throbber
Reengineering the Curriculum: Design
`and Analysis of a New Undergraduate
`Electrical and Computer Engineering Degree
`at Carnegie Mellon University
`
`STEPHEN W. DIRECTOR, FELLOW, IEEE, PRADEEP K. KHOSLA, FELLOW, IEEE,
`RONALD A. R O H R E R , FELLOW, IEEE, AND ROB A. RUTENBAR, SENIOR MEMBER, IEEE
`
`Invited Paper
`
`In the Fall of 1991, after approximately two years of devel-
`opment, the department of Electrical and Computer Engineering
`(ECE) at Camegie Mellon University (CMU) implemented a new
`curriculum that differed radically from its predecessor. Key fea-
`tures of this curriculum include: Engineering in the Freshman
`year, a small core of required classes, area requirements in place
`of most speciJc course requirements, mandated breadth, depth,
`design, and coverage across ECE technical areas, a relatively
`large fraction of free electives, and a single integrated Bachelor
`of Science degree in Electrical and Computer Engineering. In
`this paper we review the design of this curriculum, including a
`taxonomy of problems we needed to address, and a set of general
`principles we evolved to address them. The new curriculum is
`described in detail, including new data from an ongoing analysis
`of its impact on students’ curricula choices.
`
`I. INTRODUCTION
`Current engineering practice has, by necessity, evolved
`to keep pace with technology: witness the rate at which
`fundamentally new ideas are introduced into new products.
`One might suppose, then, that current engineering education
`has also evolved to track such new developments. However,
`we argue that engineering education has really evolved only
`to the extent that individual engineering courses have been
`updated-usually with increased density of content-to
`reflect new developments. The prevailing philosophy of
`engineering education-teach
`first the basics in mathe-
`matics and science, follow with exposition of engineering
`applications-has
`remained unchanged and unchallenged
`for more than four decades. While contributing to the cre-
`ation of engineers who are current in specific technologies,
`
`Manuscript received December 7, 1994; revised April 12, 1995.
`The authors are with the College of Engineering, Carnegie Mellon
`University, Pittsburgh, PA 15213 USA.
`IEEE Log Number 9413199.
`
`we believe that teaching of unmotivated math and science
`followed by incrementally updated technical courses is
`fundamentally flawed. It contributes little to the education
`of engineers who can acquire new knowledge as necessary,
`cope with dynamically changing work environments, or
`excel in nontraditional jobs. We believe that real impact
`in engineering education will be made only by looking
`at the curriculum as a whole, in the context of present
`technological and societal needs, and not just by constant
`repolishing of aging courses. It is not our intention to
`imply that engineering education has completely failed in
`its goals. Rather, we wish to drive home the point that there
`are advantages to be found in taking a fresh, unfettered look
`at the undergraduate curriculum.
`Of course, curricula have tremendous inertia, and often
`resist all but the most incremental and cosmetic of changes.
`Unfortunately, many of the problems faced by engineering
`educators are not amenable to simple, incremental fixes.
`In October 1989, the college of engineering at Carnegie
`Mellon University (CMU) instituted a review process across
`all engineering departments. The goal was to evaluate how
`well the educational mission of the college was being
`conducted, with an eye toward redefining both collegewide
`and department-specific curriculum requirements. Because
`of the breadth of this undertaking, each engineering depart-
`ment was allowed to consider the best possible curriculum
`changes, not merely those that could be wedged conve-
`niently into its current web of requirements, prerequisites,
`constraints, and customs. This paper describes the design
`and implementation of the new Electrical and Computer
`Engineering Bachelor’s degree program that emerged from
`this process. This curriculum, which took approximately
`two years to design fully, was implemented within the de-
`
`1 246
`
`~
`
`0018-9219/95$04.00 @ 1995 IEEE
`
`PROCEEDINGS OF THE IEEE, VOL. 83, NO. 9, SElTEMBER 1995
`
`HP Exhibit 1007 - Page 1
`
`

`

`partment of Electrical and Computer Engineering (ECE) in
`the Fall of 1991, and produced its first four-year graduates
`in the Spring of 1995.
`Within ECE, the curriculum was designed by a committee
`whose quickly adopted name reflected the spirit of process:
`the Wipe-the-Slate-Clean Committee. Composed of eleven
`faculty from across the breadth of the department’s research
`and teaching areas, the committee interviewed both students
`and faculty, and worked aggressively for roughly one year
`to dissect, analyze, disassemble, and finally redefine the
`ECE undergraduate curriculum. The new curriculum that
`resulted from this process hinges on a few key ideas:
`Engineering courses begin in the Freshman year, con-
`current with mathematics and science.
`The core of required “essential” engineering classes is
`extremely small.
`Area requirements across a spectrum of electrical and
`computer engineering topical areas replace most spe-
`cific course requirements.
`Breadth, depth, and coverage are mandated across this
`spectrum of technical areas, but individual courses are
`not prescribed; students flexibly choose from among
`available topical areas.
`Nearly a full year of the curriculum is unconstrained.
`At completion, the curriculum offers a single, unified
`Bachelor’s degree in Electrical and Computer Engi-
`neering.
`The end result of our exercise is a curriculum which has
`been recently reviewed by ABET for accreditation under
`the ABET “innovative cumculum” clause that permits
`thoughtful experimental curricula that diverge from existing
`ABET standards to be considered on their merits. While
`the final outcome of the accreditation process will not be
`known until late 1995, comments made by the visiting team
`were favorable. Also, initial analysis of the three groups of
`freshman entering ECE in 1991, 1992, and 1993 (about
`150 students in each group) indicates that the students
`are enthusiastic about starting engineering classes in the
`Freshman year and that these are helpful to the student
`when selecting their major. There is also evidence to
`show that the flexibility in the choice of electives has not
`resulted in a mass exodus to “easier” courses. In general
`students continue to elect challenging courses to suit their
`interests.
`In this paper we share some of the details of the design
`process for this new curriculum, and an analysis (ongoing)
`of its implementation and impact.’ Of course, we were
`not alone among universities as we embarked on our
`reengineering efforts; for example, Drexel, Rose-Hulman,
`and Texas A&M were already restructuring their curricula
`as we began our efforts, and as well the US National
`Science Foundation was organizing Engineering Education
`Coalitions with similar intent. Nevertheless, we did not
`join any of these efforts for fear of diluting our own
`efforts. So rather than attempt a broad survey of competing
`
`‘See [l] for a more detailed, contemporaneous account of this process,
`and [2] for a more recent review.
`
`curriculum strategies, we focus entirely and closely on our
`own redesign effort, from beginning to end. We offer this
`as one case study for how one department reengineered its
`curriculum.
`The remainder of the paper is organized as follows.
`Section I1 begins by summarizing our motivations for
`undertaking this effort. Section I11 offers a taxonomy of
`the basic problems faced by any electrical or computer
`engineering department as it struggles to keep pace with
`technology, students, and society. Section IV describes
`the design principles for the new Carnegie Mellon ECE
`curriculum that we evolved in response to these problems.
`Section V describes the details of the new curriculum, and
`some of its novel characteristics. Section VI describes its
`implementation, and recent efforts to analyze its impact
`on students. Finally, Section VI1 offers some concluding
`remarks.
`
`11. MOTIVATIONS
`
`A. Why Change?
`By any traditional measure in 1991, the ECE department
`was doing well educating its students. The department as
`a whole was consistently ranked among the country’s top
`EE departments [3] (Components of the graduate program
`were likewise being ranked highly [4]). The department
`attracted outstanding undergraduate students: ECE was the
`first choice among engineering departments of most en-
`tering Freshman. Our graduates were recruited heavily by
`US companies, and the ECE department was on the list of
`must-visit departments for many companies that recruited
`only among a select set of elite schools. Our graduates
`who chose to pursue an advanced degree went on to elite
`graduate schools.
`So, why did we undertake a substantial reorganization
`of our curriculum? The answers are not simple, nor are
`they independent. We categorize our broad concerns in the
`following subsections, beginning with a quick overview of
`the original ECE curriculum as it stood in the 1990-1991
`academic year. These concerns can be regarded as the
`beginnings of a set of “specifications” for a new curriculum.
`
`B. Original CMU ECE Curriculum
`In 1991, the ECE department offered two four-year
`ABET-accredited Bachelor of Science degrees: the Bache-
`lor of Science in Electrical Engineering (BSEE) and the
`Bachelor of Science in Computer Engineering (BSCE).
`Both curricula shared a common Freshman year empha-
`sizing mathematics, science, and computer programming.
`They also shared a common core of engineering classes,
`emphasizing linear circuits, electronics, solid state devices,
`digital logic design, and microprocessors. In addition, these
`curricula (as did all curricula in the colleges of engineering
`and science) shared common requirements for humanities
`and social science courses (called H&SS) that amounted to
`roughly one such course per semester. An overview of the
`curricula appears in Table 1.
`
`DIRECTOR er al.: REENGINEERING THE CURRICULUM: DESIGN AND ANALYSIS OF A NEW DEGREE
`
`1241
`
`HP Exhibit 1007 - Page 2
`
`

`

`Table 1 Original Carnegie Mellon EE and CE Cumcula
`Electrical Engineering
`Mathematics & Sciences
`Calculus
`Differential Equations
`Linear Algebra
`Probability
`Physics
`Chemistry
`Computer Programming
`
`Electrical & Computer Engineering
`Intro Digital Systems
`Linear Circuits
`Intro Electronic Devices
`Electromagnetics
`Signals & Systems
`Analog Circuits
`Digital Integrated Circuits
`EE Elective
`Senior Design Elective
`
`Electives
`Freshman
`Engineering Science
`Technical
`Free
`Humanities & Social Sciences
`
`Courses
`
`1
`
`1
`1
`
`2
`2
`5
`1
`8
`
`I Computer Engineering
`I Mathematics & Sciences
`Calculus
`Differential Equations
`Linear Algebra
`Probability
`Modem Math
`Physics
`Chemistry
`Computer Programming
`Electrical & Computer Engineering
`Intro Digital Systems
`Linear Circuits
`Into Electronic Devices
`Computer Architecture
`Concurrency & Real Time Systems
`Digital Integrated Circuits
`Logic & Processor Design
`Computer Science
`Fundamentals of CS
`CS Elective
`Electives
`Freshman
`Engineering Science
`Technical
`Free
`Humanities & Social Sciences
`
`Courses
`
`1
`1
`
`1
`1
`1
`1
`1
`1
`2
`
`2
`1
`
`2
`2
`5
`1
`8
`
`After this common core, the two curricula diverged. The
`BSEE emphasized traditional electrical engineering topics
`such as electromagnetics, analog circuits, and signals and
`systems. The BSCE emphasized computer hardware and
`software topics such as computer architecture, processor
`design, data structures, and concurrency. Both curricula
`required several technical electives, and a capstone design
`elective.
`In 1991, about 40% of our students pursued the BSEE,
`and about 50% pursued the BSCE. Roughly 10% of our
`students chose to double major in both electrical engi-
`neering (EE) and computer engineering (CE). This was
`accomplished at the sacrifice of most elective classes:
`Students completed the core requirements of one curriculum
`using the elective slots provided in the other. Also, a few
`of our students double-majored in computer engineering
`and computer science (which is in a separate college at
`Carnegie Mellon). This essentially required that all elective
`classes in the BSCE curriculum were chosen to complete
`computer science core requirements.
`
`C. Remove Structural Impediments to
`Accommodate Incremental Change
`Curricula usually evolve by accretion, with new require-
`ments and constraints often layered incompatibly on top
`of existing structures. The resulting rigid course sequences
`connected by spaghetti-like chains of prerequisites are
`difficult to modify. This was certainly true of our orig-
`inal EE and CE curricula, and by extension, likely true
`in many similar Electrical Engineering departments that
`evolved over the last two decades to become departments
`of Electrical and Computer Engineering, or departments
`
`of Electrical Engineering and Computer Science. In our
`own case, the end result was that even incremental changes
`became difficult to implement.
`In the original parallel BSEE and BSCE curricula, even
`modest changes rippled in undesirable ways throughout
`the two programs. An example makes this concrete. As
`a result of an ABET accreditation visit, we were asked to
`add a linear algebra class as a graduation requirement. We
`responded enthusiastically, on the assumption that we could
`migrate the course into the early years of the curriculum,
`and thus make it a prerequisite for our linear circuits class.
`In this position, it would strengthen the background of all
`EE students in our circuits and electronics courses, and
`broaden the background of our CE students by exposing
`them to more noncalculus mathematics.
`Unfortunately, this ideal proved impossible to implement.
`There was no small-scale alteration of the BSEE and BSCE
`course sequences that could permit the linear algebra class
`to be taken by all students before the courses that would use
`it as a prerequisite. This problem derived from the slight
`differences in the first few years of the BSEE and BSCE
`requirements. The BSCE student began to take computer
`science classes fairly early, so that Junior and Senior
`computer engineering courses were correctly synchronized
`with their computer science prerequisites. In contrast, the
`BSEE student had no such requirements. The end result was
`that we required our students to take a linear algebra class,
`but we did essentially nothing to exploit this background in
`other ECE core classes. This simple example makes clear
`how difficult it can become to achieve the goal of uniform
`mathematics, science, and engineering core preparation for
`both BSEE and BSCE students.
`
`1248
`
`PROCEEDINGS OF THE IEEE, VOL. 83, NO. 9, SEPTEMBER 1995
`
`HP Exhibit 1007 - Page 3
`
`

`

`D. Rationalize Requirements for Topical
`Coverage and Workload
`As has become amply clear over the last decade, the
`disciplines of electrical and computer engineering are ex-
`panding rapidly as new technical discoveries are made and
`applied. Likewise, society is placing increasing demands
`on our graduates to apply their skills in new contexts, and
`to appreciate and manage intelligently the consequences
`of their technical decisions. Consequently, the number of
`“critical” topics to which ECE students could profitably
`be exposed is also expanding. What is not cxpanding is the
`time we have to educate someone to level of the Bachelor’s
`degree.2 Coming to grips with this accelerating problem was
`at the heart of our motivation for a significant restructuring
`of our curriculum.
`The original ECE curriculum required a large number
`of core classes, designed to ensure familiarity with a
`substantial subset of traditional EE and CE topics. After a
`great deal of argument and discussion, we came to believe
`that this approach, which implicitly assumes all students
`need exposure to (almost) all areas, was no longer credible
`as the core of a curriculum for the 21st century. Such a
`strategy mandates that we compress ever more material
`into the same number of classes. Many of our courses
`had already fallen victim to “units-creep,’’ i.e., challenging
`classes meant to require 12 hours of work per week had
`inflated to require 15 or 18 hours of work from even the
`best of students. This was caused by well meaning faculty
`working hard to give students the best, most thorough
`view of as many topical areas they could-usually with
`the assumption that this was the only opportunity students
`would ever have to see the material.
`While certainly not opposed to demanding classes, we
`concluded that the overall strategy of putting more material
`into the curriculum had become decreasingly effective.
`Students were being asked to absorb increasing amounts
`of material, which left less time for reflection, for alter-
`native perspectives on similar technical problems, and for
`revisiting background material to ensure comprehension.
`The unpredictable preparation of entering students only
`exacerbated this problem: we kept discovering that many
`of our students had never seen material fundamental to the
`background of our core courses. The end result was that by
`forcing students to juggle too many topics with too little
`time to master these topics, many students were learning
`even less material, less well.
`
`E. Emphasize Engineering Ideas Over Techniques
`A related consequence of the explosion of material was
`that many students came to view their courses as a set
`of unrelated hurdles to be overcome. As a result, many
`students were acquiring only a bag of seemingly unre-
`
`*An altemative is, of course, to extend the amount of time required to
`educate students to some minimum level of professional competence. Such
`an approach was advanced by the Massachusetts Institute of Technology
`in [SI which proposed a five-year accredited Master’s program as the
`principal mechanism for educating entry-level engineers. We retum to this
`idea in Section VI.
`
`lated problems and solution techniques, without ever really
`understanding the big ideas that bind and inform these
`techniques.
`Conventional wisdom suggested that after first teaching a
`vast body of fundamental mathematics and science-which
`students absorbed like sponges-we were free to teach
`engineering principles, drawing as necessary on the deep
`well of basic knowledge internalized by the student in
`these early studies. This was (and is) a lovely idea, but
`depressingly unrealistic. Students often had weak or wildly
`varying preparation in K-12 mathematics and science, and
`hence uncertain motivation to master the rigorous college
`level versions of these fundamentals. When a flood of
`engineering ideas was introduced on top of this precarious
`foundation, the outcome was often less than satisfactory.
`Too often, students only had time to focus on the me-
`chanical problem-formula-solution aspects of the topics,
`without developing a deeper sense of the fundamentals, the
`interconnections, and the real ideas.
`This is especially unfortunate in a fast-moving discipline,
`where the half-life of a Bachelor’s degree is probably less
`than a decade, and a solid understanding of the “big picture”
`is the most successful base from which to acquire new
`skills. As educators, we do our students a disservice if we
`fail to impart a coherent, connected view of the ideas that
`define our discipline.
`
`F. Support Interdisciplinary Studies
`The most creative and far-reaching contributions are
`often made by individuals at the boundaries of sev-
`eral disciplines. Likewise, society is placing increasing
`value on engineers who can apply their skills across
`disciplines, and can evaluate intelligently the broader
`consequences of their actions. ECE is an extremely wide
`field, and many of its most exciting frontiers-very
`large
`scale integrated circuits (VLSI), microelectromechanical
`systems (MEMS), electronic materials, computer-aided
`manufacturing, telecommunications networks, supercom-
`puting-have
`strong and established interdisciplinary
`linkages. However, our original curriculum did little
`to encourage the creation of engineers who could
`work comfortably across the boundaries of several
`disciplines.
`The original curriculum implicitly assumed that there
`were only two sorts of engineers: EE’s and CE’s. These
`were produced by completion of a large, rigid core of EE or
`CE engineering classes. Although industry specifically, and
`society generally might have valued highly a student who
`had completed, say, 60% of the EE core classes and 40% of
`the CE core classes, we had no mechanism for giving this
`broad individual a degree. Nor did we have any mechanism
`for coping with an even broader individual who might have
`wished to complete, say, 30% of the EE core, 30% of the
`CE core, then a dozen classes in mechanical engineering,
`operations research and Japanese language, in preparation
`for a career in computer-aided manufacturing. Indeed, a
`key conclusion of the early discussion of the Wipe-the-
`
`DIRECTOR et 01.: REENGINEERING THE CURRICULUM: DESIGN AND ANALYSIS OF A NEW DEGREE
`
`1249
`
`HP Exhibit 1007 - Page 4
`
`

`

`Slate-Clean committee was that we would like not only to
`tolerate such individuals, but to encourage them.
`
`111. CURRICULUM DESIGN: PROBLEMS
`A central tenet of any engineering education is that
`no elegant solution is likely to be found for a problem
`that lacks a crisp definition. Unfortunately, curricula are
`complex, often unwieldy creations subject to conflicting
`demands from the university, from faculty, from students
`and their parents, and from the industries that employ grad-
`uates. Nevertheless, over the course of its deliberations, our
`committee kept returning to several specific problems which
`crystallized as the basic issues to address. We summarize
`these here.
`
`0
`
`0
`
`A. Student Preparation is Incomplete
`American K-12 education can be blamed for the incom-
`plete mathematics and science preparation of many of our
`students. Nevertheless, allocating blame does nothing to
`improve the preparation of our students after they arrive.
`Moreover, entering students are simply different than they
`were in past decades: less homogeneous, more diverse in
`their personal goals and career aspirations. Any curriculum
`redesign must deal with the following facts:
`Students have less facility and depth in the techni-
`cal areas we expect all students to have seen, for
`example, algebra and geometry. Some unremarkable
`mathematical manipulations that appear frequently in
`introductory science and engineering classes severely
`tax many students.
`seemingly
`Students-even the best students-have
`random gaps in their backgrounds. In the course of
`our meetings, the Wipe-the-Slate-Clean Committee
`talked to a superb Senior EE student, a straight-A
`student who was being aggressively pursued by elite
`graduate schools. Yet she mentioned to us that she was
`very uncomfortable in her first circuits class, having
`never seen a complex variable before.
`Most students have almost no basic laboratory skills
`when they enter the department, for example: how to
`keep a lab notebook; how to observe an experiment;
`how to deal with significant digits and experimental
`error; how to use orders of magnitude and quick-and-
`dirty calculations to estimate whether a measured result
`is in the right ballpark or has gone badly awry, etc.
`A related point: students have virtually no hardware
`tinkering skills. Previous generations of EE’s were
`notorious tinkerers, with radios and motors and the
`like. Upon entering college, they knew what a wire
`was, and a battery. They knew how to solder and read
`the resistor color code. This is no longer true, and
`the most elementary of hardware skills-what
`a wire
`is, what it does, how it can and cannot connect to a
`battery-must now be taught explicitly. (This is not ex-
`actly surprising, given the inaccessibility of the insides
`of most electronic products these days.) Our students
`are now much more likely to have software tinkering
`
`0
`
`0
`
`0
`
`experience. However, many students, especially those
`from less well off high schools, arrive without any
`exposure to programming ideas or hardware concepts.
`Student expectations and faculty expectations often
`differ. Roughly speaking, we tend to assume students
`have the background, energy and motivation to go
`acquire whatever mathematics, science, lab skills, etc.,
`that they lack, if we send them off in the right direction.
`(This has always been true of the best students.) In
`contrast, many students tend to assume that we will
`teach them every topic-the
`big ideas as well as the
`basic mechanical skills, the central topics as well as the
`peripheral background material-without
`independent
`initiative on their part.
`Any solution here must reconsider how and where in the
`cumculum to teach these fundamentals, and to what level
`of detail.
`
`B. Student Perspective on EE, CE, and
`Subdisciplines is Lacking
`By the time they are Seniors, faculty usually expect
`students to make intelligent choices when they have the op-
`portunity to choose an engineering elective course. Students
`are expected to ask their faculty advisers for guidance here,
`and to listen to whatever advice is offered. Our experience
`as educators suggests that it is already questionable whether
`this works for Seniors, who have a fairly extensive technical
`background. However, it is clear that students taking their
`very first course in a core ECE area like solid state
`devices or computer architecture are usually not clear
`about how this area connects to the rest of ECE as a
`whole.
`This was a particular problem in our original 1991
`curriculum. At this time, ECE offered two parallel cumcula:
`the BSEE and BSCE tracks, one of which students had to
`choose sometime during the Sophomore year. The problem
`was how to educate students to make an informed choice.
`Certainly, some students arrived absolutely decided on one
`track or the other. But many relied on our introductory
`courses to paint a sufficiently broad picture of the discipline
`for them to make a choice. Unfortunately, these introduc-
`tory courses concentrated almost entirely on packing in
`as much engineering material as possible. As faculty, we
`were often surprised when, after a few weeks in class, in
`the middle of some intricate technical discussion, a brave
`Sophomore would ask something like this:
`Exactly what does a computer engineer do? And how
`does this material help me to be a computer engineer?
`Is this different from computer science? Is the difference
`that we do hardware and they do software? When I
`graduate will I only be able to design big computers,
`or do computer engineers do something else as well?
`And why am I taking all these circuits classes-isn’t
`that for the electrical engineers?
`The emphasis on maximizing technical content in those few
`hours per week left little time to address all these questions
`satisfactorily. And as the breadth of the discipline continues
`
`1250
`
`PROCEEDINGS OF THE IEEE, VOL. 83, NO. 9, SEPTEMBER 1995
`
`~
`
`-~
`
`~
`
`HP Exhibit 1007 - Page 5
`
`

`

`to expand, we must confront this problem directly if our
`students are to make informed curriculum choices.
`
`C. Appreciation of Underlying Ideas is Weak
`We are not alone in observing that students often acquire
`only the mechanical aspects of the topics that we teach,
`without understanding the underlying ideas. The problem is
`pervasive in the teaching of technical material. For example,
`a National Science Foundation article on the teaching of
`college calculus relates this story [6]:
`A mechanical engineering professor mentioned in pass-
`ing to a class of sophomores that an integral is a sum. He
`simply assumed that the students had learned this basic
`idea from first-year calculus. But the students stared
`uncomprehendingly back at the professor. “Students
`seem to have a facility for doing things,” [the professor]
`concludes, “but they lack a sense of ideas.”
`Similar stories were easy to come by in our own de-
`partment. For example, in [7], one of our own faculty
`observed:
`[In several ECE courses] I’ve worked hard to help
`students achieve a rich and insightful understanding of
`fundamental material. Most of them seem to think I do
`a good job; they say on their FCE’s [Faculty Course
`Evaluations, a survey of each student’s evaluation of and
`reactions to the course, conducted by Carnegie Mellon
`itself] that I make even difficult and abstract concepts
`seem clear.
`Yet, when I look at the reality of their understand-
`ing, as gauged through exams and discussions in and
`out of class, it’s grossly disappointing. The majority
`simply don’t get it. Their survival skills allow many
`to get through with C’s and D’s, based mostly upon
`regurgitation of techniques I’ve shown them repetitively,
`as both they and I have forced ourselves through a
`distasteful process of pounding in material which they
`find mysterious and useless and which I find beautiful
`and important.
`Students’ varying technical preparation, the increasing
`diversity of their backgrounds, the divergence of student
`and faculty expectations and the widespread practice of
`packing ever more material into the same number of
`classes all compound the problem. We argue that curriculum
`designers must now address this problem directly. The mere
`mechanical skills that a student acquires in “survival” mode
`have a disturbingly short half-life in our rapidly moving
`discipline. The question is how to motivate students to
`appreciate the connectedness among abstract ideas, concrete
`applications, their classes, and their careers.
`
`D. Breadth in All Relevant Topical Areas is Impossible
`A foundation of many ‘‘classical’’ engineering curricula
`is the notion that every engineer must know something
`about every area of the discipline. There was certainly
`an era in which this was a reasonable assumption for
`electrical engineers. We argue that this is no longer a vi-
`able assumption-especially for an ECE department whose
`
`faculty engage in a broad program of research ranging
`from basic physics to advanced computer science. Dis-
`tancing a curriculum from this notion is difficult, since
`it tramples on nearly every faculty member’s most cher-
`ished subjects. Any attempt to reach consensus on a min-
`imum set of advanced topics to mandate in a curricu-
`lum rapidly yields a huge and unwieldy set of essential
`classes.
`One approach that we had already tried in 1991 was
`to partition the undergraduates into separate degree tracks
`leading to different BS degrees. As Carnegie Mellon’s EE
`department evolved into an ECE department, its degree
`offerings evolved into parallel BSEE and BSCE tracks.
`Computer engineering faculty argued that many required
`electrical engineering classes were inessential to the edu-
`cation of a computer engineer, and should be replaced by
`more relevant course requirements. Electrical engineering
`faculty countered that if students did not take the full
`complement of required electrical engineering classes, they
`should not be graduated as “electrical engineers.” So, a
`separate computer engineering degree was an agreeable
`solution.
`In hindsight, this debate nicely crystallizes a key problem
`for curricula as they try to evolve: what is essential to earn a
`degree with the words “electrical engineer” (or, in our case
`“computer engineer”) in the title? Slicing off portions of the
`curriculum to award them separate degrees whenever they
`attain some sort of “critical mass” is not a viable long-term
`strategy: New technologies and ideas become candidates for
`core topics in the curriculum faster than old topics expire.
`Any serious attempt at curriculum redesign must address
`the necessarily contentious issue of which topics are truly
`essential.
`
`E, Interdisciplinary Studies are DifJicult
`Many engineering curricula are based on a large number
`of required engineering classes and restricted technical
`electives. This was certainly true of our original ECE
`curriculum. The problem was how to deal with a student
`who wished to trade some technical depth for breadth.
`In 1991, one of our BSEE students could take some
`computer engineering courses, just as a BSCE student could
`take some electrical engineering courses. But we had no
`mechanism to give a degree to someone who chose to be
`broad, who chose to take, for example, 50% of

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket