Adaptive Instructional Aids for Teaching a Blind ... - ACS Publications

Jul 18, 2012 - Department of Chemistry and Biochemistry, University of Mount Union, Alliance, Ohio 44601, United States. ABSTRACT: Adaptive tools and ...
4 downloads 0 Views 2MB Size
Article pubs.acs.org/jchemeduc

Adaptive Instructional Aids for Teaching a Blind Student in a Nonmajors College Chemistry Course Debra Boyd-Kimball* Department of Chemistry and Biochemistry, University of Mount Union, Alliance, Ohio 44601, United States ABSTRACT: Adaptive tools and techniques for lecture instruction were developed for a blind student in a nonmajors college chemistry course. These adaptive instructional aids assisted the student in writing and balancing chemical reactions, calculating unit conversions and concentrations, drawing Lewis dot structures, understanding structural representations of molecules with three-dimensional models, and identifying organic functional groups.

KEYWORDS: First-Year Undergraduate/General, High School/Introductory Chemistry, Curriculum, Computer-Based Learning, Hands-On Learning/Manipulatives, Enrichment/Review Materials, Minorities in Chemistry, Molecular Properties/Structure, Nomenclature/Units/Symbols, Student-Centered Learning tools,21 which have now been compiled into “The Tactile Adaptations Kit” marketed through MDW Educational Services, LLC.29 However, depending on the curriculum, the cost of the kit is considerable, many adaptive tools still have to be manufactured by the instructor, and the full contents of the kit are not disclosed. Overall, the body of literature relevant to the chemical education of blind or low-vision students remains limited and largely focused on laboratory adaptations. Consequently, I wish to share the experience of modifying and developing tools and techniques to aid a blind student in a lecture-based liberal arts chemistry course. The overall goal of this liberal arts chemistry course is for students to establish an understanding and appreciation of chemistry in the world around them by learning about chemistry in the context of societal issues. Students are expected to develop skills in writing, reading, and understanding chemical formulas; balancing chemical equations; unit conversions; calculation of concentrations; reading and interpreting graphs and figures; drawing Lewis dot structures; understanding and using structural representations including condensed structural formulas, skeletal structures, and lineangle drawings; and identifying organic functional groups. What follows are the trials and errors of my experience in making modifications for lecture instruction to teach a blind student in a liberal arts chemistry course.

V

isual observations and representations underlie both the research and learning activities of chemistry. The significant dependence on the use of visual representations to impart chemical concepts to sighted students poses a challenge when an educator encounters a student who is blind or lowvision. Ten days prior to the start of the fall 2007 semester, this educator, having no previous experience with teaching blind or low-vision students, learned that a student enrolled in a nonmajors, lecture-only chemistry course was blind. A call to the American Chemical Society led to the discovery of its publication Teaching Chemistry to Students with Disabilities: A Manual for High Schools, Colleges, and Graduate Programs,1 an essential resource that includes information on instruction and assessment modifications and assistive technology. A search of the chemical education literature in August of 2007 yielded very little with respect to lecture instruction2−6 and only slightly more with respect to laboratory adaptations and safety issues.7−15 Outside of the chemical literature, very little information was found. However, what was discovered placed more emphasis on laboratory adaptations and accommodations rather than lecture instruction.16−19 Through a Web-based search for more information, the Independent Laboratory Access for the Blind (ILAB) project20 Web page was located. ILAB was a Research in Disabilities Education project supported by the National Science Foundation that developed and tested low-cost adaptive tools and procedural modifications with the intent to allow blind or low-vision students to work independently in the laboratory. Additionally, the ILAB Web site describes classroom instruction aids. Since August 2007, several articles with respect to blind or low-vision students have been published21−28 including the publication of some ILAB classroom and laboratory adaptive © 2012 American Chemical Society and Division of Chemical Education, Inc.

Published: July 18, 2012 1395

dx.doi.org/10.1021/ed1000153 | J. Chem. Educ. 2012, 89, 1395−1399

Journal of Chemical Education



Article

WRITING AND BALANCING CHEMICAL REACTIONS

identities were made from magnetic sheets, hot glue, and felt from a craft store. These were then labeled with puff paint so that the student could identify the atom represented by feeling the raised paint symbol. Plus signs and reaction arrows also had to be constructed out of magnetic sheets and felt. Puff paint was used so that the student could identify the plus sign or arrow by feeling the raised paint symbol. Using these tools, the student could reproduce or construct a chemical equation on the magnetic board. Once the chemical reaction was written, the student would locate letters of the appropriate elemental symbol and would place these under the reaction arrow (Figure 2). The student would then count the number of each element on the reactant side and find the respective number. This number would be placed to the left of the elemental symbol below the reaction arrow. The student would then count the number of each element on the product side and find the representative number. This number would be placed to the right of the elemental symbol below the reaction arrow. The student would then think through the process adding coefficients and changing the number of elements present on the reactant or product side, respectively, until the equation was balanced.

Execution

Writing and balancing chemical reactions first depends on the student being able to write a chemical formula. For ionic compounds, students are traditionally taught to first write the chemical symbol of the most metallic element with its associated charge and then to write the chemical symbol of the less metallic element with its associated charge. The students are then told to criss−cross the magnitude of the charges to a subscript position of the opposite element. Alternatively, students are taught to balance the charges by increasing the number of cations or anions present to give an electronically neutral compound. In both instances, visual representations are used to explain the concept. Figures are given in textbooks and the instructor will go over examples on the board. Typed lecture notes prepared for this particular student were of little use. According to the student, the program she used to convert from text to Braille did not translate the subscripts and superscripts in a manner that she could understand. This left her feeling more confused than she had been during the explanation in lecture. Adaptive tools for writing chemical formulas and for writing and balancing chemical reactions have been described in the literature.3,4,24,30 Totally blind students can learn with Braillelabeled magnetic letters and numbers. Blind and low-vision students who know the shapes of print letters and numbers do not need the Braille labels. With this in mind, a magnetic board (22 in. x 16 in.) and common magnetic letters and numbers (refrigerator magnets) were utilized. The magnetic letters and numbers were used to demonstrate both methods of how to write chemical formulas and the superscript and subscript positioning (Figure 1). The next task was to use chemical formulas to write and balance chemical reactions. An attempt was made to use the same magnetic letters and numbers to write the reactions on the magnet board, but the letters and numbers were too large for the board. Consequently, “atoms” of different sizes and

Issues

This process worked well for keeping track of the number of atoms of each element on both sides of the equation, but several sets of magnetic letters and numbers were necessary to write and balance fairly simple reactions. Additionally, boundaries need to be clear to ensure that the student is able to feel the entire equation and does not miss anything on the board.



CONVERSIONS AND CALCULATIONS

Execution

Quantitative problem solving is an essential skill to develop, especially in the study of chemistry. At the basis of quantitative problem solving is dimensional analysis that, along with abstract terminology, such as “mole,” is often one of the more difficult concepts for a chemistry student to grasp.31−35 In this particular course, students are required to understand the relationship between atoms or molecules, amount (moles), mass (grams), and molarity (mol/L), and be able to convert between them. To accomplish this, Microsoft Excel was coupled with Job Access with Speech (JAWS) screen reading software for Windows by Freedom Scientific.36 The JAWS screen reading software gives the student the location of a selected cell prior to reading what is in the cell. To aid the student, a grid similar to a spreadsheet in Excel was created using a poster board and glued-on-yarn so that the student had a tactile way to follow along with the spreadsheet. It is particularly important when doing unit conversions that the student keep track of the units associated with each value. Microsoft Excel is not able to do a calculation if there is any text included within a cell that is part of the calculation. Consequently, a way to keep track of the units was improvised. Small objects were selected to represent the units and the objects were placed in the cell within the tactile grid corresponding to the location of the value it was associated with in Excel (Figure 3). Table 1 gives a key of some of the objects used. Issues

It helped that the student had some familiarity with Microsoft Excel. Boundaries on the tactile grid must be clear so that the

Figure 1. Writing chemical formulas using a magnetic board and common magnetic letters and numbers. 1396

dx.doi.org/10.1021/ed1000153 | J. Chem. Educ. 2012, 89, 1395−1399

Journal of Chemical Education

Article

Figure 2. Balancing chemical equations using a magnetic board and common magnetic numbers and atoms created from felt and puff paint.

Table 1. Objects Used To Represent Units of Measure Unit

Object

Molecule Atom Mole Gram Milliliter Liter Molarity

Small ruler Pop tab Pencil Rubber Band Small Paper Clip Large Paper Clip Bobby Pin

student, Microsoft Excel 2007 was used with JAWS for Windows version 9.0.519. Recently, the University has transitioned to Microsoft Excel 2010 and the JAWS version previously used was found to be incompatible. JAWS for Windows 13.0.718 32-bit demo version, available for download from Freedom Scientific, was tested and found to be compatible with Microsoft Excel 2010.



Figure 3. Unit conversions with Microsoft Excel and tracking units with a tactile grid.

LEWIS DOT STRUCTURES

Execution

student starts in the right place in the calculation and has the correct correspondence with the cells in Excel. Although the yarn served as a dividing line between cells, it was not thick enough to prevent some of the objects from being shifted into the wrong cell. It might be useful if the tactile grid was developed into a box structure with each cell a distinct box with sides high enough that objects could not be shifted. Also, cells of the tactile grid could be labeled with puff paint to correspond to the cell locations read by the JAWS program. Software compatibility and cost is also an issue. For this particular

Lewis dot structures are common representations of molecular structure in two-dimensions and are the basis of understanding three-dimensional molecular geometry. Magnetic sheets, hot glue, and felt from a craft store were used to produce atoms of different sizes and identities based on a modification of adaptive tools described in the literature.3,4,20,25 These were then labeled with puff paint so that the student could identify the atom represented by feeling the raised paint symbol. Bonds and electrons using different shapes and raised paint symbols were also constructed. Again, the magnetic board was used as the 1397

dx.doi.org/10.1021/ed1000153 | J. Chem. Educ. 2012, 89, 1395−1399

Journal of Chemical Education

Article

surface on which the student could slide the magnets around to construct the correct structure and determine how the electrons are distributed in a given molecule (Figure 4).

Figure 6. (A) Tactile model kit and (B) 3D structure of methane (CH4).

structures. For the line−angle drawings, the bonds were used as the lines that were angled to create the vertex point representing carbon atoms and, where appropriate, the magnetic atoms were used to represent functional groups (Figure 7).

Figure 4. Tactile magnets packaged for constructing Lewis dot structures and other structural representations.

Issues

Some magnets were sized too small and were harder to handle. Because this student had been sighted and knew the shapes of the letters, chemical symbols were represented with letters painted on the atoms; however, these could easily be used with Braille labels. Additionally, boundaries need to be clear to ensure that the student is able to feel the entire structure and does not miss anything on the board such as the potential stray electron that the student does not know is there.



MOLECULAR GEOMETRY Using the Lewis dot structure, the student could learn about the molecular geometry of a compound by building a model of the compound (Figure 5). To accomplish this, a traditional

Figure 7. Structural representations of propanal (C3H6O).

For the student to learn the eight functional groups that were covered, keys were made using puff paint on plain letter paper. On each key, the atoms in the functional group and the bonds that connect them were painted. At the lower right-hand corner of every page, a 2-in. line was painted so that the student would always know how to orient the paper. Additionally, the Disability Support Services created Braille labels for each of the functional groups that were affixed to their respective key in the lower left-hand corner. The student used the keys to learn to recognize the functional groups and to identify them in a structure. When structures were given in class, on homework assignments, quizzes, or exams, they were painted with puff paint onto plain letter paper with a 2-in. line in the lower righthand corner to ensure correct orientation and the structure was assigned a letter which was painted in the upper right-hand corner (Figure 8). In this way, homework that could be done independently was assigned to the student.

Figure 5. Lewis dot structure, skeletal structure, and 3D model of methane (CH4).

ball-and-spring model kit was modified with tactile puff paint so that the student could distinguish atoms as suggested by Tombaugh.4 The atoms were separated into bags and the bags were labeled on the outside with puff paint to increase organization and efficiency (Figure 6). Because the balls and springs could easily roll away from the working area, a box was used to place them in as the student worked to build the model. Once the model was built, the student would describe the bonding angles and geometry of the molecule.





STRUCTURAL FORMULAS, LINE-ANGLE DRAWINGS, AND FUNCTIONAL GROUPS In a chapter on organic chemistry and drug chemistry, students are required to learn how to generate skeletal structures and line−angle drawings. For this, the magnetic board was again utilized. The complete structural formulas were created using the bonds and atoms constructed previously for the Lewis dot

CONCLUSION I described the tools I created to help a blind student succeed in a nonmajors, lecture-based chemistry course. I found that it was best to develop the tool and meet with the student ahead of time to teach her how to use the tool prior to the lecture in which the tool was needed to learn the content. It is my hope that other instructors may be able to use or adapt the tools I 1398

dx.doi.org/10.1021/ed1000153 | J. Chem. Educ. 2012, 89, 1395−1399

Journal of Chemical Education

Article

(28) Bonifácio, V. D. J. Chem. Educ. 2012, 89, 552−554. (29) MDW Educational Services, LLC. http:// MDWeducationalservices.com (accessed Apr 2012) (30) Staff., J. C. E. J. Chem. Educ. 2009, 86, 167. (31) Canagaratna, S. G. J. Chem. Educ. 1993, 70, 40−43. (32) McClure, J. R. J. Chem. Educ. 1995, 72, 1093−1094. (33) Cohen, J.; Kennedy-Justice, M.; Pai, S.; Torres, C.; Toomey, R.; DePierro, E.; Garafalo, F. J. Chem. Educ. 2000, 77, 1116−1173. (34) Wagner, E. P. Sch. Sci. Math. 2001, 101, 10−22. (35) Bruck, L. B.; Towns, M. H. J. Chem. Educ. 2011, 88, 440−442. (36) Freedom Scientific. http://www.freedomscientific.com (accessed Apr 2012). Figure 8. (Left) Homework identifying organic functional groups and (right) tactile structure examples.

have described. However, it is essential to remember that each student is different and accommodations and teaching style should be adapted to the individual student.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Notes

The authors declare no competing financial interest.



REFERENCES

(1) Teaching Chemistry to Students with Disabilities: A Manual for High Schools, College, and Graduate Programs, 4th ed.; Miner, D., Nieman, R., Swanson, A. B., Woods, M., Carpenter, K., Eds.; American Chemical Society: Washington, DC, 2001. (2) Crosby, G. A. J. Chem. Educ. 1981, 58, 206−208. (3) Smith, D. J. Chem. Educ. 1981, 58, 226−227. (4) Tombaugh, D. J. Chem. Educ. 1981, 58, 222−226. (5) Riendl, P. A.; Haworth, D. T. J. Chem. Educ. 1995, 72, 983−986. (6) Supalo, C. J. Chem. Educ. 2005, 82, 1513−1518. (7) Lunney, D.; Morrison, R. C. J. Chem. Educ. 1981, 58, 228−231. (8) Swanson, A. B.; Steere, N. V. J. Chem. Educ. 1981, 58, 234−238. (9) Flair, M. N.; Setzer, W. N. J. Chem. Educ. 1990, 67, 795−796. (10) Lunney, D. J. Chem. Educ. 1994, 71, 308. (11) Ratliff, J. L. J. Chem. Educ. 1997, 74, 710−711. (12) Gupta, H. O.; Singh, R. J. Chem. Educ. 1998, 75, 610−612. (13) Pence, L. E.; Workman, H. J.; Riecke, P. J. Chem. Educ. 2003, 80, 295−298. (14) Neppel, K.; Oliver-Hoyo, M. T.; Queen, C.; Reed, N. J. Chem. Educ. 2005, 82, 607−610. (15) Lunsford, S. K.; Bargerhuff, M. E. J. Chem. Educ. 2006, 83, 407− 409. (16) Wang, L. Chem. Eng. News 2007, 85 (30), 36−40. (17) Hinton, R.; Hinton, D. J. Visual Impairment Blindness 1999, 93, 429−433. (18) Watson, S. W.; Johnston, L. Sci. Teacher 2004, 71, 30−35. (19) Ruby, M. T. Sci. Teacher 2001, 68, 80−82. (20) Independent Laboratory Access for the Blind (ILAB). http:// ilab.psu.edu/index.html (accessed Apr 2012). (21) Winograd, M. D.; Rankel, L. A.; Speer, L. RE:view: Rehabilitation Educ. Blindness Visual Impairment 2007, 39, 122−128. (22) Neely, M. B. J. Chem. Educ. 2007, 84, 1697−1701. (23) Bromfield-Lee, D. C.; Oliver-Hoyo, M. T. J. Chem. Educ. 2007, 84, 1976−1978. (24) Poon, T.; Ovadia, R. J. Chem. Educ. 2008, 85, 240−242. (25) Supalo, C. A.; Mallouk, T. E.; Rankel, L.; Amorosi, C.; Graybill, C. M. J. Chem. Educ. 2008, 85, 243−247. (26) Supalo, C. A.; Mallouk, T. E.; Amorosi, C.; Lanouette, J.; Wohlers, H. D.; McEnnis, K. J. Chem. Educ. 2009, 86, 587−591. (27) Pereira, F.; Aires-de-Sousa, J.; Bonifácio, V. D. B.; Mata, P.; Lobo, A. M. J. Chem. Educ. 2011, 88, 361−362. 1399

dx.doi.org/10.1021/ed1000153 | J. Chem. Educ. 2012, 89, 1395−1399