course objectives.
    • Assessing the time management of the students and propose adjustments to protocol or timeline accordingly.
  • Instructors can foster a supportive environment by integrating best practices developed from student focus groups (Faulkner et al., 2021) including the following:
    • Contacting students: consider emailing or connecting with students before the semester begins. This initial contact is not only meant to offer practical information, such as class time and location, it also opens the door for a connection between students and instructor. Activate your class website before the class begins. This allows you to post messages for students and gives them the opportunity to reach out with questions.
    • Learning about students: you can promote inclusivity by learning about the identities, circumstances, and concerns of your students. You can do so through get-to-know-you surveys as well as start-of-the-semester conferences. This is a good time and place to ask students their pronouns, names, and the pronunciation of the latter.
    • Setting the right tone for the class: the first day of class sets the tone for the rest of the semester. As such, it is a critical time to establish an inclusive environment. Creating an inclusive classroom necessitates that each student feels like you see them as a unique individual. It also requires the classroom environment to reflect respect and care for all students. You should set expectations for class discussions and respect of everyone’s background. You also need to encourage everyone to express their ideas and learn from each other. Communicate that you will not tolerate any discriminatory attitudes or behaviors in the classroom.
    • Encouraging introductions and setting group atmosphere: creating a group atmosphere that cultivates student collaboration and support throughout the semester is essential for inclusive pedagogy. It helps students create relationships with other students in the classroom and encourage collaboration and support. You can facilitate this collaboration by having students introduce themselves and exchange contact information with the people they will work with.
    • Explaining syllabus and expectations: transparency is critical to a supportive class environment. Going over the CURE in detail on the first day of class is a useful way to make students feel welcome and helps you establish course expectations.
    • Self-disclosure: the classroom can feel more inclusive when you share aspects from your own life. In particular, it may help students feel less vulnerable. Students see self-disclosure both as a way to get to know you and a sign of mutual respect.
    • Being approachable: see your students as the people they are by using their name and correct pronouns, treat them as capable learners, respect them, encourage them, check in on them, make them feel welcome, and do not reinforce rigid power hierarchies. Explicitly tell students how you prioritize their learning and needs as well as your desire to help and support them.
    • Staying engaged: pay attention to both the verbal and nonverbal responses of students. Notice the silences and apathy as much as the participation. Hold office hours and consider requiring students to set up a five-minute meeting several times throughout the semester. Holding hybrid office hours both in-person and through online tools like Zoom or Teams can facilitate attendance.
    • Providing resources: give students information about campus and community resources, not only on the first day of class, but also throughout the semester. Provide specific directions for resources including academic services, clubs, tutoring, special interest groups, as well as mental health and wellness support.
    • Encourage reflection: reflect on your experience in the course, how you engage students in dialogue, how you keep communications with students open, and how you include everyone in the class discussions. Examine your own positions of power, privilege, and vulnerabilities. Encourage your students to do the same and reflect upon their roles in the classroom environment and their interactions with others.
  • Consider also the guidelines from Cramer and Prentice-Dunn (2007) for an alternative framing of the elements above into a mentorship model “caring for the whole person”.
  • There is an important role for peer-mentoring among students engaged in the CURE. Such mentoring mimics the interactions between researchers and, along with mentorship from more experienced researchers (instructor of record, postdoctoral researchers, graduate teaching assistant, etc.), can help mitigate the frustration inherent to the failures and setbacks of authentic research experiences (Hanauer et al., 2012).
  • The expert-novice divide is well recognized across disciplines (Inglis & Alcock, 2012; National Research Council, 2000; Newman et al., 2021; Stofer, 2016) and can sometimes represent an obstacle to mentorship; it may be overcome by involving peer teaching assistants. Undergraduates with prior experience of the CURE or mentored research experience outside of the classroom may be perceived as more approachable than more senior instructor(s); they also have recent experience with the struggles and challenges associated with the knowledge and skills taught in the CURE. They can provide very valuable feedback, including on writing assignments (Cho et al., 2006). Research shows that undergraduate peer assistants are valued by students enrolled in CUREs and facilitate their success (Olson et al., 2022).
  • Although peer assistants can sometimes struggle with their identity and role in the classroom (Terrion & Leonard, 2007), training and mentorship of peer assistants can help overcome these difficulties (Handelsman et al., 2005).
  • Requiring members of the instructional team to schedule dedicated time to mentoring in their week leads to more successful experiences (Shanahan et al., 2015; Terrion & Leonard, 2007). This is because time scarcity is a major barrier to effective research mentorship (Gentile et al., 2017).
  • Communication with postdoctoral researchers, graduate student instructors, or peer undergraduate assistants around the support they need and the resources that would enhance their experience (in both content and format) is important in creating a successful teaching experience for them (BrckaLorenz et al., 2020).
  • Goodwin et al. (2021) identified different mentorship roles for CURE graduate student instructors and provided evidence that instructors who embrace their functions as “student supporters” and “research mentors” are more likely to see value for themselves in the CURE and engage in teaching CUREs again. Goodwin et al. (2021) defined “student supporters” as mentors who “[provide] emotional support to students” and “research mentors” as instructors who “[develop] student[s’] autonomy and competence as researcher[s]” (Goodwin et al., 2021: Fig. 3).
  • Good mentorship should incorporate socioemotional support, culturally relevant mentoring, and appropriate personal interest in students (Haeger & Fresquez, 2016; Robnett et al., 2018; Shanahan et al., 2015). Consider also sharing your own stories of struggles and failures with research (Jayabalan et al., 2021).
  • The affective-motivational research competence model (Wessels et al., 2018) has identified six situations that mentees need support with to engage in research as well as dispositions that can be fostered to help overcome the challenges of these situations (Table 16). These dispositions can be brought about through interventions and scaffolding activities of the CURE aimed at fostering specific experiences and tasks that promote knowledge integration (Linn et al., 2015).

 

Instructors should develop experiences to foster adaptive dispositions that help students face situations like developing a research interest, making decisions, enduring setbacks, unravelling irritations, making use of feedback and critiques, and audience- appropriate communication of research.
Table 16. Experiences and tasks that mentors should use to foster adaptive dispositions in students that help them through (modified from Linn et al., 2015) challenging situations (modified from Wessels et al., 2018).

 

Box 10: Important points

  • A CURE is an opportunity to move away from a model of teacher-identity centered around the delivery and assessment of knowledge to one of a mentor.
  • Successful CUREs empower students to think for themselves and enable them to make mistakes in a safe and supportive environment.
  • A CURE is an authentic research experience and as such can introduce students to collaboration by including outside researchers, laboratory personnel, and staff members.
  • Every member of the CURE has goals, needs, and expectations. This is true of the students and instructor(s) of course, but also applies to any other member of the team. Consider carefully the help and means everyone requires to achieve their mission.

 

11. How will the success of the CURE be assessed?

Overview: Just like the assessment of the students’ work in the CURE enables them to reflect upon their progress and make corrections, the assessment of the CURE itself provides opportunities for redesigns, corrections, and reflection by the instructor(s). Assessing students’ work provides a basis for an often-necessary grade. Similarly, the evaluation of the CURE is integral to the instructor’s annual evaluations, their eligibility for promotions, tenure, and awards. Because CUREs often require financial and/or logistical support, the assessment of the experience enables the demonstration of its efficacy and can facilitate the renewal or expansion of the program. When a CURE is an integral component of the curriculum, particularly as an element of introductory courses, general education requirements, or program prerequisite, its evaluation is an important part of the overall educational experience of students that may become part of program reviews, external evaluations, and validation of the course as satisfying specific certifications or endorsements. There are three overlapping components to the assessment of CUREs: course, instructor, and student outcomes (Brownell & Kloser, 2015) summarized in Figure 4. Section F presents a discussion of the existing approaches in evaluating all three of these elements of CUREs from a programmatic and scholarly point of view; many of these approaches are aimed at incorporating data from large number of students; several may require the involvement of education research collaborators or institutional representatives. Here, the focus is on data collection and analyses that individual instructors can engage in to reflect upon the research progress and the pedagogical framework.

 

Instructor, course, and student outcomes of CUREs are showed to cover thirteen different categories, including publications/grants, academic enthusiasm, self-efficacy, career options, collaboration, and mentorship skills.
Figure 4. CURE assessment “meta-framework” (redrawn and modified from Brownell & Kloser, 2015). Dashed lines were added to reflect the importance and value of outputs (publications and grants) to students, and the roles of graduate students, peer-assistants, and postdoctoral researchers as instructors.

This section enables you to determine:

  1. The value of the individual assignments and activities implemented
  2. The usefulness of the scaffold
  3. The fit of the assessment and learning goals
  4. The research outcomes of the CURE and their significance to stakeholders
  5. The success of the CURE in promoting the success of all students

Education starts here:

  • Are the assessments aligned with the learning goals of the course?
  • Are the formative and summative assessments of the students’ learning also assessed for the success of the course activities and scaffold?
  • Does the scaffold for the CURE need to be revised?
    • Are there bottlenecks to learning that remain to be addressed by the CURE scaffold?
    • Are some parts of the course “over-scaffolded”? Is the preparation of the students underestimated for specific activities?
  • Are there concept inventories or professional society standards that can be used to assess students’ proficiency on certain topics of the CURE?
  • Did students encounter problems with specific activities?
    • Are there assignments that were unclear?
    • Were there homework assignments that multiple students asked for clarifications about?
  • What are the perceived gains from the CURE that students express?
  • Can specific assignments’ rubrics be edited to include elements useful to assessing the efficacy of the activity?
  • Does your institution have survey tools (including some incorporated into student evaluations) that can be leveraged to assess your pedagogy of the course?
  • Can you engage peer instructors in reviewing your course?
  • Can you present the CURE you developed at an education conference or in the education session of the annual meeting of your professional society to get feedback and thoughts from colleagues?
  • Can you gather data on the success of the CURE in closing the equity gap and promoting the learning and success of all?

Scholarship starts here:

  • Can the research findings of the CURE be presented to the research community for feedback?
    • Is the work ready to be submitted in the form of a proposal or manuscript?
    • Is it appropriate to present the work at a conference?
  • Are there other stakeholders of the research who could provide critiques of the work and assess its significance?

Research and Education together:

  • The efficacy of the scaffold and activities can be in part assessed through the performance and success of students. Each assessment of the students’ work is also an assessment of the work of the instructor. In many ways, some of the issues associated with CURE assessment can be answered through student assessment, an issue discussed earlier in the document.
  • A key element of the assessment of the activities and assignments of the CURE is to define clear goals for each of them. This participates in increasing transparency for students and enables the determination of their success and therefore yours.
  • Assessments of students’ work and survey tools can be used to identify knowledge/skill bottlenecks to address in iterations of the CURE.
  • There are numerous forms of formative assessment that can be implemented throughout the CURE to help determine the efficacy of the CURE activities and interventions:
    • Minute papers or muddy point papers (Anderson & Burns, 2013; Stead, 2005) can help assess whether or not a particular activity fulfilled its learning goals. Example of prompts are available from multiple sources (e.g., Tufts University; University of Wisconsin)
    • Concept maps and Venn diagrams can help students synthesize information and be assessed by instructor(s) for learning and efficacy of scaffolding activities (Bauman, 2018; McConnell et al., 2003).
    • Group and individual presentations of research updates enables the instructor to assess the efficacy of the scaffold in supporting the research progress.
  • Group reviews and self-reviews ([Activity 19]) can help assess group dynamics and the effect of team building efforts.
  • Questions can be included in the reflection workbook ([Activity 26]) to enable the assessment of the students’ understanding of concepts presented in the CURE activities, thus determining their efficacy. Student reflections can be evaluated with students to get additional insights in the student experience (McLean et al., 2022).
  • Many institutions allow the design of a few custom questions to add to student evaluations of instruction. These custom questions can be used to assess the pedagogy of the course. Examples of questions can be found in the project ownership survey (Hanauer & Dolan, 2014), the classroom and school community inventory (Rovai et al., 2004), the laboratory course assessment survey (Corwin et al., 2015), the classroom undergraduate research experience survey (Denofrio et al., 2007), the science process skills inventory (Arnold et al., 2013), or a combination of these tools (Lo & Le, 2021).
  • Some surveys have been developed to specifically assess the efficacy of CUREs, or even specific activities of the course (e.g., Satusky et al., 2022)
  • Online survey tools (e.g., Qualtrics and SurveyMonkey) or paper survey tools (including some standard student evaluation forms) administered as part of midterm evaluations and end-of-term evaluations can enable instructors to poll students on affect towards specific activities, perceptions of learning gains, and self-efficacy.
  • Concept inventories (Table 6) as well as professional standards (e.g., American Psychological Association) can be used to devise summative assessments.
  • Many indicators of learning outcome satisfaction can be assessed by integrating appropriate criteria into grading rubrics (Chamely-Wiik et al., 2014).
  • Extensive notetaking of class observations, student behaviors, reflections on pedagogy, problems encountered in the classroom, failures, and successes can be used to revise individual activities as well as their organization.
  • Interviews and focus-groups by the instructor or by a neutral third party can enable students to share their thoughts and feelings about the CURE and its outcomes (e.g., (Brownell & Kloser, 2015; Turner et al., 2021; Wooten et al., 2018). This approach can also be used to assess the experience of graduate student instructors and peer assistants to ensure positive and improved experience for all members of the CURE team (see Heim & Holt, 2019). Essays can also be used to gather the thoughts of CURE participants (Wooten et al., 2018).
  • Partnering with colleagues and education researchers can be helpful in determining the impact of the CURE on all students and the success of the CURE in mitigating systemic biases and enabling the learning of a diverse student body.

Box 11: Important points

  • Take abundant notes during the implementation of the CURE to guide redesigns, expansions, and modifications to the course
  • Ask yourself numerous questions when grading and/or assessing student work, for example:
  • Did this assignment fulfill its goal of assessing the associated ELOs?
  • Did this activity support the associated learning goal(s)?
  • Does your institution enable custom questions on student evaluations?
  • Can you use surveys or include questions within assignments to gather data from your students on the activities and assignments of the CURE?
  • Reviews by peers of manuscripts, conference presentations, and grant proposals are a measure of the success and progress of the research of the CURE